Você está na página 1de 5

TECHNICAL NOTE

Revisions to BS8006 for


reinforced soil what do
these mean for the industry?
By Stephen P Corbet, Aecom, (Chairman BSi B526/4 Reinforced
and Strengthened Soils), Chris Jenner, Tensar International UK,
Graham Horgan, Huesker UK.
Presented at the Ground Engineering Slopes Conference on 16 November 2010.

ver the past 15 years, the design


of reinforced soil structures has
evolved from the principles set
out in the first edition of BS8006,
published in 1995. The introduction
of Eurocodes made a revision to
BS8006:1995, essential as EN19971 specifically excludes the design of
reinforced soil structures and slopes.
The BSi has completed the revision of BS8006 and Part 1, which
covers the design of reinforced fills
was published at the end of 2010,
while Part 2, which covers the
design of reinforced insitu soils, is
due to be published this year.

Introduction

Since the publication of BS8006 in


1995 after 10 years of drafting in
committee, the design of reinforced
soil structures has evolved as the
results from research have improved
the understanding of the behaviour
reinforcement in soil.
The publication of the Eurocodes
for design has resulted in the need
to withdraw or revise all existing
design codes in the UK and Europe.
The Eurocode EN1997-1 Geotechnical Design does not include any
provision for the design of reinforced soil structures and the UK
NAD makes a clear statement to
this effect.
In particular, the factors to be
applied to the strength of the reinforcing elements cannot be determined from the Eurocode.

Structure of the code

The standard will be published in


two parts: Part 1: Reinforced Fills
Retaining walls, slopes, basal reinforcement and transfer platforms;
and Part 2: Strengthened Insitu
Soils Soil nails.
The two parts recognise that the
design of insitu soils reinforced with
soil nails is completely different to
the reinforcement of soils placed
26

as fills. Part 2 will be based on the


CIRIA C637 Soil Nailing Best
Practice Guidance, but with more
emphasis on the design of reinforced soil slopes

The changes

The BSi committee of experts has


reviewed the whole document and
the following changes have been
made. The changes were made public in a draft for public comment,
published in July 2009, and the
comments received were incorporated into the final document.
n Clauses and information which
conflicted with BSEN1997-1, BSEN
14475 & BSEN 14490 have been
removed, but we have not revised
all terminology to agree with BSEN
1997-1.
n The correction of some outstanding errors found in text (at least two
found in the last published version).
n The partial factors were reviewed.
The review was to look at the possibility that the partial factors in
BS8006 could be the same or very
close to the partial factors used in
EN1997-1. The current situation is
that the partial factors in BS8006
will remain unchanged.
n A new section on reinforced modular block walls has been added to
cover this new technology.
n The design of facings for slopes
and walls has been reviewed.
n All text previously included by
the Highways Agency in BD70, has
been moved to BS8006-1 for bridge
abutments and bank-seats.
n The construction details have
been revised, with more information included for information.
n Advice on the use of vegetation on
steep slopes is included. The advice
includes information about the best
methods for getting grass and other
vegetation to grow and survive.
n The problems associated with lateral forces at the edges of embank-

ments with basal reinforcement


and with piles or other foundation
improvements have been addressed.
n In Annex A the derivation of partial factors for polymeric reinforcement has been updated to follow the
principles described in PD ISO CR
20432: 2007.
n A new Annex has been added
with seismic effects on reinforced
soil structures, based on experiences
form Japan and US, with cross reference to BSEN 1998.
n Other Annexes giving test procedures have been reviewed and
removed where now described in
testing standards.
n New features include: The use
of recycled materials is encouraged
using the Highways Agency document HD35/04 as a basis; cohesive
fills can benefit from inclusions
of drains within the soil layers to
reduce the drainage paths for excess
pore water pressure; new reinforcements can be considered, such as
electro-kinetic geosynthetics, used
to set up an electro osmosis system
for rapid drainage; low strain/high
strength geogrids.

Sections 6 and 7 reinforced


soil walls and steep slopes
concepts and fundamental
principles (Jenner)

To satisfy the requirements of a


limit state code, partial load factors
are applied to the actions and material factors are applied to the material properties.
Once the factors have been
applied, the requirement for stability is that the design restoring forces
and moments should be greater
than the design disturbing forces
and moments. In a particular application, the load factors applied to
dead loads and live loads can vary
depending on the load combination
under consideration; in some circumstances, the partial load factors

for live loads may be set at zero to


produce a worst-case combination
for the design load.
The loads, both dead and live, are
calculated in their unfactored form
as characteristic values, which then
have the appropriate partial factor
applied to give the design load.
The resisting forces in any application are generated from the available shear strength of the soil and
the tensile strength of the reinforcing elements.
The shear strength of the soil,
allowing for any pore pressure,
should be a characteristic value
determined as a cautious estimate
of the value recognising the limit
state under consideration. For walls
and slopes, the peak friction angle
is used. In all applications, the consideration of external, internal and
compound stability is required. For
internal and compound stability,
the resistance provided by the soil is
supplemented by the strength of the
reinforcement.
The reinforcement strength is
derived following a procedure
described in an Annex of the document.
Reinforcement is defined as being
extensible if the design strength is
sustained at a total axial strain value
>1%, and inextensible if the design
strength is sustained at a total axial
strain <1%. This definition can
determine the actual design method
adopted for a particular application.
The Code of Practice is applicable to all types of geosynthetic and
steel soil reinforcement.

Vertical walls and bridge


abutments

The principles described in the


previous section apply to both
reinforced soil walls and bridge
abutments, which are designed as
vertical structures.
Other structures with facing

ground engineering

April 2011

Effects

Reinforced slopes

Combinations
A

Dead load of the structure

ffs = 1.5

ffs = 1.0

ffs = 1.0

Dead load of the fill on top of the structure

ffs = 1.5

ffss = 1.0

fs = 1.0

Dead load of bridge and bank seat

ff = 1.2

ff = 1.0

ff = 1.0

Backfill pressure behind the bank seat

ffs = 1.5

ffs = 1.5

ffs = 1.0

Backfill pressure behind the structure

ffs = 1.5

ffs = 1.5

ffs = 1.0

Horizontal loads due to creep and shrinkage

ff = 1.2

ff = 1.2

ff = 1.0

Traffic loading

Over the
entire
structure,
fq = 1.5

Behind the reinforced


zone, fq = 1.5

HA

fq = 1.5

fq = 1.5

HA and HB

fq = 1.3

fq= 1.3

HA

fq = 1.25

fq = 1.25

HA and HB

fq = 1.1

fq = 1.1

fq = 1.3

fq = 1.3

Bridge vertical live load


Braking dynamic load
Temperature effects

Table 1: Partial load factors for walls

Soil material
factors

To be applied tan f fms = 1.0

fms = 1.0

fms = 1.6

fms = 1.0

To be applied to c
To be applied to cu

Soil/
reinforcement
interaction
factors

fms = 1.0

fms = 1.0

Sliding across
surface of
reinforcement

fs = 1.3

fs = 1.0

Pull-out resistance
of reinforcement

fp = 1.3

fp = 1.0

Partial factors Foundation bearing fms = 1.35


of safety
capacity: to be
applied to qult
Sliding along base
of structure or any
horizontal surface
where there is
soil-to-soil contact

fs = 1.2

n/a
n/a

Table 2: Partial load factors for abutments


angles within 20 of the vertical can
also be designed as vertical structures to the Code. Table 17 (Table 1
above) of the Code gives the partial
load factors for load combinations
associated with walls and Table 18
(Table 2 above) gives the same information relating to bridge abutments.
Soil material factors can be
extracted from Table 16 (Table 3
overleaf) with a number of partial
factors of safety relating to soil/
reinforcement interaction, foundation bearing capacity and sliding
along the base of the structure.
The most adverse loads likely to
occur should be considered.

ground engineering

April 2011

Design methods for walls

The wall design approach is


based on the classification of the
reinforcement materials, whether
the reinforcement is extensible or
inextensible.
Extensible
reinforcement
is
designed using the Tie-Back Wedge
method where the stress condition
within the reinforced soil block is
taken to be an active stress state and
wall friction on the back of the block
is zero (see Figure 1).
When inextensible reinforcement
is used, the Coherent Gravity design
method is used with a stress condition within the reinforced soil block

that varies between the at-rest


condition in the upper zone reducing to active conditions deeper in
the structure with wall friction calculated (see Figure 2).

Ultimate limit state

The appropriate partial factors are


applied to the loads and materials
depending on the Load Case under
consideration. Load Case A, with
maximum factors applied throughout, is normally the critical condition for reinforcement capacity and
foundation bearing pressure and
sometimes governs reinforcement
anchorage.
Load case B, with minimum mass
of the structure but maximum overturning forces, is normally critical
for sliding along the base of the reinforced soil block and reinforcement
anchorage.

Serviceability limit state

Load Case C, with dead loads only


and partial factors of unity, is used
to determine the serviceability limit
states of foundation settlement and
reinforcement tension for internal
strains. The limit value of reinforcement tension in the serviceability check is defined by a theoretical
post-construction strain value.
For a wall, the limit is 1% and
therefore the SLS allowable loadcarrying capacity is the value at
which the reinforcement would
strain 1% between the end of construction and the end of the design
life (Figure 3, overleaf). For an abutment, the limiting post-construction
strain is 0.5%.

The design methods for reinforced


slopes, generally defined as those
slopes with face angles greater than
20 to the vertical, are all derived
from limit equilibrium methods,
originally derived for unreinforced
slopes. The Code does not limit the
methods that can be used for design,
provided that they can be adapted
to a limit state format. However,
it does describe the most common
methods in some detail.
The partial factors are defined
in Table 4 (overleaf). with only
two load cases, ULS and SLS. The
friction angle of the fill should be
defined as the peak value and, as
with the wall design, there are partial factors of safety applied to soil/
reinforcement interaction and sliding along the base.
The requirement for the inclusion
of the Moment Correction Factor c
has been removed in the revised version of the BS. Any potential failure
surface that passes around the reinforced soil zone without cutting a
layer of reinforcement is now analysed using the approach defined in
Eurocode 7.
For reinforced slopes where the
face angle is 45 or less the need for
a formal facing may not be required.
In those situations the reinforcement capacity close to the front face
may be limited because of bond,
although higher shear strengths may
be mobilised because of the low
confining stress. A check of superficial stability can be carried out to
assess the stability of a free face in
resistance to sliding on a plane that
is very close, and parallel, to the
slope surface.
It is noted in the code that compound
stability
considerations
where a slip surface passes partially
through the reinforced fill and partially through the unreinforced fill
may well be the critical situation.

Section 8 basal
reinforcement and load
transfer platforms (Horgan)

UK BS8006-1 2009 Section 8 basal


reinforcement and basal reinforcement over piles (aka load transfer
platforms), updates the previous
guidance and attempts to expand/
clarify guidance considered ambiguous in the previous draft and to
reflect changes in best practice since
BS8006 was first published in 1996.
The main additional issues
addressed in the updates relate to:
n Consideration of pile cap
geometry;
n Consideration of alternative
arching mechanisms;
n More emphasis on assessing/controlling the settlements
between piles;
n Greater emphasis on collabo27

TECHNICAL NOTE
ration between geosynthetics
and pile designers;
n Time dependant consolidation of
soil between piles and consideration
of partial support between piles;
n Additional guidance on anchorage details etc.

Selection of soil parameters


and partial factors

The British standard differentiates


between soils/fills used in slopes
and walls and those used in embankments. Since the strains induced in

multiple layer reinforcement such as


walls and slopes are deemed small,
the frictional strength is represented
by tanp, the peak effective angle of
internal shearing resistance. Larger
strains are allowed in other scenarios, for example an embankment
subject to differential settlement.
In these cases, the frictional
strength is represented by large
strain values. For cohesionless soils
this is the value when the soil shears
at constant volume.
Consideration of foundation

Figure 2: Coherent gravity stress state


0

Ko

Ka

Kj

J
H
Arc tan 0.3

6m

Ws2

Ws1

z
Soil 1:
c1 1 1 1

Soil 2:
c2 2 2 2

hj

Tpj
H

b = 0 for tie back


wedge method.
b = (1.2 L/H) f2
for coherent gravity
method

L-2e

Soil material
factors

Reinforcement
material factor

Soil
reinforcement
interaction
factors
Partial factors
of safety

Partial Factors

Ultimate limit state

Serviceability limit
state

Soil unit mass eg slope fill

ffs = 1.5

fs = 1.0

External deal loads eg line


or point loads

ff= 1.2

ff = 1.0

External live loads eg traffic


loading

fq = 1.3

fq= 1.0

To be applied tan fr

fms = 1.0

fms = 1.0

To be applied to c

fms = 1.6

fms = 1.0

To be applied to the
reinforcement base strength

The value of fm should be consistent


with the type of reinforcement to be
used and the design life over which the
reinforcement is required (see 5.3 and
Annex A)

Sliding across surface or


reinforcement

fs = 1.3

fs = 1.0

Pull out resistance of


reinforcement

fp = 1.3

fp = 1.0

Sliding along base of


structure where there is
soil-to-soil contact

fs= 1.2

NA

Table 3: Partial factors for reinforced slopes


28

Isochrone for end of construction


Isochrone for end of design life
Prescribed post-construction strain limit

Figure 3: Assessment of post-construction strain

Figure 1: Stressed imposed by the soils and surcharges

Load factors

Tcs

soils: BS8006-1 assumes that all of


the embankment loading will be
transferred through the piles down
to a firm stratum either by direct
arching on the pile caps or by the
load carried to the piles by the geosynthetic reinforcement.
Consequently, the characteristics
of the soft foundation soil is considered only with regard to the type
of piles used and their installation.
However, the time dependant consolidation of soil between piles is
subsequently discussed.
BS 8006:1996 was one of the first
national codes to adopt limit state
principles and these are maintained
in the updated code.

Consideration of pile
cap geometry

The original guidance did not differentiate between circular or square


pile caps of equal dimension yet the
area ratio of the square pile cap (a2)
is different to that of a circular cap
(D2/4). Yet this dimension is used
subsequently used to determine the
amount of arching (Marstons ratio)
or stress redistribution that occurs
within the piled embankment. The
new guidance recommends reducing the design dimensions of cir-

ground engineering

April 2011

Surcharge Ws

Embankment
Outward
shear stress
Lo

Reinforcement

Fill: ,cv

Pfill

Lb

Optional geotextile separator


(dependant on embankment fill material)

Minimum cover
500mm to
reinforcement

Anchor bond length


(varies)

BRP layer
depth can vary

Lds

Lp

Pile
caps

Soft foundation

Edge trench
option detail

Piles

Figure 4: anchorage length (Lb) at the edge of the piled area

Partial factor

Ultimate limit
state

Piles with pile caps


(pre-cast driven, vibro concrete columns
driven/cast insitu, continuous flight auger)

Serviceability
limit state

Figure 5: Reinforcement anchorage at edge of fill

Load factors
Soil unit mass,
eg embankment fill

ffs = 1.3

ffs = 1.0

Pile arrangement

Arching coefficient

External dead loads,


e.g. line or point loads

ff = 1.2

ff = 1.0

End-bearing piles
(unyielding)

Cc = 1.95H/ a -0.18

External live loads, e.g.


traffic loading

fq = 1.3

fq = 1.0

Friction & other piles

Cc = 1.5H/ a -0.07

Soil material factors

mined by using Marstons formula:

To be applied to tan

fms = 1.0

fms = 1.0

To be applied to c

fms = 1.6

fms = 1.0

To be applied to cu

fms = 1.0

fms = 1.0

Soil/reinforcement interaction factors


Sliding across surface
of reinforcement

fs = 1.3

fs = 1.0

Pull-out resistance of
reinforcement

fp = 1.3

fp = 1.0

Table 4: BS8006 Section 8 Partial factors


cular pile caps to consider a square
pile of an equivalent area. Where
circular pile caps are to be used,
the diameter of the pile should be
reduced to produce an effective pile
cap width aequ,
aequ = (D2/4)0.5 hence aequ = 0.886D
where:
a is the size of the pile caps
(assuming full support can be generated at the edges of the caps);
D is the diameter of the pile cap

Determination of load acting


across the reinforcement

Due to the significant differences


in deformation characteristics that
exist between the piles and the surrounding soft foundation soil, the
vertical stress distribution across the
base of the embankment is assumed
to be non-uniform. Soil arching
between adjacent pile caps induces
greater vertical stresses on the pile
caps than on the surrounding adja-

ground engineering

April 2011

Table 5: Arching coefficient Cc

cent soil.
BS8006-1 identifies a minimum
height whereby partial arching
begins to develop but additional
loads placed at the surface of the
embankment still influence the load
carried by the reinforcement. This
minimum height is :
H 0.7(s a)
where:
a is the size of the pile caps (or aequ
considering circular pile caps)
s is the spacing between adjacent
piles
BS8006-1 also identifies a critical height concept above which any
additional embankment weight or
surcharge loading placed at the surface of the embankment is deemed
to pass directly to the pile caps and
not influence the reinforcement.
The ratio of the vertical stress
exerted on top of the pile caps to
the average vertical stress at the base
of the embankment may be deter-

[ ]

Pc = Cca
sv
H

where:
Pc is the vertical stress on the
pile caps;
sv = (ffsH + fqws) and is the factored average vertical stress at the
base of the embankment;
is unit weight of the embankment fill;
H is the height of the embankment;
ws is the uniformly distributed
surcharge loading;
a is the size (or diameter) of the
pile caps;
Cc is the arching coefficient.
Where the value of the arching coefficient will vary depending on the
type of piles.
Several authors have studied and
compared the phenomena of soil
arching using a variety of physical,
analytical and numerical models to
try to gain a better understanding
and quantify the load acting across
the reinforcement and distributing
directly to the pile caps (Alexiew
2002, Eekelen 2008, Kempton et al
1998, Love and Milligan 2003, Rogbeck 1998, Stewart and Filz, 2005).
One criticism of the original code
is that the distributed load acting
across the reinforcement was very
sensitive to embankment heights
around the critical height
H 1.4(s a)
Embankment heights just under

this limit would result in considerably more design load across


the reinforcement than embankment heights just above this level.
Additional criticism related to the
fact the ratio of the vertical stress
exerted on top of the pile caps to
the average vertical stress at the base
of the embankment was dependant
solely on pile type and independent
of the type of fill used within the
embankment.
Finally a criticism that vertical
equilibrium is not satisfied (Eekelen,
2008). However, neither the load acting across the reinforcement nor the
increased load acting on the pile caps
are calculated directly, but rather
indirectly from the ratio of increase
in stress concentration on the pile
caps to the average vertical stress at
the base of the embankment.
BS8006-1 2009 Section 8 allows
for an alternative theoretical solution which may be used to determine the vertical load acting across
the reinforcement based on work
presented by Hewlett and Randolph, 1998, which was based on
the observed failure mechanism
from model tests and considers a
series of hemispherical domes. The
theory determines the efficacy E as
the proportion of the embankment
weight carried by the piles, hence
the proportion of the embankment
weight carried by the geosynthetic
reinforcement may be determined
(1 E).
The original guidance utilises
Marstons formula to determine the
distributed load acting across the
reinforcement
29

TECHNICAL NOTE
WT =

1.4 sffsg (s a) 2 2
s a (p/sv)
s2 a2

where:
WT is the distributed vertical load
acting on the reinforcement between
adjacent pile caps;
ffs is the partial load factor for soil
unit weight (see Table 1);
fq is the partial load factor for
external applied loads (see Table 1).
To satisfy vertical equilibrium
then the remaining load must be
carried directly by the pile caps, WP
WP= (f fsgH + fqws ) s2 WT (s2 a2)
(although this formulation is not
provided in BS8006 -1 2009). Once
the distributed load WT acting across
the reinforcement is determined,
then for an extensible reinforcement the tensile load Trp per metre
run, generated in the reinforcement
resulting from the distributed load
WT is determined by
1
Trp = WT (s a) 1 + 6

2a
where:
Trp is the tensile load in the reinforcement;
is the strain in the reinforcement.
The above equation has two
unknowns Trp and , and may be
solved for Trp by assuming a maximum allowable strain in the reinforcement and by an understanding
of the load/strain characteristics of
the reinforcement at different load
levels. Initial tensile strain in the
reinforcement is needed to generate a tensile load; BS8006 stipulates
that a practical upper limit of 6%
strain should be imposed to ensure
all embankment loads are transferred to the piles.
In addition to the load on the
reinforcement generated by the
redistributed vertical embankment
load, the reinforcement needs to
resist the outward thrust from the
embankment. The reinforcement
tensile load Tds needed to resist the
outward thrust of the embankment
may be taken as
Tds = 0.5Ka (ffsH + 2fqws)H
where:
Tds is the tensile load in the reinforcement per metre run needed
to resist the lateral thrust of the
embankment fill;
Ka is the active earth pressure
coefficient [ = tan2(45 fcv/2)];
H is the height of the embankment;
is the unit weight of the embankment fill;
ws is the surcharge intensity on
top of the embankment;
30

ffs is the partial load factor for soil


unit weight (see Table 1);
fq is the partial load factor for
external applied loads (see Table 1).

Reinforcement details

The design approach essentially


determines the requirement for two
orthogonal layers of geosynthetic
reinforcement at the base of the
piled embankment, one longitudinal
layer parallel along the centre line of
the embankment and one transverse
layer perpendicular across the line
of the embankment.
The longitudinal reinforcement is
designed to resist the redistributed
vertical load Trp.
The transverse reinforcement is
designed to resist both redistributed
vertical load Trp and to resist the lateral thrust of the embankment Tds.
To mobilise these loads the reinforcement should achieve an adequate bond with the adjacent soil
at the extremities of the piled area
to ensure that the maximum limit
state tensile loads can be generated
between the outer two rows of piles.
Across the width of the embankment the reinforcement should
extend a minimum distance (Lb)
beyond the outer row of piles, (see
Figure 4).
The reinforcement may be
extended around the row of gabions
and returned into the embankment
fill to develop the necessary bond
length (see Figure 5).
Depending on the geometry of
the embankment, it may be difficult
to achieve an adequate bond length
at the extremity of the piles by
maintaining the reinforcement in a
horizontal alignment as depicted in
Figure 4. One solution that may be
considered is to use a row of gabions, as a thrust block along the top
of the outer row of piles.
Another detail that may be considered is the inclusion of a small
periphery trench just beyond the
edge piles, running parallel to the
centreline of the embankment; the
trench is typically only as deep as
the piling mat or pile cap depth. The
reinforcement can be extended into
the trench and when backfilled, will
return into the embankment fill to
develop the necessary bond.

Time dependant consolidation


of soil between piles

BS8006-1 assumes that all of the


embankment loading will be transferred through the piles down to a
firm stratum either by direct arching on the pile caps or by the load
carried to the piles by the geosynthetic reinforcement. In reality, the
soil between the piles will need to
undergo some initial degree of consolidation to enable tensile forces in
the geosynthetic to be mobilised.

This settlement can be related to


an increase in stress from the installation of the temporary working
platform or as a result of initial fill
placement or by external factors
such seasonal fluctuations in the
ground water level increasing the
effective stress on the insitu soil.
Hence the time dependant consolidation of soil between piles is
important in reaching the assumed
end design condition of no partial
support between piles.
Consideration can also be given
to the introduction of a compressible layer between the pile caps to
ensure deflection of the geosynthetic during the initial placement of the
overlying fill.
The maximum mid-span deflection y of extensible reinforcement,
spanning between pile caps, may be
determined from the formulation
below (after Giroud)
y = (s a)

3
8

where:
a is the size of the pile caps;
s is the spacing between adjacent
piles;
is the strain in the reinforcement.
The thickness of the compressible
layer can be such that it can accommodate the deflection of the geosynthetic during the early stages of
embankment construction, ensuring
that the assumed design deflections
are achieved.

Conclusions

BS8006 was first published in 1995.


Section 8 of the code has been used
to successfully design a vast number
of piled supported embankments
worldwide.
BS8006-1 2009 updates the previous guidance to reflect better industry understanding of the behaviour
of such complex earthworks and to
reflect changes in best practice.
Currently there is a variety of different design approaches, which differ mainly in the amount of arching
considered and the degree of support offered by the existing sub soil.
The BS8006-1 assumption that
all of the embankment loading will
be transferred through the piles or
by the geosynthetic reinforcement
onto the piles can be considered
conservative.
However, the potential use
of compressible layers beneath
the geosynthetic reinforcement
between pile caps can ensure that
the end design condition of no
partial support between piles can
be achieved during construction.
This would make consideration
of the time dependant consolidation of soil largely academic.

References

BS 8006:-1:2010 Code of Practice for


Strengthened/reinforced soils and fills,
BSi London 2010.
BS EN 1997-1:2005 Geotechnical
Design, BSi London 2005
CIRIA C637 Soil Nailing best practice guidance CIRIA 2005.
PD ISO CR 20432 Guidelines for the
derivation of the longterm strength of
geosynthetics for soil reinforcement
HD 35/04 Conservation and the Use
of Secondary and Recycled Materials
AASHTO (1998), Interim Standard
Specifications for Highway Bridges,
AASHTO, Washington DC, 16th
Ed
Alexiew, D., Piled embankment
design: Methods and case studies, Proc.
XV Italian Conference on Geosynthetics, Bologna,October 2002. In:
Lingegnere e larchitetto, Special Issue
1-12/2002. pp. 32-39. 2002.
DIN 1054:2005-01 Baugrund; Sicherheitsnachweise im Erdund Grundbau.
Eekelen, van SJM, Van, MA, Bezuijen, A. Design of piled embankments
considering the basic starting point of
the British Standard BS8006. Proceedings of the Fourth European Geosynthetics conference. Edinburgh,
FHWA(1996) Mechanically Stabilized
Earth Walls and Reinforced Soil Slopes
Design and Construction Guidelines
(FHWA-SA-96-071).
Giroud, JP, (1995). Technical Note:
Determination of geosynthetic strain
due to deflection, Geosynthetics International, Vol. 2, No.3, pp.635-641
Hewlett, WJ, Randolph, MA.
Analysis of Piled Embankments.
Ground Engineering. April pp.12-18.
1988
Kempton, G.T., Russell, D., Pierpoint, N.D., Jones, C.J.F.P. Two and
three-dimensional numerical analysis of
the performance of piled embankments.
Sixth international conference on
geosynthetics, Atlanta, Georgia.
pp.767-772. 1998.
Love J, Milligan G. Design methods for basally reinforced pile-supported
embankments over soft ground, Ground
Engineering, Vol. 36, No. 3. 2003.
Rogbeck, Y, Gustavsson, S,
Sdergren, I Lindquist, D. Embankment support over piles using geogrids.
Sixth international conference on
geosynthetics, Atlanta, Georgia.
pp.755-762. 1998.
Stewart, ME, Filz, GM. Influence
of Clay compressibility on Geosynthetic
Loads in Bridging Layers for ColumnSupported Embankments, GSP 131
Contemporary Issues in Foundation
Engineering, Geofrontiers 2005,
Austin

ground engineering

April 2011

Você também pode gostar