You are on page 1of 82

design guide

G E O T E X T I L E S

stronger by design
FM 23905

A GEOTEXTILES DESIGN GUIDE


First Edition Second Edition Revision Third Edition Revision Fourth Edition 1984 1989 1992 1994 1997 2003 Back-up theory is available for those wishing deeper examination. Whilst the intention of this guide is to be as helpful as possible and the information is offered in all good faith to enable a reasonable assessment of the practical performance of a geotextile to be made the authors, publishers and geotextile manufacturers cannot accept responsibility as the conditions of use are beyond their control. Purchasers and speciers of any geotextile must statisfy themselves that it is t for the intended use. This guide follows the logical sequence of design: the relevant chapter, Select Assemble required input data, Follow the the design procedure, Write the specication.

CONTENTS
NOTATION 1
1.1 1.2 1.3 1.4 1.5 3

UNPAVED ROADS
Design Theory Input Information Design Procedures Model Specication References

5 5 8 9 11 17 18 18 18 19 23 24 25 31 34 35 38 39 39

2
2.1 2.2 2.3 2.4

PAVED ROADS
Recommendations Model Specication Design Considerations References

3
3.1 3.2 3.3 3.4 3.5 3.6 3.7

DRAINAGE
Geotextiles as Filters Design Procedure Model Specication Drainage Applications Example Typical Lotrak Drainage Geotextiles References

4
4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8

EROSION CONTROL
Introduction Design Theory Design Procedure Model Specication Soil Erosion Example Typical Lotrak Erosion Control Geotextiles References

40 40 40 47 50 51 52 53 53

5
5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8

SOIL REINFORCEMENT
Introduction Embankments Walls and Slopes Design Example Model Specication Remedial Slope Works Typical Lotrak Reinforcing Geotextiles References

54 54 54 63 72 73 73 74 74 75

GEOTEXTILE DATA

NOTATION
B b C Cv CBR c cu Cuo D De d dk d1 d90, d60 E e F Gs H Ho h ho ho h K Ka K Kb kg ks L M N Nf n width drain width embankment crest width coefficient of consolidation California Bearing Ratio effective cohesion undrained shear strength undrained shear strength at ground surface depth of drain equivalent stone diameter depth to aquifer or length of drainage path channel diameter required draw down a size which 90%, 60% etc. of soil particles are smaller than equivalence factor for axle loads or soil erodibility factor precipitation rate factor of safety specic gravity of rock height of embankment, wall, slope or waves upper limit of pavement thickness reduction granular layer thickness or artesian pressure head granular layer thickness unreinforced pavement thickness reduction in pavement thickness using reinforcement or hydraulic head geotextile secant modulus or general coefficient of earth pressure coefficient of active earth pressure damage coefficient coefficient of permeability for granular drainage blanket geotextile permeability coefficient of permeability for soil annual soil loss by erosion or soil reinforcement length turning moment (kN/m x m) number of passes of a standard axle number of stress application number of axle passes or number of log-cycles of time or geotextile porosity or embankment slope dened as 1 vertical to n horizontal 090, 060 etc P Pstandard pt p a size which 90%, 60% etc. of geotextile pores are smaller than axle load or erosion protection factor standard axle load tyre pressure pressure or stress m kN kN kPa N/mm2 4 m m m m2/year % kPa kPa kPa m m m m m m m/sec m mm mm m mm mm mm mm kN/m m/sec m/sec m/sec tonnes/ha m kN

Q QL Qp q q0 R r ru S

drain capacity line load point load unit ow rate surcharge load rainfall erosity factor or radius rut depth pore water pressure ratio slope length and steepness factor or drain spacing or vertical reinforcement spacing

m3/sec kN/m kN m3/sec/m2 kPa m mm m m kN/m m years mm kPa m/s m/s kN tonnes kg m m kN/m3 tonnes/m3 % kPa m kPa years

T t tg U U u V v W w x z s n w
1

load or tensile force in geotextile thickness of drainage blanket or time thickness of geotextile coefficient of uniformity linear coefficient of uniformity pore water pressure velocity velocity weight of soil mass or weight of armour stone weight of armour stone variable length along base of embankment depth below top of ll or ground level angle of stress distribution or bond adhesion factor slope angle unit weight or partial factor specic weight of rock armour angle of friction between geotextile and ll tensile strain in geotextile permeability reduction factor for unwoven geotextiles permeability reduction factor for woven geotextiles angle between vertical and slip circle radius angle between basal reinforcement and tangent to slip circle rate of increase of undrained shear strength with depth mathematical constant settlement beneath centre of embankment soil stress effective peak angle of shearing resistance of soil effective angle of shearing resistance of soil at constant volume undrained angle of shearing resistance of soil factor related to rut depth and wheel geometry time from end of embankment construction when basal reinforcement load becomes zero

cv
1

u
1

1 UNPAVED ROADS
Unpaved roads are dened as roads in which material is unbound. Pavements with concrete or bituminous surfaces are excluded unless at an intermediate stage of construction the unbound granular sub-base is to be trafficked. In this case the pavement to the top of the sub-base can be considered unpaved. Special conditions apply to the trafficking of sub-base for Department of Transport projects and these are considered separately. Unpaved roads may be purpose built to give temporary or permanent access depending upon conditions. Haul roads for temporary site use or longer term use, as in mineral extraction or refuse disposal, are usually unpaved.

1.1 DESIGN THEORY


Unreinforced Roads
Where the terrain to be trafficked comprises granular soils, the formation may be strong enough to carry substantial traffic after minimal treatment such as grading to alignment. However, this situation is exceptional and usually the bearing capacity of the virgin soil is inadequate to support continued, imposed wheel loadings. To minimise the stresses induced at the surface of the formation it is common practice to lay down a carpet of competent material such as a granular ll to spread the wheel loading and so reduce stresses at the formation level to an acceptable value. If consideration is given to an axle load P kN on a granular carpet ho metres thick, then the vertical stress at the surface of the formation is po, Figure 1.1. Assuming the strength of the formation to be represented by its undrained shear strength Cu, then the maximum vertical stress which can be applied is (2 + ) Cu; the classical value of the undrained bearing capacity of the formation.

Figure 1.1

However, the stresses applied to the surface of the formation are not purely vertical since they include a horizontal shear stress that acts radically outwards from each wheel load. The effect of this shear stress is to reduce the maximum value of vertical stress po that can be transmitted to the surface of the surface of the formation without causing substantial deformation. Classical theory indicates the onset of large deformation once po > 12 (2 + ) Cu, however Giroud and Noiray (1981) have suggested a slightly higher limiting value of po = Cu. The magnitude of po for the reinforced road, can be related to the axle load and the contact area of the wheels B x L. po = P 2(B+2hotan o) (L+2hotan o) + ho

The dimensions of the contact area can be related directly to tyre pressure pt. For on highway vehicles:- L = B/2, B = P pt For off highway vehicles:- L = B/2, B = 2P pt For a wide range of applications tan o 0.6 With knowledge of the axle load P , tyre pressure pt. and undrained shear strength of the formation Cu it is possible to determine ho the theoretical thickness of granular ll required to prevent overstressing of the formation. 6

Reinforced Roads
Benecial affects acrue when a geotextile is introduced at the interface of the formation and the ll. Three broad mechanisms can be identied. i) The geotextile acts as a separator and so maintains the original thickness of ll. Without a geotextile the effective ll thickness is reduced by intermixing and punching of the ll into the formation soil under the action of the wheel load. ii) Provided there is good bond between the ll and upper surface of the geotextile and provided the geotextile has adequate axial tensile stiffness, it will absorb part or all of the outwards horizontal shear stress which would otherwise be transmitted from the ll to the formation. By absorbing this shear stress the geotextile increases the vertical load which can be applied to the formation before substantial vertical deformation occurs. If the outwards horizontal shear stress is completely absorbed the geotextile restore the bearing capacity to its classical value of (2 + ) Cu. iii) If the bearing capacity of the formation is exceeded, ruts will develop on the surface of the ll and these will be reected in the formation. This deformation will cause a tensile force T, Figure 1.2, to develop in the geotextile. This force will have two effects. Firstly it will apply an inwards horizontal shear stress to the surface of the formation which will increase its bearing capacity to a value slightly above (2 + ) Cu. Secondly, the axial force T will have an upwards component dened as p1, Figure 1.2. This component is due to so-called Membrane action.

Figure 1.2

At present no comprehensive analytical methods have been developed, however, the design method assumes that the vertical pressure applied at the surface of the geotextile is p2 and that is sufficient to cause deformation of the geotextile, which invokes membrane action. This membrane action is associated with an upward pressure p1 that reduces the pressure assumed to be transmitted to the surface of the formation. If the pressure exerted on the formation is limited to its classical bearing capacity (2 + ) Cu, then a limiting value of (p2 - p1) which is dened such that: p2 p1 = (2 + ) Cu The two pressures p2 and p1 can be evaluated in terms of known parameters:p1 = K Where K is the geotextile modulus. is the geotextile strain y is a function of rut depth and axle geometry The pressure p2 can be determined in the same manner as for an unreinforced road:p2 = Ph 2(B + 2h tan ) (L + 2h tan ) + ho

Solution of these equations leads to a value of h with geotextile reinforcement, and a subsequent saving in the required thickness of aggregate. h = ho h. 7

Dynamic Design
Both the foregoing analyses are static. No account is taken of the fact that the traffic load is dynamic and would therefore impose more severe loading. Using an empirical approach it is possible to estimate the required thickness of aggregate, ho, for an unreinforced road:ho = 119.24logN + 470.98logP 279.01r 2283.34 (0.77 Cu)0.63

Where ho is expressed in metres N is the number of axle passes P is the axle load in Newtons r is the rut depth in metres Cu is the undrained shear strength of the formation in kPa No similar design approach exists for a geotextile-reinforced road, however, it may be assumed that the saving in aggregate thickness h would be the same under either static or dynamic conditions. Aggregate thickness unreinforced:- ho m Aggregate thickness reinforced:- (ho h) m

Design Charts
The design of geotextile reinforced roads can be simplied using the design charts, at the end of this section. Each design chart is for a xed value of rut depth and for a single standard axle load of 80kN. On the horizontal axis is the undrained shear strength, Cu, of the formation soil. If only CBR values are available then these can be converted into an equivalent undrained strength using the following approximation where CBR is in percentage. Cu = 23 x CBRkPa There is only one scale on the vertical axis. However, this can represent ho which is the required thickness of granular ll under dynamic traffic load without geotextile reinforcement or the same scale can be used to determine h which is allowed with a geotextile reinforcement. This leads to a total thickness of granular ll. h = (ho - h) when a geotextile is employed. The magnitude of h for a given value cu will vary in accordance to the value of the geotextile modulus K with h increasing as K increases. There is an upper limit to the value of h and this is dened by the broken line Ho on the design curve.

Highways Agency Design Guide


The above method should not be applied directly to Department of Transport contracts where the sub-base is to be trafficked by construction traffic. Design note HA 25/94 points out that pavement foundations are intended to be sufficiently robust during the construction period to provide a working platform for the construction of succeeding layers based on an average loading of 1000 standard axles. TRRL Lab Report 1132 presents the following equation which gives a relationship between the thickness of subbase, h, required to limit the rutting to 40mm, the subgrade CBR during construction, as given in Figure 1.4, and the number of passes of standard axle:h(mm) = 190 (logN + 0.24) / (CBR (%) )0.63
Figure 1.3 Thickness of sub-base required to carry construction traffic

This relationship is shown graphically in Figure 1.3. The present specication places a practical upper limit of 225mm on the thickness of a single sub-base layer. The analysis presented in LR.1132 demonstrates that this thickness of well graded granular material should be satisfactory providing the length of road base and surfacing under construction with materials carried over the sub-base is less than about 1 km in length and the soil CBR is not less than 5%. Below this level of CBR a capping layer is normally specied. Otherwise an additional thickness of sub-base is needed and it would have to be compared in two layers. Capping layer thicknesses are related to equilibrium CBR values in Departmental Standard HD25/94. It is recognised in Design Note HD25/94 that geotextiles have been shown to be useful for temporary haul roads when used with granular materials over soft ground and that the need for capping may be reduced or removed by the use of subgrade stabilisation. However, no guidance is given as to how this might be achieved using geotextiles. Design Note HD25/94 also points out that some degree of contamination into granular capping can be expected with weak cohesive subgrade soils and that the capping design thicknesses allow for this. Again no guidance is given in the use of geotextiles, however, there may be scope for using a reduced thickness of capping if a geotextile separator is installed at the capping/subgrade interface.

1.2 INPUT INFORMATION


The following data is to be supplied by the designer to establish the necessary design parameters

Sub-grade Strength
The sub grade strength is normally expressed in terms of its CBR value. For cohesive sub grades, strength can be expressed in terms of undrained shear strengths, the strengths used in design should be those prevailing at the time of construction. Where no directly measured data are available, the following approximations may be considered. For plastic soils TRRL Lab Report 1132 gives the relationship between Plasticity Index and CBR shown in Figure 1.4 As can be seen for a given Plasticity Index the design value of CBR will vary according to sub grade drainage conditions and the extent to which the sub grade has wetted. An estimate of the remoulded undrained shear strength, CU, of a clay formation can be made from its remoulding CBR value using the empiricism given TRRL Lab Report 889:cu of 23kPa = CBR of 1%, For non-plastic soils, such as sands and gravels the CBR is generally greater than 20%

CBR%

PI Figure 1.4

Traffic
The number of passes of a standard 80kN axle. Determine each axle load and the number of the passes then reduced to equivalent standard axles using the equivalent factor:-

E =

{ {
P Ps

3.95

or the graph, Figure 1.5

Multiply actual number of axle passes by E to obtain equivalent number of passes of a standard axle.

Figure 1.5

The following table services as a guide to axle loads. The gures in kN are to be considered approximate as actual loads vary according to the make of vehicle. Example 1 What is the number of passes, N, of an equivalent standard axle for a road to be built for 800 deliveries of concrete by full 6 cu m truck mixers.
-axle load kN 65 92 24 44 No of passes (n) E (from graph) 800 1600 800 1600 0.6 1.8 0.01 0.1 Exn 480 2880 8 160 Total 3528esa

front 1 8 ton truck 24 ton truck 6 cu m concrete truck mixer Dump truck (75 tonne) full empty full empty full empty full empty 35 20 65 35 65 24 625 360 front 1 Motor scraper (24m3)
Table 1.1

rear 1 45 20 95 20 92 44 710 180 front 1 280 150

rear 2 95 20 92 44 rear 2 240 100

full empty

240 150

Rut Depth
Assess the maximum permissible rut depth for the type of traffic to use the road. 75mm 150mm 225mm 300mm or 450mm

Geotextile Properties
If a particular geotextile is to be used then its secant modulus K kN/m is required. Alternatively calculate the maximum possible saving in construction thickness which will in turn determine the necessary geotextile modulus. In general the stiffer the geotextile, that is the greater its modulus, the larger the saving in aggregate thickness. The modulus of the geotextile can be expressed as a secant modulus, thus at a strain x the load is Tx which gives a modulus K = Tx/x kN/m Obviously if the load vs. strain characteristics of the geotextile are non-linear the modulus of the geotextile will vary with strain being high at small strains and lower at high strains. Under no circumstances should the failure load or failure strain of the geotextile be exceeded. Note - x is the actual engineering strain and not the percentage strain For an initial calculation it is advisable to use the 10% secant modulus.
Figure 1.6

1.3 DESIGN PROCEDURES


Select the design chart with the appropriate rut depth. Enter the graph at cu, move vertically to intercept the curve for the number of passes N. Interpolation may be necessary. Move from this point horizontally to the unreinforced construction thickness required. (ho) Repeat the procedure intercepting the appropriate curve for the geotextile modulus selected and read off the reduction in thickness h. Actual construction thickness is ho - 4h Check that the strain induced in the geotextile is not greater than its failure strain. 10

For rut depths 75mm and 150mm the maximum strains are 1.2% and 4.2% respectively. For greater rut depths interpolate between the lines of iso strain for the shear strength cu and the modulus K. Refer to the chart (in the data section) plotting K against strain for the particular grade of geotextile chosen to establish actual K at given strain and factor of safety against tensile failure. Alternatively to maximise the possible savings in construction thickness, select the graph with the appropriate rut depth, enter at cu and move vertically until intercepting the broken line. Ho read off h from the vertical axis and interpolate the required geotextile modulus ho is determined as above. The construction thickness = ho h. For practical construction considerations the construction thickness should not be less than 150mm, it should also extend well beyond the line of the outside rut. For Highways Agency and similar public sector works the sub-base of the permanent road may be used during construction as a temporary haul road. In this case special requirements apply. These are set out in the section dealing with theory. Example 2 Determine the thickness of granular ll required on a clay formation with cu = 40kPa if the traffic is to comprise 1000 passes of an axle load of 80kN and the rut depth is limited to 300mm. Under the same conditions what would be the construction thickness if the road is reinforced with a geotextile having K=200kN/m? First select the design chart with the required rut depth of 300mm. The chart is entered at cu = 40kPa and a line is drawn vertically upwards to a point where it intersects the curve for N=1000. A horizontal line from this point intersects the vertical axis to give a value of ho = 450mm. Thus with no geotextile a construction thickness of 450mm is required. To determine the value of h the chart is again entered at cu = 40kPa. A vertical line is extended to the point where it intersects the curve for K = 200kN/m. Moving horizontally from this point gives h = 170mm. Thus with the geotextile there is a reduction in construction thickness of 170mm leading to a reduced construction thickness of 450 170 = 280mm.

Figure 1.7

At this stage it is necessary to check that there are not excessive strains or loads induced in the geotextile. Some indication of geotextile strain is given by the broken lines. These give lines of iso strain for 7% and 10% strain with the Ho line indicating a strain of approximately 13%. In this example the strain is approximately 12.5%. Referring to the typical load strain curve (see annex) for the specied geotextile with K=200kN/m it can be seen that the induced strain of 12.5% is less than the failure strain of 20%. This geotextile is therefore acceptable. A more detailed assessment can be made by plotting K against for the geotextile. This shows that at 12.5% strain the modulus is 325kN/m with a factor of safety of approximately 1.3 against tensile failure. In this case it would be reasonable to assume a higher modulus than 200kN/m however for cu = 40kPa the cut-off, shown by the Ho line, shows no advantage in having K greater than approximately 250 kPa at a strain of 13%. The illustrated geotextile would cope with this and allow h = 190mm. Alternatively a more extensible geotextile could be used provided the required modulus is available at, or below, the strain dictated by the design curve.

11

Example 3 A haul road is to be constructed over a formation with cu = 30kPa with an allowable rut depth of 300mm. Traffic comprises 1000 passes of a truck with a rear tandem axle load of 180kN and a single front axle load of 85kN. What geotextile modulus is required to give a maximum saving in construction thickness? Give construction thickness with and without geotextile. a) With the rst vehicle there are the following axle loads:1000 at (180/2) 1000 at (180/2) 1000 at 75 These can be converted into a number of passes of a standard 80kN axle:1000 at 90 = 1000 x (90/80)3.95 at 80 = 1592 at 80 1000 at 90 = 1000 x (90/80)3.95 at 80 = 1592 at 80 1000 at 75 = 1000 x (75/80)3.95 at 80 = 775 at 80 Equivalent loading = 3959 passes of standard 80 kN axle. b) For the second vehicle:200 at 160 = 200 x (160/80)3.95 at 80 = 3091 at 80 200 at 85 = 200 x (85/80)3.95 at 80 = 254 at 80 Equivalent loading = 3345 passes of a standard 80 kN axle. For design the total equivalent loading is 7304 passes of a standard 80kN axle. To determine the various construction thicknesses use the design chart for the required rut depth of 300mm. This is entered at cu = 30kPa and ho is read off as 700mm for 7300 passes. This gure is approximate since it involves interpolation between N = 1000 and N = 10,000. The maximum saving in construction thickness is dened by the broken line ho. For cu = 30 kPa the maximum h is 2700mm. This is achieved using a geotextile with K = 320kN/m in which the strain does not exceed 13%. The construction thickness with this geotextile is (700 270) = 430mm. Chose suitable geotextile from K against strain data.

1.4 MODEL SPECIFICATION


Clause XY3: Geotextile for unpaved roads between Point A and Point B. The geotextile shall be manufactured from durable synthetic polymers and have the following properties. The load vs. strain characteristics shall be determined in accordance with EN 10319. The failure load must not be less than.....kN/m. The failure strain must not be less than.....%. At a strain of.....% the modulus most not be less than.....kN/m. The minimum water permeability normal to the plane shall be.....m/sec across the geotextile determined in EN ISO 11058. For contracts subject to the Design Manual for Roads and Bridges, Specication for Highway Works geotextiles used separate earthworks shall comply with Clause 609 and the contract specic Appendices. Please refer to pages 76 78 to determine appropriate Lotrakgrade required. 12

13

14

15

16

17

1.5 REFERENCES
Black, W.P.M., & Lister, N.W. 1979 The strength of clay ll subgrades: its prediction in relation to road performance. TRRL Laboratory Report 889. Transport and Road Research Laboratory, Crowthorne, Berkshire. Giroud, J.P., & Noiray, L. 1981 Design of Geotextile Reinforced Unpaved Roads. Proc. ASCE Geot.Engrg.Div. Vol. 107, No. GT9. Powell, W.D., Potter, J.F., Mayhew, H.C., & Nunn, M.E 1984 The Structural Design of Bituminous Roads. TRRL Laboratory Report 1132. Transport and Road Research Laboratory, Crowthorne, Berkshire. Design Manual for Roads and Bridges Vol 7 Section 2 Part 2 HD 25/94 Foundations - HMSO London. Design Manual for Roads and Bridges Vol 1 Specication for Highway Works - HMSO London. Design Manual for Roads and Bridges Vol 2 Notes for Guidance on the Specication for Highway works - HMSO London.

18

2 PAVED ROADS
There are two primary methods of permanent road construction: i) Flexible pavements ii) Rigid pavements Flexible pavement construction is still the most common and comprises upper layers of bituminous bound materials over a lower unbound granular sub-base. In contrast rigid pavements comprise plain and reinforced concrete over unbound granular sub-base. A combination of these two construction techniques is the semi rigid pavement, which usually comprises a concrete road base capped with bituminous bound material with an unbound granular sub-base. Geotextiles may be incorporated in permanent pavements in a number of ways, the most common being: i) At the sub-base/formation interface ii) As part of a surface overlay The function of the geotextile varies according to its type and siting within the pavement construction.

2.1 RECOMMENDATIONS
The current state of the art is one of continuing investigations and trials. In the UK it is still considered that despite trials over several years, sufficient data is not available to allow authorative designs to be formulated. However some of the results allow the following recommendations to be made: i) At the sub-base/formation interface The principal benet of geotextiles at the sub-base/formation boundary is to prolong the life of the pavement by minimising contamination of the sub-base by nes migrating from the formation. ii) As a surface overlay The geotextile is laid on the surface of an existing pavement prior to laying an asphaltic overlay. To be effective the geotextile has to act as in (ii) above to extend the life of the overlay and in addition present an impermeable barrier to water percolating through cracks in the overlay. Consequently the geotextile must be capable of absorbing sufficient bitumen or emulsion to render it impermeable.

2.2 MODEL SPECIFICATION


Clause XY1: Geotextile covering to formation between point A and point B. The geotextile shall be manufactured from durable synthetic polymers. The minimum tensile strength shall be.....kN/m at a strain of.....% determined in accordance with EN 10319. The water permeability normal to the plane of the geotextile shall not be less than.....m/sec determined in accordance with EN ISO 11058. The 090 pore size shall be in the range.....microns to.....microns determined in accordance with EN ISO 12956. Clause XYZ: Geotextile covering to existing pavement between point A and point B. The geotextile shall be manufactured form durable synthetic polymers. The minimum tensile strength shall be.....kN/m at a strain of.....% determined in accordance with EN 10319. The geotextile shall be installed in accordance with the manufacturers recommendations. Please refer to pages 76 78 to determine appropriate Lotrakgrade required.

19

2.3 DESIGN CONSIDERATIONS


Geotextiles at the Sub-base Formation Interface
Among other things the service life of a pavement is a function of the magnitude of surface deection under loading. Black & Lister (1978) have indicated an inverse power law relation between pavement deection and life thus it is vital that deections are limited to acceptable levels. Analytical work by Brown (1978, 1981) has indicated that the deformation modulus and resilient modulus of bound road construction materials is high. With adequate connement unbound granular material can achieve high moduli, up to 700 MN/m2. In practice unbound layers are not so well conned and in consequence the deformation modulus is much lower with insitu investigations indicating a modular ratio of 2.5 between unbound sub-base and formation. It follows from these observations that pavement deection and hence design life is strongly inuenced by the modulus of the subbase and particularly the resilient modulus which controls permanent deformation. Since the deections suffered in paved roads are small compared to those in unpaved, geotextiles, unless extremely stiff, will not generate any direct reinforced effect. Analytical work by Thompson has shown that in the case of a railway ballast a geotextile decreased horizontal strain at the base of the ballast which consequently reduced vertical strains at the top of the formation by approximately 30%. Brown (loc cit) extended this analysis to a layered pavement system and concluded that such reductions in strain could treble the modular ratio. It is theoretically possible for geotextiles to reduce horizontal strains, and therefore vertical strains in the sub-base, geotextiles are more likely to prolong life through reducing the contamination of the sub-base by nes migrating from the formation. Trials reported by De Groot (1986) using geotextile reinforced with steel cord, so giving high tensile stiffness, show that they can impart effective reinforcement when applied over soft formations. As well as acting as a structural element in a pavement the sub-base should also act as a drain to intercept any downward ows and so prevent damage to the formation. If the sub-base does not have high permeability the damage will still occur. In consequence there is a risk of water standing at the sub-base/ formation interface. Under dynamic traffic loading this free water, as well as the pore water at the surface of the sub-base generates high pressure. This results in weakening of the sub-grade and a jetting of nes into the sub-base. The pumping action leads to progressive contamination of the sub-base and a correspondence reduction of permeability. As well as decreasing shear strength the enhanced nes content reduces the resilient modulus of the sub-base and therefore increases the permanent strain caused. The increased strain in the sub-base is ultimately reected through the overlying bound materials which can fail prematurely. Reduced performance under dynamic load has been investigated experimentally. As can be seen Figure 2.1 a contaminated sub-base, subject to undrained loading, suffers permanent strains several times larger than its uncontaminated counterpart where drained loading prevails. The change from drained to undrained loading is largely a function of sub-base permeability. If a saturated sub-base is highly permeable than even when a rapid load is applied the resulting pore water pressures can be dissipated rapidly with the pore water owing from the loaded areas thus giving rise to drained loading. Conversely if the sub-base is contaminated its permeability is reduced and the sub-base does not have the same facility to drain rapidly.

Figure 2.1

20

This migration of nes can be reduced by the introduced of a suitable geotextile at the sub-base formation interface. Work by Bell et al 1982, has indicated that the sub-base contamination, measured as grams/m2 of nes (clay sizes) passing per square metre can be reduced as shown in Figure 2.2.

Figure 2.2

The work by Bells might lead to the conclusion that the pumping of nes can only be reduced if the pore size of the geotextile is small enough to prevent the passage of the nes. Considering the geotextiles tested it was found that none had small enough pore sizes to eradicate contamination completely. In fact the smallest effective opening size 095, was achieved using a granular lter. The result conrmed in workpublished by Schober & Teindl (1979) who presented a relationship between the channel diameter dk and coefficient of uniformity U for a granular soil. Taking dk090 a granular lter can achieve a very small pore size. For example a well compacted sand with U20m and 09520m (such small pore sizes are usually only to be found in specialist geotextiles). As well as pore size the thickness of the lter appears to have bearing on its effectiveness with the granular lter represented in Figure 2.2 being an order of magnitude thicker than the geotextile lters. Increased thickness would give rise to more even contract pressures between sub-base and sub-grade as well as having a threedimensional ltering effect. More research and eldwork is warranted, however it appears that a thick brous composite geotextile may form the basis of a useful sub-base sub-grade lter.

Figure 2.3

21

Geotextiles Incorporated in the Pavement Structure


The mechanism of failure in the bound elements of a pavement vary according to the structure of the pavement, in exible pavements, comprising layers of bituminous material over an unbound granular sub-base, failure of the bound layers is either indicated by rutting or fatigue. Fatigue generally results in the propagation of a tensile crack from the bottom to the top of the bound layers. Once a crack is exposed at the surface of the pavement it allows ingress of water to the sub-base and sub-grade which deteriorates rapidly under trafficking. Rigid pavements do not usually suffer from rutting but are prone to fatigue cracking. The cracks, once developed, allow ingress of water which reduces sub-grade support and more cracking is induced. Where a pavement structure comprises several layers of material the fatigue related cracking is generally initiated at the bottom of the lowest layer. Cracks migrate vertically upwards and pass into the layer above, until the cracks eventually reach the surface. Development of such cracks is termed reective cracking. The same phenomenon occurs in cracked pavement structures which are resurfaced. The use of geotextiles to resist in reective cracking is still developing. It is accepted that rutting failure is related to the magnitude of vertical strain in the pavement. The rutting life, determined by the cumulative number of standard axles passes to cause failure, is generally not the governing criterion, this is fortunate, since geotextiles seem to have little or no effect in reducing vertical strain. As far as fatigue failure is concerned, it is the maximum horizontal strain in the bound layers that must be limited to prevent the generation of cracks. Work by Majidzadeh et al (1982) has shown that fatigue life, dened by the cumulative number of stress applications, or standard axle loads, Nf, to reach a failure strain h, increases when a geotextile is incorporated, Figure 2.4.

Figure 2.4

The analogy is similar to that between the crack resistance of a mass concrete and reinforced concrete beam. Unfortunately even the strongest and stiffest geotextiles are not adequate to act as a reinforcement at the base of a thick pavement construction. in view of this, geotextiles should be installed higher up in the pavement construction, for example between the surfacing and road base as shown in Figure 2.5.

Figure 2.5

22

In this position a geotextile would be expected to arrest reective cracking and so extend service life. The location of the geotextile just below the surface is well suited for new construction as well as for surfacing overlays on existing pavements where cracks have, just between the surfacing, already reached the surface. The actual fatigue life obtained will be affected by the stiffness of the geotextile as shown in Figure 2.6

Figure 2.6

Horizontal strain is to be limited, the geotextile modulus used in Figure 2.6 is the 5% secant modulus. Depending upon the polymer used, its treatment during the geotextile manufacture and the temperature of the hot bituminous overlay, shrinkage or stretching of the geotextile may occur which may subsequently induce cracking in the surface. Button et al (1982).

Geotextiles as Surface Overlays


In existing pavements that have cracked, geotextiles can be used to advantage in resurfacing. When applying such bituminous overlays the geotextile can prevent the migration of cracks from the old pavement to the new surfacing as well as acting as an impermeable barrier if sufficient binder has permeated the fabric. Reducing permeability can generally be achieved by rst applying a tack coat to the existing road surface before laying the geotextile. If the geotextile is not impregnated then the service life of the road can be reduced. Cracks will eventually reach the surface whether the new surfacing is reinforced or not. Once the cracks open, water will ingress immediately if the geotextile permeable. If however the geotextile is impermeable it will prevent water ingress until the geotextile itself fails. The failure of the geotextile generally stems from its inability to absorb differential vertical movement across cracks, Hugo et al (1982).

23

2.4 REFERENCES
Bell A.L., McCullough L.M. & Snaith M.S.1982 An experimental investigation of sub-base protection using geotextiles. Proceedings 2nd International Conference on Geotextiles, Las Vegas. I.F.A.I. Vol.2. Black W. & Lister N. 1978 The Strength of clay ll sub-grades:- its prediction and relation to road performance. Proceedings Conference of Clay Fills, Institution of Civil Engineers. I.C.E London (pub. 1979). Brown S.F. 1978 The potential of fabrics in permanent road construction. Proceedings Conference on Textiles in Civil Engineering, UMIST. Brown S.F 1981 The structural role of granular materials in exible pavements. Proceedings Symposium on Unbound Aggregates in Roads. Dept of Civil Engineering, University of Nottingham. Button J.W., Epps J.A., Lytton R.L. & Harman W.S. 1982 Fabric interlay for pavement overlays. Proceedings 2nd International Conference on Geotextiles, Las Vegas. I.F.A.I Vol. 2. De Groot M.T 1986 Woven steel cord networks as reinforcement of asphalt Proceedings 3rd International Conference on Geotextiles, Vienna Vol. 1. Hugo F., Strauss P. & Schnitter O.1982 The control of reection cracking with the use of a geotextile:- a ten year case history. Proceedings 2nd International Conference on Geotextiles, Las Vegas. I.F.A.I. Vol. 2. Leaive E., Morel G. & Khay M. 1982 The use of geotextiles in exible pavement surface dressing. Proceedings 2nd International Conference on Geotextiles, Las Vegas. I.F.A.I. Vol. 2. Majidzadeh K., Luthur M.S. & Skylut H. 1982 A mechanistic design procedure for fabric reinforced pavement systems. Proceedings 2nd International Conference on Geotextiles, Las Vegas. I.F.A.I. Vol. 2. Murray C.D. 1982 Stimulation testing of geotextile membranes for reection cracking. Proceedings 2nd International Conference on Geotextiles, Las Vegas. I.F.A.I. Vol. 2. Schober W. & Teindl I.H. 1979 Filter Criteria for geotextiles. Proceedings 7th Euro Conference Soil Mechanics and Foundation Engineering. Brighton. Vol. 2. Thompson M.R. 1976 Mira 140 Fabric in Conventional Railway Track Support Systems. Res. Rep University of Illinois.

24

3 DRAINAGE
In drainage geotextiles can be used as lters between the naturally occurring soil and the conducting medium. Such applications are: Linear Drainage Trenches Here the geotextile encapsulates a free draining ll in a shallow trench.

Fig. 3.1 Linear Drains

Figure 3.2 Drainage Blanket

Horizontal Drainage Blankets Normally placed beneath a layer of ll material of various types including wastes.

Fin Drains These consist of a composite system of geotextile lters wrapped around a water conducting mesh core.

Figure 3.3 Fin Drain

25

3.1 GEOTEXTILES AS FILTERS


Introduction
Aggregate or rubble drains of many types are used in all manner of civil engineering and agricultural projects. The most common form of aggregate lled drains is the French drain that comprises a trench lled with a free draining aggregate.

Figure 3.4 Cross-section of French Drain

French drains can be used to collect surface water run-off that enters at the top of the drain. However, these drains are more commonly used to control ground water ow or to depress high ground water tables. Either of these applications involves the ow of ground water from the base soil into the aggregate of the drain whence the water ows through the aggregate in the trench to a suitable point of discharge. As the ground water ows towards the drain it is possible for it to carry ne particles of the base soil which can ow into the drain and block the aggregate. The continued transmission of nes from the base soil is called piping. If this is allowed to continue it causes internal erosion of the base soil. In a hydraulic structure such as an earth ll dam, this can lead to uncontrolled seepage and ultimately failure. To obviate these hazards the aggregate in the drain must full two criteria: i) Maintain permeability for transmission of water ii) Act as a lter to the base soil to prevent piping The criteria can be met by using aggregates with special gradings related to the grading base soil. Meeting these requirements often proves to be an extremely expensive. The same result can be achieved at a fraction of the cost by using geotextiles that act as lters while maintaining the required system permeability. Before the design of geotextile lters can be considered, it is necessary to review some basic properties of geotextiles and the interaction of these properties with the prevailing soil and hydraulic regimes.

Geotextile Types
One of the rst considerations in assessing a geotextile as a lter is the construction of the geotextiles and especially the effects of structure on the magnitude and distribution of pores within the fabric. To this end geotextiles are classied into three categories: i) Wovens ii) Thin non-wovens iii) Thick non-wovens As the same name implies woven geotextiles comprise two elements, a weft element running across the width of the fabric and a warp element running the length of the fabric. These are woven using conventional techniques to give an orthogonal structure having a relatively regular mesh opening evenly distributed over the surface of the fabric. The elements of the fabric are commonly at tapes, tape yarns or monolaments of circular cross-section.

26

Figure 3.5 Monolament/tape woven

Figure 3.6 Tape/tape woven

When the weave is not square the mesh opening is rectangular. In this case the mesh opening is taken as the smaller dimension. Although mesh sizes can be measured by direct optical methods they are more assessed by a dry sieving technique. In this case the performance of the fabric is commonly related to the 095 pore size, which corresponds to the average diameter of the sand fraction of which 5% by weight passes through the fabric. Thin non-woven geotextiles are fabrics with a thickness of less than 1mm. These fabrics are commonly constructed using thin continuous laments that are initially laid to form a loose web. On rolling and heating the web is compressed and the laments caused to fuse together at their contact points. It is from this melt bonding process that the fabric gains its integrity. As a result of the random distribution of the various contact points the pore openings in non-wovens have a larger variation in size than those in a woven fabric.

Figure 3.7 Heat bonded (thin) non-woven

Figure 3.8 Needle punched (thick) non-woven

Thick non-wovens are commonly felts that are constructed using short staples or bres. These are initially laid down as a loose web, which is penetrated by a bank of reciprocating barbed needles. The needling action causes mechanical entanglement of the bres and so imparts a degree of integrity. Often the product is nished by rolling to give a comparative smooth surface. By virtue of their thickness needle punched felts are often considered to be three-dimensional lters. As with the thin non-wovens the pore size can often vary over a wide range of sizes.

27

For both the thick and thin non-wovens, the pore size distribution is assessed by using the wet sieving technique according to EN ISO 12956. From these test results, it is possible to construct a pore size distribution curve for a fabric. As can be seen the mesh or pore sizes for a woven geotextile are comparatively uniform whereas those for a non-woven vary over a wider range. In practice the largest pore size for a non-woven is 098, although performance is generally considered to be controlled by the 095 pore size, however, some ltration rules will be based on 090 or 050 pore sizes.

Figure 3.9

Filter Mechanism
Filter criteria vary according to the hydraulic regime involved. For the drainage applications presented in this chapter the hydraulic regime involves steady state unidirectional laminar ow. The importance of this becomes apparent when the lter mechanism is considered. In woven geotextiles where mesh opening is comparatively uniform and evenly distributed over the fabric, the coarser soil particles bridge over the mesh openings, allowing ner particles immediately in contact with the geotextile to pass through the mesh openings and to be ushed away by the ow of water. Unless the soil is subject to suffusion, that is migration of very ne soil particles within the soil matrix, the system stabilises very rapidly and there is no change in the soil structure remote from the soil particle interface. Once the system is established there is no further disturbance to the soil structure unless the hydraulic regime changes. Thus under conditions of steady state unidirectional ow the bridging network remains stable once it has been established. If the ow rate changes or the ow alternates then a bridging network will not be formed a nd a stable state will not prevail. It is often found that there can be a slight reduction in ow rate during the initial development of the bridging network due to the larger particles blocking the mesh openings in geotextile. Blocking does not involve soil particles lodging in the mesh openings but merely their partial covering of the openings. A similar bridging network mechanism is attributed to non-woven fabrics where the size of pores tends to be much more variable. However immediately upstream of this it has been suggested that a lter cake or lter zone develops. This zone comprises a band of soil particles of the coarser fraction forming the bridging network and ner material that has migrated from further upstream in the soil. Once formed this zone prevents further migration, however it does present a zone of lower permeability than that of the soil with its original grading. Also in non-wovens there is more of a tendency to clog. Clogging is somewhat different to blocking in as much as it involves soil particles actually lodging in pores either at the surface of the geotextile or within the thickness of the geotextile.

Figure 3.10

28

For either woven or non-woven geotextiles there is some drop off in the through permeability of the system until a stable condition is reached. A stable condition will only be reached if the geotextile is correctly designed to suit the system and the hydraulic regime is steady thus allowing the development of the bridging, and where appropriate, the lter zone.

Filter Criteria
To function as an effective lter a geotextile must prevent piping yet the pore sizes must not be that small that they clog and therefore reduce permeability. The geotextile must: i) Prevent piping within the soil ii) Maintain the permeability of the soil system

Piping Criteria
To prevent piping, the mesh or apparent opening size has to be related to the average particle size of the base soil to dene a ratio 090/d50, known as the concept of positive retention. The retention concept is based upon the premise that a non-compressible geotextile lter, under unidirectional ow conditions, should be able to retain enough large-sized soil particles in order for a natural soil lter to be formed at the soil/geotextile interface (Giroud 1982). When selecting a geotextile for use as a lter it is necessary to take into account the relationship between the effective opening size (EOS) of the geotextile, the coefficient of uniformity of the soil, U, and the average soil particle size, d50. For example, if the soil has a very uniform grading, then the geotextiles signicant pore size, either O90, O95 or Dw, should only be a low multiple of d50; whereas, if the soil is well graded, then the signicant geotextile pore size could be many times d50. This basic idea has the support of a number of researchers, such as Giroud (1982), Heerten and Wittmann (1985), Ingold (1985), John (1987), Schober and Teindl (1979), and Watson (1995). On this basis, the Highways Agency carried out tests at the Transport and Road Research Laboratory (TRRL), (Murray & McGown 1992). The following design criteria based on O90, as detailed in Table 3.1 below was the result (Highways Agency 1993). Uniformity Coefficient U 1 to 5 More than 5 Woven and Meltbonded Geotextiles O90/d50 1 to 3 Less than 3 or O90/d90 < 1 Needle-punched Geotextiles O90/d50 4 to 6 Less than 6 or O90/d90 < 1.8

Table 3.1 U.K. Filter Criteria as recommended by the Highways Agency

From Table 3.1, it can be seen that there is a distinction is between non-compressible and compressible geotextiles rather than woven and non-woven geotextiles. The approach is logical as the pore size of a compressible geotextile when buried on site will be smaller than that measured in the laboratory under low or no conning stress. Table 3.1 depicts the U.K. criteria in the tabular form, alternatively the criteria can be presented in chart form as shown in Figure 3.11. It should be noted that in very ne-grained soils, these criteria could lead to very low ow rates through the geotextile. Also for some coarse soils, the criteria may lead to problems with the internal stability of the soil. Limits are therefore placed on the design values for O90 of 0.1 mm and 1.0 mm respectively in the HA Notes for Guidance. If the soil is gap-graded, the coarser fraction of the particle size distribution curve is neglected in the calculation, this can again result in a high risk of clogging (Farrar 1993), so care must be exercised with any drainage design.

29

Fig 3.11 U.K. Filter Criteria Represented in Chart Form

Permeability Criteria
For steady state laminar ow, the ow rate Q per unit area is related to Darcys coefficient of permeability, k, and the hydraulic gradient, i, across the section being considered. Q = ki (per unit area) Although it is possible to dene a coefficient of permeability for a geotextile it must be remembered that k is likely to be affected by several parameters. If the thickness of the geotextile is tg, then the hydraulic gradient may be dened as the head loss across the geotextile H, divided by the thickness of the geotextile, such that: i = H/t

Thus from a measurement of the ow rate Q across a geotextile, it might be assumed that: kg = Q/i = Qtg/H which suggests that kg is a constant of the geotextile. Generally not the case, since in the Darcy formulation it is assumed that velocity and permeability are directly proportional to hydraulic gradient i.e. laminar ow. The denition of permeability is further complicated by the fact that the thickness of the geotextile, tg, is especially pronounced for the thicker non-woven geotextiles which are particularly compressible. In view of this variability, it is common practice to quote the hydraulic conductivity or water permeability of a geotextile at a given normal stress level for a given head loss, for example 30 ltrs/m2/sec for a head loss of 50 mm of water at a normal stress level of 40 kN/m2. In this manner, it is possible to design for a head loss to maintain a desired ow rate rather than blindly applying what may be an inappropriate permeability. Test results presenting water permeability data should always include the head at which the test should be carried out, normally 100mm. The relationship between water ow and head is not normally linear, as shown in Figure 3.12. The non linearity is particularly true at low heads, since for certain geotextile types a minimum head of water is required to overcome surface tension effects and to initiate ow.

30

Figure 3.12 Graph showing drop off of Flow Rate with Load

The considerations about ow, are limited to a geotextile tested in isolation, that is of a geotextile alone and not in the geotextile-soil situation in which it will be expected to function. Another order of complexity is introduced by considering the effects of soil since as described earlier this is likely to affect long term conductivity with ow rate falling with time until a steady state condition can be achieved.

Figure 3.13 Graph to determine stable condition of a geotextile

If a geotextile lter is to be used in situations where the ramications of malfunction are high, then the entire geotextile-soil system should be evaluated by testing including all relevant variables. The end result is to determine if a long term stable condition can be reached i.e. dQ/dt = 0 and at what value of hydraulic conductivity, as shown in Figure 3.13. For less critical situations, it has been suggested that the permeability of a geotextile be expressed relative to the permeability of the soil. These values ranged from kg = 0.1ks to kg = 10ks. With lack of any other information it would be reasonable to assume kg = 5ks. The approximation applies to pure water where the inservice reduction of permeability is solely a function of the geotextile-soil system. Where groundwater is biologically or chemically active, the permeability of the geotextile-soil system may be radically impaired by deposits forming at the geotextile-soil interface or even direct chemical attack of the geotextile. Such problems are encountered in groundwater rich in ferrous iron and are described by Puig et al (1986).

31

3.2 DESIGN PROCEDURE


Design Input Information
The following data has to be supplied by the designer to establish the design parameters. SOIL PROPERTIES: A particle size grading curve for the soil to be ltered is required. Should an actual grading not be available the following typical curves may be used as a guide.

Figure 3.14 Typical particle-size distribution curves for different soils

SOIL PERMEABILITY: Soil permeability ks is required in m/sec. The following approximations may be used as a guide. ks (m/s) 10-11

10-1

10-2

10-3

10-4

10-5

10-6

10-7

10-8

10-9

10-10

pebble beds

gravel sands

silts clays Cemented sandstones ssured mudstones

Figure 3.15 Typical soil permeability values

DRAIN CAPACITY: The required discharge capacity of the drain Q in m3/s for the expected prevailing conditions is needed.

32

Geotextile Filter Design


From the particle size distribution curve (grading curve) for the soil (see Figure 3.14), determine both the mean particle size d50, and the uniformity coefficient of the soil, U. Using the lter criteria recommended by the Highways Agency (see Figure 3.11), and the soil properties determined, calculate the maximum value of the O90 pore size for the lter geotextile. Ensure that if the soil is either very ne-grained soils, or coarse, that the appropriate limit on the design values for O90 of either 0.1mm or 1.0mm is used. Select an appropriate geotextile. Check the permeability of the chosen geotextile: 5ks kg (geotextile permeability) (soil permeability)

Drain Capacity Design


For a drainage medium comprising single size aggregate only, the ow velocity v m/s is determined from Figure 3.16 for a given aggregate at a given hydraulic gradient. The required discharge capacity Q m3/s divided by v m/s will give the necessary cross-sectional drain area.

Figure 3.16 Determination of Flow Velocity at a known Trench Gradient

Figure 3.17 Determination of Flow Velocity at a known Pipe Gradient

Where a carrier pipe is to be used its size is determined from Figure 3.17 knowing the gradient and the required discharge capacity.

33

Depth of Drain Design


Where two parallel drains are required to cope with precipitation between them and to prevent the ground water table rising above original ground level the depth is determined from:

where: D = drain depth S = distance between drains (m) e = precipitation rate (m/s) ks = soil permeability (m/s) b = drain width (m) V = ow velocity as determined in Figure 3.16 or 3.17 (m/s) L = drain length (m) N.B:

Figure 3.18

For a series of parallel drains the factor 12 is omitted

Thickness of Horizontal Drainage Blanket

Figure 3.19 Nomenclature for Horizontal Drainage Blanket

For upward ow from artesian aquifer: where: h = excess pressure head d = depth to aquifer For download ow due to precipitation: where: e = precipitation rate (m/s) The maximum ow from blanket to French drain is: per linear metre of French drain which each of the French drains must be capable of carrying.

Fin Drain Design


The selection of the geotextile lter, the carrier pipes and the drain depths are the same as for a French drain. CORE CAPACITY: The system will always use a carrier pipe for linear drainage so the maximum capacity required of the core is: m3/s per linear metre of drain where: ks = soil permeability (m/s) where: d = depth of n drain For a cut off drain where ow is from one side only: m3/s 34

For a horizontal drainage blanket controlling upward ow as in section 3.4.5 above, the required capacity is: m3/s/m length For a horizontal drainage blanket intercepting rainfall as in section 3.4.5 above, the required capacity is: m3/s/m length For a combined system the required capacity is: m3/s/m length Check the core capacity available at the expected soil conning pressure, data from the EN 1897 Compressive Creep Test can be used for checking the core capacity.

3.3 MODEL SPECIFICATION


Encapsulated Drainage
The linear drainage shall be constructed to the sizes, depths and levels as indicated in the drawings and shall comprise geotextile lter plus granular ll pipe (if required). The geotextile shall be manufactured from durable synthetic polymers. The water permeability normal to the plane of the geotextile shall not be less than.....m/sec determined in accordance with EN ISO 11058. The O90 pore size shall be in the range.....microns to.....microns determined in accordance with EN ISO 12956. The CBR push through resistance shall not be less than.....N determined in accordance with EN ISO 12236. The drainage ll shall be a hard clean crusted rock, crushed slag or gravel having a nominal single size of.....mm. The carrier pipe where used shall conform to the Specication for Highway Works, Clause 501 and shall have an internal diameter of.....mm.

Fin Drainage
The linear n drainage shall be constructed to the depths and levels as indicated on the drawings and shall comprise geotextile lter plus core plus carrier pipe. The geotextile shall be manufactured from durable synthetic polymers. The water permeability normal to the plane of the geotextile shall not be less than.....m/sec determined in accordance with EN ISO 11058. The 090 pore size shall be in the range.....mm to.....mm determined in accordance with EN ISO 12956. The CBR push through resistance shall be not less than . . . N determined in accordance with EN ISO 12236. The core shall be manufactured from durable synthetic polymers. The in-plane water permeability shall be not less than.....m/sec at a hydraulic gradient of.....and at a conning pressure of.....kN/m2 determined in accordance with EN ISO 12958. The carrier pipe shall be uPVC performed or slotted and manufactured in accordance with the appropriate BS and have an internal diameter of.....mm. 35

Horizontal Drainage Blanket - Granular


The drainage blanket shall be constructed in the positions and to the sizes and thicknesses shown on the drawings. It shall comprise a layer of granular material with a geotextile lter on one or both sides. The geotextile lter shall be manufactured from durable synthetic polymers. The water permeability normal to the plane of the geotextile shall not be less than.....m/sec determined in accordance with EN ISO 11058. The 090 pore size shall be in the range.....mm to.....mm determined in accordance with EN ISO 12956. The CBR push through resistance shall be not less than.....N determined in accordance with EN ISO 12236. The granular layer shall be hard clean crushed rock, crushed slag or gravel having a permeability not less than.....m/sec after placing.

Horizontal Drainage Blanket - Fin Drain


The n drain blanket shall be constructed in the positions and to the sizes shown on the drawings. It shall comprise geotextile lter plus conducting core. The geotextile lter shall be manufactured from durable synthetic polymers. The water permeability normal to the plane of the geotextile shall not be less than.....m/sec determined in accordance with EN ISO 11058. The 090 pore size shall be in the range.....microns to.....microns determined in accordance with EN ISO 12956. The core shall be manufactured from durable synthetic polymers. The in-plane water permeability shall not be less than.....m/sec at a hydraulic gradient of.....and at a conning pressure of.....kN/m2 determined in accordance with EN ISO 12958. Please refer to pages 76 78 to determine appropriate Lotrakgrade required.

3.4 DRAINAGE APPLICATIONS


Draw Down
French drains, with or without a carrier pipe, have a wide range of applications in both civil engineering and agriculture. One particular application is to draw down a high ground water table, to do this the suppression of ground water can be achieved between two parallel drains at a spacing S.

Figure 3.20 Nomenclature for Draw Down of Groundwater

36

If the ground water regime is never recharged then almost any drainage system will eventually draw down the ground water table. In reality any exposed ground will be subject to recharge by rain or other precipitation. Therefore, to be effective the French drains must be able to cope with a given rate to prevent the ground water table rising. For drains at spacing S, ow rate per metre run of drain will be m3/s where e is the precipitation rate in m/sec The portion of the drain depth, d, required to carry ow can be determined using the graph of velocity versus trench gradient as shown in Figure 3.16. The total depth of drain, D, is selected to give the required draw down, d1.

Generally the draw down, d1, is set to zero hence :

Alternatively for a trench breadth, b, this can be expressed as :

where V is the ow velocity determined directly from the design curve Figure 2.16.

Horizontal Drainage Blanket


Horizontal drainage blankets are generally employed to intercept vertical ow. This may be upward vertical ow that might extend from an artesian aquifer or spring or downward ow in the form of precipitation. Although a horizontal blanket intercepts the vertical ow, this ow is generally borne by a carrier, in the form of a French drain.

Figure 3.21 Nomenclature for Horizontal Drainage Blanket

Figure 3.22

37

Figure 3.21 shows a horizontal drainage blanket of thickness t and permeability kb resting in a soil with permeability ks. The excess artesian head is h and therefore the upward hydraulic gradient is i = h/d. This gives rise to an upward ow rate q m3/s per m2.

This will enter the horizontal drainage blanket from which it will discharge by horizontal ow into French drains. The thickness of the horizontal blanket, t, must be sufficient to completely intercept the vertical ow. This is achieved if the seepage line is not allowed to rise more than t above formation level. This condition is achieved when:

Since the ow rate into the drainage blanket is known:

A similar analysis can be employed to intercept inltration from rainfall. In this case, the ow rate q is equal to the inltration rate, e, where the required thickness of the drainage blanket is:

The rate of ow from the horizontal drainage blanket reaches a maximum value of:

N.B.: For the interception of either precipitation or upward hydraulic ow the French drains must each be capable of carrying this discharge. Basically, the French drains reduce the vertical ow rates. An indication of the degree of pressure control possible is given in the following design chart (see Figure 3.23), which is applicable to the control of artesian uplift pressures beneath impermeable pavements.

Figure 3.23 Design Chart for Control of Artesian Uplift

38

3.5 EXAMPLE
A 20 m wide road is to traverse a silty sand of the above grading. The permeability of the sand is ks = 10-4 m/s. The sand is 14 m thick and overlies an artesian aquifer giving an excess head of 3 m of water. What is the 090 of a geotextile to lter the soil and what thickness of horizontal drainage blanket is required to intercept vertical ow. The drainage blanket is a coarse sand of permeability kb = 1.0 m/s. The particle size distribution curve for the silty sand is shown below in Figure 3.24.

Figure 3.24 Particle Size Distribution Curve for a Silty Sand

From the grading curve determine the d50, d60 and d10 particle size values. d50 = 0.090 mm d60 = 0.101 mm d10 = 0.022 mm Therefore, the uniformity coefficient of the silty sand is:

From the HIghways Agencys design criteria, see Figure 3.11, the maximum pore size for a lter geotextile is: For needle-punched geotextiles: Maximum value of O90/d50 = 5.6 Maximum value of O90 = 5.6 x 0.090 mm = 0.50 mm

39

For woven and melt-bonded geotextiles: Maximum value of O90/d50 = 2.8 Maximum value of O90 = 2.8 x 0.090 mm = 0.25 mm The required thickness of drainage blankets is given by:

Therefore, thickness, t = 46 mm

3.6 TYPICAL LOTRAK DRAINAGE GEOTEXTILES


Lotrak Geotextile Pore Size O90 ner than (microns) EN ISO 12956 Water permeability (m/sec x 10-3) EN ISO 11058 CBR Push through Resistance (N) EN ISO 12236
Table 3.2

1800 225 16 1800

2300 200 22 2500

2800 260 20 2800

HF 550 380 75 2000

HF 400 260 45 2000

3.7 REFERENCES
Highways Agency Manual of Contract Documents for Highway Works, March 1998: Specication for Highway Works, Clause 501, Vol.1, Series 500, Notes for Guidance, Clause 501, Vol.2, Series 500. Farrar, D.M. Geotextiles in the Design of Filtration and Drainage for Highways, in Geotextiles in Filtration and Drainage - edited by S. Corbet and J. King, Thomas Telford, London 1993, 13-28. Giroud, J.P. Filter Criteria for Geotextiles, Proc. Second Int. Conf. on Geotextiles, Las Vegas 1982, Vol.1, 103-108. Heerten, G. and Wittmann, L. Filtration Properties of Geotextile and Mineral Filters Related to River and Canal Bank Protection, Geotextiles and Geomembranes, Vol.2, No.1, 1985, 47-63. Ingold, T.S. A Theoretical and Laboratory Investigation of Alternating Flow Filtration Criteria for Woven Structures, Geotextiles and Geomembranes, Vol.2, No.1, 1985, 31-46. John, N.W.M. Geotextiles, Blackie, Glasgow 1987. Murray, R.T. and McGown, A. Ground Engineering Applications of Fin Drains for Highways, Department of Transport Application Guide 20, Transport and Road Research Laboratory, Crowthorne, Berks., 1992. Puig, J., Gouy, J.L. and Labroue, L. Ferric Clogging of Drains, Proc. Third Int. Conf. on Geotextiles, Vienna 1986, Vol.4, 1179-1184. Schober, W., and Teindl, H. Filter Criteria for Geotextiles, Proc. Seventh Euro. Conf. on Soil Mechanics and Foundation Engineering, Brighton 1979, Vol.2, 121-129. Watson, P.D.J., Geotextile Filter Design and Particle Bridge Formation, Ph.D. Thesis, University of London, June 1995.

40

4.0 EROSION CONTROL


4.1 INTRODUCTION
Water is the most aggressive of the elements causing soil erosion. The degree of damage caused by erosion will vary enormously with the erosive power of the water and the erodibility of the soil, or system, under attack. Erosion is a natural process forming part of the geological cycle and it is this process which erosion control works presume to modify. This section makes an introduction to two aspects of erosion control in which geotextiles and related products can play a useful role. The rst of these is erosion control in inland waterways where the erosive power of the water has two main components. Firstly there is plucking and subsequent transportation of soil particles which occurs when owing water exceeds a critical velocity related to soil particle size. Unchecked this can lead to mass movement in the form of landslip. The second, and usually more damaging, component is wave action and propeller scour caused by vessels. Wave action from the passing of the vessels is particularly damaging to unprotected banks at mean water level. Other than impact, waves cause an oscillation in water level that gives rise to rapidly reversing ow in the soil. These dynamic loadings are highly erosive and can only be resisted by adequately designed revetments incorporating appropriate anti-erosion ltration systems. Where inland waterways such as rivers, lead to the sea, there may be a need for more substantial hydraulic works such as breakwaters. Although this subject is touched upon, such structures require specialist design. The second aspect of erosion control considered is soil erosion. As will be seen this is a function of many variables. However, prime among these is the erosion power of the rainfall and the erodibility of the soil. Since erosion control measures required for soil conversation must have a long service life, they frequently include the use of vegetation. Although this is intended primarily to give protection to soil slopes subject to intermittent rainfall, certain techniques can be extended to give protection to the banks of minor watercourses or riverbanks where ow velocities are not excessive.

4.2 DESIGN THEORY


Introduction
In inland waterways current velocity and the passage of vessels are the prime causes of erosion. If these effects can be modelled by a current velocity, then Figure 4.1 would give some indication of the erodibility of the banks and the beds of the waterway. It is interesting to note from Figure 4.1 that the most erodable soil is sand, which suffers erosion at a current velocity of approx. 0.2 m/s.

Figure 4.1 Erosion as a Function of Current Velocity and Particle Size

41

To minimise erosion the banks or bed of the waterway may be protected with armour stone of sufficient weight and size to resist the current. A possible detail is indicated in Figure 4.2, which shows an outer armour layer, two courses thick, founded on a bedding layer of ner granular material. This is turn is underlain by a geotextile lter placed on the river bank/bed after preparation.

Figure 4.2 Protection of Inland Waterways

Filter Design
The design of a geotextile lter in erosion control is complicated by the fact that ow is often turbulent and subject to frequent reversal. For example, in a non-tidal canal the ground water level adjacent to the canal might equal the water level in the canal. On the passage of a vessel the bow wave created would locally raise the water level in the canal causing a ow of water into the bank. Following this, the trough of the bow wave lowers the water level in the canal and induces water to ow out of the bank. Due to this alternating ow, there is often no possibility for the formation of a bridging network of soil particles adjacent to the geotextile, and as such the soil cannot generate a stable ltrating conguration within its structure. Although research has been carried out on this problem it is fragmented, save for the work of Heerten who looked at both woven and non-woven fabrics for lters in coastal and inland waterways. This study also encompassed both sands and silty sands, which are the soil types more prone to erosion. Heerten considers the dual problems of designing a lter that is ne enough to prevent serious loss of soil particles i.e. to ensure sand tightness, whilst ensuring that the geotextile maintains sufficient high permeability during its service life to prevent the build up of over pressure. The ltration rules to full the sand tightness requirement may be summarised as follows: a) Non-cohesive soils: STATIC LOAD CONDITIONS: If U 5 then and If U < 5 then and

O90 < 10d50 O90 < d90 O90 < 2.5d50 O90 d90

DYNAMIC LOAD CONDITIONS: O90 < d50 b) Cohesive soils: STATIC/DYNAMIC LOAD CONDITIONS: O90 < 10d50 and O90 d90 and O90 100m

42

Static load conditions are considered to be laminar ow, but can include a change of ow direction. Dynamic load conditions are created by high turbulent ow, wave action or pumping phenomenon. Also, for both soils types there will be variations in the above criteria according to soil silt content and the precise coefficient of uniformity (Heerten 1986 and Saathoff 1988). The permeability criterion requires the permeability of the geotextile, kg, to always be greater than the permeability, ks, of the bank soil that is being protected. If the permeability of the geotextile is measured in isolation in the laboratory, at the appropriate normal stress level, it may appear to be adequate i.e. kg > ks. However, when the geotextile is placed in contact with the soil its permeability is likely to decrease. Heerten has suggested that blocking, which is the phenomenon of soil particles either projecting over, or seating in the mesh openings of the geotextile, reduces the permeability of woven geotextiles. The reduction in woven fabric permeability can be modelled by a reduction factor w, which is a function of the geotextile permeability kg and the d10 size of the soil being ltered. The requirement is then that: w kg > ks The value of hw may be determined from Figure 4.3 below with knowledge of kg and d10.

Figure 4.3

In contrast to woven fabrics, the permeability of non-woven fabrics is affected by clogging, which is the lodgement of the soil particles within the geotextile. Clogging is inuenced by several parameters including: n = fabric porosity (typical 0.8 to 0.9 for needle punched felts) t = fabric thickness measured at a normal stress level of 2kN/m2. If the permeability of the geotextile in isolation, kg, is measured at a normal stress level compatible with that prevailing under service conditions then the non-woven reduction factor n can be applied such that: n kg > ks Values of n can be obtained from Figure 4.4 for values of the parameter kg2/n.t.O90.

43

Figure 4.4

For non-wovens there is an additional criterion related to n, which denes an upper bound when the O90 size is small in comparison to the soil particles. This may be as stated as: If O90 < 0.5d10 then n = 1

The design procedure to be followed is best summarised by the following ow chart, Figure 4.5.

Figure 4.5 Flow Diagram to check the Filtration Properties of Geotextiles

44

Geotextile Uplift
Uplift of the geotextile and its protective covering can be prevented by ensuring that the downward force from the cover material exceeds the upward force from the water pressure beneath the geotextile (Mauw et al. 1986).

Figure 4.6 Notation for Uplift Analysis

In Figure 4.6 the upward force perpendicular to the slope over length l is: Fu = w.h.l The downward force perpendicular to the slope is: Fd = pc.cos .l where: w pc h is the unit weight of water (kN/m3) is the submerged weight per unit area of the protective covering (kN/m2) is the head loss across the geotextile (m) is the angle of slope (degrees)

Hence, for stability: Fd > F u pc cos > w h h < pc cos /w

Wave Action
Waves breaking on a waterway revetment induce several different types of loading, as shown in Figure 4.7 (Tutuarima & van Wijk 1984). In Figure 4.7, the loading types are: a) Forces due to down rush b) Uplift pressures due to water in the lter c) Uplift pressures due to approaching wave front d) Change in velocity eld e) Wave impact f) Uplift pressure due to mass of water falling on slope g) Low pressures on slope due to air entrainment h) Forces due to up-rush

45

a = forces due to down-rush d a b c b = uplift pressures due to water in lter c = uplift pressures due to approaching wave front d = change in velocity eld

g e f

h f

e = wave impact f = uplift pressures due to mass of water falling on slope g = low pressures on slope due to air entrainment h = forces due to up-rush

Figure 4.7 Wave forces

The stability of the armour in coastal revetments composed of large rocks, has traditionally been assessed using Hudsons formula or updated versions of this. Experience indicates that for waterway revetments incorporating geotextiles, an alternative approach is justied (Pilarczyk 1984). In these cases, the stability of the armour is related to the stability number, SN. SN = H / RD where: H R D W A and R = ( A - W) / W

is the wave height (m) is the submerged relative unit weight of the armour (dimensionless) is the depth of the armour (m) is the unit weight of water (kN/m3). is the unit weight of the armour (kN/m3).

Usually R lies between 1.24 and 1.38 (Pilarczyk 1984). The minimum depth of the protective covering required to withstand the wave action can be determined from the table of required stability numbers, as shown in Table 4.1 (Pilarczyk 1984 and Tutuarima & van Wijk 1984). Protective Covering Unbonded Rip-Rap Free Blocks Asphalt Grouted Open Aggregate Sand Filled Mattresses Articulated Blocks Grouted Articulated Blocks Required Stability Number < 2 < 2 < 4.3 < 5 < 5.7 < 8

Table 4.1 Required Stability Numbers for Waterway Revetment Systems

The lower limit of 2 is compatible with Hudsons formula and can therefore be safely applied to any revetment system not listed above. The higher values for the upper limit stability number for the other systems permit the adoption of a thinner depth for the protective covering and result from the greater integrity that these systems case, it is assumed that the permeability of the geotextile exceeds that of the soil. If the geotextile permeability is only equal to that of the soil, then the above required stability numbers should be reduced by 40% or to 2.0, whichever gives the higher value (Tutuarima & van Wijk 1984). 46

Overlap Stability
Whenever possible any overlaps in waterway revetment mattresses should be arranged so that the exposed edge of the upper sheet is on the downstream side of the overlap. If this is not possible, or if the current direction reverses, then the situation shown in Figure 4.8 may occur.

Figure 4.8 Forces on an Overlapping Edge

Flow separation over the exposed overlap edge will cause an upward force attempting to lift the edge of the mattress. Lifting is found to occur (Tutuarima & van Wijk 1984) above a critical velocity, Vc, given by: Vc = s R.d.g

where: s R d g is a factor depending on the shape of the edge and the ow conditions (between 1.4 and 2.0). relative density of the revetment. average thickness of mattress gravity

Provided the edge of the mattress rests directly on the underlying mattress and the ow is not turbulent, then s can be taken as 2.0. This value gives the design chart shown in Figure 4.9.

Figure 4.9 Critical ow velocity for the overlapping edge

The same method of analysis can be used to investigate whether the edge of the mattress is likely to lift off the underlying soil. In this case, s can be taken as 1.4.

47

4.3 DESIGN PROCEDURE


Design Input Information
The following data has to be supplied by the designer to establish the design parameters. SOIL PROPERTIES: A particle size distribution curve is required from which the d10, d50, d60 and d90 are used. If an actual grading curve is not available the following typical curves shown in Figure 4.10 below, may be used as a guide.

Figure 4.10 Typical Particle-Size Distribution Curves for different soils

SOIL PERMEABILITY: Soil permeability ks is required in m/s. The following approximations may be used as a guide. ks (m/s) 10-11

10-1

10-2

10-3 gravel

10-4

10-5

10-6

10-7

10-8 silts

10-9

10-10

pebble beds

sands cemented sandstones ssured mudstones

clays

Figure 4.11 Typical soil permeability values

EROSION ACTION: Determine whether erosion is by laminar ow including reversing ow direction i.e. static condition or by high turbulent ow, wave action or pumping i.e. dynamic condition.

48

Erosion Control Design


This design procedure is concerned principally with those soils most susceptible to erosion, i.e. sands and silty sands, and to constructing a control system similar to that shown in Figure 4.12 below. 1 Determine maximum 090 lter A) NON-COHESIVE SOILS Determine uniformity coefficient

For static conditions: If U 5 If U < 5 then O90 < 10d50 then O90 < 2.5d50 and O90 < d90 and O90 d90 For dynamic conditions: 090 < d50 B) COHESIVE SOILS For both static and dynamic conditions 090 < 10d50 and and 090 d90 090 100m
Figure 4.12 Typical Erosion Control Application

2) Check Geotextile: Select probable geotextile lter A) WOVEN GEOTEXTILES If where w. kg > ks then kg is sufficient

w is the permeability reduction factor for woven geotextiles kg is the geotextile permeability m/s

The permeability reduction factor w is then determined from Figure 4.3 as shown in section 4.2.2, using the appropriate d10 line. B) NON-WOVEN AND COMPOSITE GEOTEXTILES If where n. kg > ks then kg is sufficient
Figure 4.13

n is the permeability reduction factor for non-woven geotextiles 090 < 0.5d10 then n = 1 090 > 0.5d10 then calculate the value of:

If If

where: n = fabric porosity t = fabric thickness (m) under pressure of 2kN/m2 090 = fabric pore size (m) at the 90% ner level Determine the reduction factor n from Figure 4.4, as shown in section 4.2.2. 49

Figure 4.14

3 Determine the Required Armouring This design relates to loose dumped rock on a bedding layer. There are linked precast systems available that do not require bedding layers. A) ARMOURING AGAINST WAVE ACTION For the anticipated wave height read off the required armour weight, W tonnes, at the appropriate slope angle from Figure 4.15. Note the graph relates to the following typical properties: s Gs K Specic weight of rock = 2.73 tonnes/m3 Specic gravity of rock = 2.73 Damage coefficient = 3.20 i.e. no damage or overtopping Slope angle Figure 4.15 is based upon the Waterways Experiment Station formula for breakwaters, which in turn is based on the work of Iribarren and Hudson (1959). Comparison of the various formulae shows wide variations in armour stone weights for very steep or very at slopes. Therefore, the results should be used with caution in large scale projects or on locations with unusually severe wave action.

Figure 4.15

50

B) ARMOURING AGAINST WATER FLOW For the anticipated water velocity read off the required armour weight, w kg, at the appropriate ow from Figure 4.16. The equivalent stone diameter:

where: w = armour weight in kg The armouring layer should be laid in two courses to a minimum thickness of 2De. 4 Determine the required bedding This layer is not always necessary, e.g. where hand pitching or precast systems are used. The main purpose of the layer is to provide a bridge between armour and lter and protect the latter from damage. Size (bedding) d100 Thickness (bedding) < 0.5 De (armouring) De (armouring)

Figure 4.16

4.4 MODEL SPECIFICATION


For erosion control: The geotextile shall be manufactured from durable synthetic polymers. The minimum tensile strength shall be....kN/m at a strain of..... % determined in accordance with EN ISO 10319. The water permeability normal to the plane of the geotextile shall not be less than.....m/s determined in accordance with EN ISO 11058. The O90 pore size shall be in range.....microns to.....microns determined in accordance with EN ISO 12956. Please refer to pages 76 78 to determine appropriate Lotrakgrade required.

51

4.5 SOIL EROSION


Soil erosion is not normally associated with civil engineering activity; however, the fact is that the stripping of natural groundcover such as grass or forest can cause an enormous increase in rates of erosion. This is reected in the following Table 3.2, which relates to a 1:14 slope in silty loam. Relationship Between Annual Soil Loss and Ground Cover Vegetative Cover Forest Grass Crops None
Table 4.2

Soil Loss (tonnes/hectare) 0.01 0.04 40.00 240.00

The prime agent in soil erosion is rainfall. The impact of raindrops on bare soil will detach soil particles and subsequently transport them down slope by a combination of rain splash and overland run-off. Where high run-off velocities are manifested there will be additional detachment of transportation of soil particles. Soil loss will be a function of soil erodibility and the erosive power of the rainfall, as well as the length and steepness of the soil slope. For a given set of these parameters, the only ameliorating factor is the protection given to the bare soil. The relationship between these factors can be quantied in terms of the simple Universal Soil Loss Equation: L = P . R.E.S where L P R E S = = = = = soil loss per annum (tonnes/hectare) a protection factor P < 1, except for bare soil where P = 1 rainfall erosive power soil erodibility slope length and steepness

For a given location all of the above factors, with the exception of the protection factor, P , are xed. The objective is to provide cover that gives an appropriately low value of protection factor and, therefore, limits the annual soil loss to an acceptable level. The various parameters involved are usually determined by direct observation for large scale soil conservation programmes (Morgan 1986). However, this is not generally the case for civil engineering works where standard topsoil and grass seeding techniques are adopted. Grass cover, particularly in temperate regions, provides adequate protection. However, for this protection to be effective, the grass must rst be established. Until this time the soil is still bare and subject to erosion. It should be remembered that once the bare slope is seeded the erosive elements could also affect the ungerminated seed, which may be transported to the toe of the slope. Short term slope protection and the establishment of a protective sward can be assisted by the use of natural bre or synthetic bre matting placed over the slope. Depending on the product used, topsoiling and seeding may precede or follow on after placing the matting. In either case appropriate matting can offer three broad advantages. Firstly it will give short term protection against erosion by absorbing kinetic energy of raindrops which would otherwise detach soil particles. Secondly it will impede run-off and thereby reduce the ability of the run-off to transport detached soil particles. Thirdly it will retain heat so generating a microclimate conducive to grass seed germination and growth. Matting and meshes can also be used to advantage to reduce soil erosion in small watercourses where the prime cause of erosion is owing water. Grass lined water courses can withstand signicant ow velocities, of the order of 2 m/s (Whitehead 1976). Improvements on this performance may accrue where matting similar to that employed for slope erosion control reinforces the grass root system. This topic has been the subject of research by CIRIA (Hewlett et al. 1987).

52

4.6 EXAMPLE
A river is subject to laminar ow in one area where the water velocity rises to a maximum of 3 m/s. The riverbank consists of a silty sand with the following properties: d10 d50 d60 d90 = = = = 0.06 mm 0.20 mm 0.25 mm 0.80 mm

and approximate permeability, ks = 1 x 10-5 m/s. The bank slope is about 1:2 . What are the characteristics required for the geotextile, the bedding and the armour?

a) Geotextile Design
Non cohesive soil under static conditions:

2.5 x d50 = 0.5 d90 = 0.8 Therefore, maximum O90 is 0.5 mm Select appropriate geotextile: Woven, with a permeability, kg = 2 x 10-4 m/s From Figure 4.3: Reduction Factor, w = 0.09 Therefore, w.kg = 0.09 (2 x 10-4) = 1.8 x 10-5 m/s Hence, geotextile is OK as permeability in-situ, w.kg, is greater than soil permeability, ks, at 10-5 m/s.

b) Armouring Design
From Figure 4.16, an armour stone weight for laminar ow at a velocity of 3 m/s can be determined. Equivalent Stone Diameter,

Therefore, total thickness > 710 mm

c) Bedding Layer
Size Thickness d100 < 0.5 x 355 < 178 mm = 355 mm

53

Solution Summary:
Geotextile: Woven O90 kg

500 microns 2 x 10-4 m/s minimum

Bedding layer:

200 mm down stone Minimum thickness 355 mm 20 kg stones Minimum thickness 750 mm

Armour:

4.7 TYPICAL LOTRAK EROSION CONTROL GEOTEXTILES


Lotrak Geotextile Pore Size O90 ner than (microns) EN ISO 12956 Water permeability (m/sec x 10-3) EN ISO 11058 Tensile Strength (kN/m) EN 10319
Table 4.3

4000 400 26 35 30

25R 250 12 25 25

50R 250 16 52 50

70R 225 16 72 72

warp weft

4.8 REFERENCES
Heerten, G. Functional Design of Filters using Geotextiles, Proceedings 3rd International Conference on Geotextiles, Vienna, 1986, Vol.4. Hewlett, H.W.M., Bootman, L.A. & Bramley, M.E. Design of Reinforced Grass Waterways, CIRIA Report 116, 1987. Iribarren & Hudson, R.Y Laboratory Investigation of Rubble Mound Breakwaters, Proceedings of the Waterways and Harbours Div., A.S.C.E., September 1959. Morgan, R.P.C. Soil Erosion and Conservation, Longman Scientic and Technical, London, 1986. Mouw, K.A.G. et al. Geotextiles in Shore and Bottom Protection Works, Proceedings 3rd International Conference on Geotextiles, Vienna, 1986, Vol.2., 349-354. Pilarczyk, K.W. Discussion on Revetment Design, Proceedings of the International Conference on Flexible Armoured Revetments Incorporating Geotextiles, London, 1984, Thomas Telford, 209-215. Pilarczyk, K.W. Prototype tests on Slope Protection Systems, Proceedings of the International Conference on Flexible Armoured Revetments Incorporating Geotextiles, London, 1984, Thomas Telford, 239-254. Saathoff, F. Examinations of Long Term Filtering Behaviour of Geotextiles, Durability of Geotextiles, Chapman & Hall, London, 1988. Whitehead, E. A guide to the Use of Grass in Hydraulic Engineering Practice, Technical Note 71, CIRIA, London, 1976. Tutuarima & Van Wijk Prole Mattresses - An Alternative Errosion Central System, Proceedings of the International Conference on Flexible Armoured Revetments Incorporating Geotextiles, London, 1984, TTL.

54

5.0 SOIL REINFORCEMENT


5.1 INTRODUCTION
Practice is now established in using geotextiles and geogrids in three principal soil reinforcing applications: i) Embankments over soft ground. ii) Vertical walls and steep-sided ll slopes. iii) Reinstatement of slips in cuttings and embankments. Design techniques for embankments, walls and slopes are illustrated through the presentation of design examples and theorectical aspects of design are expanded in later sections. The vast majority of civil engineering structures and earthworks are designed always to subject the soil to compressive loading. A prime example of this is a retaining wall where both the vertical and horizontal stresses are compressive. This compressive loading regime is reected in standard laboratory tests. For example the triaxial compression tests involve a vertical compressive stress and a radial compressive stress. Despite this compressive loading regime tensile strains are still developed. In the triaxial test these strains are radial and account for barreling in the sample. One way of reducing these tensile strains, and so increasing the strength of the sample, would be to increase the externally applied radial stress. A more discrete way of achieving a similar effect would be to implant tensile reinforcement in the form of horizontally inclined discs of reinforcing material placed at suitable vertical centres. Provided that such reinforcement is sufficiently rough, strong and inextensible, then through the agency of soil-reinforcement friction the reinforcement will reduce tensile soil strains and increase strength. The compatibility of soil and reinforcement strains is the basis of design of any reinforced soil structure.

5.2 EMBANKMENTS
Any embankment consists of two parts:- The embankment ll which may be selected by the engineer and placed to give specied engineering properties and the foundation soil on which the embankment is constructed. If the foundation soil has inadequate strength to carry the embankment then its properties may be improved by consolidation or other geotechnical strengthening processes. In extreme cases the foundation soil may be removed and replaced by more suitable soil. An alternative now available to the engineer is the use of geotextile reinforcement. The following analytical techniques apply to embankments on weak ground such as soft and very soft clays of medium to high compressibility where settlement will not exceed 20% of initial embankment height. On highly compressible soils such as organic clays and peats other methods of analysis are required. Consideration is limited here to using a full width basal layer of geotextile to reinforce the embankment and raise the short/ intermediate term factor of safety to an acceptable level. The required factor of safety to be selected by the designer. A rst stage in the analysis is to determine an embankment geometry to give a satisfactory long term factor of safety. This is achieved using conventional methods of analysis such as the Bishop Routine Method. A second stage is to analyse short term stability to determine what forces are required for equilibrium. Simple force equilibrium methods are employed and three main aspects of stability are investigated: i) Internal stability. ii) Lateral stability. iii) Deep seated rotational stability.

Internal Stability
This involves assessments of the stability of the embankment ll resting on top of the basal reinforcement which extends from one toe of the embankment to the other. For simplicity the ll is, assumed to be cohesionless and to have zero porewater pressure. Analysis is carried out in terms of cv the angle of shearing resistance at constant volume. Typical failure modes are indicated in Figure 5.1. These involve substantially rotational movement in which the lower section of the failure surface may be at the interface of the ll and basal 55

reinforcement. The reinforcement may prove to be a plane of weakness if the angle of ll-geotextile friction, S, is substantially less than cv. Conventional analytical techniques such as circular and non-circular methods can be employed. Provided S>23cv failure along the ll-geotextile interface is not generally critical. In this case the stability of the ll slope can be rapidly assessed using innite slope analysis where for a 1 : n batter FF = n.tancv.

Figure 5.1 Internal Stability

Lateral Stability
Two aspects of lateral stability must be considered. a) Lateral sliding of ll over geotextile. b) Lateral extrusion of soil beneath geotextile. Both components will induce axial tension in the basal geotextile layer and these components are additive. The various forces involved are shown in Figure 5.2 for an embankment resting on a clay foundation soil with an undrained shear strength cuo at the surface. In this manual the undrained shear strength is taken to increase linearly with a depth at a rate kN/m2 per metre depth. Consequently the undrained shear strength at any depth, cu, is a function of the depth Z below original ground level such that cu = cuo + Z. Other distributions of undrained shear strength with depth may be encountered in practice and the effects of these on stability must be analysed accordingly.

Figure 5.2 Lateral Stability

A) LATERAL SLIDING The disturbing force may be taken as the simple active thrust developed in the ll. This is 12KaH2 where for a dry cohesionless ll Ka = (1-sincv)/(1 + sincv). This force must be resisted by the geotextile which is put into tension. The tension is transmitted from the ll to the upper surface by friction. If the angle of ll geotextile friction is then the bond resistance is 12 H2tan and the factor of safety with respect to bond failure is n.tan.Ka. This factor of safety is usually large and would only be realised if the geotextile was able to resist the tensile load involved in mobilising this high bond stress. Consequently it is more economic to apply a lower factor of safety against lateral sliding, Fs, governed by tensile failure. This leads to the need for an available geotextile force Fs x 12.Ka. H2. Of course a check should always be made to ensure that the available bond force is greater than the available tensile force so conrming that bond failure is not critical.

56

B) LATERAL EXTRUSION As an embankment is constructed over weak foundation soil so the soil will compress vertically downwards and laterally outwards. The horizontal forces inducing lateral extrusion will be resisted by passive pressure developed in the soil near the toe of the embankment and horizontal shear stresses mobilised beneath the batter and shoulder of the embankment. A simple analysis can be carried out by assuming that these horizontal shear stresses act on a potential failure surface at some depth Z and on the underside of the geotextile where the shear stress is limited to .cuo. The length of the horizontal plane on which these restoring shear stresses act is nH and consequently lateral extrusion considerations may govern the required side slope 1 : n. A simplied distribution of forces acting on a block of foundation soil is shown in Figure 5.2. If the factor of safety against extrusion is dened as the ratio of horizontal restoring forces to horizontal disturbing forces, then at any generalised depth the factor of safety, FE, is:

FE =

cuo nH [4 + (2z + H) + (1 + )] cuo z H

For the undrained shear strength distribution with depth given in Figure 5.2, the minimum factor of safety occurs at a critical depth:

Zc =

nH (1 + ) Cuo 3

If this factor of safety falls below the required minimum for a given embankment geometry then either the side slope must be attened by increasing n or reducing the height. The maximum available bond force on the underside of the geotextile will be cuo.nH/FE, where is an adhesion factor determined by laboratory testing. For this bond force to be developed the required geotextile force is TE = cuo.nH/FE. The available geotextile force, for a factor of safety FE, should be FExTE. This force must be added to the available geotextile force to resist lateral sliding, i.e., FSxTS = FSx12KaH2. Consequently the total force required for lateral stability is FE TE + FS TS. The simplied method above should only be used to make an initial assessment of stability with respect to extrusion. More detailed analyses are necessary to determine how the required force TE varies with distance along the base of the embankment. This can be estimated using modied conventional circular slip analysis in which the embankment section is assumed to be cracked, Figure 5.3(a), Jewell (1987). Each trial circle will render a value of TE which will be found to vary with the distance x measured from the toe of the embankment, Figure 5.3 (a). As with conventional circular slip analysis many trial circles must be employed to determine the maximum value of the required force TE, Figure 5.3 (b). The envelope of required force, Figure 5.3 (b) must then be compared with the available force. The available force may be limited by the bond force developed at the soil geotextile interface or the available tensile force in the reinforcement.

Figure 5.3 (a)

Figure 5.3 (b)

57

The available bond force will vary with the distance x and will have a value x cuo/FE. To prevent extrusion failure by bond the locus of available bond force must fall above the envelope for required force, TE, Figure 5.3 (c). It must be remembered that the theoretical bond force can only be realised if the available tensile force in the geotextile equals or exceeds TE. When a factor of safety FE is applied the available geotextile force, at the end of construction, must not be less than FETE. To account for lateral stability with respect to sliding of the ll over the geotextile, then the available geotextile force FETE must be increased by an amount FSTS as described previously.

Figure 5.3 (c)

Figure 5.3 (d)

Deep Seated Rotational Stability


Without reinforcement the embankment will have an inadequate short term factor of safety against deep seated rotational failure. Indeed the reason for installing basal reinforcement is to raise the short and intermediate term factor of safety to an acceptable level. This can be achieved by installing a basal layer of geotextile which generates a force TR. The tangential component of this force is TR.sin which leads to an incremental increase in restoring moment M=R.TR.sin, Figure 5.4. The objective is to determine the maximum required value of TR for a required factor of safety, FR, against rotational failure. As with conventional circular slip analysis, this factor of safety applies to the shear strengths of the ll and foundation soil. For a given trial slip circle the required value of TR can be determined directly using modied Bishop circular slip analysis, Ingold (1986).

Since the required value of FR is known in advance the above solution for TR is not reiterative. As with conventional circular slip analysis a search must be made to determine the maximum value of TR. Checks should be made to ensure that there is adequate bond strength. This should be checked for the length of reinforcement on either side of the point where it is intersected by the slip circle. Generally a more critical condition applies beneath the batter and shoulder of the embankment. The bond force may be taken as the product of the bond length, L. Figure 5.4, and the shear stresses acting on both sides of the geotextile, i.e., L (Wtan + cuo).

Figure 5.4 Deep Seated Rotational Stability

58

It should be noted that the value of TR needed to give a required factor of safety FR will decrease with time as the foundation soil consolidates and gains strength. Since TR is a function of time, t, it should be denoted TR(t). Once a certain degree of consolidation is achieved in the foundation soil the factor of safety FR is achieved without the aid of geotextile at which time TR(t) = 0. Between this time and the end of construction when TR has a maximum value TR(t=0) the generalised value is TR(t) which can be determined from simple one dimensional consolidation theory, Ingold (1986b).

t TR(t) = TR(t = 0) (1 + )
where: TR(t=0) is the maximum required tensile force at the end of construction (kN/m) t is the time measured from the end of construction (years) is the time from the end of construction when TR(t)= 0 (years)

d2 ru(max) - ru(0) 4cv ru(max) - ru(min)

d is the length of the vertical drainage path (m). cv is the coefficient of consolidation (m2/ year) ru(max) is the foundation porewater pressure ratio at the end of construction using u = 0 analysis. ru(min) is the foundation porewater pressure ratio at the end of primary consolidation. ru(0) is the foundation porewater pressure at which TR(t) = 0 and the factor of safety is FR. The variation of required force with time can be plotted in the form of TR(t) vs t using a logarithmic time scale, Figure 5.8. For stability with respect to rotational failure the available geotextile force must at all times exceed TR(t). A more complicated analysis can be carried out to take account of the consolidation which occurs during construction. This will lead to a lower required force TR(t) however, to be meaningful such an analysis requires high quality data for soil consolidation characteristics and detailed knowledge of the construction programme. The additional expense involved in such analysis may be more than the materials cost involved in over-design resulting from a simple design approach. Consequently, an economic appraisal should be made before embarking on a more complicated design approach.

TR(t)

log time (t) years


Figure 5.5 Variation of TR(t) with time after end construction

59

Geotextile Selection
Geotextiles used as basal reinforcement must lll two basic criteria: a) There must be no tensile failure leading to instability. b) Required forces must be mobilised at geotextile strains compatible with embankment serviceability limits. To avoid tensile failure the available geotextile force must at all times equal or exceed the required tensile force Tr(t) Figure 5.7. In mobilising this force the geotextile must not suffer unacceptably large axial tensile strains leading to a serviceability failure in the embankment. The ability of a geotextile to meet these two requirements can be assessed using stress relaxation or isometric creep plots. Stress relaxation data are gathered by submitting a geotextile to a constant tensile strain and measuring how tensile force decays with time. In contrast an isometric creep plot is obtained from a family of strain vs time plots for various load intensities, Ingold (1986b). Both techniques will render a plot of tensile load against log-time for a given constant strain, Figure 5.6, however, for a given geotextile the relationship between load and time may vary according to the test method, e.g., stress relaxation or creep loading tests.

Figure 5.6 Variation of load at constant strain and creep rupture load with time

For embankments on soft ground it would be expected that serviceability would be maintained provided the geotextile can provide the required restoring force at an axial tensile strain of 5% or less. There will be other sources of strain induced by construction. It is not possible to calculate these, however, an assessment can be made of the axial tensile strain, s, induced by a settlement, P , at the centre of the embankment.

s = [1 + (/nH)2 - 1] x 100%
Unless the settlement is particularly high, e.g. /H>0.2, or the embankment side slope is particularly steep, e.g. n<2, then the average settlement-induced strain would not be expected to exceed 0.5% However, it is prudent to ensure that the sum of the strain induced in providing the required tensile restoring force and the settlement induced strain does not exceed 50% of the failure strain of the geotextile at any time. Provided the axial tensile strain in the geotextile does not exceed any critical threshold tensile failure is unlikely to occur. However, to gauge the margin between available tensile load and the failure tensile load it is useful to have data on the variation of rupture load with time. Such information can be gathered from a series of constant load creep tests with samples loaded at different intensities. The higher the load intensity the shorter the time to failure. By recording the times at which tensile rupture occurs at different load intensities it is possible to construct a locus of creep rupture loads as shown in Figure 5.6. Data on creep rupture and stress relaxation are generally obtained by laboratory testing of undamaged samples in a neutral environment at constant temperature. Such conditions are unlikely to prevail on site. Consequently the geotextile is likely to be damaged by the construction process, e.g., tears and perforations, and may be exposed to aggressive environments such as highly alkaline ll. In addition temperatures may rise 60

above those used for laboratory testing. This may occur in lls subject to spontaneous heating. Any one of these factors is likely to reduce the tensile load which the geotextile can support without distress. It is important that the total magnitude of these effects are known. In most cases manufacturers will be able to supply base data for consideration. Fortunately, many forms of environmental attack such as chemical or microbial are time dependant and may, therefore, not be signicant on the time scale involved. In contrast construction induced damage will have an immediate effect. Consequently in assessing the available geotextile force reduction factors must be applied to the data from laboratory tests performed on undamaged samples in a neutral laboratory environment. Some indication of the reduction in the short term strength of a lightweight woven fabric is given in Table 1, Koerner et al (1987). Similar damage may occur if wick drains are installed through the geotextile.

Damage Hole Diam. (mm) 25 50 75

Strength* Reduction % 11 24 38

*Based on 200mm wide sample


Table 5.1 Effects of damage

Having made assessments of the available geotextile force with time at constant strain and ultimate geotextile force with time, these data are plotted and compared with the required force TR(t) Figure 5.7. Provided available forces mall times exceed required forces and that checks have been made to ensure that no form of bond failure is critical then stability and serviceability should be maintained.

TR(t)

Figure 5.7 Available and Required Forces

If any drainage is assumed to take place cross the geotextile, care should be taken to check that the water permeability of the geotextile is adequate. For example if an aggregate basal drain is constructed to accelerate porewater pressure dissipation in the foundation soil the permeability of the geotextile should be at least ve times that of the foundation soil.

Embankments on Soft Ground


The short term stability of embankments on soft ground is controlled largely by the undrained shear strength of the foundation soil. Where construction is rapid the embankment is at its most unstable at the end of the construction period. During and following construction there will be an increase in strength of the foundation soil which ceases once the foundation soil is fully consolidated. This is the long term situation where the stability of the embankment is assessed using drained shear strength parameters. In general an embankment of a given geometry will have a higher factor of safety in the long term than in the short term. Consequently the geometry of an embankment consistent with long term stability may not be consistent with short term stability. 61

Input Information: Embankments


This information has to be supplied by the engineer to establish design parameters. i) Embankment height H (m) and crest width C (m). ii) Shear strength parameter for the ll: cv (assuming dry cohesionless ll). iii) Unit weight of ll (kN/m3) iv) Short term shear strength parameters for foundation soil, cu (Kn/m2), including variations with depth. v) Long term shear strength parameters for foundation soil, c (kN/m2), (degrees). vi) Unit weight of foundation FS soil (kN/m2). vii) Coefficient of consolidation of foundation soil Cv. (m2/year). viii) Location of initial groundwater table in foundation soil. ix) Details of any external loading or surcharge. x) Geotextile properties including soilgeotextile adhesion factor, , ll-geotextile friction angle , and variation of creep rupture strength with time.

Design Procedures: Embankments


i) Determine embankment geometry to give adequate long term factor of safety with no reinforcement. Use long term soil strength parameters c, , and long term porewater pressure distribution assuming full primary consolidation. Re-analyse embankment geometry determined in (i) to determine short term factor of safety. Use short term soil strength parameters, cu, for foundation soil. Determine maximum required tensile restoring force at the end of construction, using u = 0 analysis, based on considerations of internal stability, lateral stability and deep seated rotational stability. Determine how required tensile restoring force decays with time due to consolidation of foundation soil. Select basal geotextile reinforcement which provides required tensile restoring at all times consistent with an acceptable strain being developed in the geotextile.

ii) iii) iv) v)

Design Example
Consider the embankment geometry and soil properties shown in Figure 5.8. Using conventional circular slip analysis, such as the Bishop Routine Method, the long term factor of safety against deep seated rotational failure is 1.4, but in the short term the factor of safety is only 0.7. The short fall in short and intermediate term factor of safety can be compensated for by a basal layer of geotextile reinforcement. Supporting theory is provided in later sections. Three main aspects of stability must be investigated: i) Internal stability. ii) Lateral stability. iii) Deep seated rotational stability. i) INTERNAL STABILITY: Assume a side slope of 1: n with n = 2.5 and check stability of unreinforced ll slope: FF = ntancv FF = 2.5 x tan30 = 1.4 OK.

Figure 5.8 Embankment Design Example

62

ii) LATERAL STABILITY: a) Check lateral extrusion of foundation soil using circular slip analysis assuming cracked embankment section. Maximum required restoring force, TE = 94 kN/m. Available bond force, xcuo/FE = 94 kN/m.....OK Required geotextile tension with a factor of safety FE = 1.2* x 94 = 113 kN/m. *NB. In this example the factor of safety against extrusion, FE, has been given a value of 1.2. Larger factors of safety can be obtained using a atter slope. b) Check lateral sliding of ll over geotextile. Required restoring force TS = 12KaH2 = 83 kN/m. Available bond force, 12nH2tan/FS = 243 kN/m*.....OK. Required geotextile tension to provide restoring force with factor of safety, FS of 1.2 = 100 kN/m.....OK. *NB. The factor of safety against bond failure is (1.2 x 243)/83 = 3.5. To provide adequate tensile resistance the required geotextile tension would be 292 kN/m. This is excessive hence the available geotextile tensile force is reduced to 100 kN/m thus reducing the factor of safety against sliding to 1.2. c) Determine total required tensile force to resist lateral sliding of ll and lateral extrusion of foundation soil with a factor of safety of, say, 1.2. 1.2 (TE + TS) = 1.2 (94 + 83) = 212 kN/m. iii) DEEP SEATED ROTATIONAL STABILITY: Check maximum required restoring force against deep seated rotational failure. Carry out circular analysis taking factor of safety, say 1.2, on soil and ll shear strength parameters: This gives TR(t ) = 340 kN/m. Check minimum pull-out resistance, Wtan + cuoL = 595 kN/m..... iv) REQUIRED DESIGN FORCE: Total required force to resist lateral instability = 212 kN/m. Total required force for deep seated rotational stability = 340 kN/m. In this example deep seated rotational stability is the worst case. Determine variation of TR(t) with time:

t TR(t) = TR(t = 0) (1 - ) TR(t = 0) = 340kN/m = d2 ru(max) - ru(0) 4cv ru(max) - ru(min)

d = 4.5m (drainage path length taken as depth to base of slip circle). Cv = 10m2/year. ru(max) = 0.70 (t = 0 and unreinforced FS = 0.7). ru(min) = 0.27 (t = 0 and GWT at OGL). ru(o) = 0.47 (t = w and unreinforced FS = 1.2).
Figure 5.9 Variation of TRftf with time

= 0.84 years (10 months). TR(t) = 340 (1 - t/0.84) kN/m. Plot TR(t) against log-time, Figure 5.9.

63

v) MODEL SPECIFICATION Clause XYZ: Geotextile

beneath embankment point A and point B.

The geotextile shall be manufactured from durable synthetic polymers. Under a constant axial tensile strain of.....% at a temperature of.....C the geotextile shall have the minimum load vs time characteristics shown in Figure 00. Due account must be taken of any constructioninduced damage caused by using ll of the grading shown in Table 00, and the construction plant indicated in Table 00. The proposed construction method involves..... The polymer of the geotextile shall be resistant to all forms of environmental attack for a period of up to..... months/years or due allowance made for environmental attack. Samples of the ll and foundation soil can be made available for chemical analysis. The maximum pH value for the ll is..... The tensile rupture strain in the geotextile, under operational conditions, shall at no time during the service life of.....months/ years fall below a value of.....%. The angle of ll-geotextile friction under normal stresses in the range.....kN/m2 to.....kN/m2 shall not be less than..... when tested under consolidated drained conditions in a 300mm x 300mm shear box. The soil-geotextile adhesion factor under normal stresses in the range ...kN/m2 to.....kN/m2 shall not be less than.....When tested under unconsolidated-undrained conditions in a 300mm x 300mm shear box the minimum permeability of the geotextile shall be m/sec measured at a head lost of 50mm across the geotextile and a maximum normal stress level of... kN/m2.

5.3 WALLS AND SLOPES


Unlike embankments on soft ground the tensile forces required for equilibrium remain constant during the design life of the structure. The design life may be a few months or years for temporary structures up to 120 years for permanent structures. For the latter class of structure, compliance with Department of Transport recommendations is advised. Design equations and theoretical aspects of design are expanded in later sections.

Walls and Slopes


In designing reinforced soil walls and slopes, consideration must be given to both external and internal stability. To evaluate external stability the reinforced mass may be treated as monolithic in which case conventional design methods may be employed. Checks must be made to ensure adequate margins of safety against failure by: i) ii) iii) iv) v) Forward sliding on the base of the wall/ slope. Overturning about the toe of the wall/slope. Bearing capacity failure. No tension to be developed along base of reinforced mass. Overall stability against rotational failure.

In considering internal stability there must at all times during the design life of the structure be adequate margins of safety against failure by: i) Tensile rupture of the reinforcement. ii) Pull-out or bond failure of the reinforcement. In addition there will be a serviceability requirement which limits the strain in the reinforcement and the deection of the wall or slope.

64

For permanent walls, and bridge abutments, reference should be made to the Department of Transport Technical Memorandum BE3/78 which was revised in 1987 to include polymeric soil reinforcement. However, the following gives an approach to assessing internal stability for both temporary and permanent structures. i) FILL/REINFORCEMENT STRAIN COMPATIBILITY Both the ll and the reinforcement must strain to mobilise their respective strengths. Assuming the ll to be cohesionless its peak strength, dened by the effective internal angle of shearing resistance, , will be fully mobilised at lateral strains of about 1 %. As lateral strains exceed this value the mobilised strength of the ll decreases until it reaches a constant minimum value known as the constant volume angle of shearing resistance, cv. This strength can be relied on even at large lateral strains. In contrast the axial tensile strain at which polymeric reinforcement, such as geotextiles or geogrids, rupture will be much larger, typically in excess of 10%, and as such will be about ten times larger than the 1% strain at which the ll is likely to mobilise its peak strength. Consequently when considering tensile failure of the reinforcement it is advisable to calculate design forces generated in the reinforced ll mass using cv.

ii) TENSILE RUPTURE OF THE REINFORCEMENT Both vertical walls and steep sided embankments rely on the continued tensile integrity of the soil reinforcement for stability. The tensile forces required to maintain stability in a wall or slope subject to dead load only remain constant throughout the design life of the structure. Additional forces due to line load will vary with time, however, for the purposes of design the worst expected loading should be used to calculate the required tensile forces which should be assumed to remain constant with time. At all times during the design life of the structure the design strength of the reinforcement must be greater than, or at least equal to the design force. iii) DESIGN FORCES: WALLS An idealised wall cross-section is shown in Figure 5.10.

Figure 5.10 Idealised Wall Cross-Section

The wall is reinforced with a horizontal layer of geosynthetic (geotextile or geogrid) at a constant vertical spacing S. The wall is of height H and has a uniform surcharge of intensity qo. The total length of each reinforcing layer is L. Selected ll is used within the reinforced zone behind which is common ll which exerts a lateral thrust 12KIaH2. The tensile force induced in a layer of reinforcement at some depth z beneath the top of the wall is taken as: Tz = hS

65

where h is the horizontal effective stress at the wall face at depth z and S is the area, per metre run of wall, served by the layer of soil reinforcement. This is numerically equal to the soil reinforcement spacing S. The horizontal force at the face is made up of up to ve components. Shz = S(Kaz+Kav+Kaqo+hp+ hl) where: S is the vertical reinforcement spacing Ka v is the coefficient of active earth pressure in the selected ll based on Ocv. is the unit weight of the selected ll or common ll. is the vertical stress at the face induced by the lateral thrust of the common ll. = Kaz3/L2 where Ka relates to the common ll. is the uniform surcharge. is induced by one, or any number, of line loads and is derived from Figure 5.11(a). is induced by one, or any number, of line loads and is derived from Figure 5.11(b).

qo hl hp

It is not recommended that geosynthetic reinforced walls with wrap-round facings be designed to take horizontal loading at the top of the wall. The objective is to determine TZ for each layer and to ensure that this does not exceed Tpermissible, where Tpermissible relates to the design life of the wall. In many cases the self weight of the ll will dominate. This is clearly the case when hp = hl = 0. For this case the maximum tensile force will occur at the base of the wall and have a magnitude Tmax.

Tmax = SKa(qo + h 1 +

K|aH2 ) L2

(a)
Figure 5.11 Lateral Earth Pressures due to surface Load

(b)

A very rapid assessment of Tmax can be made using the following assumptions: i) qo << H ii) L = 0-8H III) common ll 30 KIa2 = 0.33

This leads to: Tmax = 1.5 SKaH For a given value of Tpermissible the above expression can be used to determine a maximum constant reinforcement spacing:

S =

Tpermissable 1.5KaH
66

It is not prudent, with soft wrap-round facings to use a reinforcement spacing greater than 500mm. This allows a rapid assessment of the required value of Tpermissible for a given wall height. Tpermissible (required) = 0.75 KaH In summary, for simple loading cases and a constant vertical reinforcement spacing tensile failure is governed by the force in the bottom layer of reinforcement. For complex loadings each layer of reinforcement must be checked to ensure TZ < Tpermissible. For high walls two different constant spacings may be used. One in the top half of the wall and a smaller spacing in the bottom half of the wall. iv) DESIGN FORCES: STEEP SIDED SLOPES For an unsurcharged vertical reinforced soil wall the maximum tension is assumed to be developed in the bottom layer of soil reinforcement. With a constant vertical reinforcement spacing S. Tmax = KaHS If the slope of the wall was to be slackened then for a cohesionless ll the value of earth pressure coefficient K decreases from Ka, when the wall slope is 90, to zero when = . Clearly there is a corresponding variation in Tmax For steep sided ll slopes on competent foundation soil the design procedure follows that for vertical walls. The same objectives apply, namely, to nd a value for Tmax, which must be less than Tpermissible, and to nd a value of L which obviates pull-out failure. Although it is rarely critical in vertical walls, the value of L should also be large enough to obviate overturning, forward sliding, or the development of tension along the base of the wall. The problem dened above can be solved rapidly using the design chart given in Figure 5.12. In all cases the ll is assumed cohesionless and to have no positive porewater pressure operating. The chart pertains generally to ll materials with cv in the range 20 to 40 in increments of 5. The chart gives values of the earth pressure coefficient K, plotted against slope angle , for a range of values of cv. Where a surcharge is required this can be modelled as a head of ll which is added to the true slope height. This gives an equivalent height which is used for design.

Figure 5.12

67

v) PERMISSIBLE REINFORCEMENT DESIGN LOAD The design forces generated within the reinforced mass are assumed to be constant throughout the design life of the wall or slope. In contrast the tensile strength of polymeric reinforcement will decrease with time. For example Figure 5.13 shows an idealised relationship between tensile creep rupture load and time. The axial strain at which creep rupture occurs in presently available geotextiles and geogrids will generally be greater than 10% For constant strains less than the rupture strain the load that the reinforcement can sustain will also decrease with time. This is illustrated in Figure 5.13 for isometric plots for constant strains of 10% and 5% based on constant load tests. Slightly different results may be obtained from stress relaxation tests. Consequently the permissible reinforcement design load may be based on creep rupture strength or, in a more creep prone reinforcement, it may be based on a limiting value of strain. For design purposes it is the lower of these two loads which is used as the basis of the permissible design load. Permissible reinforcement design loads based on creep rupture strength provide a margin of safety against collapse of the structure through tensile failure of the reinforcement during the design life. Permissible reinforcement design loads based on a limit in value of strain provide a margin of safety against the structure exceeding some predetermined serviceability limit.

Figure 5.13 Idealised Wall Cross-Section

vi) DETERMINATION OF PERMISSIBLE DESIGN LOAD The permissible design load of the reinforcement governed by tensile rupture must at all times during the design life of the structure be greater than or at least equal to the worst expected design force exerted by the ll and any superimposed loading. It is necessary to assess how the tensile rupture strength will decrease with time. This can be achieved by loading different samples of the reinforcement at different load intensities so that the times to failure fall inside a predetermined range of time at a standard test temperature of, say 20C. The test temperature should at least equal the maximum operational temperature of the reinforcement. Where the duration of the longest test is less than the required design life extrapolation of results can be facilitated by testing at higher temperatures, subject to certain limitations, to accelerate time. A very clear distinction must be drawn between undamaged control samples tested in a benign medium, such as air at constant temperature, and operational samples which will be damaged during construction and be subject to environmental attack through agencies such as water and chemicals or bacteria in the ll. For a product subject to strict quality control there should be little variation in the extrapolated characteristic strength from batch to batch. However, the degree of mechanical damage and aggressiveness of the ll will vary from ll to ll as might the operational temperature. It is vital that

68

tensile rupture tests are carried out on operational samples to determine how mechanical damage and environment will reduce the time-dependent rupture strength. This will allow the determination of various partial factors to be applied to the characteristic control strength to reduce it according to the nature of the particular ll to be employed. Ideally laboratory tests on operational samples, which have been predamaged should be carried out in an aggressive environment since the combined effects of environmental and mechanical damage may be synergistic. This means that the combined effects of environmental and mechanical damage may be greater than the sum of the effects of testing damaged samples in a benign environment and undamaged samples in an aggressive environment. To determine permissible design load the following minimum partial factors are suggested. These would be used to reduce the tensile rupture strength determined at a time equal to the design life of the structure. Those factors relating to the effects of mechanical damage and environment should be determined by exhaustive testing along the lines described above. a) Material factor: m This relates to the probability that the control strength of the soil reinforcement may occasionally fall below the specied characteristic strength. The suggested value of m is 1.2. b) Test data extrapolation factor: t This relates to the decreasing degree of condence in extrapolated data as extrapolation is made over increasing time intervals. No test data should be extrapolated at the design temperature without the aid of accelerated testing at appropriate higher temperatures. Ideally extrapolation should not exceed one cycle of common log-time, that is logarithmic time to the base 10. Extrapolation should never exceed two log-cycles of time. For n log-cycles of extrapolation where 1 < n < 2 the suggested value of t is 1.1n. Laboratory tests should be conducted at the design temperature which should equal the maximum operational temperature in the soil. For temperate climates a standard test temperature of 20C should be adequate. c) Construction induced damage factor: c This relates to the long term effect of mechanical damage suffered by the reinforcement during installation. Among other things it will be a function of ll type, layer thickness and type of compaction plant. The effects of mechanical damage should be assessed using long term tensile rupture tests such as those employed to assess the long term tensile rupture strength of intact and undamaged control samples. The minimum suggested value of c is 1.2. d) Environmental attack factor: e This relates to the long term affect of the ll environment on tensile rupture strength. Both chemical and bacteriological attack must be considered and their effects quantied by carrying out long term tensile rupture tests in an appropriate aggressive environment at the design temperature of higher temperatures as appropriate. The minimum suggested value of e is 1 1. e) Overall factor of safety: r The partial factors m, t, c and e are applied to the long term characteristic tensile rupture strength to reduce this to the basic design value. Where the ramications of attaining the ultimate limit state of tensile rupture of the reinforcement are more serious the basic design strength may be reduced by applying an overall factor of safety of r to arrive at the permissible design load. The suggested minimum values of r are in the range 1.0 to 1.2.

69

The design strength of the reinforcement for permanent structures is the 120 year characteristic tensile rupture strength, determined for intact control samples in a benign environment at the design temperature, divided by the partial factors m, t, c, e and r. Minimum values of these factors have been suggested. Actual values of c and e due to construction damage and environmental effects are product and ll specic and must be determined directly by long term testing. As tensile rupture test data are gathered over longer test periods, the uncertainty of extrapolation decreases and therefore t may be decreased as longer term data become available. It is recommended that permanent structures be designed in accordance with the requirements set out in Technical Memorandum BE/78 (Revised Edition 1987). Among other things this requires a British Board of Agrement Roads and Bridges Certicate for proprietary soil reinforcement. It is further recommended that design forces with respect to internal stability be based on cv and that test temperatures equal or exceed the maximum operational temperature of the reinforcement. Although not advocated here others have suggested that the permissible load under prolonged loading may be based on the short term tensile strength determined using constant rate of strain testing. In particular it has been suggested that the permissible load for polyester is of the order of 50% of the short term strength or, less than 25% for polypropylene and polyethylene, Van Zanten (1986), Leshchinsky and Perry (1987). If such an approach was adopted, for say short term temporary structures, then additional allowance would need to be made for effects of construction induced damage and environmental attack. vii) BOND FAILURE: WALLS The reinforced soil mass can be considered to comprise two zones:- an active zone near the front of the wall and a passive or anchorage zone remote from the face of the wall. These two zones can be considered to be divided by a planar failure surface, Figure 5.10. When any layer of reinforcement at depth z is subject to a tensile load TZ the length of the reinforcement Lb, Figure 5.10, embedded in the restraint zone must be long enough not to pull out. If the angle of ll/ geosynthetic bond stress is then the pull out resistance is 2ztanS.Lb since the shear stress ztan acts on both sides of the fabric. With a factor of safety Fb against bond failure it follows that:

Lb =

FbTz 2ztan

To determine the total reinforcement length a quantity (H z)tan(45 /2) must be added to Lb. Since tan(45 /2) is numerically equal to Ka this leads to:

Lb =

FbTz (H - Z) + 2ztan Ka

As a rule of thumb start with L = 0.8H. Also the value of Lb should never be less than 1 m. In theory use can be made of variable L. In practice it is simplest, and in the long run cheapest and safest, to set L to a constant value equal to the maximum value determined. Clearly a value of Lb and therefore L must be determined at each level. For simple loading cases the worst case occurs at the top of the wall, but rarely proves critical if the rule of thumb L = 0.8H is used. Bond failure can also occur at the front of the wall if the short wrap-round lap length Lo Figure 5.10, is inadequate. To a good approximation: Lo = 12Lb As before, Lo should not be less than 1m. viii) BOND FAILURE: SLOPES The total length of reinforcement in each layer will be governed by pull-out from the anchorage zone, outward sliding and the need to prevent the development of tensile stress along the base of the reinforced mass. Design curves have been produced by Jewell et al (1984) and these have been reproduced in Figure 5.10. Where wrap-round geotextile facing is used the lap length, Lo, at the front of the wall should be a minimum of 1 m for slopes up to 5m high. For higher slopes a longer lap length may be required.

70

ix) SERVICEABILITY The Highways Agency Memorandum BE3/78 (DTp 1987) requires strains in walls to be limited to less than 1% from the end of construction to the end of the design life of 120 years. A gure of 0.5% is quoted for abutments. Both gures apply to vertical walls constructed using rigid partial height facing panels such that each layer of the wall undergoes deformation prior to placing of subsequent layers. When full height facing panels are used and propped in a way that does not allow this early deformation to take place, the time span used for the strain limits is taken as 120 years from the start of construction. For various reasons calculations of wall movements and reinforcement strains generally lead to an overestimation and at present there is no proven method of calculation. Measured strains, at least in the short term, have been typically less than 1 for geogrids, Yamanouchi et al, 1986, Berg et al 1986, Carroll and Richardson 1986. For a steep sided embankment reinforced with woven polypropylene tape geotextile Du Bois et al (1982) reported mean strains in the range 0.8% to 2.8%. However, the majority of the strain occurred during placement and compaction of the ll with subsequent strains being very small. Provided a satisfactory alignment is achieved at the end of construction creep strains developing in simple temporary works retaining walls are not likely to give rise to serviceability problems provided there are adequate margins of safety against attaining ultimate limit states.

Input Information: Walls and Slopes


This information has to be supplied by the engineer to establish design parameters. i) Wall height H metres; ii) Surcharge loadings; iii) Constant volume angle of shearing resistance of selected ll cv; iv) Effective angle of shearing resistance of common ll ; v) Unit weights of ll kN/m3; vi) Maximum operational temperature of reinforcement; vii) Design life; viii) Parameters to assess external stability.

Design Procedure: Walls


In designing reinforced soil walls consideration must be given to the external stability and internal stability of the wall. To evaluate external stability the wall is treated as a gravity retaining structure in which case conventional design methods are employed. Checks must be made to ensure adequate margins of safety against failure by: 1. 2. 3. 4. Forward sliding on the base of the wall. Overturning about the toe of the wall. Bearing capacity failure. Overall stability against rotational failure.

To evaluate internal stability the geotextile should be designed to guard against: 1. Tensile failure of the geotextile. 2. Pull-out or bond failure of the geotextile. 3. Excessive elongation of the geotextile or deformation of the wall. Assuming a constant vertical reinforcement spacing S, the maximum design force, for an unsurcharged wall, is taken to occur at the base of the wall and can be approximated using the expression: Tmax = 1.5KaHS. By rule of thumb the length of the reinforcement, L, in each layer is approximated using the expression: L = 0.8 H.

Design Example: Walls


A 5m high wall with no superimposed loading is to be constructed with ll having cv = 40 to retain a 5m height of ll with = 30. Both lls have a unit weight of 20 kN/m3. What is the required permissible design load of the reinforcement, at the end of the design life of 3 years, and what is the total length of each reinforcement? The maximum operational temperature is 20C. 71

1. Rapid Assessment a) Maximum design force, Tmax: Assume a maximum reinforcement spacing S = 0.5m. Tmax = 1.5 x 0.217 x 20 x 5 x 0.5. Tmax = 20kN/m To prevent tensile failure the reinforcement should be able to carry a minimum permissible tensile load of 20 kN/m over a design life of 3 years at 20C. (b) Reinforcement length, L: L 0.8H L 4m. 2. Detailed check: By inspection the maximum tensile force, Tmax will be at the base of the wall.

Tmax = SKah 1 +

K|aH2 L2

= 0.5 x 0.27 x 20 x 5 Tmax = 20.5kN/mOK

1+

0.33 x 52 42

Use reinforcement with permissible design load 20.5 kN/m at 500mm vertical spacings. Now check critical value of Lb at top layer assuming

Fb = 2, = 30 Lb =
Now

2Tz 2ztan Lb = 0.27 (0.33 x 20 x 53) 42 2 x 1.4 2 x 20 x tan30

Tz = S [Kaz + Ka v] Tz = 0.5 0.27 x 20 x 0.5 + Tz = 1.4kN/m


L * = 3.3m (This is less than 0.8H = 4m).

Lb = 0.24m, say 1m minimum Lb = (H - Z) Ka + Lb = 4.5 x 0.27 + 1

*NB. If wrap-round facing is used then the cut length of geotextile must allow for the lap length Lo, and layer thickness S.cosec. By inspection overturning O.K. Assume rotational failure O.K. Check forward sliding taking = 30

Fsliding =

L x H x x tan 1 /2 x Ka x x H2 4 x 5 x 0 x tan30 = 1 /2 x 0.33 x 20 x 52

L = (H - 2) Ka + Lb = 4.5 x 0.27 + 1 Lb = 2 x 1.4 2 x 20 x 0.5 x tan30

Tsliding = 2.8OK
NB. In selecting a geotextile it is advisable to specify a water permeability which allows through drainage of any vertical inltration. Construction details to prevent any development of positive porewater pressure in selected ll or common ll. 72

Design Procedure: Slopes


As with reinforced soil walls consideration must be given to external and internal stability of the slope. The maximum design force, Tmax, is assumed to occur near the base of the slope and is approximated using the expression Tm = KHS The earth pressure coefficient will vary with I and slope angle as shown in Figure 5.14, for dry cohesionless soil. Jewell et al (1984). The length of the reinforcement, L, also varies with cv and , Figure 5.3, and takes into account bond failure, forward sliding and suppression of vertical tensile forces on the base of the reinforced mass. Separate checks must be made on overall stability.

Figure 5.14

5.4 DESIGN EXAMPLE


A dry cohesionless ll of unit weight 20 kN/m3 and cv = 30 is to be used to construct a steep sided embankment, 5m high, over competent foundation soil to a side slope of 60. Assuming a maximum vertical soil reinforcement spacing of 500mm, what permissible tensile strength is required of the geosynthetic reinforcement at the end of its design life? What is the total length of each soil reinforcement layer? Solution: Go to Figure 5.14 for = 60, cv = 30. This renders K = 0.18. Tpermissible = Tmax = KaHS = 0.18 x 20 x 5 x 0.5. Tpermissible = 9 kN/m. Go to Figure 5.14 to determine L/H. For = 60, = 30. This renders L/H = 0.7. For H = 5m this gives L = 3.5m. L = 3.5m. This leads to the use of 10 layers (0.5m vertical spacing over 5m) of geosynthetic with a minimum permissible design life strength of 9 kN/m each 3.5m long. (Note the cut length should be L + Lo + wrap-round = 3.5m + 1 m + 0.5 cosec 60 = 5.1 m).

73

5.5 MODEL SPECIFICATION


Clause ABC: Geotextile reinforcement to retaining wall/slope placed in layers as shown on the drawings. The geotextile shall be manufactured from durable synthetic polymers (Polypropylene, Polyethylene, Polyester, etc.,) and have the following properties. Under a constant axial tensile strain of.....% at a temperature of.....C, the geotextile shall sustain a minimum tensile load of.....kN/m for a period not less than.....months/years. Due account must be taken of any construction induced damage caused by using ll of the grading shown in Table , and the construction plant indicated in Table The proposed construction method involves..... The polymer of the geotextile shall be resistant to all forms of environmental attack for a period up to..... months/years or due allowance made for environmental attack. Samples of the ll and foundation soil can be made available for chemical analysis. The maximum pH value for the ll is..... The tensile rupture strain in the geotextile, under operational conditions, shall at no time during the service life of.....months/years fall below a value of.....%. The tensile rupture strength of the geotextile under operational conditions shall at no time during the service life of.....months/years fall below a value of.....kN/m. The angle of ll-geotextile friction under normal stresses in the range.....kN/m2 to.....kN/m2 shall not be less than..... when tested under consolidated drained conditions in a 300mm x 300mm shear box. The minimum permeability of the geotextile shall be.....m/sec measured at a head loss of 50mm across the geotextile and a maximum normal stress level of.....kN/m2. Please refer to pages 76 78 to determine appropriate Lotrakgrade required.

5.6 REMEDIAL SLOPE WORKS


At the design stage of highways it is impossible to accurately analyse and design every cutting slope. Consequently there are failures, especially in overconsolidated clays. In the past the procedure has been to remove all the slip debris, excavate below the slope and reinstate with free draining granular ll. Whilst this is a satisfactory solution from the technical viewpoint, it is expensive and can cause inconvenience since the slip debris has to be carted away using one lane of the highway as a haul road. If an attempt was made to reinstate the slope using the slip debris this would almost certainly result in future failure since in the process of slipping the soil becomes disturbed and often loses strength. With the advent of geotextiles there is now a more economic alternative. This is depicted in Figure 5.15. As shown in (a) the slip debris is initiated
Figure 5.15

74

and fails, (b) the slip and the slip plane must be excavated. The extent of the excavation is often governed by the bond length required for subsequent reinforcement. (c) Generally the excavated material can be stockpiled on the land above the cutting. This can of course only be done with permission of the land owner. Also the stockpile must be constructed so as not to cause a stability hazard to the intact parts of the slope. Once the temporary works excavation is complete a basal drainage blanket is laid. The downstream outlet must not be covered with topsoil or blocked in any way. Following this the slope is reinstated using the slip debris reinforced with layers of geotextile, (d), which wrap round the face of the ll. The design of such reinforcement is a matter for a qualied geotechnical engineer. Since the available bond length space is limited, it is often advantageous to employ a mesh or grid rather than a geotextile fabric as they can often generate a higher pull-out resistance.

5.7 TYPICAL LOTRAK REINFORCING GEOTEXTILES


Lotrak Geotextile Pore Size O90 ner than (microns) EN ISO 12956 Water permeability (m/sec x 10-3) EN ISO 11058 Tensile Strength (kN/m) EN 10319
Table 5.2

25R 250 12 25 25

50R 250 16 52 50

70R 225 16 72 72

warp weft

5.8 REFERENCES
Berg, R. R. Bonaparte, R. Anderson R. P., and Chouery, V. E., 1986. Design, construction and performance of two geogrid reinforced soil retaining walls. Proc. III Int. Conf. on Geotextiles, -Vienna, Vol. 12. Bishop, A.W. 1955. The use of the Slip Circle in the Stability Analysis of Slopes. Geotechnique, Vol. 5, pp. 7-17 or Proceedings Euro Conference on Stability of Earth Slopes, Vol. 1. Caroll, R. G., and Richardson, G. N. 1986. Geosynthetic reinforced retaining walls. Proc. III Int. Conf. on Geotextiles, Vienna, Vol. 2. Department of Transport 1987. Reinforced earth retaining walls and bridge abutments for embankments. Technical Memorandum (Bridges) BE3/78 (Revised 1987). Du Bois, D. D., Bell, A. L., and Snaith, M. 1982. A fabric reinforced trial embankment. Q. J. Eng. Geol. Vol. 15, No. 3. Ingold, T. S. 1986a. Analysis of geotextile reinforced embankments over soft clays. Jrnl. of Inst. Highways and Transportation, Vol. 33, No. 3. Ingold, T. S. 1986b. Short, intermediate and long term stability of geotextile reinforced embankments over soft clays. Proc. III Int. Conf. on Geotextiles, Vienna, Vol. 2. Jewell, R. A., Paine, N., and Woods, I. 1984. Design methods for steep reinforced embankments. Polymer Grid Reinforcement, pub. Thomas Telford Ltd., London. Jewell, R. A. 1987. Reinforced Soil Mechanics and Design. Short Course, Oxford University. Koerner, R. M. Hwu B. L. and Wayne, M. K. 1987. Soft soil stabilisation designs using geosynthetics. Geotextiles and Geomembranes Vol. 6 Nos. 1-3 Pub Elsevier Applied Science Publishers, London. Leshchinsky, D., and Perry, E. B. 1987. A design procedure for geotextile reinforced walls. Proc. Geosynthetics -87, Vol 1, IFAI. Van Zanten, R. V. 1986. Geotextiles and geomembranes in civil engineering. pub. A. A. Balkema, Rotterdam. Yamanouchi, T., Fukuda, N., and Ikegami, M. 1986. Design and techniques of steep reinforced embankments without edge supportings. Proc. III Int. Conf. on Geotextiles, Vienna, Vol. 1.

75

6.0 GEOTEXTILES DATA


The technical data given in this section refers to the range of LotrakGeotextiles available at the time of writing. The gures quoted are the mean of a number of tests. For further information please contact: Don & Low Ltd. Newfordpark House Glamis Road Forfar, Angus DD8 1FR Telephone: 01307 452200 Fax: 01307 452300 Website: www.lotrak.com Email: lotrak@donlow.co.uk The manufacturer reserves the right to vary product specications at any time and without notice.

76

G E OT E X T I L E S

standard grades
physical properties of Lotrak

Test

Standard

Advance

1800

2300

2800

4000

Tensile Strength (kN/m)

EN 10319

Warp Weft

12 9

12 12

20 17

23 22

35 30

Elongation at max. load (%)

EN 10319

Warp Weft

28 20

28 16

28 18

28 22

30 25

CBR Puncture Resistance (N)

EN ISO 12236

1650

1800

2500

2800

4000

Cone Drop Penetration (mm)

EN 918

17

15

12

12

Pore Size 90% ner than (microns)

EN ISO 12956

300

225

200

260

400

Water Permeability (m/sec)

EN ISO 11058

22x10-3

16x10-3

22x10-3

20x10-3

26x10-3

Effect of UV Light

T h e Po l y p r o p y l e n e u s e d c o n t a i n s a U V i n h i b i t o r

Weight (g/m2)

80

95

110

135

205

Roll Size

Width Length

4.5 100

4.5 100

4.5 100

4.5 100

5.2 100

The gures quoted above are the mean of several tests. Specic information on tolerances is available on request. Don & Low Ltd operate a policy of continuous improvement so attention is drawn to the issue date on this sheet. A potential specier/purchaser is advised to check that this information has not been superseded either by contacting Don & Low directly or by visiting the Lotrak website. The information included herein is offered in good faith to enable a reasonable assessment of the practical performance of a geotextile to be made. However, Don & Low cannot accept resposibility as the conditions of use are beyond their control. Purchasers and speciers of geotextile should satisfy themselves that it is t for the intended use.

77

G E OT E X T I L E S

reinforcing grades
physical properties of Lotrak

Test

Standard

25R

50R

70R

Tensile Strength (kN/m)

EN 10319

Warp Weft

25 25

52 50

72 72

Tensile Strength @ 5% Extension (kN/m)

EN 10319

Warp Weft

13 16

27 45

45 58

Elongation at max. load (%)

EN 10319

Warp Weft

11 10

12 7

9 7

CBR Puncture Resistance (N)

EN ISO 12236

3100

6000

8500

Cone Drop Penetration (mm)

EN 918

12

Pore Size 90% ner than (microns)

EN ISO 12956

250

250

225

Water Permeability (m/sec)

EN ISO 11058

12x10-3

16x10-3

16x10-3

Effect of UV Light

T h e Po l y p r o p y l e n e u s e d c o n t a i n s a U V i n h i b i t o r

Weight (g/m2)

120

240

330

Roll Size

Width Length

5.0 100

5.0 100

5.0 100

The gures quoted above are the mean of several tests. Specic information on tolerances is available on request. Don & Low Ltd operate a policy of continuous improvement so attention is drawn to the issue date on this sheet. A potential specier/purchaser is advised to check that this information has not been superseded either by contacting Don & Low directly or by visiting the Lotrak website. The information included herein is offered in good faith to enable a reasonable assessment of the practical performance of a geotextile to be made. However, Don & Low cannot accept resposibility as the conditions of use are beyond their control. Purchasers and speciers of geotextile should satisfy themselves that it is t for the intended use.

78

G E OT E X T I L E S

highow grades
physical properties of Lotrak

Test

Standard

HF550

HF400

Tensile Strength (kN/m)

EN 10319

Warp Weft

27 16

27 17

Elongation at max. load (%)

EN 10319

Warp Weft

28 14

28 14

CBR Puncture Resistance (N)

EN ISO 12236

2000

2000

Cone Drop Penetration (mm)

EN 918

14

12

Pore Size 90% ner than (microns)

EN ISO 12956

380

260

Water Permeability (m/sec)

EN ISO 11058

75x10-3

45x10-3

Effect of UV Light

T h e Po l y p r o p y l e n e u s e d c o n t a i n s a U V i n h i b i t o r

Weight (g/m2)

120

130

Roll Size

Width Length

4.5 100

4.5 100

The gures quoted above are the mean of several tests. Specic information on tolerances is available on request. Don & Low Ltd operate a policy of continuous improvement so attention is drawn to the issue date on this sheet. A potential specier/purchaser is advised to check that this information has not been superseded either by contacting Don & Low directly or by visiting the Lotrak website. The information included herein is offered in good faith to enable a reasonable assessment of the practical performance of a geotextile to be made. However, Don & Low cannot accept resposibility as the conditions of use are beyond their control. Purchasers and speciers of geotextile should satisfy themselves that it is t for the intended use.

79

80

EFFECTS OF VARIOUS CHEMICALS UPON POLYPROPYLENE FOUND IN LOTRAK GEOTEXTILES


This information is complied from various publications and indicates any long term degradation of polypropylene under fully submersed conditions at 20C. Chemical Acids Acids Alkalis Alkalis (diluted) (concentrated) (diluted) (concentrated) Polypropylene U U U U U U F U U U U F F U U U U U U U U U U U SA U U U U U U U U U U SA F = unaffected = slight attack = fails 1. 2. 3. 4. 5. 6. 7. 8. 9. Acetone Amyl Acetate Dimethyl Formamide Ethanol Ethyl Acetate Isopropanol Methanol N-Butanol Perchlorethylene U U U U SA F U F F U F SA Organic Solvents: in the ph range 2 to 11 Chemical Methyl cellosolve Mineral oils Nickel salts O Chlorephenol Pesticides Petrol (aromatic) Phenol 90% Selenium salts Sodium Hypochlorite (15% available chlorine) Trichlorethylene Turpentine Urine Zinc salts F F U U U Polypropylene U SA U U U F F U

Ammonia liquid (0.88) Arsenic salts Benzene Bleach Brine Bromine Water Cadmium salts Carbon Tetrachloride Chloroform Chlorine Water Chromium Salts Copper Salts Cresols Diesel Oil Detergents Developers, photographic Edible oils Ethylene Glycol Fatty Acids Formaldehyde Fuel oils Glucose Glycerine Iron salts Lead salts Lead tetraethyl Lubricating oils M Cresol Manganese salts Mercury salts

10. Pyridine 11. Toluene 12. Xylene

Combinations of certain chemicals although unaffecting plastics when in isolation may cause degradation when in an emulsion.

81

stronger by design

Don & Low Ltd. Newfordpark House Glamis Road, Forfar Angus DD8 1FR Scotland Tel: +44(0)1307 452200 Fax: +44(0)1307 452300

Email lotrak@donlow.co.uk

Visit our website www.lotrak.com

Contact our technical helpline +44 (0)1307 452632

You might also like