You are on page 1of 15
2. Soil Liquefaction Photo 2.1 Soil liquefaction 2.1 Introduction During earthquakes the shaking of ground may cause a loss of strength or stiffnes that results in the settlement of buildings, Iandshdes, the failure of earth dams, or other hazards. The process leading to such loss of strength or stiffies 1s called soil liquefaction Itis a phenomenon associated primarely, but not exclusively, with saturated cohesionless soils (Committee on Earthquake Engineering, 1985). Some effects of liquefaction can be catastrophic, such as flow faitures of stopes or earth ams, and settlement and tipping of buildings and bridge piers Others are less dramatic. such as lateral spreading of slightly inclined ground and deformation and settlement of the ground surface, Even these latter effects, though, have been seen to cause severe damage to highways, railroads, pipeline and buildings in many earthquakes. ‘Attention was drawn to the effects of liquefaction for the first time during the Niigata Earthquake in Japan, 1964. This event caused exstensive damages mainly due to soil falures connected with liquefaction. During the Alaska earthquake the same year, further examples of the destructivness of liquefaction were produced. Since then, great progress has been made in understanding liquefaction phenomena, and in evaluating the risk for liquefaction on a site. In section 2,2 an attempt is made to describe what today is known about liquefaction, its processes and its manifestations Methods of evaluating the potential for liquefaction on a site are discussed in section 2.3. Section 2.4 shortly describes how a zonation of liquefaction susceptibility using geological criteria can be made, ‘Much of the content in this chapter is based on the publication Liquefaction of Soils During Earthquakes (Committee on Earthquake Engineering, 1985). This is a state-of-the- art publication with its origin m a workshop held in Dedham, Massachusetts, on March 1985, at which specialists on liquefaction from the United States, Japan, Canada, and the United Kingdom panicipated 2.2 The Phenomenon 2.2.1 Manifestations of Liquefaction ‘The manifestation most commonly associated with liquefaction is probably the formation of sand boils. These small sand volcanos indicates elevated pore pressures at depth, caused by the compactation of granular soils during the seismic shaking. From zones of high pore pressure water makes channels and transports Sediments to the surface. The sediments are deposited as a cone around the opening of the channel See fig 2./ and photo 2.1 The formation of sand boils is ome dependent on the soil overlying the liquefied material. A thick overlying Fig. 2.1 Sand boil, the manifestation most layer gives fewer and larger boils commonly associated with liquefaction If the overlying material is cohesive it cracks up and the water-sand mixture vents to the surface through the cracks, forming irregular and elongated sand boils. The amount of sand deposited on the surface depends on the depth of the liquified zone An overlying layer of gravel may prevent the pressurised mixture to reach the surface, ifit can be expended in the pores of the coarser material Liquefaction at depth often cause the overlying soil to break up into blocks bounded by fissures. Floating on the liquefied zone, these superficial blocks are allowed to move in relation to eachother. See fig 2.2 On honsontal ground, the movement of the blocks during an earthquake is called. Fig 2.2 The mechanism of ground oscillation. ground oscillation, and are often seen by observers as ground waves. On slightly inclined ground, gravitational forces and inertial forces generated by the earthquake make the superficial blocks move laterally. This is known as lateral spread and commonly involve displacements ranging up to several meters, but depending on the slope and the seismic shaking, it can extend up to several tens of meters. The movement is towards a free face, such as a river channel ‘and commonly results in x= the formation of fissures, horsts, and grabens. ‘While ground oscillation seldom is very destructive, lateral spreads have caused more damage than any other liquefaction-induced soil s+ failure, Damages typically include the rupture of INITIAL SECTION foundations of buildings, ‘BEPORMED SECTION the tilting of bridge abutments and buckling of bridge decks, Fig 2.3 Lateral spread due to liquefaction. the rupture of burried tubes, and fissuring of road embankments. ‘When a soil that supports a structure liquefies, it will result in a foss of bearing capacity, allowing the structure to settle and tip A classic example of this phenomenon occured during the Niigata Earthquake where apartment buildings tipped as much as 60 degrees. Another phenomenon caused by liquefaction is the buoyant rise of burried structures. ‘When burried structures are lighter in weight than the surrounding soil it “floats” up to the surface This might lead 10 unexpected emergencies of burried tanks, tubes, and cut-off timber piles. When the excess pore pressures connected with liquefaction dissipates, it will be accompanied by a densification of the soil leading to settlements of the ground surface. ‘Uneven settlements may cause damage to roads, railroad tracks, and other structures, ‘The most catastrophic ground failures caused by liquefaction are flow failures These flows develops on slopes greater than 3 degrees and are composed of completely liquefied soil or intact blocks of soil riding on a layer of liquefied material. This type of failures commonly displace great volumes of of soil for tens of meters or more, often at high velocities Man-made earth structures such as earth dams and mine-waste tailings dams have been seen to be susceptible to flow failures In coastal areas, flow failures under water have carried away sections of port facilities Fig 24 Flow failure ‘The increase of pore pressures associated with liquefaction brings about a corresponding. decrease in effective stresses, which in turn not only decreases the strength of the soil but also reduces its stiffness, Reduction in strength and stifiness are the reasons for soil failures and soil deformations in the phenomena described above 2.2.2 Knowledge From Laboratory Tests Laboratory tests performed on soil samples have shown that cyclic loading causes soil particles to rearrange towards a state of closer packing (fig. 2.5) B= B ig 2.5 Compactation duc to a cyclic load. The compactation can not occur if the soil is saturated with an incompressible fluid that is prevented to move within or from the soil. Instead, the gravity loading is transferred from the soil skeleton to the pore water, i, with a constant total stress, a decrease in effective stress will be counteracted by an increase in pore pressure (eg. 2.1). c=c+U Qn Since both the strength and the stiffness of a soil are proportional to its effective stress, the capacity of the soil to resist loading will be reduced Tests have shown that saturated granular soils (ie, some silts, sands, and even gravels) are the most susceptible to pore pressure buildup. The greater the content of fine particles, the less the susceptibility to pore pressure buildup, since the plasticity of clays and other fines prevent the rearrangemen of particles. Coarse materials allow pressurised pore water to dissipate more easily, thus preventing pore pressures of larger magnitude. Tsuchida (1970), proposed boundaries for the grain size distribution of soils susceptible to liquefaction The boundaries were the result of sieve analyses on soils that did or did not liquefy during past earthquakes (fig 2.6) Percent finer by weight Grain size (mm) Fig 2.6 Boundaries inthe gradation curves for souls susceptible to liquefaction. After Tsuchida (1970). Among the factors affecting the rate of pore pressure buildup in cohesionless soils are the density of the soil and the magnitude of the cyclic strain. In a loose sand, pore pressures build up rapidly, and the soil may loose its initial shear resistance. In dense sands the pore pressure builds up at a slower rate, and even if strains due to the cyclic load and permanent strains may develop, the soil can retain much of its resistance to shear. The process of densification in a dry soil, or the buildup of pore pressures in a saturated soil, is also controlled by the amplitude of the cyclic loading Several authors have shown that there is a threshold shear strain below which no densification takes place Though the amount of densification is controlled by the cyclic strain, in practical use, itis often predicted from the cyclic stress In a soil susceptible to pore pressure buildup and subjected to a cyclic loading, the pore pressure increases with each cycle until it equals, or nearly equals, the total stress acting upon the soil, ie. a condition of zero, or nearly zero, effective stress is reached (eq. 2.2). o=0 2.2) The first time this condition occures is sometimes called initial liquefaction. Fig. 2.7 shows the result of a cyclic triaxial test on sand, performed by Seed and Lee (1966). COMPRESSION LOAD (ke) EXTENSION COMPRESSION DEFORMATION (in.) ° EXTENSION, os = PORE WATER PRESSURE 5 oa : ooo Se ees Fig. 2.7 Cyclic triaxial test on loose sand with a zero initial deviator stress. After Committee on. Earthquake Engineering (1985). ‘When the c'= 0 condition is reached in a test like this there is an increase in cyclic strain. For looser sands this increase is quite sudden. The number of cycles required to reach the o'= 0 condition is primarely a function of the density and the magnitude of the cyclic loading, The results illustrated in fig. 2.7 comes from a test with isotropic consolidation, ie, with a zero initial deviator stress. Several authors have shown that unisotropic consolidation, giving static shear stresses within the sample, greatly can change the pore pressure buildup In fig. 2.8 are the results from a cyclic triaxial test on an isotropically and unisotropically consolidated spiecemen respectively illustrated in the q, p\-plane. Fig. 2.8a Stress path for isotropically Fig, 2.8b Stress path for unisotropically consolidated fine sand in a triaxial test. consolidated fine sand. After Committee on Earthquake Engineering (1985), Starting at the right hand side in the diagrams the stress path progressively moves towards the left as the cycling continues, the distance representing the increase in pore pressure. After the o’= 0 (p'= 0) condition is reached in the test with the isotropically consolidated sample the stress path follow the failure lines up and down passing through the origin twice each cycle (fig. 2.8a). In the case of the unisotropically consolidated sample, the stress path stabilizes when the failure line is reached so that the pore pressure does not build up further. In this case a o’= 0 condition is never reached (fig. 2.85). Even if a sand develops a o'= 0 condition during the cyclic loading it may still possess a substantial resistance to shear during a following undrained monotonic loading. After a certain amount of undrained deformation the soil reaches a condition of steady-state of deformation with a corresponding steady-state strength. As shown by Castro (1969) and Castro et al. (1982), this strength is only a function of the void ratio of the soil, Fig. 2.9 illustrates the stress-strain curve for a soil with an initial shear stress and subjected to a cyclic loading followed by a monotonic undrained loading, BEFORE eYCLING 8 8 / # wonoronte g i 5 tgacing E / wma & aprer i t SHEAR —5. CYCLING § ms STRE = £ / ss § e foaerer B/S ounns eveums 8 De SauSep By Evecie Loan STRAIN STRAIN @ 0) Fig. 2.9 Undramned stress-strain curves for medium dense or dense sand with and without cychc loading. After Commitee on Earthquake Engineering (1985), It can be seen that the undrained steady-state shear strength is greater than the initial stressO applied to the soil In the field this type of soil would remain stable following a cyclic loading, even if the deformation during the cycling might be quite large. The result in fig. 2.9a corresponds to a dense or medium dense sand with dilative characteristics. If, on the other hand, the sand is very loose the steady strength might be smaller than the applied initial stress (fig. 2.105). In this case the strains due to the cyclic loading could carry the soil past the peak in shear resistance typical for a contractive material (fig 2. !0a) FT| ounine evcuine g wma 2 a SHEAR —@. Qe STRESS 3 COLLAPSE. 2 wean sTaain @ o) Fig, 2.10 Undrained stress-strain curves for very loose sand with and without cyclic loading. After Commuttee on Earthquake Engineering (1985), The result would be a collapse with continous deformation, in the field manifested as a flow failure. It has been argued by Casagrande (1975), Castro (1975), and Castro and Poulos (1977) that the soil-responce displayed in fig. 2.10a is the only true form of liquefaction, whereas the behaviour in fig. 2.92 is a problem of deformation and should be termed cyclic mobility. In this report, however, the word liquefaction is used in a broader sense, denominating deformations caused by elevated pore pressures as liquefaction phenomena as well. In what has been written above could be noted the influence of intial static shear stresses on liquefaction. While an initial shear stress may reduce the pore pressure buildup it may in the same time, depending on the soil density, cause soil collapse after or during the dynamic loading. 2.2.3 Assessing the Susceptibility of Liquefaction ‘When assessing the susceptibility of liquefaction on a site the task is to determine the ‘magnitude and duration of ground shaking that will cause liquefaction in the soil deposit in question In this report it is assumed that the shaking of ground has its origin in an earthquake, but liquefaction is known to have occured due to other dynamic sources, such as detonating explosives and even sea waves A site characterized by a horizontal surface of large extent and without any structures that impose stresses into the ground 1s termed Jevel ground. The problem in this case will be to determine the cyclic load that will bring about a o'= 0 condition in the soil profile, On the other hand, if the soil on a site possesses static shear stresses, which is the case in a slope or beneath a building, both the pore pressure buildup and the undrained steady-state strength of the soil must be analyzed There are two general methods in use 1. Zones where a o'= 0 condition is likely to occur are determined from cyclic load tests or in-situ penetration resistance. These zones are given a steady-state strength, evaluated from correlations between observed in-situ penetration resistance and steady- state strength. Finally a static stability test is performed 2. The potential for a flow failure is focused as the in-situ steady strength is determined from laboratory tests complemented with the correlations between field index tests and the steady strength, Ifa potential for flow failure exists the conditions required to trigger the failure is determmed using laboratory cyclic tests or empirical correlations. Ifa design ground motion is used in the calculations, a safety against liquefaction can be determined. Though this frequently is done following the schemes described above there are simplified empirical methods of assessing the liquefaction potential to be used in ‘engineering practise. The many documented instances of liquefaction and nonliquefaction in the field have enabled investigators to establish correlations involving the intensity of ‘ground shaking and the resistance of the soil from which a factor of safety can be evaluated 2,3 Empirical Procedures to Evaluate the Potential of Liquefaction 2.3.1 Introduction ‘The mass of field data concerning the occurrence and nonoccurrence of liquefaction obtained during earthquakes since the Niigata event has been the basis for the development of empirical correlations between the tensity of ground motion and in-situ evaluations of the soil resistance. By using these correlations when assessing the liquefaction potential on a site one circumvents the problems connected with taking undesturbed samples and laboratory testing. The in-situ soil resistance is commonly evaluated from the standard penetration test (SPT) which is a dynamic penetration test carried out using a standard procedure described in Test 19 of British Standard 1377: 1975. The testis performed using a split barrel sampler, 50 mm extemal diameter, 35 mm intemal diameter, and with a length of 650 mm, connected to the end of drill rods. The sampler is driven into the ground at the bottom of a cased borehole by a 65 kg hammer falling from a height of 760 mm onto the top of the drill rods. Initially the sampler is driven 150 mm into the ground to seat the sampler. The umber of blows required to drive the sampler a further 300 mm is then recorded and called the standard penetration resistance (N). Though the main purpose is to obtain an indication of the relative density of the soil there is much published information correlating the results of the test with other soil parameters. It seems as the SPT is an appropiate way of evaluating the resistance to liquefaction since the test is sensitive to several factors that also affects the liquefaction resistance (eg, relative density and ratio of horizontal to vertical stress). In the empirical methods, the intensity of cyclic load is described by the peak ground acceleration at the surface of the soil deposit caused by an earthquake It should be remembered that determining the ground motion on a site due to an earthquake (or an expected earthquake) is a problem as complex as liquefaction itself. A site specific ground acceleration is commonly evaluated from the earthquake magnitude and the distance to the epicentre of the earthquake, or from rough correlations between peak accelerations and earthquake intensity scales such as the Modified Mercalli Scale (see also section 3.2.1) In the following two empirical methods of assessing the potential of liquefaction on a site will be described. The data from which these methods are developed correspond to level ground, 10 2.3.2 Empirical Method After Seed et al In fig. 2.// each point corresponds to site specific conditions obtained during some particular earthquake, The resistance of the soil is represented on the horizontal axis by the (Ni)so value, which is the blow count value in the SPT-test corrected for, as will be discussed later, the overburden pressure and the energy delivered to the drill rod. On the vertical axis the inensity of ground motion is represented by the average cyclic stress ratio, ‘tavlo’s, av is the average shear stress mobilized by the earthquake and o's the initial vertical effective stress tav/c'o is calculated from the peak surface acceleration as. Tay any © we ogg ems So % Bo @3) where amax = maximum acceleration at the ground surface 0 = total overburden stress at the depth in question effective overburden stress at the same depth stress reduction factor decreasing with depth The curve in fig. 2.11 is intended to divide zones of liquefaction and nonliquefaction for an earthquake of magnitude 7.5. In order to construct the corresponding curves for other earthquake magnitudes the abscissa of the curve in fig. 2.11 should be corrected with the factors in table 2.1. ° a ad i Earthquake Number of Factorto ° r magnitude (M) representative correct ig cycles at abscissa of : : 0.65 tmx of curve y in fig, 2.11. . ols a5 26 0.89 £ 15 15 1.0 e 6.75 10 113 fines coments 60 oe 132 onan saan con geet # | 5.25 23 Ls Table 2.1 Representative number of cycles and corresponding correction factors (Seed and Idriss, 1982). After Committee on Earthquake ig 2.14 Relationship between stress ratios causing Engineering (1985) liquefaction and (Ni) values for clean sands valid for magnitude 7.5 earthquakes (Seed etal, 1984), After Commutiee on Earthquake Engineering (1933) n ‘The curves dependecy on magnitude is due to the fact that a certain earthquake magnitude is assumed to produce a certain number of representative loading cycles. The larger the magnitude, the larger the number of loading cycles. In order to make the results from the standard penetration test truly stanardized at least two corrections of the blow count value, N, must be made. The first correction is made to normalize the observed blow count value to a value, Ni, which would be measured at an effective overburden pressure of | ton/f? Fig. 2.2 and eg. 2.4, suggest how to chose the value of the correction factor Cx. 1 Gn eS % es The second correction relates to the differences in the amount of energy delivered to the drill rod by different types of SPT-equipments. Table 2.2 shows the fraction of energy, ERm, reaching the rod when using different hammer types and release mechanisms. The blow count is corrected to a value corresponding to a situation where 60% of the energy is delivered to the rod. Thus the corrected blow count value is calculated as: ER = 60 (Neo es) Entering fig. 2.1/ with the (Ni)éo value and using the appropiate curve, the cyclic strength of the soil will be obtained. Comparing this strength with the design earthquake loading calculated with eq. 2.3, the potentialfor liquefaction can be expressed =Ca & Neve y $0 Bros ear Fig. 212 Chart for values of correcting factor Cs for overburden pressure (Seed etal. 1984, based on data and analyses from Marcuson and Biegenousky 1977), After Committee on Earthquake Engineering 1985) 16 i Hummer Country Type Hammer Release Japan’ Doout ‘Freel Donut Rope and pulley with special throw release Unted Safety Rope and pelley States Donut? Rope and pulley Argentina Donut Rope and pulley China Donut Freefall Donut __ Rope and pulley Corecuon Factor Estimated Rod for 60 Percent Energy (Percent) Rod Eneisy 7% 78160 = 1.30 o 6760 = 1:12 « 00 = 1.00 45 45160 = 075 4s 48160 = 075 « 60180 = 1.00 50 $0160 = 083 “Tepanee SPT resale fave addonalconecuons for Borchale Gamer tad eqoency EH ‘Prevalent method nthe Une State today conse hammers aeveop an eneray Fa of abou 4 percent SOURCE” Seed etal (1984) ‘Table 2.2 Hammer types and release mechanisms with different delivery of energy to the rod in the SPT- test, After Committee on Earthquake Engineering (1985). 2 It should be remembered that the data underlying the chart in fig. 2.11 correspond to level ‘ground level with no initial shear stresses and to an effective overburden pressure less than 150 kPa, When these conditions are violated, corrections of the soil strength obtained fromthe 9} chart must be made. As have been shown in section 2.2.2, initial shear stresses may have different influence on the ost development of liquefaction depending on if the soil is contractive or dilative. In fig. 2.13 a correction factor, Ko, is presented as a function of c= ta/o'o (where tat equals the D, = 45-50% | | | horizontal static shear stress) and Fig. 2.13 Ranges in Ka for relative soil density, Dr. With Ke. various relative densities (Seed and Harder, 1990) the effect of initial shear stresses ‘After Committee on Earthquake Engincering (1985) can be taken into account. A large effective overburden pressure bas been seen to give a substantial reduction in resistance to cyclic loading. For effective overburden pressures other than in the range 100- 150 kPa, a correction factor Ka, is used. Fig. 2.14 shows values of Ko, for Folsom gravels Finally, the chart in fig, 2.11 is valid for sands with a fines content less than 5%. As previously mentioned, fines tend to stabilize the soil skeleton and decrease the susceptibility of liquefaction. Seed at el (1988) proposed that the content of fines could be taken into account by adding increments to the (Ni)so value, With the increments in fable 2.3, (Ni)éo values measured in silty sands could be corrected to equivalent clean sand values, Fines Content (%) 10 1 25 2 30 4 75 5 ‘Table 2.3 Increments to measured (Nis for fines content. After Committee on Earthquake Engineering (1985) Fig, 2.14 Ko factors for Folsom gravels (Seed and Harder, 1990) After Committee on Earthquake Engineering (1985). B The method outlined above is in the litterarure commonly known as the Seed method, A more detailed discussion of the method could be found in Comittee on Earthquake Engineering (1985) and Finn (1991), 2.3.3 Empirical Method After Iwasaki et al In 1981 Iwasaki, Tokida and Tatsuoka proposed a procedure to evaluate the potential of liquefaction termed "Simple Geotechnical analysis". This procedure is outlined in the Japanese Bridge Code and based on data obtained during a number of Japanese earthquakes, The method is similar to the Seed Method in that a soil strength against liquefaction, here called R, is calculated along with a load, L, induced by an earthquake The ratio of these two factors gives the liquefaction resistance factor, Ft: R 2 eo The soil strength, R, is made up by a sum of three components, reflecting different properties of the soil, R=Rit+R+R 27 The first component, RI, is a function of the blow count value: R= 0.0882 a o'+07 as) where o'o 1s the effective overburden pressure expressed in (kgfem’) R2 depends on the mean particle diameter, Ds0 (mm) R2=019 if0.02 < Dso< 005 0.225 log(0 35/D50) —_if 0.05 < Ds0< 06 -0.05 if 0.6 40 (2.10) “ L, the cyclic load, is calculated almost as in the Seed method max a me Sg So 8 Mo au where amax = maximum acceleration at the ground surface 0 = total overburden pressure at the depth in question '0 = effective overburden pressure at the same depth d= stress reduction factor considering that the soil stiffness increases with depth and calculated by eg. 2.12 d= 1-0 015z (2.12) where 2= the depth in meters below the ground surface The only difference between the Seed and Iwasaki methods when calculating the cyclic load is the factor 0 65, which in the Seed method yields the average cyclic stress ratio ‘tav/o'o In the Iwasaki method the load 1s characterized by the peak cyclic stress ratio maven, The blow count value, N, is the value measured with the SPT-equipment most commonly used in Japan The influence of the effective overburden pressure on the N-value is taken into account since oo is present in the denominator in the expression for Ru. Corrections to be used when the level ground condition is violated are not included in this method For more information about the Iwasaki method the reader is directed to Iwasaki et al (1981) and TC-4 Committee of ISSMFE (1993), 2.4 Zonation of Liquefaction Potential Using Geological Criteria ‘The two empirical methods described in the previos section are examples of procedures that can be used when assessing the risk for liquefaction on a specific site, Sometimes (e.g, in planning activities), an evaluation of the liquefaction potential for a hole area would be desired. Regional studies of this kind commonly delineates areas where liquefaction could occur given a sufficiently large earthquake. The studies would not be a substitute for site-specific investigations, but they could indicate areas where more detailed investigations should be made In order to find any mappable properties for liquefaction susceptibility, several investigators have tried to correlate liquefaction potential with geological and geomorphological criteria. The investigations point out that fluvial and eolian processes help to sort and sediment granular soils in a loose state, thereby making them susceptible to liquefaction. It has also been seen that sediments generally gain resistance to liquefaction with progress of age. Table 2.4 shows different sedimentary deposits susceptibility to liquefy according to Youd and Perkins (1978). Using criteria as those displayed in fable 2.4 a zonation of the liquefaction susceptibility for an area can be made, based upon the geological and geomorphological setting 15 General dis: | Likelihood thar Cohestoniess sediments, inbution of | When Saturated, Would Be Susceptible ‘Type of deposi Cohesionless to Liguely (by Age of Deposit) Sediments Pleis- | Prepleis in deposits _|-

You might also like