Sie sind auf Seite 1von 12

Geotechnical Testing Journal, Vol. 26, No. 4 Paper ID GTJ11056_264 Available online at: www.astm.

org

Jinyuan Liu,1 Magued G. Iskander,2 and Samer Sadek3

Consolidation and Permeability of Transparent Amorphous Silica

ABSTRACT: The consolidation and permeability of transparent amorphous silica are studied in this paper. Transparency is achieved by matching the refractive indices of the amorphous silica particles and the pore fluid. The fundamental premise of this research is that transparent synthetic soils made of amorphous silica can be used in model tests to study three-dimensional deformation and flow problems, using nonintrusive optical visualization techniques, if amorphous silica can be produced with geotechnical properties similar to natural soils. The amorphous silicas studied in this paper exhibit consolidation behavior similar to that of organic clays and peat. The permeability of the material falls within the range typically reported for clays and peats. KEYWORDS: Clay, Peat, Consolidation, Permeability, Image Analysis, Optical, Transparency, Digital Image Correlation, Model Tests.

Introduction Visualization of flow and deformation in soils can provide insight for solving many geotechnical problems. Conventional imaging techniques, such as radiography (X-ray, -ray), computerized axial tomography (CT), and nuclear magnetic resonance imaging (NMR), have been attempted in visualizing flow and deformation in soils (Lorenz and Heinz 1969; Mandava et al. 1990; Desrues et al. 1991; Orsi et al. 1992; Kantzas and Trigg 1996; Posadas et al. 1996; Wong 1999). However, routine application of those techniques is limited mainly by the high cost of the experimental setup, and also by the experimental limitations of these techniques. Transparent surrogates made of glass beads or quartz powder and matched refractive index pore fluids have been used to help visualize deformation and flow inside models (Chen and Wada 1986; Brock and Orr 1991; Konagai et al. 1992; Allersma 1998). However, glass and quartz surrogates are limited by their inability to represent the geotechnical properties of natural soils, and also by their poor transparency. Transparent synthetic soils made of amorphous silica and matched refractive index fluids were shown to represent the macroscopic geotechnical properties of natural soils (Iskander et al. 1994; Mannheimer and Oswald 1993). These materials are truly transparent, not merely translucent, thus permitting complete light intrusion. Two families of transparent materials have been developed for modeling sand and clay. The first family, which models sand, is made of transparent silica gels (Sadek et al. 2002). The second family is made of amorphous silica powder and was found to model the geotechnical properties of natural clays (Iskander 1997, 1998). Both families have the same refractive index and use the same pore fluids, thus permitting their use in the same model. Simple optical
Received December 12, 2001; accepted for publication December 18, 2002; published November 17, 2003. 1 Senior Geotechnical Engineer, STV, Inc., 469 Seventh Avenue, New York, NY 10018, Jliu@mta-esa.org. 2 Associate Professor, Department of Civil Engineering, Polytechnic University, 6 Metrotech Center, Brooklyn, NY 11201, Iskander@poly.edu. 3 Tunnel Engineer, Parsons, Brinckerhoff, Quade, and Douglas, Inc., 469 Seventh Avenue, New York, NY, 10018, Ssadek@mta-esa.org.

techniques have already been used to measure the response of transparent soils in model tests. For example, Welker et al. (1999) used color saturation to study flow of contaminants into perforated vertical (wick) drains. Gill and Lehane (2001) used target tracking to study pile penetration in clays. Nevertheless, the full potential of transparent synthetic soils will only be realized when methods being developed to measure three-dimensional deformation and flow within transparent soil models are available. The objective of this paper is to present the consolidation and permeability of transparent amorphous silica powders suitable for modeling clays. Physical and shear properties are also briefly introduced. Transparent Synthetic Soils The transparent synthetic soils used in this study were made by consolidating suspensions of amorphous silica powder and liquids with matched refractive indices. The clarity of transparent synthetic soils depends on the perfect matching of the refractive indices and the absence of entrapped air and impurities. The transparency of the produced materials is demonstrated in Fig. 1, where the de-aired suspension was consolidated in a transparent Plexiglas mold. The amorphous silica powder used in this research is a commercial product of PPG. An electron microscope image of amorphous silica is shown in Fig. 2. Amorphous silica consists of ultra fine particles with individual diameters on the order of 0.02 m. These particles combine together to form larger aggregates, which are porous. The produced soils benefit from two characteristics of the amorphous silica (Mannheimer 1990). First, the particles constituting the silica aggregates do not scatter light, because their diameters are shorter than the wavelength of visible light. Therefore, any defect in the particles does not affect the transparency of the synthetic soil. Secondly, porous aggregates are capable of adsorbing pore fluid, thus displacing air, which is a major source of transparency degradation. Several types of amorphous silica were tested in this study (Table 1). The tested materials encompass the range of available silica sizes and differ only by the size of the silica aggregates. The pore fluid was a 1:1 blend, by weight, of Drakeol 35

Copyright by2003 ASTM (all rights reserved); Wed Apr 11 09:58:47 2012 Copyright byInt'l ASTM International, 100 Barr Harbor Drive, EDT PO Box C700, West Conshohocken, PA 19428-2959. Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

GEOTECHNICAL TESTING JOURNAL

mineral oil and Norpar 12 paraffinic solvent. The viscosity of the oil blend at room temperature (24C) was 5.0 cP. More information about the materials and the procedure for sample preparation can be found in Iskander et al. (1994). Physical Properties Amorphous silica has a bulk density of 56230 kg/m3 (3.514 pcf) and a moisture content of 67 %. The reported specific gravity is 2.02.1, and the median aggregate size ranges between 1.4 and 175 m (Table 1). The consolidated specimens exhibit a high apparent total void ratio, e, due to the internal porosity of the silica aggregates. The inter-aggregate void ratio, ei, is more representative than the total void ratio, e, for geotechnical purposes, since it only considers the volume in between the aggregates. The interaggregate void ratio can be derived as follows, assuming full saturation: Vv Vv i e as ei Vs Vv i 1 as (1)

FIG. 1No smoking sign viewed through 50-mm-thick transparent soil model. The lower half of the word SMOKING is visible through amorphous silica, while the top half is visible through pore fluid. The word NO is viewed through the mold only.

Where, Vv is the total volume of voids, Vvi is the volume of voids inside the aggregates, Vs is the volume of solids, s is the unit weight of solids, and a is the adsorption factor. The adsorption factor is defined as the volume of pore fluid absorbed per unit weight of solids, and was estimated by Mannheimer and Oswald (1993) to be 2.1 cm3 of pore fluid per gram of amorphous silica. Mineral oil in the pore fluid has a high boiling point; thus, it does not totally evaporate in conventional moisture (oil) content tests. A correction factor, j, can be used to obtain oil content as follows: jWl wc Wt jWl (2)

Where wc is the oil content, Wt is the total weight of the specimen, and Wl is the measured weight loss due to evaporation. The value of j depends on the percentage of solvent in the mineral oil blend, and it was found to be 2.0 for specimens dried for 48 h. At least 48 h was needed for all the solvent to evaporate at 105C. Mineral oil did not evaporate at that temperature. A mass balance can also be used to estimate oil content, assuming full saturation after sedimentation.
FIG. 2The microstructure of FGSP amorphous silica (from PPG 2000).

Shear Properties Amorphous silica exhibited typical stress strain behavior in conventional triaxial compression tests in Fig. 3, where the deviator stress, (1 3), and the pore water pressure, u, are normalized by the preconsolidation pressure, 3. The behavior of amorphous silica is plotted in Fig. 4 along with typical data of well-known clays from Henkel (1956) and Simons (1960). All tests showed that shear properties of transparent amorphous silica are consistent with those of natural clays. However, amorphous silica exhibits a higher strength, but lower modulus, than the majority of natural clays. A detailed discussion of the shear strength properties of amorphous silica can be found in Iskander et al. (2002). Consolidation Properties One-dimensional (Ko) consolidation tests were carried out according to ASTM D 2435, Test Method for One-Dimensional Consolidation Properties of Soils (Table 2). Samples were 38.1 mm

TABLE 1Some physical properties of amorphous silica powders Median Aggregate Size (m) 1.4 10 25 175 Surface Area m2/g 150 180 220 150 Oil Absorption mL/100g 150 210 260 200350 Bulk Density kg/m3 56 128 144 230

Material Hi-Sil T600 (HST600) Flo-Gard FF (FGFF) Flo-Gard SP (FGSP) Hi-Sil SC-72 (SC72)

Specific Gravity 2.1 2.0 2.0 2.0

From PPG Industries (2000)

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

LIU ET AL. ON TRANSPARENT AMORPHOUS SILICA

(2.5 in.) in diameter and 19 mm (0.75 in.) high. Tests were conducted using conventional oedometer apparatus, using fixed weights. Isotropic consolidation tests were also carried out in a triaxial cell using computerized apparatus, where confining pressure was applied using a computer-controlled pump, which also measured volume change with an accuracy of 0.01 mL.

Consolidation Indices Typical e-log p curves from oedometer tests are shown in Fig. 5 for three types of amorphous silica. Void ratios were taken 24 h after loading because of the difficulty in identifying the end of primary consolidation. The compression and recompression indices

TABLE 2Consolidation properties of amorphous silica powders. Sample No. FGFF-1 FGFF-2 FGSP-1 FGSP-2 FGSP-3 FGSP-4 FGSP-5 SC72-1 SC72-2 SC72-3* Cc 1.630 1.800 2.651 2.582 2.306 2.020 2.628 2.544 2.933 2.060 Cr 0.180 0.199 0.232 0.227 0.260 0.162 0.204 0.231 0.300 Ca (Ca/Cc) 0.061(0.037) 0.064(0.036) 0.094(0.035) 0.079(0.030) 0.084(0.036) 0.087/(0.043) 0.101/(0.038) 0.103(0.040) 0.073(0.025) 0.105(0.051) Void Ratio, e 6.96 6.81 8.82 7.10 8.99 8.00 7.90 10.55 8.60 6.65 Preconsolidation Pressure, KPa 132 243 88 176 176 138 138 132 243 176

Computer-controlled loading system was used. * The pore fluid was calcium bromide brine.

FIG. 3Typical normalized stress strain and pore pressure diagrams of undrained overconsolidated FGSP.

FIG. 5Typical e-log p curves for amorphous silica from oedometer tests.

FIG. 4Variation of shear strength, pore pressure, and pore pressure coefficient for FGSP at Failure.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

GEOTECHNICAL TESTING JOURNAL

(Cc and Cr) are 1.63.0 and 0.160.3, respectively (Table 2). Both indices are within the range typically reported for clay Montmorillonite minerals (Lambe and Whitman 1968). The ratio Cr/Cc is around 0.1, which is within the normal range of 0.020.2 for natural soils (Terzaghi et al. 1996). Two separate consolidation tests are presented in Fig. 6, where data is repeatable and compression and rebound behavior is consistent with natural clays. Conventional Casagrande method for determination of preconsolidation pressure was found to be applicable. The measured compression index is consistent with that of natural clays and peats having water contents of 80130 % (Terzaghi et al. 1996). Consolidation Behavior Typical consolidation volume change versus time for both Ko and isotropic consolidation tests is shown in Fig. 7. Volume change continued to increase with time, and the end of primarily consolidation was not readily identifiable. Conventional curve-fitting methods based on Terzaghis theory of consolidation did not predict the observed volume change due to the occurrence of large secondary consolidation (Fig. 8). As indicated in Fig. 2, amorphous silica consists of aggregates made of ultrafine particles. The change in void ratio during consolidation represents the change in

both the voids within and between the silica aggregates. Two consolidation processes take place when a load is applied on a specimen. The first consolidation process involves the pores between the silica aggregates, while the second involves the pores inside the silica aggregates. Since only the first consolidation process is accounted for in the conventional theory of consolidation (Terzaghis), therefore, the second consolidation process appears as secondary consolidation. Similar behaviors are reported in the consolidation of peat and organic clays (Buisman 1936; Adams 1965; Hobbs 1986; Hanrahan 1954). Peat and organic clays have a macro-pore structure corresponding to the voids within the inorganic component in them, and a micro-pore structure corresponding to the pores within the organic matter component. When a load is applied to peat, inorganic matter consolidates first, followed by organic matter. Peat also exhibits a high void ratio typically on the order of 520 (Head 1994), like amorphous silica. Reduction in permeability due to consolidation also accounts for variation from Terzaghis theory of consolidation. Settlement Components Strain versus logarithm of time exhibited three main components: initial strain, primary consolidation, and secondary con-

FIG. 6Two consolidation tests on FGSP.

FIG. 8Measured and calculated (Terzaghis) settlement of FGSP.

FIG. 7Strain versus time for FGSP from Ko and isotropic tests.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

LIU ET AL. ON TRANSPARENT AMORPHOUS SILICA

sure sensor at the base of the specimen (not at the center) could have also contributed to this effect (Berre and Iversen 1972). Compression Isochrones For materials with large secondary consolidation, such as peat, time is typically introduced in the e-log p relationship using isochrones, where the strains measured at different times are plotted against the corresponding applied total stress (Head 1994). Hobbs (1986) reported that peat and organic clay exhibit parallel isochrones. Amorphous silica exhibits an increasing isochronal compression index, i.e., the slope of settlement against log pressure increases with time (Fig. 15). For example, the isochronal compression index changed from 2.623.10 at 1 min to 5.445.87 at 1 day for the tested amorphous silicas. Ko vs. Isotropic Consolidation
FIG. 9Initial strain versus pressure for three silicas from oedometer tests.

solidation. The initial strain, measured 0.25 min after loading, increased with the applied load (Fig. 9). Initial strain accounted for approximately half the strain measured in one day. The primary and secondary consolidation could not be clearly distinguished from settlement curves for both oedometer (Fig. 10) and triaxial consolidation tests (Fig. 11). Strain continued to increase, while load was maintained constant for up to 78 days in an oedometer test, as shown in Fig. 12. Similar behavior was reported in peats by Dhowian and Edil (1980). The value of the secondary compression index normalized by the compression index (Table 2) was approximately C /Cc 0.030.05, which is consistent with the values reported by Mesri et al. (1999) for 18 soft clay deposits. Pore Pressure Dissipation The dissipation of pore pressure during consolidation was studied in a typical specimen (FGSP-5 in Table 2). The test was performed using a computer controlled loading frame consisting of a stepper motor and sensors for measuring load, displacement, and pore pressure. Load was increased to the desired stress level in 60 s, and maintained constant thereafter. The system allowed for maintaining the stress with an accuracy of 0.1 %. Pore pressure was measured at the bottom of the specimen with an accuracy of 0.25 %. Next, load was doubled and the procedure was repeated. Typical pore pressure and volume change under load are shown in Fig. 13. It is evident that settlement continued to increase, even after full pore pressure dissipation. The time for pore pressure dissipation increased with increasing vertical stress, from less than 1 min for the 383766 kPa loading stage to almost 1 day for the 30646029 kPa. A similar phenomenon has been observed frequently in tests on peat (De Jong 1968; Dhowian and Edil 1980; and Lefebvre et al. 1984). A lag between development of pore pressure and applied load was noticed (Fig. 14). The lag increased with increasing applied pressure. A similar lag has been reported in tests on peat (Edil and Haan 1994; Lefebvre et al. 1984). This was probably caused in our tests by incomplete saturation with the degree of saturation typically within 9095 %. However, the location of the pore pres-

Volume change is plotted against logarithm of time in Fig. 16, for both Ko and isotropic consolidation. In Ko consolidation, the volume change increased linearly with time and the rate of settlement increased with pressure. During isotropic consolidation, a larger nonlinear increase in volume change was observed at corresponding loading stages. Similar behavior was reported by Yamaguchi et al. (1985) for undisturbed fibrous peat. The relation between the slope of volume change and pressure was not apparent because of the different back pressures applied in each loading stage, which also contributed to the lower volume change in isotropic consolidation compared to that of oedometer tests for the same pressure. Permeability Properties The permeability reported in this paper refers to the permeability of pore fluid through transparent soil samples. All tests were carried out using a constant head setup. ASTM D 5084, Test Method for Measurement of Hydraulic Conductivity of Saturated Porous Materials Using a Flexible Wall Permeameter, (FW) was used for tests on HST600, FGFF, and FGSP. FW tests were carried out on specimens with 37.2-mm (1.5-in.) diameter and height around 76 mm (3 in.). A pneumatic pressure panel board was used to provide a constant head at the inflow line. A back pressure of 300400 kPa (4358 psi) was used to facilitate saturation. Two computer-controlled pumps were used to keep the pressure head constant, and measure volume change. The first pump was connected to the outflow line. It provided back pressure and measured the volume of outflow of the specimen. The second pump was connected to the permeability cell. It controlled the cell (confining) pressure and measured the volume change of the specimen during consolidation and permeability testing. The pumps have an accuracy of 0.001 mL/min. Hydraulic conductivities ranging between 2 108 and 1 105 m/s were measured. These conductivities correspond to intrinsic permeabilities in the range of 136500 millidarcy. Natural soils with intrinsic permeabilities in this range usually exhibit hydraulic conductivities to water at room temperature ranging between 1.3 107 and 6.5 105 m/s. ASTM D 2434, Test Method for Permeability of Granular Soils, was used for tests on specimens of HST600 and FGSP consolidated in a rigid wall (RW) permeameter. Tests were carried out on specimens 64-mm (2.5-in.) in diameter and approximately 77-mm (3-in.) long. After consolidation, a pneumatic pressure panel board was used to provide a constant head at the inflow line. The volume

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

GEOTECHNICAL TESTING JOURNAL

FIG. 10Strain versus both log and square root of time from Ko consolidation tests.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

LIU ET AL. ON TRANSPARENT AMORPHOUS SILICA

FIG. 11Strain versus both log and square root of time from isotropic consolidation tests.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

GEOTECHNICAL TESTING JOURNAL

FIG. 12Secondary consolidation of FGSP (load kept constant in an oedometer).

FIG. 13Pore pressure and volume change in oedometer tests.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

LIU ET AL. ON TRANSPARENT AMORPHOUS SILICA

FIG. 14Pore pressure lags the applied load in oedometer tests.

FIG. 15Isochrones of consolidation for amorphous silica from oedometer tests.

of outflow was used to compute the hydraulic conductivities. Conductivities ranging between 2.3 109 and 2.5 107 m/s were measured, which correspond to intrinsic permeability in the range of 1.5160 millidarcy. Natural soils with intrinsic permeability in this range usually exhibit hydraulic conductivities to water ranging between 1.5 108 and 1.6 106 m/s. Permeability with Void Ratio The logarithm of permeability decreased linearly with void ratio, as shown in Fig. 17, where the correlation coefficient, R2, ranges between 0.974 and 0.995. This behavior is consistent with the behavior of clay minerals reported by Mesri and Olson (1971). Permeability with Vertical Pressure The logarithm of permeability decreased with increasing pressure, as shown in Fig. 18, where RW refers to the rigid wall tests and FW refers to the flexible wall tests. The behavior was linear for RW tests, but not for FW tests. The decrease in permeability in FW tests became stable at approximately 400 kPa, and no appreciable change in permeability was noticed after that.

Permeability with Material Type The permeability varied from 2 108 to 1 105 m/s in FW tests. Values ranging between 2.3 109 and 2.5 103 m/s were measured in RW tests. The smaller the aggregate size, the lower the permeability. These permeabilities fall within the range reported for clays (Casagrande and Fadum 1940; Terzaghi et al. 1996). Use of Transparent Synthetic Soil in Modeling Soil Structure Interaction Experiments are performed using the optical setup shown schematically in Fig. 19. The model is sliced optically using a laser light sheet. The interaction of laser light and transparent synthetic soils produces a distinctive speckle image. A sequence of speckle images is captured during loading using a digital camera. Digital image correlation is used to generate the two-dimensional deformation field within each slice from consecutive images. This technique is illustrated in Fig. 20, where displacement below a model footing is shown using arrows, and the speckle image is shown in the background (Sadek 2001).

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

10

GEOTECHNICAL TESTING JOURNAL

FIG. 16Comparison of volume change during consolidation between Ko and isotropic consolidation tests for FGSP.

FIG. 17Permeability versus void ratio from flexible wall permeability test.

FIG. 18Change in permeability with pressure for both flexible and rigid wall tests.

FIG. 19Experimental setup for measuring deformation in transparent soil model.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

LIU ET AL. ON TRANSPARENT AMORPHOUS SILICA

11

FIG. 20Soil deformation under the shallow foundation predicted by digital image correlation using transparent synthetic soil.

Conclusions The consolidation behavior of transparent amorphous silica is similar to that of natural peat and organic clays, where secondary consolidation dominates the behavior. Amorphous silica has an internal porosity consisting of both macro and micro pore structures, like peat. The similarity in structure is believed to contribute to similar consolidation behavior. The permeability of amorphous silica is consistent with that of natural clays. Amorphous silica can be customized to simulate specific geotechnical properties of clay in model tests. It provides a good medium for studying deformations and soil structure interaction inside transparent soil models. Acknowledgments This work is supported through NSF Grant No: CMS 9733064 CAREER: Modeling 3D Flow & Soil Structure Interaction Using Optical Tomography. References Adams, J. I., 1965, The Engineering Behavior of a Canadian Muskeg, Proceedings, 6th International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, pp. 37. Allersma, H., 1998, Using Imaging Technologies in Experimental Geotechnics, Proceedings, 2nd International Conference on Imaging Technologies: Techniques and Applications in Civil Engineering, American Society of Civil Engineers, pp.19. Berre, T. and Iversen, K., 1972, Oedometer Tests With Different Specimen Heights on a Clay Exhibiting Large Secondary Compression, Geotechnique, Vol. 22, No.1, pp. 5370. Brock, D. C. and Orr, F. M., 1991, Flow Visualization of Viscous Fringering in Heterogeneous Porous Media, SPE Annual Technical Conference and Exhibition, pp. 211222. Buisman, A. S. K., 1936, Results of Long Duration Settlement Tests, Proceedings, 1st International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, pp.103105.

Casagrande, A. and Fadum, R. E., 1940, Notes on Soil Testing for Engineering Purposes, Harvard University Graduate School of Engineering, Publication 268. Chen, J. D. and Wada, N., 1986, Visualization of Immiscible Displacement in A Three-Dimensional Transparent Porous Medium, Experiments in Fluids, Vol. 4, No. 6, pp.336338. De Jong, G. D. J., 1968, Consolidation Models Consisting of an Assembly of Viscous Elements or a Cavity Channel Network, Geotechnique, Vol.18, No.2, pp. 195228. Desrues, J., Mokni, M., and Mazerolle, F., 1991, Tomodensitometrie et la localisation sur les sables, Proceedings, 10th International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, pp. 6164 (in French). Dhowian, A. W. and Edil, T. B., 1980, Consolidation Behavior of Peats, Geotechnical Testing Journal, Vol. 3, No. 3, pp.105114. Edil, T. B. and Haan, E. J., 1994, Settlement of Peats and Organic Soils, Proceeding of Settlement94, Vertical and Horizontal Deformation of Foundations and Embankments, American Society of Civil Engineers, pp.15431572. Gill, D. and Lehane, B. M., 2001, An Optical Technique for Investigating Soil Displacement Patterns, Geotechnical Testing Journal, Vol. 24, No. 3, pp. 324329. Hanrahan, E. T., 1954, An Investigation of Some Physical Properties of Peat, Geotechnique, Vol. 4, pp.108123. Head, K. H., 1994, Manual of Soil Laboratory Testing, Vol. 2, Permeability, Shear Strength and Compressibility Tests, 2nd ed., Pentech Press. Henkel, D. J., 1956, The Effect of Over Consolidation on The Behavior of Clays During Shear, Geotechnique, Vol. 6, pp. 139150. Hobbs, N. B., 1986, Mire Morphology and the Properties and Behavior of Some British and Foreign Peats, Quarterly Journal of Engineering Geology, London, Vol. 19, No. 1, pp. 780. Iskander, M., Lai, J., Oswald, C., and Mannheimer, R., 1994, Development of a Transparent Material to Model the Geotechnical Properties of Soils, Geotechnical Testing Journal, Vol. 17, No. 4, pp. 425433.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

12

GEOTECHNICAL TESTING JOURNAL

Iskander, M., 1997, A Transparent Material to Model the Geotechnical Properties of Soils, Proceedings, 14th International Conference on Soil Mechanics and Foundation Engineering, Hamburg, Germany, Vol. 1, pp. 315319. Iskander, M., 1998, Transparent Soils to Image 3D Flow and Deformation, Proceedings, 2nd International Conference on Imaging Technologies: Techniques and Applications in Civil Engineering, American Society of Civil Engineers, pp. 255264. Iskander, M., Liu, J., and Sadek, S., 2002, Transparent Amorphous Silica to Model Clay, ASCE Journal of Geotechnical and Geoenvironmental Engineering, Vol. 128, No. 3, pp. 262273. Kantzas, A. and Trigg, A., 1996, Laboratory Investigation of Pipeline/Soil Interactions Using X-Ray Computer Assisted Tomography, Proceedings of the International Pipeline Conference, IPC, Vol. 2, pp. 12371247. Konagai, K., Tamura, C., Rangelow, P., and Matsushima, T., 1992, Laser-Aided Tomography: A Tool For Visualization of Changes in The Fabric of Granular Assemblage, Proceedings of JSCE, No 455 I-21 in Structural Engineering/Earthquake Engineering, Vol. 9, No. 3, pp. 193s201s. Lamb, T. and Whitman, R., 1968, Soil Mechanics, John Wiley, New York. Lefebvre, G., Langlois, P., and Lupien, C., 1984, Laboratory Testing and In Situ Behavior of Peat As Embankment Foundation, Canadian Geotechnical Journal, Vol. 21, pp. 322337. Lorenz, H. and Heinz, W. F., 1969, Change of Density in Sands Due to Loading, Proceedings, 7th International Conference on Soil Mechanics and Foundation Engineering, Vol. 1, pp. 267273. Mandava, S. S., Watson, A. T., and Edwards, C. M., 1990, NMR Imaging of Saturation During Immiscible Displacements, AIChE Journal, Vol. 36, No. 11, pp. 16801686. Mannheimer, R., 1990, Slurries You Can See Through, Technology Today, p. 2. Mannheimer, R. J. and Oswald, C., 1993, Development of Transparent Porous Media with Permeabilities and Porosities Comparable to Soils, Aquifers, and Petroleum Reservoirs, Ground Water, Vol. 31, No. 5, pp. 781788. Mesri, G. and Olson, R. E., 1971, Mechanism Controlling The Permeability of Clays, Clays and Clay Minerals, Vol. 19, No. 3, pp.151158.

Mesri, G. and Feng, T. W., and Shahien, M., 1999, Coefficient of Consolidation By Inflection Point Method, ASCE Journal of Geotechnical and Geoenvironmental Engineering, Vol. 125, No. 8, pp. 716718. Orsi, T. H., Anderson, A. L., Leonard, J. N., Bryant, W. R., and Edwards, C. M., 1992, Use of X-Ray Computed Tomography in The Study of Marine Sediments, Proceedings, CEO V, American Society of Civil Engineers, College Station, TX, pp. 968982. Posadas, D. A. N., Tannus, A., Panepucci, H., and Crestana, S., 1996, Magnetic Resonance Imaging as a Non-Invasive Technique For Investigating 3-D Preferential Flow Occurring Within Stratified Soil Samples, Computers and Electronics in Agriculture, Vol.14, No. 4, pp. 255267. PPG Industries, 2000, Silica Products, One PPG Place, Pittsburgh, PA. Sadek, S., Iskander, M., and Liu, J., 2002, Geotechnical Properties of Transparent Silica, Canadian Geotechnical Journal, Vol. 39, No. 1, pp. 111124. Sadek, S., 2001, Soil Structure Interaction in Transparent Synthetic Soils Using Digital Image Correlation, Ph.D. Dissertation, Polytechnic University, New York. Simons, N. E., 1960, The Effect of Over Consolidation on The Shear Strength Characteristics of An Undisturbed Oslo Clay, Proceedings, Research Conference on Shear Strength of Cohesive Soils , American Society of Civil Engineers, pp. 747763. Terzaghi, K., Peck, R. B., and Mesri, G., 1996, Soil Mechanics in Engineering Practice, 3rd ed., A Wiley-Interscience Publication. Welker, A., Bowders, J., and Gilbert, R., 1999, Applied Research Using Transparent Material With Hydraulic Properties Similar to Soil, Geotechnical Testing Journal, Vol. 22, No. 3, pp. 266270. Wong, R., C. K., 1999, Mobilized Strength Components of Athabasca Oil Sand in Triaxial Compression, Canadian Geotechnical Journal, Vol. 36, No. 4, pp. 718735. Yamaguchi, H., Ohira, Y., and Kogure, K.,1985, Volume Change Characteristics of Undisturbed Fibrous Peat, Soils and Foundations, Vol. 25, No. 2, pp.119134.

Copyright by ASTM Int'l (all rights reserved); Wed Apr 11 09:58:47 EDT 2012 Downloaded/printed by Univ Estadual De Montes Claros pursuant to License Agreement. No further reproductions authorized.

Das könnte Ihnen auch gefallen