Acessibilidade / Reportar erro

Geotechnical and other characteristics of cement-treated low plasticity clay

Abstract

This research work examines the utilization of cement in order to improve low plasticity clay soil. The soil samples treated with 2, 4 and 6% cement percents and cured for different curing times extended to 90 days. Laboratory investigations include unconfined compression, indirect tensile, gas permeability and microstructural tests, which were conducted on the tested samples. The soil-water retention behavior has been also investigated. The test results showed that the cement addition improved both the compressive and tensile strength properties of soil specimens. These strength properties were also increased with curing times. pH and electrical conductivity values were good indicators for the enhancement in the strengths properties. The results of micro structural tests illustrated that the natural soil specimens contain voids and the open structure. Further, these tests showed the cementation of soil grains and filling the voids among soil grains with cementing compounds. Gas permeability and soil-water retention behavior of soil specimens are strongly related to the variations in the soil structures. Further examination illustrated that in the case of low cement content, the pore size distribution (PSD) and the efficiency of gas permeability are more sensitive to curing times.

Keywords:
Clay soil; Cement stabilization; Strength behavior; Microstructure; Curing time; Water retention behavior

1. Introduction and past studies

The successive urban development in various parts of the world necessitated further improvement of the infrastructure accompanying the constructed facilities. Compacted fine-grained soils are used in the infrastructure earthworks such as the construction embankment of roads, highways, road foundations. Fine-grained soils (especially clayey soils) consider as a problematic soil and can induce damages to roads founded on them, due to their volume changes, higher water content and/or low bearing capacity. The use of ordinary Portland cement; its components or residues; has been widely used in stabilizing cohesionless and some types of problematic soils like clayey soil. Studies conducted in this field may be classified into three main categories: use byproduct from cement production operations, direct use of cement alone or mixed with other materials, and recycling of cement as concrete waste. The use of cement byproduct, especially cement kiln dust to stabilize or improve clay soil was cover by many studies (Adeyanju & Okeke, 2019Adeyanju, E.A., & Okeke, C.A. (2019). Clay soil stabilization using cement kiln dust. IOP Conference Series: Materials Science and Engineering, 640, 012080. http://dx.doi.org/10.1088/1757-899X/640/1/012080
http://dx.doi.org/10.1088/1757-899X/640/...
; Amadi & Osu, 2018Amadi, A.A., & Osu, A.S. (2018). Effect of curing time on strength development in black cotton soil: quarry fines composite stabilized with cement kiln dust (CKD). Journal of King Saud University - Engineering Sciences, 30(4), 305-312. http://dx.doi.org/10.1016/j.jksues.2016.04.001.
http://dx.doi.org/10.1016/j.jksues.2016....
; Miller & Azad, 2000Miller, G.A., & Azad, S. (2000). Influence of soil type on stabilization with cement kiln dust. Construction & Building Materials, 14(2), 89-97. http://dx.doi.org/10.1016/S0950-0618(00)00007-6.
http://dx.doi.org/10.1016/S0950-0618(00)...
; Naseem et al., 2019Naseem, A., Mumtaz, W., Fazal-e-Jalal, & De Backer, H. (2019). Stabilization of expansive soil using tire rubber powder and cement kiln dust. Soil Mechanics and Foundation Engineering, 56(1), 54-58. http://dx.doi.org/10.1007/s11204-019-09569-8.
http://dx.doi.org/10.1007/s11204-019-095...
). The mixing of cement with fly ash become commonly used to reduce the amount of cement used or improve specific geotechnical properties of soil (Amu et al., 2008Amu, O., Fajobi, A., & Afekhuai, S. (2008). Stabilizing potential of cement-fly ash mixture on expansive clay Soil. Journal of Technology and Education in Nigeria, 12(2), http://dx.doi.org/10.4314/joten.v12i2.35698.
http://dx.doi.org/10.4314/joten.v12i2.35...
; Chenari et al., 2018Chenari, R.J., Fatahi, B., Ghorbani, A., & Alamoti, M.N. (2018). Evaluation of strength properties of cement stabilized sand mixed with EPS beads and fly ash. Geomechanics and Engineering, 14(6), 533-544. http://dx.doi.org/10.12989/gae.2018.14.6.533.
http://dx.doi.org/10.12989/gae.2018.14.6...
; Khemissa & Mahamedi, 2014Khemissa, M., & Mahamedi, A. (2014). Cement and lime mixture stabilization of an expansive overconsolidated clay. Applied Clay Science, 95, 104-110. http://dx.doi.org/10.1016/j.clay.2014.03.017.
http://dx.doi.org/10.1016/j.clay.2014.03...
). Portland cement was also used with other stabilizing materials to improve the soil engineering properties. Lime is used with cement to improve the soil strength and reduce the swelling and settlement (Amu et al., 2008Amu, O., Fajobi, A., & Afekhuai, S. (2008). Stabilizing potential of cement-fly ash mixture on expansive clay Soil. Journal of Technology and Education in Nigeria, 12(2), http://dx.doi.org/10.4314/joten.v12i2.35698.
http://dx.doi.org/10.4314/joten.v12i2.35...
; Joel & Agbede, 2010Joel, M., & Agbede, I.O. (2010). Cement stabilization of Igumale shale lime admixture for use as flexible pavement construction material. The Electronic Journal of Geotechnical Engineering, 15, 1661-1673.; Lemaire et al., 2013Lemaire, K., Deneele, D., Bonnet, S., & Legret, M. (2013). Effects of lime and cement treatment on the physicochemical, microstructural and mechanical characteristics of a plastic silt. Engineering Geology, 166, 255-261. http://dx.doi.org/10.1016/j.enggeo.2013.09.012.
http://dx.doi.org/10.1016/j.enggeo.2013....
; Mousavi & Leong Sing, 2015Mousavi, S., & Leong Sing, W. (2015). Utilization of brown clay and cement for stabilization of clay. Jordan Journal of Civil Engineering, 9(2), 163-174.; Riaz et al., 2014Riaz, S., Aadil, N., & Waseem, U. (2014). Stabilization of subgrade soils using cement and lime: a case study of Kala Shah Kaku, Lahore, Pakistan. Pakistan Journal of Science, 66(1), 39-44. Retrieved in March 1, 2014, from http://aplicacionesbiblioteca.udea.edu.co:3653/ehost/detail/detail?vid=11&sid=9827d053-420e-4669-9575-5e6211a04bbc@sessionmgr112&hid=125&bdata=Jmxhbmc9ZXMmc2l0ZT1laG9zdC1saXZl#db=a9h&AN=98360899
http://aplicacionesbiblioteca.udea.edu.c...
; Saeed et al., 2015Saeed, K.A., Kassim, K.A., Nur, H., & Yunus, N.Z.M. (2015). Strength of lime-cement stabilized tropical lateritic clay contaminated by heavy metals. KSCE Journal of Civil Engineering, 19(4), 887-892. http://dx.doi.org/10.1007/s12205-013-0086-6.
http://dx.doi.org/10.1007/s12205-013-008...
; Sharma et al., 2018Sharma, L.K., Sirdesai, N.N., Sharma, K.M., & Singh, T.N. (2018). Experimental study to examine the independent roles of lime and cement on the stabilization of a mountain soil: a comparative study. Applied Clay Science, 152, 183-195. http://dx.doi.org/10.1016/j.clay.2017.11.012.
http://dx.doi.org/10.1016/j.clay.2017.11...
; Umesha et al., 2009Umesha, T.S., Dinesh, S.V., & Sivapullaiah, P.V. (2009). Control of dispersivity of soil using lime and cement. International Journal of Geology, 3(1), 8-16.; Wei et al., 2014Wei, D., Zhu, B., Wang, T., Tian, M., & Huang, X. (2014). Effect of Cationic Exchange Capacity of Soil on Strength of Stabilized Soil. Procedia: Social and Behavioral Sciences, 141, 399-406. http://dx.doi.org/10.1016/j.sbspro.2014.05.070.
http://dx.doi.org/10.1016/j.sbspro.2014....
). Nayak & Sarvade (2012)Nayak, S., & Sarvade, P.G. (2012). Effect of cement and quarry dust on shear strength and hydraulic characteristics of lithomargic Clay. Geotechnical and Geological Engineering, 30(2), 419-430. http://dx.doi.org/10.1007/s10706-011-9477-y.
http://dx.doi.org/10.1007/s10706-011-947...
used cement and quarry dust to improve the shear strength and hydraulic features of lithomarge clay. Ayeldeen & Kitazume (2017)Ayeldeen, M., & Kitazume, M. (2017). Using fiber and liquid polymer to improve the behaviour of cement-stabilized soft clay. Geotextiles and Geomembranes, 45(6), 592-602. http://dx.doi.org/10.1016/j.geotexmem.2017.05.005.
http://dx.doi.org/10.1016/j.geotexmem.20...
utilized fiber, and liquid polymer to enhance the strength of cement-soft clay blends. The fibers and liquid polymers displayed a notable mechanically, economically and environmentally prospects to be used as an additive to cement in improving the soft clay. Also, organic soils have become the target of many studies that have addressed improving the properties of these soils by adding cement and other materials (Kalantari & Huat, 2008Kalantari, B., & Huat, B.B.K. (2008). Peat soil stabilization, using ordinary portland cement, polypropylene fibers, and air curing technique. The Electronic Journal of Geotechnical Engineering, 13, 1-13.; Kalantari & Prasad, 2014Kalantari, B., & Prasad, A. (2014). A study of the effect of various curing techniques on the strength of stabilized peat. Transportation Geotechnics, 1(3), 119-128. http://dx.doi.org/10.1016/j.trgeo.2014.06.002.
http://dx.doi.org/10.1016/j.trgeo.2014.0...
). Moreover, Osinubi et al. (2011)Osinubi, K.J., Oyelakin, M.A., & Eberemu, A.O. (2011). Improvement of black cotton soil with ordinary portland cement - locust bean waste ash blend. The Electronic Journal of Geotechnical Engineering, 16, 619-627. used ordinary Portland cement –Locust bean waste ash mixture to enhance the engineering properties such as (UCS) and ‎California bearing ratio (CBR) for black cotton clayey soil. Crushed concrete waste, which represents the last form of cement used, has been used in many studies to improve the properties of clay soils (Abdulnafaa et al., 2019Abdulnafaa, M.D., Cabalar, A.F., & Arabash, Z. (2019). Shear strength characteristics of clay with waste solid construction and demolition materials. In Shear Strength Characteristics of Clay with Construction and Demolition Solid Waste Materials, Turkey.‎; Cabalar et al., 2016Cabalar, A.F., Abdulnafaa, M.D., & Karabash, Z. (2016). Influences of various construction and demolition materials on the behavior of a clay. Environmental Earth Sciences, 75(9), 841. http://dx.doi.org/10.1007/s12665-016-5631-4.
http://dx.doi.org/10.1007/s12665-016-563...
, 2017Cabalar, A.F., Hassan, D.I., & Abdulnafaa, M.D. (2017). Use of waste ceramic tiles for road pavement subgrade. Road Materials and Pavement Design, 18(4), 882-896. http://dx.doi.org/10.1080/14680629.2016.1194884.
http://dx.doi.org/10.1080/14680629.2016....
, 2019aCabalar, A., Abdulnafaa, M., & Isik, H. (2019a). The role of construction and demolition materials in swelling of a clay. Arabian Journal of Geosciences, 12(11), 361. http://dx.doi.org/10.1007/s12517-019-4552-4.
http://dx.doi.org/10.1007/s12517-019-455...
, bCabalar, A., Zardikawi, O., & Abdulnafaa, M. (2019b). Utilisation of construction and demolition materials with clay for road pavement subgrade. Road Materials and Pavement Design, 20(3), 702-714. http://dx.doi.org/10.1080/14680629.2017.1407817.
http://dx.doi.org/10.1080/14680629.2017....
; İşbuğa et al., 2019İşbuğa, V., Çabalar, A., & Abdulnafaa, M. (2019). Large-scale testing of a clay soil improved with concrete pieces. Proceedings on Engineering Sciences B, 2, 95-100.). The main purpose of adding ordinary Portland cement to cohesionless soils is to provide strong bonds between soil particles (Consoli et al., 2011Consoli, N., Cruz, R., & Floss, M. (2011). Variables controlling strength of artificially cemented sand: influence of curing time. Journal of Materials in Civil Engineering, 23(5), 692-696. http://dx.doi.org/10.1061/(ASCE)MT.1943-5533.0000205.
http://dx.doi.org/10.1061/(ASCE)MT.1943-...
, 2017Consoli, N., Faro, V., Schnaid, F., & Born, R. (2017). Stabilised soil layers enhancing performance of transverse-loaded flexible piles on lightly bonded residual soils. Soils and Rocks, 40(3), 219-228. http://dx.doi.org/10.28927/SR.403219.
http://dx.doi.org/10.28927/SR.403219...
). Al-Aghbari et al. (2009)Al-Aghbari, M.Y., Mohamedzein, Y.E.A., & Taha, R. (2009). Stabilisation of desert sands using cement and cement dust. Proceedings of the Institution of Civil Engineers: Ground Improvement, 162(3), 145-151. http://dx.doi.org/10.1680/grim.2009.162.3.145.
http://dx.doi.org/10.1680/grim.2009.162....
used cement and cement dust to stabilize desert sands. The results showed that the cement and cement by-pass dust could be used to improve the compressibility and shear strength characteristics of desert sands. Also, Saberian et al. (2018)Saberian, M., Moradi, M., Vali, R., & Li, J. (2018). Stabilized marine and desert sands with deep mixing of cement and sodium bentonite. Geomechanics and Engineering, 14(6), 553-562. http://dx.doi.org/10.12989/gae.2018.14.6.553.
http://dx.doi.org/10.12989/gae.2018.14.6...
studied the stabilization of marine and desert sands with deep mixing of cement and sodium bentonite and found an improvement in the geotechnical properties of these soils. Shooshpasha & Shirvani (2015)Shooshpasha, I., & Shirvani, R.A. (2015). Effect of cement stabilization on geotechnical properties of sandy soils. Geomechanics and Engineering, 8(1), 17-31. http://dx.doi.org/10.12989/gae.2015.8.1.017.
http://dx.doi.org/10.12989/gae.2015.8.1....
reported that the use of cement to stabilize sandy soils resulting in increased strength parameters, reduced strain at failure, and changed soil behavior to a noticeable brittle behavior. Iravanian & Bilsel (2016)Iravanian, A., & Bilsel, H. (2016). Strength characterization of sand-bentonite mixtures and the effect of cement additives. Marine Georesources and Geotechnology, 34(3), 210-218. http://dx.doi.org/10.1080/1064119X.2014.991463.
http://dx.doi.org/10.1080/1064119X.2014....
studied the sand-bentonite landfill barrier material with and without cement additive, at different periods of aging. The strength characterization of mixtures was a marked improvement with cement inclusion and that the effect of aging has been very effective.

The clay-cement reaction produce primary and secondary cementations materials in the soil-cement matrix (chew et al., 2004Chew, S.H., Kamruzzaman, A.H.M., & Lee, F.H. (2004). Physicochemical and engineering behavior of cement treated clays. Journal of Geotechnical and Geoenvironmental Engineering, 130(7), 696-706. http://dx.doi.org/10.1061/(ASCE)1090-0241(2004)130:7(696).
http://dx.doi.org/10.1061/(ASCE)1090-024...
). Cement has two chemical reactions; the first one begins at the time of adding the water to the fine soil-cement mixture and the second one is the secondary reaction occurs as the calcium ions diffuse through the soil (Chen & Wang, 2006Chen, H., & Wang, Q. (2006). The behaviour of organic matter in the process of soft soil stabilization using cement. Bulletin of Engineering Geology and the Environment, 65(4), 445-448. http://dx.doi.org/10.1007/s10064-005-0030-1.
http://dx.doi.org/10.1007/s10064-005-003...
; Chew et al., 2004Chew, S.H., Kamruzzaman, A.H.M., & Lee, F.H. (2004). Physicochemical and engineering behavior of cement treated clays. Journal of Geotechnical and Geoenvironmental Engineering, 130(7), 696-706. http://dx.doi.org/10.1061/(ASCE)1090-0241(2004)130:7(696).
http://dx.doi.org/10.1061/(ASCE)1090-024...
). These chemical reactions are responsible for the strength development in cement-treated soils. The geotechnical properties of cement-treated clay soils have been investigated by different researchers (Consoli et al., 2010Consoli, N., Arcari Bassani, M., & Festugato, L. (2010). Effect of fiber-reinforcement on the strength of cemented soils. Geotextiles and Geomembranes, 28(4), 344-351. http://dx.doi.org/10.1016/j.geotexmem.2010.01.005.
http://dx.doi.org/10.1016/j.geotexmem.20...
; Goodary et al., 2012Goodary, R., Lecomte-Nana, G.L., Petit, C., & Smith, D.S. (2012). Investigation of the strength development in cement-stabilised soils of volcanic origin. Construction & Building Materials, 28(1), 592-598. http://dx.doi.org/10.1016/j.conbuildmat.2011.08.054.
http://dx.doi.org/10.1016/j.conbuildmat....
; Kalıpcılar et al., 2016Kalıpcılar, İ., Mardani-Aghabaglou, A., Sezer, G.İ., Altun, S., & Sezer, A. (2016). Assessment of the effect of sulfate attack on cement stabilized montmorillonite. Geomechanics and Engineering, 10(6), 807-826. http://dx.doi.org/10.12989/gae.2016.10.6.807.
http://dx.doi.org/10.12989/gae.2016.10.6...
; Kasama et al., 2000Kasama, K., Ochiai, H., & Yasufuku, N. (2000). On the stress-strain behaviour of lightly cemented clay based on an extended critical state ‎concept‎. Soil and Foundation, 40(5), 37-47. http://dx.doi.org/10.3208/sandf.40.5_37.
http://dx.doi.org/10.3208/sandf.40.5_37...
; Kenai et al., 2006Kenai, S., Bahar, R., & Benazzoug, M. (2006). Experimental analysis of the effect of some compaction methods on mechanical properties and durability of cement stabilized soil. Journal of Materials Science, 41(21), 6956-6964. http://dx.doi.org/10.1007/s10853-006-0226-1.
http://dx.doi.org/10.1007/s10853-006-022...
; Lorenzo & Bergado, 2004Lorenzo, G.A., & Bergado, D.T. (2004). Fundamental parameters of cement-admixed clay - New approach. Journal of Geotechnical and Geoenvironmental Engineering, 130(10), 1042-1050. http://dx.doi.org/10.1061/(ASCE)1090-0241(2004)130:10(1042).
http://dx.doi.org/10.1061/(ASCE)1090-024...
; Okyay & Dias, 2010Okyay, U.S., & Dias, D. (2010). Use of lime and cement treated soils as pile supported load transfer platform. Engineering Geology, 114(1-2), 34-44. http://dx.doi.org/10.1016/j.enggeo.2010.03.008.
http://dx.doi.org/10.1016/j.enggeo.2010....
; Park, 2011Park, S.S. (2011). Unconfined compressive strength and ductility of fiber-reinforced cemented sand. Construction & Building Materials, 25(2), 1134-1138. http://dx.doi.org/10.1016/j.conbuildmat.2010.07.017.
http://dx.doi.org/10.1016/j.conbuildmat....
; Petchgate et al., 2001Petchgate, K., Sukmongkol, W., & Voottipruex, P. (2001). Effect of height and diameter ratio on the strength of cement stabilized soft Bangkok ‎clay‎. Journal of Geotechnical Engineering, 31(3), 227-239.; Saadeldin & Siddiqua, 2013Saadeldin, R., & Siddiqua, S. (2013). Geotechnical characterization of a clay-cement mix. Bulletin of Engineering Geology and the Environment, 72(3-4), 601-608. http://dx.doi.org/10.1007/s10064-013-0531-2.
http://dx.doi.org/10.1007/s10064-013-053...
; Zhang et al., 2014Zhang, T., Yue, X., Deng, Y., Zhang, D., & Liu, S. (2014). Mechanical behaviour and micro-structure of cement-stabilised marine clay with a metakaolin agent. Construction & Building Materials, 73, 51-57. http://dx.doi.org/10.1016/j.conbuildmat.2014.09.041.
http://dx.doi.org/10.1016/j.conbuildmat....
). Common Portland cement was used to improve shear resistance and durability of clay soils. The unsaturated properties of cement-treated clay soils were observed by different studies (Attom et al., 2000Attom, M.F., Taqieddin, S.A., & Mubeideen, T. (2000). Shear strength and swelling stabilization of unsaturated clayey soil using pozzolanic material. In C.D. Shackelford, S.L. Houston & N.-Y. Chang (Eds.), Advances in unsaturated geotechnics (pp. 275-288). Reston, VA: American Society of Civil Engineers.; Azam & Cameron, 2013Azam, A.M., & Cameron, D.A. (2013). Geotechnical properties of blends of recycled clay masonry and recycled concrete aggregates in unbound pavement construction. Journal of Materials in Civil Engineering, 25(6), 788-798. http://dx.doi.org/10.1061/(ASCE)MT.1943-5533.0000634.
http://dx.doi.org/10.1061/(ASCE)MT.1943-...
; Nahlawi et al., 2004Nahlawi, H., Chakrabarti, S., & Kodikara, J. (2004). A direct tensile strength testing method for unsaturated geomaterials. Geotechnical Testing Journal, 27(4), 356-361. http://dx.doi.org/10.1520/GTJ11767.
http://dx.doi.org/10.1520/GTJ11767...
). Horpibulsuk et al. (2012)Horpibulsuk, S., Phojan, W., Suddeepong, A., Chinkulkijniwat, A., & Liu, M.D. (2012). Strength development in blended cement admixed saline clay. Applied Clay Science, 55, 44-52. http://dx.doi.org/10.1016/j.clay.2011.10.003.
http://dx.doi.org/10.1016/j.clay.2011.10...
found that the strength of clay is governed by the clay-water/cement ratio. The strength increases with the decrease in the clay-water/cement ratio. (Horpibulsuk et al., 2012Horpibulsuk, S., Phojan, W., Suddeepong, A., Chinkulkijniwat, A., & Liu, M.D. (2012). Strength development in blended cement admixed saline clay. Applied Clay Science, 55, 44-52. http://dx.doi.org/10.1016/j.clay.2011.10.003.
http://dx.doi.org/10.1016/j.clay.2011.10...
) studied the microstructural characteristics of cement-stabilized soils and found that soil behavior enhanced significantly. The optimum dosages of cement added to clay soil to improve some geotechnical properties were investigated by Rojas-Suárez et al. (2019aRojas-Suárez, J.P., Orjuela-Abril, M.S., & Prada-Botía, G. (2019a). Study of low plasticity clay for optimum dosage of the soil-cement. Journal of Physics: Conference Series, 1386(1), 012078. http://dx.doi.org/10.1088/1742-6596/1386/1/012078.
http://dx.doi.org/10.1088/1742-6596/1386...
, bRojas-Suárez, J.P., Orjuela-Abril, M.S., & Prada-Botía, G.C. (2019b). Determination of the adequate dosage of the soil-cement, using clay of high plasticity. Journal of Physics: Conference Series, 1386(1), 012077. http://dx.doi.org/10.1088/1742-6596/1386/1/012077.
http://dx.doi.org/10.1088/1742-6596/1386...
). Korf et al. (2017)Korf, E.P., Prietto, P.D.M., & Consoli, N.C. (2017). Hydraulic and diffusive behavior of a compacted cemented soil. Soils and Rocks, 39(3), 325-331. observed the hydraulic and diffusive behavior of compact clay soil, with and without cement addition. The results of the reactive behavior analysis showed that the retention by adsorption increased with the increase of pH, but it was not affected by the applied static load.

In summary, many interesting results indicating the potential of the use of ordinary Portland cement to improve clayey soils have been reported. This study aims to extend and increase the knowledge of the clayey soil-cement stabilization technique.

2. Materials and testing methods

2.1 Materials

Two materials were used in this experimental work: clayey soil and cement. The soil was obtained from a depth of 1.5 m from the ground surface. The soil samples were oven-dried for 2 days at 60 C° and passed through a 4 mm sieve before use in various tests. The soil specific gravity was 2.68, the liquid limit was 35% and the plasticity index was 16%. All the chemical and physical properties tests were carried out following the ASTM standards and test procedures adopted by Aldaood et al. (2014a)Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014a). Geotechnical properties of lime-treated gypseous soils. Applied Clay Science, 88–89, 39-48. http://dx.doi.org/10.1016/j.clay.2013.12.015.
http://dx.doi.org/10.1016/j.clay.2013.12...
. The sample is categorized as CL following the Unified Soil Classification System (USCS). The X-ray diffraction test results presented that the main clay mineral was kaolinite and the non-clay minerals were quartz and calcite. Table 1 presents some properties of the clay soil used in the experimental program.

Table 1
Chemical and physical properties of the natural soil.

The stabilizing agent used for this study was ordinary Portland cement. The specific gravity is 3.13 and the specific surface is 3790 (cm2/gm). The main composition of cement is (CaO is 63.1%, SiO2 is 19.4% and Al2O3 is 5.4%). The loss on the ignition of the cement is 2.33%.

2.2 Sample preparation

An oven-dried soil was mixed with a pre-determined quantity of ordinary Portland cement (2%, 4% and 6% of dry soil weight) in dry condition.‎ The soil specimens were prepared at the optimum moisture content of natural soil (i.e. 11%). The formation of lumps was avoided when the water was added to the soil-cement mixture. The soil-cement mixture kept in the plastic bags then left for 10 minutes for homogeneity (Khattab & Aljobouri, 2012Khattab, S.I., & Aljobouri, M.M. (2012). Effect of Combined Stabilization by Lime and Cement on Hydraulic Properties of Clayey Soil Selected From Mosul Area. Al-Rafidain Engineering Journal, 20(6), 139-153.). After that, the soil specimens were statically compacted in a specific rigid mold related to the type of the test. A standard Proctor compaction test (ASTM, 2003ASTM D 698. (2003). Standard test methods for laboratory compaction characteristics of soil using standard effort (12 400 ft-lbf/ft3 (600 kN-m/m3)). In ASTM International, The annual book of ASTM standards (Vol. 3, pp. 1-11). ASTM, West Conshohocken, PA. https://doi.org/10.1520/D0698-12E01.1.
https://doi.org/10.1520/D0698-12E01.1...
) was adopted in the preparation of soil-cement specimens to obtain the maximum dry density of natural soil. All treated and untreated specimens were compacted statically to dry density of (17.5 kN/m3), which is the maximum dry density of natural soil. After compaction, the treated soil specimens were wrapped in cling film and coated with paraffin wax to prevent moisture loss, then specimens were left at room temperature of 20 C° for different periods of 3, 10, 30, 60 and 90 days to be cured.

2.3 Testing methods

The pore size distribution and microstructural characteristics of the natural soil and cement-treated soil specimen were measured using a scanning electron microscope (SEM) and porosity tests. These tests were conducted on the natural and cement-treated soil specimens, following the test procedures suggested by Aldaood et al. (2014b)Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014b). Impact of wetting-drying cycles on the microstructure and mechanical properties of lime-stabilized gypseous soils. Engineering Geology, 174, 11-21. http://dx.doi.org/10.1016/j.enggeo.2014.03.002.
http://dx.doi.org/10.1016/j.enggeo.2014....

To conduct the UCS, a cylindrical (50 mm diameter × 100 mm height) soil specimens were statically compacted at the optimum moisture content and maximum dry unit weight obtained from the standard compaction curve of natural soil. The rate of compaction was (1 mm/min) to obtain a uniform unit weight of the soil sample. The UCS has been determined according to the ASTM D-2166 and D-1633 (ASTM 2000ASTM (2000). Annual book of ASTM standards. Soil and rock, Vol. 04.08. Philadelphia: American Society for Testing and Materials.). procedures for untreated and cement-treated soil samples, respectively. Before testing, the wave velocity of the soil specimens was determined using A PUNDIT device with a frequency of 82 kHz.

The commonly used ‎alternative procedure for the determination of tensile strength is the Brazilian tensile test, which is ‎generally referred to as the (ITS) (Das et al., 1995Das, B.M., Yen, S.C., & Dass, R.N. (1995). Brazilian tensile strength test of lightly cemented sand‎. Canadian Geotechnical Journal, 32(1), 166-171. http://dx.doi.org/10.1139/t95-013.
http://dx.doi.org/10.1139/t95-013...
). The soil specimens were prepared in a metal mold with dimensions of (50 mm high and 25 mm diameter). The soil specimens were compacted statically, at the same rate as for preparing the UCS specimens. After the preparation of the natural soil specimens, they were extracted from the stacking mold and tested. While the cement-treated specimens are encapsulated as in the UCS test and exposed to the same curing time before tested. The ITS test was performed according to the method approved by the ASTM (2011)ASTM 6931. (2011). Standard Test Method for Indirect Tensile (IDT) Strength of Bituminous Mixtures (Vol. 1, pp. 3-7). ASTM International, West Conshohocken, PA. https://doi.org/10.1520/D6931-17.2.
https://doi.org/10.1520/D6931-17.2...
, by applying compressive strength along the diameter of the model and with the rate of the unconfined compressive resistance test (1.27 mm/min) until the specimens fail. The (ITS) is calculated using Equation 1

S t = 2 P max π t d (1)

where St is the indirect tensile strength and Pmax. ; is the maximum applied load on the sample; t is the average height of the sample with d as diameter.

For For pH and electrical conductivity test (EC), a portion of failed (tested) samples in the UCS test was used to ‎determine the pH and EC values, following the tests procedures suggested by (Eades & Grim, 1966Eades, J.L., & Grim, R.E. (1966). A quick test to determine lime requirements for lime stabilization‎. Highway Research Record, (139), 61-72.; Aldaood et al. 2014aAldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014a). Geotechnical properties of lime-treated gypseous soils. Applied Clay Science, 88–89, 39-48. http://dx.doi.org/10.1016/j.clay.2013.12.015.
http://dx.doi.org/10.1016/j.clay.2013.12...
).

For gas permeability, the test procedure suggested by Aldaood et al. (2016)Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2016). Effect of water during freeze-thaw cycles on the performance and durability of lime-treated gypseous soil. Cold Regions Science and Technology, 123, 155-163. http://dx.doi.org/10.1016/j.coldregions.2015.12.008.
http://dx.doi.org/10.1016/j.coldregions....
was adopted to measure the gas permeability of cylindrical soil specimens of 50 mm diameter and 50 mm height. The soil specimens were statically compacted inside a cylindrical metal mold so that it reached the maximum dry unit weight of natural soil. The gas permeability specimens were exposed to different curing times as the specimens for UCS and ITS tests. The coefficient of gas permeability was estimated using the modified Darcy’s equation as follows:

K A = Q A × 2 μ L P a t m P i 2 P a t m 2 (2)

where: Q is the volume flow rate (m3/sec), L is the thickness of the sample (m), μ is the viscosity (1.76*10-5 Pa.s for nitrogen gas at 20 °C), Patm is the atmospheric pressure (Pa) and Pi is the injection pressure (Pa), A is the cross-sectional area of the sample (m2).

It worth noting that, the measurements of permeability were conducted in an air-conditioned room having a constant temperature of 20 °C. Each permeability test involved four measurements of apparent permeability at various injection pressures.

The soil–water retention curve (SWRC) of natural and cement-treated soil specimens was determined by using the vapor equilibrium technique, osmotic membrane, and tensiometric plates. The vapor equilibrium technique was used to evaluate the SWRC in suction pressure more than 1500 kPa. The osmotic membrane determined the SWRC in the suction pressure range of 100 kPa and 1500 kPa. The evaluation of the SWRC continued in low suction pressure ranging between 10-20 kPa by using tensiometric plates. The required time to reach the balance condition (in the determination of the SWRC) varied between 20-35 days, depending on the desired technique. More details about these techniques can be found in Aldaood et al. (2015)Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2015). Soil-water characteristic curve of gypseous soil. Geotechnical and Geological Engineering, 33(1), 123-135. http://dx.doi.org/10.1007/s10706-014-9829-5.
http://dx.doi.org/10.1007/s10706-014-982...
. It worth noting that, all the previous SWRC determination techniques were carried at room temperature of (20 °C).

‎3. Results and discussion

3.1 Assessment of pH and Electrical Conductivity (EC)

The pH values of cement-treated soil specimens before and after curing were determined. Cement addition increases the pH value from (8.2) for natural soil to 12.5 for 6% cement-treated soil specimens, which promotes cation exchange (due to increasing calcium Ca++ ions). In the literature (Al-Mukhtar et al., 2014Al-Mukhtar, M., Lasledj, A., & Alcover, J.F. (2014). Lime consumption of different clayey soils. Applied Clay Science, 95, 133-145. http://dx.doi.org/10.1016/j.clay.2014.03.024.
http://dx.doi.org/10.1016/j.clay.2014.03...
; Eades & Grim, 1966Eades, J.L., & Grim, R.E. (1966). A quick test to determine lime requirements for lime stabilization‎. Highway Research Record, (139), 61-72.; Feng et al., 2001Feng, T.W., Lee, J.Y., & Lee, Y.J. (2001). Consolidation behavior of a soft mud treated with small cement content. Engineering Geology, 59(3-4), 327-335. http://dx.doi.org/10.1016/S0013-7952(01)00021-7.
http://dx.doi.org/10.1016/S0013-7952(01)...
), it was agreed that the pH value of 12.5 represent the necessary value to get a favorable environment for producing the cementing materials, and thus, the development of acceptable mechanical performance. Table 2 shows the changes in pH and EC values of cement-treated soil specimens after various curing times. It is observed that the pH values of soil specimens decreased slightly as the curing time increased. More reduction in pH occurs for low cement content and high curing time and the value reached 11.2. At this level of pH value, the pozzolanic products such as (CSH and CAH) will continue. (Al-Mukhtar et al., 2014Al-Mukhtar, M., Lasledj, A., & Alcover, J.F. (2014). Lime consumption of different clayey soils. Applied Clay Science, 95, 133-145. http://dx.doi.org/10.1016/j.clay.2014.03.024.
http://dx.doi.org/10.1016/j.clay.2014.03...
) reported that, as calcium cation is existed and the pH is high enough (more than 10.5), the pozzolanic reaction continues. Moreover, Chen & Wang (2006)Chen, H., & Wang, Q. (2006). The behaviour of organic matter in the process of soft soil stabilization using cement. Bulletin of Engineering Geology and the Environment, 65(4), 445-448. http://dx.doi.org/10.1007/s10064-005-0030-1.
http://dx.doi.org/10.1007/s10064-005-003...
documented that, when pH value (≤ 9) low level of hardening will produce or even no hardening. The reduction in pH values of soil specimens related to the reduced amount of Ca++ and (OH) ions due to the development of the pozzolanic reactions.

Table 2
Variation of pH and electrical conductivity values of soil specimens with cement content and curing times.

The electrical conductivity values (EC) of soil specimens followed the same trend as pH values. Cement addition causes an increasing in EC values from (0.42 mS/cm) for natural soil to (3.9 mS/cm) for 6% cement-treated soil specimens. This increasing related to the existing high calcium ions in adding cement (CaO is 63.1%). As the curing time increases, the EC value of soil specimens continues to down, but slightly. The reduction in EC values related to the consumption of calcium ions during the pozzolanic reactions. Finally, obtaining pH and EC values corroborate the next-obtained results of unconfined compressive and indirect tensile strengths, where significant cementing materials (such as CSH and CAH) were formed.

3.2 Microstructural characterization

Microstructural analyses were carried out to investigate the variations in the microstructure of the cured specimens and for natural soil as a comparison. These analyses helped in understanding the increase in strength of cemented soil specimens at a microscopic level. The analysis focused on the formation of cementing materials named calcium silicate hydrates (CSH) and calcium aluminate hydrates (CAH); which normally presented in lime and cement stabilized soils (Aldaood et al., 2014aAldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014a). Geotechnical properties of lime-treated gypseous soils. Applied Clay Science, 88–89, 39-48. http://dx.doi.org/10.1016/j.clay.2013.12.015.
http://dx.doi.org/10.1016/j.clay.2013.12...
; Mengue et al., 2017Mengue, E., Mroueh, H., Lancelot, L., & Medjo Eko, R. (2017). Physicochemical and consolidation properties of compacted lateritic soil treated with cement. Soil and Foundation, 57(1), 60-79. http://dx.doi.org/10.1016/j.sandf.2017.01.005.
http://dx.doi.org/10.1016/j.sandf.2017.0...
). Figure 1 presents the pore size distribution (PSD) of natural and cement-treated soil specimens. The natural soil specimens exhibited a tri-modal PSD with a large number of macrospores centered at (1–200 μm) and with a less pronounced peak centered at (0.01 μm). The PSD curve of natural soil supported the SEM results, where the texture of the natural soil specimens exhibited a fairly open type of microstructure, as illustrated in Figure 2. Besides, many coarse grains (sand grains) relatively well calibrate and assembled with fine grains (clay grains) in a dispersed arrangement, resulting to form many voids in different dimensions.

Figure 1
PSD of natural and cement-treated soil specimens with different cement content and curing times.
Figure 2
SEM images (a) natural soil (b, c, d) 30 days cured specimens treated with 2, 4 and 6% cement content, respectively.

Cement addition enhanced the PSD of soil specimens by decreasing the amount of macropores (> 10 μm) and increasing the micropores (≤ 0.01 μm), see Figure 1. The changes in the PSD of cemented soil specimens related to that, the pores (especially macrospores) covered and filled by the hydrated cement. During cement addition and with the presence of water, the clay and cement particles grow together to large clusters. Then cement gel is stable in macropores and micropores due to the attractive forces, leading to enhance PSD of cemented soil specimens (Horpibulsuk et al., 2009Horpibulsuk, S., Rachan, R., & Raksachon, Y. (2009). Role of fly ash on strength and microstructure development in blended cement stabilized silty clay. Soil and Foundation, 49(1), 85-98. http://dx.doi.org/10.3208/sandf.49.85.
http://dx.doi.org/10.3208/sandf.49.85...
, 2010Horpibulsuk, S., Rachan, R., Chinkulkijniwat, A., Raksachon, Y., & Suddeepong, A. (2010). Analysis of strength development in cement-stabilized silty clay from microstructural considerations. Construction & Building Materials, 24(10), 2011-2021. http://dx.doi.org/10.1016/j.conbuildmat.2010.03.011.
http://dx.doi.org/10.1016/j.conbuildmat....
). Further, as the curing times increase the hydration products grow and cause more reduction in the macropores. An investigation of the structure of the cemented soil specimens allowed to reflect the changes in the structure of specimens from open structure to denser one with fewer voids formation (Figure 2). Further, as the cement content increase, the soil structure became tighter than the structure of natural soil and the cluster of grains become more effective (Mengue et al., 2017Mengue, E., Mroueh, H., Lancelot, L., & Medjo Eko, R. (2017). Physicochemical and consolidation properties of compacted lateritic soil treated with cement. Soil and Foundation, 57(1), 60-79. http://dx.doi.org/10.1016/j.sandf.2017.01.005.
http://dx.doi.org/10.1016/j.sandf.2017.0...
). It is noting that the cement addition was more affected on the PSD of soil specimens than curing times, as presented in Figure 1.

3.3 Unconfined Compressive Strength characteristics (UCS)

The results of USC for cement-treated soil specimens were illustrated in Figure 3. This figure also presents the effect of curing time on the UCS. The results suggest that the cement content has a significant effect on the strength characteristics of soil specimens and the UCS of soil specimens increase with cement content. The increase in the UCS was approximately linearly with the increase in the cement ‎content. This finding is consistent with the findings of previous studies by (Chenari et al. (2018)Chenari, R.J., Fatahi, B., Ghorbani, A., & Alamoti, M.N. (2018). Evaluation of strength properties of cement stabilized sand mixed with EPS beads and fly ash. Geomechanics and Engineering, 14(6), 533-544. http://dx.doi.org/10.12989/gae.2018.14.6.533.
http://dx.doi.org/10.12989/gae.2018.14.6...
and Pakbaz & Alipour (2012)Pakbaz, M.S., & Alipour, R. (2012). Influence of cement addition on the geotechnical properties of an Iranian clay. Applied Clay Science, 67-68, 1-4. http://dx.doi.org/10.1016/j.clay.2012.07.006.
http://dx.doi.org/10.1016/j.clay.2012.07...
. The increase in USC with increasing cement content was attributed to the pozzolanic reactions between soil and cement mixtures. The pozzolanic reactions resulting in the formation of cementing compounds named calcium silicate hydrate (CSH) and calcium aluminate hydrates (CAH). These cementing compounds enhanced the inter-cluster bonding strength and filled the pore space between soil particles (see Figure 1). As a result, the strength values (i.e. UCS and ITS) of the soil specimens increased with an increase in cement content (Sharma et al., 2018Sharma, L.K., Sirdesai, N.N., Sharma, K.M., & Singh, T.N. (2018). Experimental study to examine the independent roles of lime and cement on the stabilization of a mountain soil: a comparative study. Applied Clay Science, 152, 183-195. http://dx.doi.org/10.1016/j.clay.2017.11.012.
http://dx.doi.org/10.1016/j.clay.2017.11...
). Moreover, as the cement content increases, the contact points among cement and soil particles increases and, upon hardening, gives a suitable amount of bonding at these points. Further, during cement addition, the flat and smooth particles of soil disintegrate into rough and crumbled portions and this behavior improves the cohesion value among the particles, which then increases the strength values. It worth noting that, the development of white cementing compounds (CSH and CAH) on the surfaces of soil particles aids as an indicator of the pozzolanic reactions, as illustrated in Figure 4. Similar results have been noticed for various types of soil (Lemaire et al., 2013Lemaire, K., Deneele, D., Bonnet, S., & Legret, M. (2013). Effects of lime and cement treatment on the physicochemical, microstructural and mechanical characteristics of a plastic silt. Engineering Geology, 166, 255-261. http://dx.doi.org/10.1016/j.enggeo.2013.09.012.
http://dx.doi.org/10.1016/j.enggeo.2013....
; Sharma et al., 2018Sharma, L.K., Sirdesai, N.N., Sharma, K.M., & Singh, T.N. (2018). Experimental study to examine the independent roles of lime and cement on the stabilization of a mountain soil: a comparative study. Applied Clay Science, 152, 183-195. http://dx.doi.org/10.1016/j.clay.2017.11.012.
http://dx.doi.org/10.1016/j.clay.2017.11...
).

Figure 3
UCS of soil specimens with (A) cement content and (B) curing times.
Figure 4
SEM images of cement-treated soil specimens cured for 90 days showing the roughness of the soil structure.

The role of curing time on the strength improvement of the cement-treated soil specimens was illustrated in Figure 3B. It is observed that as the curing times increased, the UCS increased. At specific cement content, the UCS increased significantly until a curing ‎time of 60 days. After 60 days of curing, the UCS increases gently as shown in Figure 3B. The UCS increase can be classified into two zones. As the curing times increase up to 60 days, the UCS increased and this zone is referred to as the active zone. After this zone, the UCS improvement slows down while still gradually increasing and this zone is designated as the inert zone. This behavior may be due to that kaolinite is exhausted by the pozzolanic reactions, which lead to reducing the action of pozzolanic reaction with increasing curing time. Besides, the continuous reduction in water content during curing times could affect the pozzolanic reactions. Great attention has been given to calculating the residual water content (RWC) of the soil samples, as shown in Figure 5. RWC means the water content of soil samples after the end of specific curing time. The RWC decreased with the increasing of curing time and cement percentages. Most reduction in water content occurs during the first times of curing until 60 days, after that the reduction in water content continued slightly. The reduction in the RWC could be due to the hydration process of cement and to completion of the pozzolanic reactions. It worth noting that, all the UCS curves (for all cement content) follow the same pattern with curing times.

Figure 5
RWC of cement-treated soil specimens cured for different curing times.

The stress-strain of UCS test results is presented in Figure 6. Results showed that the failure strain decreases considerably as the cement content and curing time increases. While the slope of the stress-strain curves (before and after the maximum stress value), increases with increasing both cement content and curing times. This means that the utilization of cement addition increased the UCS, reduced the strain at failure, and changed the soil behavior from ductile to brittle behavior. The influence of curing times on the stress-strain curves was more pronounced for higher cement content. Many researchers reported that the natural soil specimen exhibited ductile behavior; while the stabilized soil specimen posed brittle behavior (Horpibulsuk et al., 2012Horpibulsuk, S., Phojan, W., Suddeepong, A., Chinkulkijniwat, A., & Liu, M.D. (2012). Strength development in blended cement admixed saline clay. Applied Clay Science, 55, 44-52. http://dx.doi.org/10.1016/j.clay.2011.10.003.
http://dx.doi.org/10.1016/j.clay.2011.10...
; Mousavi & Leong Sing, 2015Mousavi, S., & Leong Sing, W. (2015). Utilization of brown clay and cement for stabilization of clay. Jordan Journal of Civil Engineering, 9(2), 163-174.). It worth noting that, all the stress-strain curves were similar, except the difference in the maximum stress values.

Figure 6
Stress-strain curves of cement-treated soil specimens cured for different curing times.

3.4 Indirect Tensile Strength characteristics (ITS)

The ITS test results of cement-treated soil specimens were shown in Figure 7. The data show a significant increase in the ITS of treated soils in comparison to the natural soil. It is also shown that the tensile strength increases linearly with the increase of both cement content and the curing times. The linear increase in ITS with a high slope at 3 days of curing was attributed to the short-term reactions and cement hydration. Further, this behavior largely depended on the cement content. As the curing times increase the ITS was likely to be more reliant upon the pozzolanic reactions. Also, Figure 7B implies that ITS improvement for all cement contents starts to moderate beyond 60 days of curing. This is consistent with the steady of the pozzolanic reaction at high curing times (more than 60 days). The most obvious explanation for this significant increase in the ITS is that this strength is indirectly calculated and is based on the compression pressure (Pmax.) used in Equation 1. Thus, the same reasons considered to explain the increase in UCS can be used to illustrate the significant increase in ITS values.

Figure 7
ITS of soil specimens with (A) cement content and (B) curing times.

3.5 Wave velocity results

A wave velocity test was performed on natural and cement-treated soil specimens, and the results were presented in Table 3. The results show that the wave velocity increases with increasing both cement content and curing times and followed the same trend as UCS. In general, the increase in wave velocity from the value of natural specimens to the 3 days of curing was more pronounced than the increase from 3 days to 10 days of curing. Sequentially, this value was more than other intervals of curing times (i.e. the interval between 10 to 30 days, etc.). As the curing times increase, the reactions between the soil particles and cement increased and result to increase the stiffness of the soil specimens. As a result, the wave velocity propagation increased with increasing both cement content and curing times. Mandal et al. (2016)Mandal, T., Tinjum, J.M., & Edil, T.B. (2016). Non-destructive testing of cementitiously stabilized materials using ultrasonic pulse velocity test. Transportation Geotechnics, 6, 97-107. http://dx.doi.org/10.1016/j.trgeo.2015.09.003.
http://dx.doi.org/10.1016/j.trgeo.2015.0...
documented similar test results. Besides, Yesiller et al. (2000)Yesiller, N., Hanson, J.L., & Usmen, M.A. (2000). Ultrasonic assessment of stabilized soils. In Soft Ground Technology Conference (Vol. 301, pp. 170-181), Noordwijkerhout, the Netherlands. http://dx.doi.org/10.1061/40552(301)14.
http://dx.doi.org/10.1061/40552(301)14...
reported that the wave velocity of cement-treated soil specimens was higher than the wave velocity of the natural specimens.

Table 3
Variation of wave velocity values of soil specimens with cement content and curing times.

Further, the cementing compounds and the unreacted cement help to filling the voids among soil particles, resulting to create other paths with short traveling times. This behavior increases the wave velocity values of soil specimens.

3.6 Gas permeability results

Gas permeability is the capacity of soil to allow air to flow in the existence of a pressure gradient. In this research, gas permeability is used as a pointer of the structural changes of soil specimens. The use of gas permeability rather than water permeability avoids the interaction of water with the soil-cement mixtures. The variations of coefficient of gas permeability (Ka) values with both cement content and curing times were illustrated in Figure 8. In general, the Ka of soil specimens decreased with increasing both cement content and curing times. The values of Ka decreased from (2.2 × 10-13 m2) of natural soil to (8.9 × 10-14, 7.6 × 10-14 and 6.8 × 10-14 m2) of soil specimens treated with 2, 4 and 6% cement content respectively, and cured for 3 days. While the values of soil specimens cured for 90 days were (4.9 × 10-15, 2.4 × 10-15, and 8.4 × 10-16 m2) of soil specimens treated with 2, 4, and 6% cement content respectively. It is well known that the voids and pores (macropores and micropores) of soil specimens play a major role in the gas permeability values (Aldaood et al., 2016Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2016). Effect of water during freeze-thaw cycles on the performance and durability of lime-treated gypseous soil. Cold Regions Science and Technology, 123, 155-163. http://dx.doi.org/10.1016/j.coldregions.2015.12.008.
http://dx.doi.org/10.1016/j.coldregions....
; Wang et al., 2017Wang, Y., Cui, Y.J., Tang, A.M., Benahmed, N., & Duc, M. (2017). Effects of aggregate size on the compressibility and air permeability of lime-treated fine-grained soil. Engineering Geology, 228, 167-172. http://dx.doi.org/10.1016/j.enggeo.2017.08.005.
http://dx.doi.org/10.1016/j.enggeo.2017....
). As discussed previously in section (3.2), the PSD of soil specimens mainly affected by cement content and curing times. Before cement addition (i.e. natural soil), the pores available for gas flow are larger (see Figures 1 and 2), resulting in larger values of Ka. Again, the Ka is a function of two parameters: the porosity and the interconnectivity between the pores (Aldaood et al., 2016Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2016). Effect of water during freeze-thaw cycles on the performance and durability of lime-treated gypseous soil. Cold Regions Science and Technology, 123, 155-163. http://dx.doi.org/10.1016/j.coldregions.2015.12.008.
http://dx.doi.org/10.1016/j.coldregions....
). When the cement was added and the soil specimens cured for different times, both porosity and the interconnectivity between the pores decreased due to the formation of cementing compounds during the pozzolanic reactions. As a result, the Ka decreased with both cement content and curing times. Further, it is observed that the decrease in Ka values from the value of natural specimens to the 60 days of curing was more pronounced than the decrease from 60 days to 90 days of curing. This behavior is attributed to the formation of most cementing compounds (as discussed previously) and these compounds will bound the soil grains and hinder the gas flow in soil specimens. Therefore, the Ka values of soil specimens decreased. Another reason to decrease the Ka values (especially at short curing times) of soil specimens was the unreacted cement particles which act as a filler and fill the voids among soil particles, leading to enclosed the voids and decreasing the gas permeability.

Figure 8
Gas permeability of soil specimens with (A) cement content and (B) curing times.

3.7 Soil-water retention behavior

The soil-water retention curves (SWRC) referring to both cement content and curing times were plotted together to comment on the general shape of the SWRCs and whether these curves affected by the cement content and curing times. Figure 9 presents the influence of the cement content and curing times on the SWRCs in terms of suction pressure and volumetric water content. In general, the cement addition and curing times have an insignificant influence on the shape of the SWRC, and all curves having an S-shape curve. For all cement contents, the SWRCs of soil specimens cured for 90 days were lie above the other curves (see Figure 9). This behavior was attributed to high capillary and absorptive forces resulting from finer soil structure (Aldaood, 2020Aldaood, A. (2020). Impact of fine materials on the saturated and unsaturated behavior of silty sand soil. Ain Shams Engineering Journal, 11(3), 717-725. http://dx.doi.org/10.1016/j.asej.2019.11.005.
http://dx.doi.org/10.1016/j.asej.2019.11...
). Moreover, the influence of curing times on the SWRCs was larger at low suction pressure than at high suction pressure. Another interesting observation from Figure 9 is that there was a continuous reduction in the volumetric water content of soil specimens with increasing suction pressure. This reduction was found to be dependent on both cement content and curing times as presented in Table 4. The differences between the volumetric water content values of soil specimens were obvious at suction pressure lower than 1500 kPa. While at suction pressure larger than 1500 kPa the differences in values were slight particularly for soil specimens cured at 60 and 90 days. This behavior is attributed to the more pozzolanic reactions in the soil specimens. Certainly, increasing curing time promotes the pozzolanic reaction within the soil mixture and resulting to the development of cementing materials (i.e. CSH and CAH), so that they help to the change in the PSD of soil specimens as discussed previously. Moreover, cement addition can help to enhance the microstructure properties of soil specimens, thus make the PSD more uniform and improving the water-holding performance of treated soil specimens (Jiang et al., 2019Jiang, X., Huang, Z., Ma, F., & Luo, X. (2019). Large-scale testing of clay soil improved with concrete pieces. Materials, 12(23), 3873. http://dx.doi.org/10.3390/ma12233873.
http://dx.doi.org/10.3390/ma12233873...
).

Figure 9
SWRCs of soil specimens with various cement content and curing times.
Table 4
Variation of volumetric water content values of soil specimens with cement content and curing times.

The key parameters of SWRC were established using the method suggested by Vanapalli et al. (1999)Vanapalli, S.K., Fredlund, D.G., & Pufahl, D.E. (1999). The influence of soil structure and stress history on the soil-water characteristics of a compacted till. Geotechnique, 49(2), 143-159. http://dx.doi.org/10.1680/geot.1999.49.2.143.
http://dx.doi.org/10.1680/geot.1999.49.2...
, as illustrated in Figure 10. The main zones (states) of the SWRC are saturated and residual zones. The saturation volumetric water content a) and the air entry value AEV a) represented the saturation state. While the residual volumetric water content r) and the corresponding residual suction pressure r) represented the residual state. As the suction pressures of soil specimens increased from 10 kPa to the AEV the volumetric water content of the soil specimens was approximately constant. Beyond the AEV, there was a continuous reduction in the volumetric water content of the soil specimens with increasing suction pressure. Further, the slope of the SWRCs of soil specimens cured for 90 days in part between the AEV and the residual water content was larger than the slope of other parts. This means that the soil structure was uniform and compact then resulting in better water holding capacity (Aldaood et al., 2014Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014). Soil-water characteristic curve of lime treated gypseous soil. Applied Clay Science, 102, 128-138. http://dx.doi.org/10.1016/j.clay.2014.09.024.
http://dx.doi.org/10.1016/j.clay.2014.09...
). Table 5 presents the saturation and residual state values of soil specimens with all cement contents and curing times. It is observed that both a) and r) increased with increasing cement content and curing times. The increase in a) was greater than the increase in r). The AEV showed insignificant changes with cement addition and curing times. Further, no obvious relationship was observed for the residual suction pressure with cement content and curing times. The difference in saturated and residual states values with cement content and curing times reveals the mineralogical and microstructural variations in soil specimens as discussed in section 3.2.

Figure 10
Typical SWCC showing the saturation, desaturation and residual zones (Vanapalli et al., 1999Vanapalli, S.K., Fredlund, D.G., & Pufahl, D.E. (1999). The influence of soil structure and stress history on the soil-water characteristics of a compacted till. Geotechnique, 49(2), 143-159. http://dx.doi.org/10.1680/geot.1999.49.2.143.
http://dx.doi.org/10.1680/geot.1999.49.2...
).
Table 5
Saturation and residual states values of soil specimens with cement content and curing times.

4. Conclusions

The following conclusions can be driven from this study:

• Increasing both cement content and curing times increased the strengths properties and wave velocity values of soil specimens. On the other hand, the gas permeability, pH, electrical conductivity values, and the failure strain were decreased with increasing curing times;

• The strength improvement of cement-treated soil specimens with curing times is divided into two zones: active and inert zones. Inactive zone the soil structure became compactness rather than the structure of natural soil and the cluster of grains become more effective. In the inert zone, there were insignificant changes in soil structure, thus there was a slight increase in strength value;

• Cement addition and curing times considerably modified the microstructural behavior of soil specimens. Cement content enhanced the volume and the morphology of pores (particularly the macropores), suggesting more cementing compounds formed and its action was more than the curing times;

• Interesting agreements between the microstructural, mechanical, and unsaturated hydraulic properties were obtained. Whereas the states of pores over time mainly affect the strength, gas permeability, and soil-water retention behavior;

• The AEV of soil specimens did not affect considerably with cement content and curing times. While the water holding capacity of soil specimens increased with these parameters. The most influences of cement content and curing times were on the part of SWRC with suction pressure smaller than 1500 kPa.

References

  • Abdulnafaa, M.D., Cabalar, A.F., & Arabash, Z. (2019). Shear strength characteristics of clay with waste solid construction and demolition materials. In Shear Strength Characteristics of Clay with Construction and Demolition Solid Waste Materials, Turkey.‎
  • Adeyanju, E.A., & Okeke, C.A. (2019). Clay soil stabilization using cement kiln dust. IOP Conference Series: Materials Science and Engineering, 640, 012080. http://dx.doi.org/10.1088/1757-899X/640/1/012080
    » http://dx.doi.org/10.1088/1757-899X/640/1/012080
  • Al-Aghbari, M.Y., Mohamedzein, Y.E.A., & Taha, R. (2009). Stabilisation of desert sands using cement and cement dust. Proceedings of the Institution of Civil Engineers: Ground Improvement, 162(3), 145-151. http://dx.doi.org/10.1680/grim.2009.162.3.145
    » http://dx.doi.org/10.1680/grim.2009.162.3.145
  • Aldaood, A. (2020). Impact of fine materials on the saturated and unsaturated behavior of silty sand soil. Ain Shams Engineering Journal, 11(3), 717-725. http://dx.doi.org/10.1016/j.asej.2019.11.005
    » http://dx.doi.org/10.1016/j.asej.2019.11.005
  • Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014). Soil-water characteristic curve of lime treated gypseous soil. Applied Clay Science, 102, 128-138. http://dx.doi.org/10.1016/j.clay.2014.09.024
    » http://dx.doi.org/10.1016/j.clay.2014.09.024
  • Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014a). Geotechnical properties of lime-treated gypseous soils. Applied Clay Science, 88–89, 39-48. http://dx.doi.org/10.1016/j.clay.2013.12.015
    » http://dx.doi.org/10.1016/j.clay.2013.12.015
  • Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2014b). Impact of wetting-drying cycles on the microstructure and mechanical properties of lime-stabilized gypseous soils. Engineering Geology, 174, 11-21. http://dx.doi.org/10.1016/j.enggeo.2014.03.002
    » http://dx.doi.org/10.1016/j.enggeo.2014.03.002
  • Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2015). Soil-water characteristic curve of gypseous soil. Geotechnical and Geological Engineering, 33(1), 123-135. http://dx.doi.org/10.1007/s10706-014-9829-5
    » http://dx.doi.org/10.1007/s10706-014-9829-5
  • Aldaood, A., Bouasker, M., & Al-Mukhtar, M. (2016). Effect of water during freeze-thaw cycles on the performance and durability of lime-treated gypseous soil. Cold Regions Science and Technology, 123, 155-163. http://dx.doi.org/10.1016/j.coldregions.2015.12.008
    » http://dx.doi.org/10.1016/j.coldregions.2015.12.008
  • Al-Mukhtar, M., Lasledj, A., & Alcover, J.F. (2014). Lime consumption of different clayey soils. Applied Clay Science, 95, 133-145. http://dx.doi.org/10.1016/j.clay.2014.03.024
    » http://dx.doi.org/10.1016/j.clay.2014.03.024
  • Amadi, A.A., & Osu, A.S. (2018). Effect of curing time on strength development in black cotton soil: quarry fines composite stabilized with cement kiln dust (CKD). Journal of King Saud University - Engineering Sciences, 30(4), 305-312. http://dx.doi.org/10.1016/j.jksues.2016.04.001
    » http://dx.doi.org/10.1016/j.jksues.2016.04.001
  • Amu, O., Fajobi, A., & Afekhuai, S. (2008). Stabilizing potential of cement-fly ash mixture on expansive clay Soil. Journal of Technology and Education in Nigeria, 12(2), http://dx.doi.org/10.4314/joten.v12i2.35698
    » http://dx.doi.org/10.4314/joten.v12i2.35698
  • ASTM 6931. (2011). Standard Test Method for Indirect Tensile (IDT) Strength of Bituminous Mixtures (Vol. 1, pp. 3-7). ASTM International, West Conshohocken, PA. https://doi.org/10.1520/D6931-17.2
    » https://doi.org/10.1520/D6931-17.2
  • ASTM D 698. (2003). Standard test methods for laboratory compaction characteristics of soil using standard effort (12 400 ft-lbf/ft3 (600 kN-m/m3)). In ASTM International, The annual book of ASTM standards (Vol. 3, pp. 1-11). ASTM, West Conshohocken, PA. https://doi.org/10.1520/D0698-12E01.1
    » https://doi.org/10.1520/D0698-12E01.1
  • ASTM (2000). Annual book of ASTM standards. Soil and rock, Vol. 04.08. Philadelphia: American Society for Testing and Materials.
  • Attom, M.F., Taqieddin, S.A., & Mubeideen, T. (2000). Shear strength and swelling stabilization of unsaturated clayey soil using pozzolanic material. In C.D. Shackelford, S.L. Houston & N.-Y. Chang (Eds.), Advances in unsaturated geotechnics (pp. 275-288). Reston, VA: American Society of Civil Engineers.
  • Ayeldeen, M., & Kitazume, M. (2017). Using fiber and liquid polymer to improve the behaviour of cement-stabilized soft clay. Geotextiles and Geomembranes, 45(6), 592-602. http://dx.doi.org/10.1016/j.geotexmem.2017.05.005
    » http://dx.doi.org/10.1016/j.geotexmem.2017.05.005
  • Azam, A.M., & Cameron, D.A. (2013). Geotechnical properties of blends of recycled clay masonry and recycled concrete aggregates in unbound pavement construction. Journal of Materials in Civil Engineering, 25(6), 788-798. http://dx.doi.org/10.1061/(ASCE)MT.1943-5533.0000634
    » http://dx.doi.org/10.1061/(ASCE)MT.1943-5533.0000634
  • Cabalar, A., Abdulnafaa, M., & Isik, H. (2019a). The role of construction and demolition materials in swelling of a clay. Arabian Journal of Geosciences, 12(11), 361. http://dx.doi.org/10.1007/s12517-019-4552-4
    » http://dx.doi.org/10.1007/s12517-019-4552-4
  • Cabalar, A., Zardikawi, O., & Abdulnafaa, M. (2019b). Utilisation of construction and demolition materials with clay for road pavement subgrade. Road Materials and Pavement Design, 20(3), 702-714. http://dx.doi.org/10.1080/14680629.2017.1407817
    » http://dx.doi.org/10.1080/14680629.2017.1407817
  • Cabalar, A.F., Abdulnafaa, M.D., & Karabash, Z. (2016). Influences of various construction and demolition materials on the behavior of a clay. Environmental Earth Sciences, 75(9), 841. http://dx.doi.org/10.1007/s12665-016-5631-4
    » http://dx.doi.org/10.1007/s12665-016-5631-4
  • Cabalar, A.F., Hassan, D.I., & Abdulnafaa, M.D. (2017). Use of waste ceramic tiles for road pavement subgrade. Road Materials and Pavement Design, 18(4), 882-896. http://dx.doi.org/10.1080/14680629.2016.1194884
    » http://dx.doi.org/10.1080/14680629.2016.1194884
  • Chen, H., & Wang, Q. (2006). The behaviour of organic matter in the process of soft soil stabilization using cement. Bulletin of Engineering Geology and the Environment, 65(4), 445-448. http://dx.doi.org/10.1007/s10064-005-0030-1
    » http://dx.doi.org/10.1007/s10064-005-0030-1
  • Chenari, R.J., Fatahi, B., Ghorbani, A., & Alamoti, M.N. (2018). Evaluation of strength properties of cement stabilized sand mixed with EPS beads and fly ash. Geomechanics and Engineering, 14(6), 533-544. http://dx.doi.org/10.12989/gae.2018.14.6.533
    » http://dx.doi.org/10.12989/gae.2018.14.6.533
  • Chew, S.H., Kamruzzaman, A.H.M., & Lee, F.H. (2004). Physicochemical and engineering behavior of cement treated clays. Journal of Geotechnical and Geoenvironmental Engineering, 130(7), 696-706. http://dx.doi.org/10.1061/(ASCE)1090-0241(2004)130:7(696)
    » http://dx.doi.org/10.1061/(ASCE)1090-0241(2004)130:7(696)
  • Consoli, N., Arcari Bassani, M., & Festugato, L. (2010). Effect of fiber-reinforcement on the strength of cemented soils. Geotextiles and Geomembranes, 28(4), 344-351. http://dx.doi.org/10.1016/j.geotexmem.2010.01.005
    » http://dx.doi.org/10.1016/j.geotexmem.2010.01.005
  • Consoli, N., Cruz, R., & Floss, M. (2011). Variables controlling strength of artificially cemented sand: influence of curing time. Journal of Materials in Civil Engineering, 23(5), 692-696. http://dx.doi.org/10.1061/(ASCE)MT.1943-5533.0000205
    » http://dx.doi.org/10.1061/(ASCE)MT.1943-5533.0000205
  • Consoli, N., Faro, V., Schnaid, F., & Born, R. (2017). Stabilised soil layers enhancing performance of transverse-loaded flexible piles on lightly bonded residual soils. Soils and Rocks, 40(3), 219-228. http://dx.doi.org/10.28927/SR.403219
    » http://dx.doi.org/10.28927/SR.403219
  • Das, B.M., Yen, S.C., & Dass, R.N. (1995). Brazilian tensile strength test of lightly cemented sand‎. Canadian Geotechnical Journal, 32(1), 166-171. http://dx.doi.org/10.1139/t95-013
    » http://dx.doi.org/10.1139/t95-013
  • Eades, J.L., & Grim, R.E. (1966). A quick test to determine lime requirements for lime stabilization‎. Highway Research Record, (139), 61-72.
  • Feng, T.W., Lee, J.Y., & Lee, Y.J. (2001). Consolidation behavior of a soft mud treated with small cement content. Engineering Geology, 59(3-4), 327-335. http://dx.doi.org/10.1016/S0013-7952(01)00021-7
    » http://dx.doi.org/10.1016/S0013-7952(01)00021-7
  • Goodary, R., Lecomte-Nana, G.L., Petit, C., & Smith, D.S. (2012). Investigation of the strength development in cement-stabilised soils of volcanic origin. Construction & Building Materials, 28(1), 592-598. http://dx.doi.org/10.1016/j.conbuildmat.2011.08.054
    » http://dx.doi.org/10.1016/j.conbuildmat.2011.08.054
  • Horpibulsuk, S., Phojan, W., Suddeepong, A., Chinkulkijniwat, A., & Liu, M.D. (2012). Strength development in blended cement admixed saline clay. Applied Clay Science, 55, 44-52. http://dx.doi.org/10.1016/j.clay.2011.10.003
    » http://dx.doi.org/10.1016/j.clay.2011.10.003
  • Horpibulsuk, S., Rachan, R., & Raksachon, Y. (2009). Role of fly ash on strength and microstructure development in blended cement stabilized silty clay. Soil and Foundation, 49(1), 85-98. http://dx.doi.org/10.3208/sandf.49.85
    » http://dx.doi.org/10.3208/sandf.49.85
  • Horpibulsuk, S., Rachan, R., Chinkulkijniwat, A., Raksachon, Y., & Suddeepong, A. (2010). Analysis of strength development in cement-stabilized silty clay from microstructural considerations. Construction & Building Materials, 24(10), 2011-2021. http://dx.doi.org/10.1016/j.conbuildmat.2010.03.011
    » http://dx.doi.org/10.1016/j.conbuildmat.2010.03.011
  • Iravanian, A., & Bilsel, H. (2016). Strength characterization of sand-bentonite mixtures and the effect of cement additives. Marine Georesources and Geotechnology, 34(3), 210-218. http://dx.doi.org/10.1080/1064119X.2014.991463
    » http://dx.doi.org/10.1080/1064119X.2014.991463
  • İşbuğa, V., Çabalar, A., & Abdulnafaa, M. (2019). Large-scale testing of a clay soil improved with concrete pieces. Proceedings on Engineering Sciences B, 2, 95-100.
  • Jiang, X., Huang, Z., Ma, F., & Luo, X. (2019). Large-scale testing of clay soil improved with concrete pieces. Materials, 12(23), 3873. http://dx.doi.org/10.3390/ma12233873
    » http://dx.doi.org/10.3390/ma12233873
  • Joel, M., & Agbede, I.O. (2010). Cement stabilization of Igumale shale lime admixture for use as flexible pavement construction material. The Electronic Journal of Geotechnical Engineering, 15, 1661-1673.
  • Kalantari, B., & Huat, B.B.K. (2008). Peat soil stabilization, using ordinary portland cement, polypropylene fibers, and air curing technique. The Electronic Journal of Geotechnical Engineering, 13, 1-13.
  • Kalantari, B., & Prasad, A. (2014). A study of the effect of various curing techniques on the strength of stabilized peat. Transportation Geotechnics, 1(3), 119-128. http://dx.doi.org/10.1016/j.trgeo.2014.06.002
    » http://dx.doi.org/10.1016/j.trgeo.2014.06.002
  • Kalıpcılar, İ., Mardani-Aghabaglou, A., Sezer, G.İ., Altun, S., & Sezer, A. (2016). Assessment of the effect of sulfate attack on cement stabilized montmorillonite. Geomechanics and Engineering, 10(6), 807-826. http://dx.doi.org/10.12989/gae.2016.10.6.807
    » http://dx.doi.org/10.12989/gae.2016.10.6.807
  • Kasama, K., Ochiai, H., & Yasufuku, N. (2000). On the stress-strain behaviour of lightly cemented clay based on an extended critical state ‎concept‎. Soil and Foundation, 40(5), 37-47. http://dx.doi.org/10.3208/sandf.40.5_37
    » http://dx.doi.org/10.3208/sandf.40.5_37
  • Kenai, S., Bahar, R., & Benazzoug, M. (2006). Experimental analysis of the effect of some compaction methods on mechanical properties and durability of cement stabilized soil. Journal of Materials Science, 41(21), 6956-6964. http://dx.doi.org/10.1007/s10853-006-0226-1
    » http://dx.doi.org/10.1007/s10853-006-0226-1
  • Khattab, S.I., & Aljobouri, M.M. (2012). Effect of Combined Stabilization by Lime and Cement on Hydraulic Properties of Clayey Soil Selected From Mosul Area. Al-Rafidain Engineering Journal, 20(6), 139-153.
  • Khemissa, M., & Mahamedi, A. (2014). Cement and lime mixture stabilization of an expansive overconsolidated clay. Applied Clay Science, 95, 104-110. http://dx.doi.org/10.1016/j.clay.2014.03.017
    » http://dx.doi.org/10.1016/j.clay.2014.03.017
  • Korf, E.P., Prietto, P.D.M., & Consoli, N.C. (2017). Hydraulic and diffusive behavior of a compacted cemented soil. Soils and Rocks, 39(3), 325-331.
  • Lemaire, K., Deneele, D., Bonnet, S., & Legret, M. (2013). Effects of lime and cement treatment on the physicochemical, microstructural and mechanical characteristics of a plastic silt. Engineering Geology, 166, 255-261. http://dx.doi.org/10.1016/j.enggeo.2013.09.012
    » http://dx.doi.org/10.1016/j.enggeo.2013.09.012
  • Lorenzo, G.A., & Bergado, D.T. (2004). Fundamental parameters of cement-admixed clay - New approach. Journal of Geotechnical and Geoenvironmental Engineering, 130(10), 1042-1050. http://dx.doi.org/10.1061/(ASCE)1090-0241(2004)130:10(1042)
    » http://dx.doi.org/10.1061/(ASCE)1090-0241(2004)130:10(1042)
  • Mandal, T., Tinjum, J.M., & Edil, T.B. (2016). Non-destructive testing of cementitiously stabilized materials using ultrasonic pulse velocity test. Transportation Geotechnics, 6, 97-107. http://dx.doi.org/10.1016/j.trgeo.2015.09.003
    » http://dx.doi.org/10.1016/j.trgeo.2015.09.003
  • Mengue, E., Mroueh, H., Lancelot, L., & Medjo Eko, R. (2017). Physicochemical and consolidation properties of compacted lateritic soil treated with cement. Soil and Foundation, 57(1), 60-79. http://dx.doi.org/10.1016/j.sandf.2017.01.005
    » http://dx.doi.org/10.1016/j.sandf.2017.01.005
  • Miller, G.A., & Azad, S. (2000). Influence of soil type on stabilization with cement kiln dust. Construction & Building Materials, 14(2), 89-97. http://dx.doi.org/10.1016/S0950-0618(00)00007-6
    » http://dx.doi.org/10.1016/S0950-0618(00)00007-6
  • Mousavi, S., & Leong Sing, W. (2015). Utilization of brown clay and cement for stabilization of clay. Jordan Journal of Civil Engineering, 9(2), 163-174.
  • Nahlawi, H., Chakrabarti, S., & Kodikara, J. (2004). A direct tensile strength testing method for unsaturated geomaterials. Geotechnical Testing Journal, 27(4), 356-361. http://dx.doi.org/10.1520/GTJ11767
    » http://dx.doi.org/10.1520/GTJ11767
  • Naseem, A., Mumtaz, W., Fazal-e-Jalal, & De Backer, H. (2019). Stabilization of expansive soil using tire rubber powder and cement kiln dust. Soil Mechanics and Foundation Engineering, 56(1), 54-58. http://dx.doi.org/10.1007/s11204-019-09569-8
    » http://dx.doi.org/10.1007/s11204-019-09569-8
  • Nayak, S., & Sarvade, P.G. (2012). Effect of cement and quarry dust on shear strength and hydraulic characteristics of lithomargic Clay. Geotechnical and Geological Engineering, 30(2), 419-430. http://dx.doi.org/10.1007/s10706-011-9477-y
    » http://dx.doi.org/10.1007/s10706-011-9477-y
  • Okyay, U.S., & Dias, D. (2010). Use of lime and cement treated soils as pile supported load transfer platform. Engineering Geology, 114(1-2), 34-44. http://dx.doi.org/10.1016/j.enggeo.2010.03.008
    » http://dx.doi.org/10.1016/j.enggeo.2010.03.008
  • Osinubi, K.J., Oyelakin, M.A., & Eberemu, A.O. (2011). Improvement of black cotton soil with ordinary portland cement - locust bean waste ash blend. The Electronic Journal of Geotechnical Engineering, 16, 619-627.
  • Pakbaz, M.S., & Alipour, R. (2012). Influence of cement addition on the geotechnical properties of an Iranian clay. Applied Clay Science, 67-68, 1-4. http://dx.doi.org/10.1016/j.clay.2012.07.006
    » http://dx.doi.org/10.1016/j.clay.2012.07.006
  • Park, S.S. (2011). Unconfined compressive strength and ductility of fiber-reinforced cemented sand. Construction & Building Materials, 25(2), 1134-1138. http://dx.doi.org/10.1016/j.conbuildmat.2010.07.017
    » http://dx.doi.org/10.1016/j.conbuildmat.2010.07.017
  • Petchgate, K., Sukmongkol, W., & Voottipruex, P. (2001). Effect of height and diameter ratio on the strength of cement stabilized soft Bangkok ‎clay‎. Journal of Geotechnical Engineering, 31(3), 227-239.
  • Riaz, S., Aadil, N., & Waseem, U. (2014). Stabilization of subgrade soils using cement and lime: a case study of Kala Shah Kaku, Lahore, Pakistan. Pakistan Journal of Science, 66(1), 39-44. Retrieved in March 1, 2014, from http://aplicacionesbiblioteca.udea.edu.co:3653/ehost/detail/detail?vid=11&sid=9827d053-420e-4669-9575-5e6211a04bbc@sessionmgr112&hid=125&bdata=Jmxhbmc9ZXMmc2l0ZT1laG9zdC1saXZl#db=a9h&AN=98360899
    » http://aplicacionesbiblioteca.udea.edu.co:3653/ehost/detail/detail?vid=11&sid=9827d053-420e-4669-9575-5e6211a04bbc@sessionmgr112&hid=125&bdata=Jmxhbmc9ZXMmc2l0ZT1laG9zdC1saXZl#db=a9h&AN=98360899
  • Rojas-Suárez, J.P., Orjuela-Abril, M.S., & Prada-Botía, G. (2019a). Study of low plasticity clay for optimum dosage of the soil-cement. Journal of Physics: Conference Series, 1386(1), 012078. http://dx.doi.org/10.1088/1742-6596/1386/1/012078
    » http://dx.doi.org/10.1088/1742-6596/1386/1/012078
  • Rojas-Suárez, J.P., Orjuela-Abril, M.S., & Prada-Botía, G.C. (2019b). Determination of the adequate dosage of the soil-cement, using clay of high plasticity. Journal of Physics: Conference Series, 1386(1), 012077. http://dx.doi.org/10.1088/1742-6596/1386/1/012077
    » http://dx.doi.org/10.1088/1742-6596/1386/1/012077
  • Saadeldin, R., & Siddiqua, S. (2013). Geotechnical characterization of a clay-cement mix. Bulletin of Engineering Geology and the Environment, 72(3-4), 601-608. http://dx.doi.org/10.1007/s10064-013-0531-2
    » http://dx.doi.org/10.1007/s10064-013-0531-2
  • Saberian, M., Moradi, M., Vali, R., & Li, J. (2018). Stabilized marine and desert sands with deep mixing of cement and sodium bentonite. Geomechanics and Engineering, 14(6), 553-562. http://dx.doi.org/10.12989/gae.2018.14.6.553
    » http://dx.doi.org/10.12989/gae.2018.14.6.553
  • Saeed, K.A., Kassim, K.A., Nur, H., & Yunus, N.Z.M. (2015). Strength of lime-cement stabilized tropical lateritic clay contaminated by heavy metals. KSCE Journal of Civil Engineering, 19(4), 887-892. http://dx.doi.org/10.1007/s12205-013-0086-6
    » http://dx.doi.org/10.1007/s12205-013-0086-6
  • Sharma, L.K., Sirdesai, N.N., Sharma, K.M., & Singh, T.N. (2018). Experimental study to examine the independent roles of lime and cement on the stabilization of a mountain soil: a comparative study. Applied Clay Science, 152, 183-195. http://dx.doi.org/10.1016/j.clay.2017.11.012
    » http://dx.doi.org/10.1016/j.clay.2017.11.012
  • Shooshpasha, I., & Shirvani, R.A. (2015). Effect of cement stabilization on geotechnical properties of sandy soils. Geomechanics and Engineering, 8(1), 17-31. http://dx.doi.org/10.12989/gae.2015.8.1.017
    » http://dx.doi.org/10.12989/gae.2015.8.1.017
  • Umesha, T.S., Dinesh, S.V., & Sivapullaiah, P.V. (2009). Control of dispersivity of soil using lime and cement. International Journal of Geology, 3(1), 8-16.
  • Vanapalli, S.K., Fredlund, D.G., & Pufahl, D.E. (1999). The influence of soil structure and stress history on the soil-water characteristics of a compacted till. Geotechnique, 49(2), 143-159. http://dx.doi.org/10.1680/geot.1999.49.2.143
    » http://dx.doi.org/10.1680/geot.1999.49.2.143
  • Wang, Y., Cui, Y.J., Tang, A.M., Benahmed, N., & Duc, M. (2017). Effects of aggregate size on the compressibility and air permeability of lime-treated fine-grained soil. Engineering Geology, 228, 167-172. http://dx.doi.org/10.1016/j.enggeo.2017.08.005
    » http://dx.doi.org/10.1016/j.enggeo.2017.08.005
  • Wei, D., Zhu, B., Wang, T., Tian, M., & Huang, X. (2014). Effect of Cationic Exchange Capacity of Soil on Strength of Stabilized Soil. Procedia: Social and Behavioral Sciences, 141, 399-406. http://dx.doi.org/10.1016/j.sbspro.2014.05.070
    » http://dx.doi.org/10.1016/j.sbspro.2014.05.070
  • Yesiller, N., Hanson, J.L., & Usmen, M.A. (2000). Ultrasonic assessment of stabilized soils. In Soft Ground Technology Conference (Vol. 301, pp. 170-181), Noordwijkerhout, the Netherlands. http://dx.doi.org/10.1061/40552(301)14
    » http://dx.doi.org/10.1061/40552(301)14
  • Zhang, T., Yue, X., Deng, Y., Zhang, D., & Liu, S. (2014). Mechanical behaviour and micro-structure of cement-stabilised marine clay with a metakaolin agent. Construction & Building Materials, 73, 51-57. http://dx.doi.org/10.1016/j.conbuildmat.2014.09.041
    » http://dx.doi.org/10.1016/j.conbuildmat.2014.09.041

Publication Dates

  • Publication in this collection
    24 May 2021
  • Date of issue
    2021

History

  • Received
    27 June 2020
  • Accepted
    21 Oct 2020
Associação Brasileira de Mecânica dos Solos Av. Queiroz Filho, 1700 - Torre A, Sala 106, Cep: 05319-000, São Paulo - SP - Brasil, Tel: (11) 3833-0023 - São Paulo - SP - Brazil
E-mail: secretariat@soilsandrocks.com