ISSN 1854-0171 C ACTA GEOTECHNICA SLOVENICA ISSN: 1854-0171 Ustanovitelji Founders Univerza v Mariboru, Fakulteta za gradbeništvo, prometno inženirstvo in arhitekturo University of Maribor, Faculty of Civil Engineering, Transportation Engineering and Architecture Univerza v Ljubljani, Fakulteta za gradbeništvo in geodezijo University of Ljubljana, Faculty of Civil and Geodetic Engineering Univerza v Ljubljani, Naravoslovnotehniška fakulteta University of Ljubljana, Faculty of Natural Sciences and Engineering Slovensko geotehniško društvo Slovenian Geotechnical Society Društvo za podzemne in geotehniške konstrukcije Society for Underground and Geotechnical Constructions Izdajatelj Publisher Univerza v Mariboru, Fakulteta za gradbeništvo, prometno inženirstvo in arhitekturo Faculty of Civil Engineering, Transportation Engineering and Architecture Odgovorni urednik Editor-in-Chief Borut Macuh University of Maribor Tehnična urednica Technical Editor Tamara Bračko University of Maribor Uredniki Co-Editors Jakob Likar Janko Logar Primož Jelušič Stanislav Škrabl Milivoj Vulic Bojan Žlender Geoportal d.o.o. University of Ljubljana University of Maribor University of Maribor University of Ljubljana University of Maribor Posvetovalni uredniki Advisory Editors Heinz Brandl Chandrakant. S. Desai Bojan Majes Pedro Seco e Pinto Vienna University of Technology University of Arizona University of Ljubljana National Laboratory of Civil Eng. Lektor Proof-Reader Paul McGuiness Naklada Circulation 220 izvodov - issues Cena Price 25 EUR/letnik - 25 EUR/vol.; (50 EUR for institutions/za institucije) Tisk Print Tiskarna Saje d.o.o. Revija redno izhaja dvakrat letno. Članki v reviji so recenzirani s strani priznanih mednarodnih strokovnjakov. Baze podatkov v katerih je revija indeksirana: SCIE - Science Citation Index Expanded, JCR - Journal Citation Reports / Science Edition, ICONDA - The international Construction database, GeoRef. Izid publikacije je finančno podprla Javna agencija za raziskovalno dejavnost Republike Slovenije iz naslova razpisa za sofinanciranje domačih periodičnih publikacij. Uredniški odbor Marx Ferdinand Ahlinhan Amin Barari Theodoros Hatzigogos Vojkan Jovičič Rolf Katzenbach Nasser Khalili Svetlana Melentijevic Seyed Hamed Mirmoradi Ana Petkovšek Borut Petkovšek Mihael Ribičič César Sagaseta Patrick Selvadurai Stephan Semprich Devendra Narain Singh Abdul-Hamid Soubra Kiichi Suzuki Antun Szavits-Nossan Kosta Urumovic Ivan Vaniček Editorial Board National University in Abomey Aalborg University Aristotle University of Thessaloniki IRGO-Ljubljana Technical University Darmstadt The University of New South Wales, Sydney Complutense University of Madrid Federal University of Rio de Janeiro University of Ljubljana Slovenian National Building and Civil Engineering Institute University of Ljubljana University of Cantabria McGill University University of Technology Graz Indian Institute of Technology, Bombay University of Nantes Saitama University University of Zagreb Croatian geological survey Czech Technical University in Prague Založnik Univerzitetna založba Univerze v Mariboru Slomškov trg 15, 2000 Maribor, Slovenija e-pošta: zalozba@um.si, http://press.um.si/, http://j ourn als.um.si/ Published by University of Maribor Press Slomškov trg 15, 2000 Maribor, Slovenia e-mail: zalozba@um.si, http://press.um.si/, http://journals.um.si/ Naslov uredništva Address ACTA GEOTECHNICA SLOVENICA Univerza v Mariboru, Fakulteta za gradbeništvo, prometno inženirstvo in arhitekturo Smetanova ulica 17, 2000 Maribor, Slovenija Telefon / Telephone: +386 (0)2 22 94 300 Faks / Fax: +386 (0)2 25 24 179 E-pošta / E-mail: ags@um.si Spletni naslov web Address http://zalozba.um.si/index.php/ags/index The journal is published twice a year. Papers are peer reviewed by renowned international experts. Indexation data bases of the journal: SCIE - Science Citation Index Expanded, JCR - Journal Citation Reports / Science Edition, ICONDA- The international Construction database, GeoRef. The publication was financially supported by Slovenian Research Agency according to the Tender for co-financing of domestic periodicals. B. Janc and Ž. Vukelič Mud-pump pressure in geothermal wells B. Janc in Ž. Vukelič Tlak izplačne črpalke pri geotermalnih vrtinah J. A. Duque Felfle in drugi Razširjena enačba nosilnosti temeljnih tal pod plitvimi temelji na grobozrnatih zasekih v mehkih tleh J. A. Duque Felfle et al. An extended bearing-capacity equation for shallow foundations on granular trenches i soft soil D. ignjatovič Stupar in drugi Opazovanje obsežnih premikov mase v kamnolomu Lipica II s senzorjem premika žarka EL D. Ignjatovic Stupar et al. Observation of extensive mass movements in the Lipica II quarry with an EL beam displacement sensor 23 A. Lazar in drugi Vključevanje merilnih tehnik, ki se uporabljajo za spremljanje plazu Laze v Sloveniji A. Lazar et al. Integration of measurement techniques in monitoring of the Laze landslide in Slovenia F A. Pal in drugi Modeliranje in analiza digitalnih površinskih modelov z uporabo visoko ločljivih UAV slik A. Pal et al. Modeling and analysis of digital surface models using high-resolution UAV images K. Doumi in drugi Eksperimentalno raziskovanje vpliva relativnega efektivnega premera na mejno strižno trdnost delno zasičenih grobo zrnatih zemljin K. Doumi et al. Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils P. Deb in S. Kumar Pal Odnosa obtežba-posedek in obtežba-delitev obtežbe za temeljno ploščo na pilotih na slojevitih tleh P. Deb and S. Kumar Pal Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils 46 71 P. Li in drugi Numerična raziskava učinkov prostorskih vplivov in podpornih konstrukcij med izkopavanjem gradbene jame P. Li et al. Numerical investigation of the influence of spatial effects and supporting structures during pit excavation F Navodila avtorjem Instructions for authors 97 1QQ. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells MUD-PUMP PRESSURE IN TLAK IZPLAČNE ČRPALKE PRI GEOTHERMAL WELLS GEOTERMALNIH VRTINAH Blaž Janc Željko Vukelic University of Ljubljana, University of Ljubljana, Faculty of Natural Sciences and Engineering Faculty of Natural Sciences and Engineering Aškerčeva c. 12, 1000 Ljubljana, Slovenia Aškerčeva c. 12, 1000 Ljubljana, Slovenia E-mail: blaz.janc@ntf.uni-lj.si E-mail: zeljko.vukelic@ntf.uni-lj.si https://doi.Org/10.18690/actageotechslov.17.1.2-11.2020 DOI Keywords drilling hydraulics, pump pressure, drilling mud, rheological models, geothermal energy Abstract Rotary drilling is a mining method for the extraction and exploration of mineral resources. A significant pressure drop occurs during deep-well rotary drilling. This paper presents a procedure and a theoretical background of the working-pump-pressure determination for exploration geothermal borehole Sob-4g, located in Murska Sobota, NE Slovenia. We determined all the partial pressure drops in the mud-circulation system when drilling the final section of a 1201.15-m-deep borehole. For this, it is important to choose the correct rheological model that follows the behaviour of the drilling fluid. In the presented case, the Bingham plastic model was used. The aim of the paper's hydraulic analysis is to provide the optimal drilling parameters and therefore the maximum effects in deep-well drilling. We show that most of the drilling mud-pressure energy is consumed within the drill string and through the bit. The fluid-flow regime in the drill pipes, collars and drill bit is turbulent, while it is laminar in the annulus. 1 INTRODUCTION Deep wells are most commonly drilled for oil and gas extraction or the use of geothermal energy. One of the most important elements in the drilling process is the drilling mud. It is a fluid that performs numerous Ključne besede hidravlika vrtanja, tlak črpalke, izplaka, reološki modeli, geotermalna energija Izvleček Rotacijsko vrtanje je rudarska metoda za pridobivanje in raziskovanje mineralnih surovin. Med rotacijskim vrtanjem globokih vrtin prihaja do občutnega padca tlaka. V prispevku je predstavljen postopek in teoretično ozadje določanja delovnega tlaka črpalke za raziskovalno geotermalno vrtino Sob-4g, ki se nahaja v Murski Soboti, SV Slovenija. Določili smo vse delne padce tlaka v izplačnem sistemu pri vrtanju končnega odseka 1201,15 m globoke vrtine. Za to je pomembno izbrati pravilen reološki model, ki sledi obnašanju izplačnega fluida. V predstavljenem primeru je bil uporabljen Binghamov plastični model. Cilj hidravlične analize prispevka je zagotoviti optimalne parametre vrtanja in s tem največje učinke pri vrtanju globokih vrtin. Večina tlačne energije izplake se porabi znotraj drogovja in preko dleta. Režim pretoka izplačnega fluida v vrtalnem drogovju, težkem drogovju in dletu je turbulenten, medtem ko je v medprostoru laminaren. functions, including carrying the rock fragments to the surface, providing hydrostatic pressure in the well and cooling the drill bit. The mud circulation is provided by mud pumps via a suitable working pressure and flow rate. The pump pressure provides sufficient energy for the mud circulation, while the flow rate enables the transport of the drilled cuttings to the surface. 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells The drilling fluid circulates in the mud system. It starts at the surface and runs downwards through the drill string and the drill bit and upwards in the annulus between the drill string and the borehole wall, back to the surface. The pressure energy from mud pump is mostly consumed in overcoming the internal friction in the drill string and for the fluid acceleration through the bit nozzles. There is a significant local reduction of the diameter in the bit nozzles, which leads to a large increase in the fluid kinetic energy and, therefore, the fluid velocity. Consequently, the bit nozzles are areas with a drastic pressure drop. Increasing the kinetic energy at the drill bit means a higher hydraulic power, which reflects in a better cutting action and a more efficient cleaning of the drilled particles beneath the bit. Determining the working pump pressure (WPP) is an important element of the drilling hydraulics and represents an indispensable part of the deep-well design. The WPP is defined as the sum of all the partial pressure drops. Therefore, all the pressure drops in the mud circulating system have to be known. The WPP increases with an increasing drilling depth. 2 DRILLING PROCESS TECHNOLOGY_ A borehole is drilled using a drilling rig, which is a set of machines, devices and elements that are necessary for the drilling process. The drilling process is a sequence of the following operations: 1) Connecting the drill pipes, drill collars and the drill bit. 2) Lengthening the drill pipes and lowering the drill bit into the borehole bottom. 3) Drilling bit action (drilling) with the simultaneous carrying of drill cuttings to the surface. 4) Connecting additional drill pipes due to the bit progress in depth. 5) Rising of the drill string from the borehole (for example, due to the wear of the drill bit). The rock at the borehole bottom is cut using the rotation of the roller cone bit, which is attached to the end of the drill collars and affected by its load. The roller cone bit rotates together with the drill string, cutting the rock beneath it. The transport of the drilled cuttings from the borehole bottom to the surface goes via the mud fluid [1]. Three main components of the deep well drilling are: 1) Weight on the bit. 2) Bit rotation. 3) Circulation of the drilling fluid. 3 DRILLING FLUID The mud or the drilling fluid is a liquid used in the drilling process that continuously circulates from the surface downwards through the drill pipes and bit nozzles and upwards to the surface through the annular space between the drill pipes and the borehole wall. We can distinguish between the major and minor functions of the drilling fluids. The major functions are, in general, the removal of the drilled cuttings, the containment of the subsurface formation fluid pressures and the borehole stabilization. Minor functions, including the cooling and lubricating of the drill string and drill bit, preventing particle sedimentation at the mud-circulation stop, and reducing the weight during the drill string operations (buoyancy), aid in the formation evaluation and cleaning of the drill bit [2]. Removing the cuttings beneath the drill bit is necessary for progress in the drilling process. This is achieved by the flow of drilling fluid through the annular space between the borehole wall and the drill string. Removal of the cutting particles depends on the annular mud velocity, the mud rheological properties, the borehole deviation, the rotation of the drill string, the borehole eccentricity, the drilling rate and the size and shape of the drilled cuttings [2]. A sufficient mud density prevents the intrusion of the formation fluids into the borehole. The mud density is achieved with additives (for example, barite). Clay is added, for example, for a higher viscosity of the mud. The drilling regime is, in most cases, over-pressured, which means that the pressure gradient is greater than 9.8 kPa per meter of well depth. 4 MUD PUMPS A mud pump provides enough energy for the pressure of the fluid across the circulating system. Mud-pump engines can produce around 1600 kW of power with flow rates of up to 5000 l/min and pressures of up to 600 ■ 105 Pa. Generally, there are two types of reciprocating mud pumps: duplex and triplex. 4.1 Duplex pump The principle of a double-acting, two-cylinder (duplex) pump is schematically shown in Fig. 1. As the piston moves forward (to the right-hand side in Fig. 1), it discharges fluid through the open discharge valve. At same time, the intake valve is opened, allowing fluid to enter the chamber behind the piston. There is a reversible principle as the piston returns (see Fig. 2) [3]. 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells Figure 1. Duplex pump (piston moves forward). Figure 2. Duplex pump (piston moves backward). Theoretical volume when the piston moves forward: nd2L V, = (1) where V is the volume of the discharged fluid when the piston moves forward (m3), d is the inner diameter of the piston cylinder (m) and L is the piston stroke (m). When the piston returns, the theoretical volume of the discharged fluid is: V, = u(d2 - df)L (2) where V2 is the volume of the discharged fluid when the piston moves backwards (m3) and dr is the piston-rod diameter (m). The total volume of discharged fluid in one crankshaft stroke is: 2u(2d2 - d2)Lr]v V = 2(V1 + V2)I1V = (3) 4.2 Triplex pump The principle of a single-acting, three-cylinder (triplex) pump is schematically shown in Fig. 3. With this pump the piston discharges fluid in only one direction. Figure 3. Triplex pump (piston moves forwards). Figure 4. Triplex pump (piston moves backwards). The total volume of discharged fluid in one crankshaft stroke is: 3nd2Lt]v V = ■ 5 RHEOLOGICAL MODELS (4) where r¡v is the volumetric efficiency of the pump (/). There are significant resistance forces to overcome in a mud-circulation system. Rheological models, used as an approximation for the fluid behaviour, are in general Newtonian (linear) and non-Newtonian (non-linear). These models are used to derive the pressure-drop equations. 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells 5.1 Newtonian model A Newtonian fluid is ideally viscous. It follows a linear relation between the shear stress (t) and the speed of the shear deformation-shear rate (y). t = hy (5) where tis the shear stress (Pa), y is the shear rate (s-1) and p is the Newtonian viscosity (Pa s). Examples of a Newtonian fluid are water, oil and gas. The linear relationship between the shear stress and the shear rate is valid only as long as the fluid moves in layers or laminae. This is true only at relatively low rates of shear. When turbulent flow occurs, pressure drops have to be determined with empirical correlations [4]. 5.2 Non-Newtonian models Non-Newtonian fluids are real. They do not exhibit a direct proportionality between the shear stress and the shear rate [4]. Shear-dependent non-Newtonian fluids are pseudoplastic (shear thinning) if the apparent viscosity decreases with an increasing shear stress and dilatant (shear thickening) if the apparent viscosity increases with an increasing shear stress. Drilling fluids and cement slurries are generally thixotropic, which means that they are pseudoplastic and have a time-dependent viscosity [4]. Fluids with plastic flow behaviour (Bingham fluids) start to flow when the limit shear stress is exceeded. When the shear flow is established, fluids with plastic flow behaviour show a linear dependence of the shear stress and the shear rate. In this case the Newtonian model is valid. The Bingham model represents a rigid matter that starts to flow as a viscous fluid when the yield strength is exceeded. After this, it behaves as a Newtonian fluid. A Bingham fluid is typical for bentonite muds. The rheo-logical model is defined by: T = UplY + To (6) where ppi is the plastic viscosity (Pa s) and t0 is the yield strength (Pa). The Ostwald-de Waele model is defined as a power law: my (7) Like in the Bingham model, the equation consists of two parameters: K and n. A higher K means a higher viscosity of the fluid. The deviation of parameter n from 1 is a criterion for the fluid deviation from a Newtonian fluid. In the case when n=1, the fluid follows the Newtonian law, if K=p. It behaves pseudoplastically (viscosity decreases with increasing shear stress) when n<1 and dilatantly (viscosity increases with increasing shear stress) when n>1. The equation is valid in the laminar-flow region. Shear rate, y Figure 5. Rheological models. 6 PRESSURE DROPS where K is the flow-consistency index (Pa s) and n is the flow-behavior index (/). 6.1 Generally about pressure drops When flowing in a pipe, a fluid losses part of its energy due to the friction/resistance forces. These forces are internal friction due to the viscosity and external friction due to pipe roughness [5]. A circulating drilling mud has an initial energy represented by the pump-discharge pressure. This energy is totally lost in the mud circuit. The mud pressure is zero when it returns to the pits. In this case, the pumpdischarge pressure represents the total pressure losses in the mud circuit [5]. When drilling, pressure drops occur in the following areas: surface equipment, drill pipes and drill collars, drill bit and annulus between the well bore and the drill string. Pressure drops in the drill string and annulus do not directly contribute to the drilling process, but cannot be avoided if the fluid is to be circulated around the system. Pressure drops in drill bit, on the other hand, do perform a useful function, since it helps to cut rock and clean the drilled cuttings from the face of the bit. 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells It is, therefore, desirable to optimize the pressure drops through the nozzles (and therefore the cleaning of the bit face) and minimize the drops in the drill string and annulus [3]. The pump pressure (pp) is expended by: frictional pressure losses in the surface equipment (ps), frictional pressure losses in the drill pipes (pdp), frictional pressure losses in the drill collars (pdc), pressure losses through the bit nozzles (pb), frictional pressure losses in the drill-collar annulus (pdca) and frictional pressure losses in the drill-pipe annulus (pdpa) [4]. Pp = Aps + Apdp + Apdc + Apb + Apdca + &Pdpa (8) The total frictional pressure loss can be represented as pj: Pp = Pb+ Pf (9) It is evident from the equation that the working pump pressure is consumed for the fluid acceleration in the bit nozzles and overcoming the flow resistance in drill pipes and the annulus. Pressure drops depend on the rheological properties of the mud, the flow type (laminar or turbulent) and the geometry of the pipes and the well bore. r r¡ ^ 1 W ( Figure 6. Mud-velocity profile for flow in a pipe (r2 is the radius of the pipe). r ' r. ( / r> t ¡ü? ) Figure 7. Mud-velocity profile for the flow in an annulus (r\ is the radius of the pipe, r2 is the radius of the borehole). 6.2 Laminar flow in the drill string and annulus The flow type, within which the fluid flows in the drill string or annulus, depends on the Reynolds number (Nrc). It is defined as (for pipe flow): Nb„_ = pvdj V- (10) where p is the mud fluid density (kg/m3), v is the fluid average velocity (m/s) and d, is the pipe inner diameter (m). For annular flow: N, Re„ pv(db - dp) (11) where db is the borehole diameter (m) and d0 is the pipe outer diameter (m). In the case when the density and viscosity of the drilling mud are constant, the Reynolds number depends only on the pipe diameter and the mud velocity. If the flow rate is constant, the Reynolds number depends only on the pipe diameter. The value of the Reynolds number is not constant across the whole mud system, but it changes. Thus, the mud flow can be laminar at one point and turbulent at another. The Newtonian fluid flow is laminar if NRe is less than 2100 and turbulent if NRe is more than 2100. Actually, when NRe values are in region 2000-4000, the flow is in a transition between laminar and turbulent flow. Fluid flow in the drill string or annulus do not have a uniform velocity. In the case of laminar flow, the fluid velocity by the wall pipe equals zero. The velocity is the highest at the maximum distance from the pipe wall, which is in the center of the pipe. Figure 8. Laminar and turbulent velocity profiles. 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells Velocity profiles for laminar flow in the pipe annulus are shown in Fig. 6 and Fig. 7. Pipe flow represents the flow from the surface to the bit and annular flow represents the flow from the drill bit back to the surface. Fig. 8 shows the velocity profile of a circular pipe (laminar and turbulent flow). The maximum velocity in the case of turbulent flow is vmax,t = and in the case of laminar flow vmax,i = 2vav , where vav represents the average velocity [6]. 6.3 Turbulent flow in the drill string and annulus High velocities of flow rates mean that the fluid does not flow in layers, but in a chaotic way. Turbulent flow can be divided into three regions: the laminar flow, the transition between laminar and turbulent flow and the turbulent core. Figure 9. Turbulent flow in a pipe (1 thin layer of laminar flow, 2 transition layer, 3 turbulent core; d is the diameter of the pipe). Figure 10. Laminar and turbulent flow patterns in a circular pipe (a - laminar flow, b - transition between laminar and turbulent flow, c - turbulent flow). When calculating pressure drops in drill pipes and annulus, the type of fluid must be known (Newtonian, non-Newtonian). The flow type is determined with a calculation of the critical velocity, which depends on the rheological parameters of the mud. There are some assumptions in calculating the pressure drops. These are: drill pipes are placed into the well concentric, drill pipes do not rotate, well bore is circular in shape with known diameter, mud fluid is incompressible, the flow is isothermal and the pipes are smooth. For this reason, the equations contain some experimentally determined factors. 6.4 Pressure drop in drill string and annulus The equations for the pressure-drop calculation in the pipe and the annular space are given as follows, according to the Fanning equations [7]. For pipe flow: A Pi = 2fpLv2 di (12) where Ap, is the pipe pressure drop (Pa), f is the hydraulic friction factor (/) and v is the flow rate (m/s). For annular flow: Apa = 2fpLv2 (13) db - d0 where Apa is the annular pressure drop (Pa). 6.5 Pressure drop in bit nozzles The purpose of bit nozzles is a better cleaning action of the drilling fluid at the bottom of the hole. Because of the small diameter of the bit nozzles, fluids reach high velocities inside the nozzle [8]. The nozzle velocity is defined as: vn= J" (14) where vn is the nozzle velocity (m/s), q is the mud flow rate (m3/s) and At is the total nozzle area (m2). Figure 11. Drill bit nozzle (v0 and p1 are the mud fluid input velocity and pressure, vn and p2 are the mud fluid output velocity and pressure). 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells Figure 12. Drill bit with three nozzles. bentonite drilling mud and in second section (580.731201.15 m) polymer drilling mud were used. The EGB Sob-4g location belongs to one of six regional numerical models of groundwater flow in Slovenia [11]. 7.1 Borehole construction The borehole is constructed in two stages. The first stage consists of drilling the first section (from 0.00 m to 580.73 m) and the casing. The second stage consists of drilling the second section (from 580.73 m to 1201.15 m). Table 1. Drilled intervals. Fluid flow through bit nozzles is shown in Fig. 11. The input velocity v0 is negligible compared to the output velocity v„. Fluid acceleration occurs due to the local reduction of the diameter in the nozzles. Kinetic energy increases, while the pressure energy decreases. Consequently, a pressure drop occurs at the bit nozzles. The pressure drop is defined by [8]: &Pb = PR 2C2A[ (15) where C is the nozzle-discharge coefficient (/). 7 EXPERIMENTAL_ An exploratory geothermal borehole (EGB) Sob-4g was constructed with the aim of exploring potential geothermal aquifers for the reinjection of cooled thermo-mineral water from the Sob-3g borehole back into the production aquifers. A reinjection system, which consist of a geothermal reinjection well and the surface reinjection unit, is a necessary part for returning the water to the production aquifer [9]. EGB Sob-4g is located in Murska Sobota close to the existing geothermal boreholes Sob-1 and Sob-2, which exploit the thermo-mineral water for district heating and balneology. In general, geothermal energy in Slovenia is utilized for individual space heating, district heating, cooling, greenhouse heating, bathing, swimming and snow melting [10]. EGB Sob-4g is 1201.15 m deep. Drilling has been made through geological formations of Mura and Lendava. Drilling has stopped at the upper part of the Murska Sobota formation. In the first section (0-580.73 m) Section Drilled interval (m) Bit diameter (mm) 1 0.00-580.73 444.50 2 580.73-1201.15 311.20 Table 2. Casing intervals. Section Casing interval (m) Casing diameter (mm) 1 0.00-578.25 339.70 2 530.30-1201.00 177.80 Table 3. Drilling equipment for second stage drilling. Drilling Equipment Length (m) drill bit 311.15 mm HUGES, IADC 135 GTX-G3; nozzles 3x10.32 mm 0.30 1. stabilizer 0 311.15 mm 1.21 drill collar 0 165.1 mm 8.47 2. stabilizer 0 311.15 mm 1.49 drill collar 0 165.1 mm 8.90 3. stabilizer 0 311.15 mm 1.39 drill collar 0 165.1 mm 78.98 transition pipe 101.6 mm IF (male) x 101.6 mm IF (female) 0.28 drill pipe 0 127.00 mm, 29 kg/m, 114.3 mm IF 524.49 transition pipe 114.3 mm IF (male) x 101.6 mm FH (female) 0.61 drill pipe 0 101.60 mm, 22.6 kg/m, 101.6 mm FH 568.83 kelly hexagonal; thread 88.9 mm IF (male) 8.2 All the pressure drops during drilling at the final depth (1201.15 m) are determined. At the final depth the maxi- 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells mum pressure drop that the mud pump has to overcome is expected. In Table 3 the drilling equipment in the borehole is listed. Individual tools are described from the borehole bottom to the top. It means that the drill bit is located at the borehole bottom, followed by sections of stabilizers and drill collars, a transition pipe, drill pipes and a kelly at the borehole surface. Table 4. Calculation input data - borehole construction. Hole/ Sec- Length Casing Drill Drill String Drill String - inner , , String - outer tion (m) diameter , , type diameter diameter (m) (m) (m) 1 525 0.318 drill pipe 0.127 0.109 2 53 0.318 drill pipe 0.102 0.085 3 516 0.311 drill pipe 0.102 0.085 4 97 0.311 drill collar 0.165 0.076 Table 4 represents the input data for calculation of the pressure drops. The borehole is divided into four sections. The first three sections are equipped with drill pipes and the fourth section with drill collars. Table 5. Calculation input data - - mud properties. p (kg/m3) 1150 T0 (Pa) 13 ^ (10-3 Pa s) 22 q (m3/s) 0.0183 For the working pump pressure determination, all the partial pressure drops have to be calculated. The pressure drop in the surface equipment is estimated to be 3-105 Pa (from Drilling Data Handbook). The calculation of all the other pressure drops is made relating to the calculation input data, shown in Table 4 and Table 5. Equations for the Bingham rheological model are used. 8 RESULTS_ The pressure drop in the drill string (drill pipes and drill collars) is 18.741-105 Pa, in the annular space 3.239-105 Pa, in the drill bit 38.427-105 Pa and in the surface equipment 3-105 Pa. The total pressure drop is 63.4-105 Pa. The mud pump has to provide a minimum of 63.4-105 kPa of pressure at a flow rate of 0.0183 m3/s (1100 l/min). Fig. 13 shows the dependency of the pressure drop versus the well bore depth and the type of equipment. Figure 13. Pressure drop versus well bore depth. The direction of the drilling fluid flow follows the black line from the left- to the right-hand side of the diagram. 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells Table 6. Pressure drop in drill pipes and drill collars (mud fluid moves in the downwards direction). Section Fluid velocity (m/s) Critical velocity (m/s) Fluid flow type Pressure drop (10S Pa) 1 1.978 1.91B turbulent B.B71 2 B.242 1.971 turbulent 1.114 B B.242 1.971 turbulent 10.844 4 4.019 2.002 turbulent B.41B Table 7. Pressure drop in annular space (mud fluid moves in upwards direction). Section Fluid velocity (m/s) Critical velocity (m/s) Fluid flow type Pressure drop (10S Pa) 1 0.270 1.686 laminar 1.444 2 0.2SB 1.670 laminar 0.128 B 0.270 1.676 laminar 1.B07 4 0.BB6 1.7B1 laminar 0.B60 Table 8. Pressure drop at drill bit (mud fluid moves in downwards direction). Nozzle number Nozzle diameter Fluid velocity Pressure drop (mm) (m/s) (105 Pa) 1 10.B2 2 10.B2 7B.11B B8.427 B 10.B2 The fluid actual and critical velocities in the drill string are shown in Table 6. It is evident that in all four sections of the borehole, the fluid flow type is turbulent because the actual velocity exceeds the critical velocity. In contrast, actual fluid velocities in the annular space are significantly lower than the critical velocities. Consequently, the fluid flow type is laminar, as shown in Table 7. The maximum pressure drop occurs in the drill pipes and the drill-bit nozzles. In the shown experimental case, the pressure drop in the drill bit equals 61 % and the pressure drop in the drill pipes equals 29 % of all the pressure drops. Other pressure drops arise in the surface equipment and the annular space. They are negligible. Fluid-flow type through the drill pipes (drill string and drill collars) and the drill bit nozzles is turbulent. The fluid flow type in the annular space is laminar. Fluid velocities in the drill pipes are in range 1 m/s to 5 m/s, depending on the inner diameter of the drill pipe. The drill bit nozzle output velocity exceeds 70 m/s. The drill bit pressure drop (61 % of all the pressure drops) is in range of hydraulics optimization and therefore suitable. 9 CONCLUSIONS In this article, deep drilling hydraulics and pressure drops have been studied in order to determine the working pressure of a mud pump. Based on the governing equations of Bingham non-Newtonian fluid flow, a numerical method has been used to calculate all the pressure drops while drilling the final section of the exploratory geothermal borehole Sob-4g, located in Murska Sobota, NE Slovenia. The results of the numerical analysis shows where the areas of maximum pressure drops in the deep well drilling are to be expected. The majority of the pressure drop occurs in the drill bit nozzles and the drill pipes. The fluid flow is turbulent in both of these regions. The pressure drops in the annular space and the surface equipment is minimal. They are negligible in comparison with the pressure drops in the pipes and the drill bit. The contribution of this article can be placed in the wider context of geothermal energy usage development. With the use of the described model for determining the pressure drops in a deep geothermal borehole, it is shown how we can predict the hydraulic state within the borehole during the process of drilling. REFERENCES [1] Hossain, M. E., Al-Majend, A. A. 2015. Fundamentals of Sustainable Drilling Engineering. Scriv-ender, Beverly, MA. [2] Azar, J. J., Samuel, G. R. 2007. Drilling Engineering. PenWell Corporation, Tulsa, Oklahoma. [3] Ford, J. (2005). Drilling Engineering. Heriot Watt University, Department of Petroleum Engineering, Edinburgh. [4] Bourgoyne, A. T. J., Millheim, K. T., Chenevert, M. E., Young, F. S. J. 1991. Applied Drilling Engineering. Society of Petroleum Engineering, Richardson. [5] Gabolde, G., Nguyen, J. P. 1999. Drilling data handbook. Technip, Paris. [6] Stropnik, J. 1999. Hidromehanika. Tehniška založba Slovenije, Ljubljana. [7] Ma, D., Li, G., Huang, Z., Niu, J., Hou, C., Liu, M., Li, J. 2012. A model of calculating the circulating pressure loss in coiled tubing ultra-short radius radial drilling. Petroleum exploration and develop- 10. Acta Geotechnica Slovenica, 2020/1 B. Janc and Ž. Vukelic: Mud-pump pressure in geothermal wells ment 39(4), 528-533. [8] Boyun, G., Gefei, L. 2011. Applied Drilling Circulation Systems. Gulf Professional Publishing, Houston, Texas. [9] Vukelic, 2., Kraljic, M., Dervaric, E. 2013. Lendava - the first geothermal city in Slovenia. 5th Balkan Mining Congress Proceedings, Ohrid, pp. 422-430. [10] Rajver, D., Rman, N., Lapanje, A. 2016. The state of exploitation of geothermal energy and some interesting achievements in geothermal research and development in the world. Geologija 59(1), 99-114. [11] Souvent, P., Vizintin, G., Celare, S., Cencur Curk, B. 2014. An expert system as a support to the decision making process for groundwater management of alluvial groundwater bodies in Slovenia. Geologija 57(2), 245-252. 10. Acta Geotechnica Slovenica, 2020/1 J. A. Duque Felfle et al.: An extended bearing-capacity equation for shallow foundations on granular trenches in soft soil AN EXTENDED BEARING-CAPACITY EQUATION FOR SHALLOW FOUNDATIONS ON GRANULAR TRENCHES IN SOFT SOIL RAZŠIRJENA ENAČBA NOSILNOSTI TEMELJNIH TAL POD PLITVIMI TEMELJI NA GROBOZRNATIH ZASEKIH V MEHKIH TLEH Jose Alejandro Duque Felfle (corresponding author) Carlos José Lascarro Estrada University del Norte, University del Norte, Department of Civil and Environmental Engineering Department of Civil and Environmental Engineering Km.5 Vía Puerto, Barranquilla, Colombia Km.5 Vía Puerto, Barranquilla, Colombia E-mail: jfelfle@uninorte.edu.co E-mail: cjlascarro@uninorte.edu.co Melany Gil Rueda University del Norte, Department of Civil and Environmental Engineering Km.5 Vía Puerto, Barranquilla, Colombia E-mail: gmelany@uninorte.edu.co Oscar Fernando García Guardo University del Norte, Department of Civil and Environmental Engineering Km.5 Vía Puerto, Barranquilla, Colombia E-mail: oguardo@uninorte.edu.co William Mario Fuentes Lacouture University del Norte, Department of Civil and Environmental Engineering Km.5 Vía Puerto, Barranquilla, Colombia E-mail: fuentesw@uninorte.edu.co https://doi.org/10.18690/actageotechslov.17.L12-22.2020 Keywords granular trenches, bearing capacity, Extended Drucker-Prager model, footings Ključne besede grobozrnati zrnati zaseki, nosilnost, Drucker-Pragerjev razširjeni model, temeljenje DOI Abstract In this article a method to estimate the bearing capacity of a shallow foundation on a granular trench in soft soil is proposed. The method is based on the interpolation of two equations: the first describing the bearing capacity of the shallow foundation on a homogeneous soft soil, and the second, on an infinitely wide layer of compacted granular fill above the soft soil. The proposed relationship was calibrated using finite-element (FE) simulations. A constitutive model for the soil, accounting for a pressure-dependent elasticity and an extended Drucker-Prager yield criterion, was considered. An analysis of the results showed that the proposed relationship gives more accurate estimations of the bearing capacity than those delivered by the conventional average-parameters method. The comparison was conducted for different dimensions of the granular trench. Finally, the relationship was used to predict the bearing capacity of a physical model test. Izvleček V članku je predlagana metoda za oceno nosilnosti temeljnih tal pod plitvem temeljem na grobozrnatem zaseku v mehkih tleh. Metoda temelji na interpolaciji dveh enačb: prva opisuje nosilnost temeljnih tal pod plitvim temeljem na homogeni mehki zemljini, druga pa na neskončno širokem sloju zgoščenega zrnatega nasipa ležečega na mehki zemljini. Predlagani odnos je bil umerjen s simulacijami z uporabo metode končnih elementov. Uporabljen je bil konstitutivni model tal, ki upošteva elastičnost kot funkcijo tlačnih napetosti in razširjeni Drucker-Pragerjev kriterij tečenja. Analiza rezultatov je pokazala, da predlagani odnos daje natančnejše ocene nosilnosti od tistih, ki jih daje običajna metoda povprečnih parametrov. Primerjava je bila izvedena za različne dimenzije grobozrnatega zaseka. Na koncu je bil odnos uporabljen za napovedovanje nosilnosti fizičnega modela preizkusa. 12. Acta Geotechnica Slovenica, 2020/1 J. A. Duque Felfle et al.: An extended bearing-capacity equation for shallow foundations on granular trenches in soft soil 1 INTRODUCTION An estimation of the bearing capacity is important for the design of shallow foundations. The first bearing capacity model was proposed by [1], and has been subjected to further enhancements in [2, 3, 4] and others. These methods consider a slip surface geometry developed for homogeneous soils, e.g., [1, 5, 6], which limits their application in layered soil profiles. This is disappointing, as in geotechnical engineering it is common to work with shallow foundations on stratified soils. In addition, many soft soils do not guarantee the required bearing capacity to withstand the design loads. Some methods to improve the mechanical characteristics of such soils have been proposed, such as soil-cement mixing [7], soil-lime mixing [8], gravel and stone columns [9, 10, 11, 12, 13], granular trenches [14, 15, 16, 17, 18, 19] among many others. Numerical investigations of these alternatives have also been reported [20, 21, 22, 23]. In particular, granular trenches is one of the most used methods due to its low cost, the availability of the granular fills and the simplicity of its construction. This is undoubtedly one of the most employed methods in regions where the required technology for other alternatives is not available. A granular trench in a soft soil consists of a bed of competent material beneath the footing that improves the bearing capacity of the system. Engineers often employ compacted fill materials, classified as SW-SP, SW-SC, SP and SC according to the USCS, or as A-1 and A-2 according to the AASHTO. The geometry of the trench, i.e., its thickness and width, are carefully designed in order to obtain the required bearing capacity. This is not a simple task, as conventional methods for estimating the bearing capacity assume a homogeneous soil. As an alternative, engineers can compute the "average (geomechanical) parameters" to be used in conventional methods for bearing capacity, following the recommendations of some authors [24]. However, this method is debatable, since the slip-surface geometry should depend on the granular trench dimensions as well. Moreover, sophisticated methods for estimating the bearing capacity considering the geometry of the granular trench have been proposed by other authors, [17, 19, 25, 26, 27]. These methods are also limited by some assumptions, such as equal unit weights for both materials, saturation conditions, among others. It seems that further research is still required to provide methods for estimating the bearing capacity of the shallow foundations supported on granular trenches in soft soils. In this work a simple relationship to estimate the bearing capacity of a foundation supported on a granular trench is proposed. For this purpose a set of finite-element models of a strip foundation was constructed. A constitutive model for soils, considering a pressure-dependent elasticity and an extended Drucker-Prager criterion, was considered. Different geometries of the granular trench were simulated, and the results were carefully analyzed. Consequently, a bearing-capacity relationship is proposed based on the interpolation between the bearing capacity of a foundation on a homogeneous soft soil (without a granular trench), and the one on an infinitely wide granular trench above the soft soil. The relationship gives accurate estimations compared to those of conventional methods with average parameters. Additionally, the results of a physical test reported by [28] are compared with the estimations of the proposed relationship. Finally, some conclusions are drawn. The notation of the article is as follows: scalar quantities are denoted with italic fonts (e.g., a, b), vectors with bold italic fonts (e.g., a, b), second rank-tensors with bold fonts (e.g., A, a), and fourth-rank tensors with special fonts (e.g., E, L). Tensors represented with the indicial notation are denoted with italic symbols and their respective lower indices (e.g., Ay, Oy). Multiplication by two dummy indices, also known as double contraction, is denoted with a colon ":" (e.g. A:B = Ay By). When the symbol is omitted, it is understood to be a dyadic product (e.g., AB = Ay By). The mean stress is defined as p=H3aii and the deviator stress as q = -^/3/2 II adev || , where adev = a - 1/3p1 is the deviator stress tensor, and 1 is the Kronecker delta tensor. 2 BRIEF DESCRIPTION OF THE FINITE-ELEMENT (FE) MODELS In this section a brief description of the finite-element (FE) models is given. The models simulate a strip footing on a granular trench surrounded by a soft soil. For the sake of generality, a soft soil is here considered as either a very loose sand or a normally consolidated clay, both of which are cohesionless c = 0. Different geometries of the granular trench were considered to construct the FE models. Figure 1 presents an illustrative scheme of the problem to solve. The width of the footing is fixed at B = 2 m and the embedment depth at Df = 1 m. The commercial software ABAQUS Standard V6.16 [29] was employed to build and solve the problem. The mechanical behavior of the soil was simulated through an elastopla-stic model, with a pressure-dependent elastic stiffness and a yield surface according to an extended Drucker Prager criterion that accounts for the Lode's angle dependence. In the following sections the Boundary Value Problem is described, and the obtained results are discussed. 12. Acta Geotechnica Slovenica, 2020/1 J. A. Duque Felfle et al.: An extended bearing-capacity equation for shallow foundations on granular trenches in soft soil B T Df Soft soil Figure 1. Sketch of the problem to solve. Values of B = 2 m and Df = 1 m are fixed. Different values of B2 and y are considered. 2.1 Geometry, mesh and boundary conditions The numerical model assumed plane-strain conditions and considered dry soil under a static analysis. Therefore, its solution is based on the linear momentum equation neglecting the inertial terms, see Equation (1): + = 0 (1) where aij is the effective stress, p is the mass density and gi is the gravity vector. Figure 2 presents the geometry and mesh of the Boundary Value Problem (BVP). Advantage was taken of the symmetry, and only one-half of the problem was simulated. The geometry was 20 m high and 30 m wide. Previous analyses showed that these dimensions avoid a boundary-dependent solution to the problem. The mesh was refined in the vicinity of the footing to achieve mesh convergence, see Figure 2. Horizontal displacements were restrained at the lateral boundaries, while vertical and horizontal displacements were restrained at the bottom boundary, see Figure 3. In order to simulate different geometries of the granular trench, some mesh partitions were introduced, as shown in Figure 3. This allowed the finite elements to be grouped according to their materials, as shown in Figure 4. In total, three materials were simulated, i.e., the concrete of the footing, the granular trench, and the soft soil, see Figure 4. The concrete was simulated with a linear elastic model, while the soil was simulated with the extended Drucker-Prager model. 2.2 Initial conditions and steps of analysis Initial geostatic stresses were defined assuming oedo-metric conditions. The initial vertical stresses av were computed by considering the densities of the materials, see Table 1. A lateral earth coefficient K0 was calculated for each material to determine the horizontal stresses ah = K0 av. The relationship proposed by Jaky K0 = 1 -sinf was used for this purpose, where f is the friction angle presented in Table 1. Considering this initial state, a geostatic step was taken to solve Equation (1). The generated displacements were reset to zero before proceeding with the loading of the footing, which was executed during the next step of the analysis. The footing load was introduced as a vertical displacement at the top boundary of the footing. The load was considered as a displacement boundary condition and not as an external load, to avoid numerical problems related to instability issues, i.e., a loss of controllability in the constitutive model [30]. The displacement was gradually increased until reaching failure. During this process, the vertical reaction force on the top of the footing was measured. For this work, the maximum vertical reaction force 20 m 30 m Figure 2. Geometry of the Boundary Value Problem (BVP). U2=0 -A. U1=U2=0 ^ U1=U2=0 _AT Figure 3. Boundary conditions and internal mesh partitions. 12. Acta Geotechnica Slovenica, 2020/1 J. A. Duque Felfle et al.: An extended bearing-capacity equation for shallow foundations on granular trenches in soft soil divided by the width of the footing is understood as the ultimate bearing capacity qu. This method is similar to those recommended by professional software packages for geotechnical numerical modeling [31, 32]. 1.0 m A" 1.0 ml Concrete Granular trench Clay Figure 4. Geometry of the materials. 2.3 Description of the elastoplastic model for the soils An elastoplastic constitutive model was used to simulate the mechanical behavior of the granular fill material and the soft soil. The constitutive model aims to describe the behavior of the effective stresses. The general equation of the constitutive model is presented in Equation (2): & = E-.(è- £f) (2) where à is the stress rate, è is the strain rate, èp is the plastic strain rate and E is the elastic stiffness (fourth-rank) tensor. For the elastic stiffness, modified Cam-Clay relationships were adopted. The elastic stiffness is isotropic and was computed using the relationship E = K11 + 2G(I-11), where 1 corresponds to the Kronecker delta tensor, I is the fourth-order tensor for symmetric tensors, K is the bulk modulus, see Equation (3), and G is the shear modulus, see Equation (4). Accordingly, the latter factors read [33]: „ (l + e0) , K =-p K r 3(1-217) G- 2(1 + v) (3) (4) where v is the Poisson ratio, k is the swelling index, both considered as material parameters, p' is the mean effective stress and eo is the initial void ratio. The plastic strain rate eP is computed by considering an extended Drucker-Prager yield surface function F, Equation (5), defined as: F -t- ptan/? -d- 0 (5) where t is a transformed deviator stress that accounts for the Lodes angle d dependence, defined in the sequel, and ft and d correspond to the modified friction angle and the modified cohesion defined in the space of t vs. p, respectively, see Equations (6) and (7). The parameters ft and d are related to the classic Mohr-coulomb parameters f and c, the latter are defined in the shear stress t vs. normal stress a space, through the relationships [33]: tan/? = d = 3V3tan

0.93 when adjusting the function f to each particular thickness y. When the function f is adjusted for all the simulation results (c1 = 0.552), the values of R2 are greater than R2 > 0.88. Table 3 provides the specific values of c1 and R2 for the evaluated cases. Table 3. Values of the parameter q and the corresponding R2. y = 1m y = 1.5m y = 2m Adjusted to each c1 0.626 0.628 0.508 thickness y R2 0.99 0.93 0.93 Adjusted for all Cl 0.552 0.552 0.552 simulations R2 0.97 0.88 0.88 The proposed relationship and the simulation results are plotted in Figures 7 to 10. As shown, a value of approximately qu ~ qu2 is obtained for B2/B > 6. This conclusion holds in all cases, independently of its thickness y. The bearing capacities qu, estimated using the interpolation constant c1 = 0.552, are shown in Figures 9 and 10. The results show a satisfactory agreement, with some small discrepancies. y 12. Acta Geotechnica Slovenica, 2020/1 J. A. Duque Felfle et al.: An extended bearing-capacity equation for shallow foundations on granular trenches in soft soil 800 750 700 650 600 550 500 450 400 350 300 Cj = 0.508 Ci = 0.628 y = 2 m y = 1.5m -X "X Cj = 0.626 y = 1 m ......................................................................... 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 B2/B[-] Figure 7. Numerical models and proposed relationship. Adjustment with different q for each thickness y. B2/B[-] Figure 8. Numerical models and proposed relationship. Adjustment with different q for each thickness y. 900 800 700 600 500 400 300 y = 2m Ct = 0.552 y = 1.5m Cj = 0.552 -X y = lm_Cj = 0.552 -X ......................................................................... 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 B2/B [-] Figure 9. Numerical models and proposed relationship. Adjustment with q = 0.552 for all cases. 1 1,2 1,4 1,6 1,8 2 2,2 2,4 2,6 2,8 3 Bj/B [-] Figure 10. Numerical models and proposed relationship. Adjustment with q = 0.552 for all cases. 12. Acta Geotechnica Slovenica, 2020/1 J. A. Duque Felfle et al.: An extended bearing-capacity equation for shallow foundations on granular trenches in soft soil 4 COMPARISON OF THE PROPOSED RELATIONSHIP WITH THE AVERAGE-PARAMETERS METHOD (APM)_ This section presents a comparison between the performance of the proposed relationship and the conventional average parameters method (APMs). The latter has been widely used by engineers, and is offered by various geotechnical software packages, e.g., [41]. According to this method, a stratified soil profile is analyzed as a homogeneous profile with "average (geomechanical) parameters", which allows the direct computation of the bearing capacity with conventional methods. The calculation of the average parameters requires the definition of a slip surface. A schematic example of such a surface is shown in Figure 11 for two cases: the first case considers a granular fill of the same width as the footing B = B2 , and the second case considers a much wider granular fill. The average friction angle f, the average cohesion c and the average unit weight y are computed as shown in Equations (14), (15) and (16), respectively, for case 1 and in Equations (17), (18) and (19), respectively, for case 2. Case 1) Footing Granular trench Slip surface Case 2) Figure 11. Sketch of a slip surface through a soft soil with granular trenches of different widths. A slip surface with constant geometry is assumed. y = /1^4 + 72(^1 + ^2+^3) lAi (16) For case 2 (see Figure 11): ^(¿i + Ls) + (¡g2(¿2 + ¿3+¿4 + ¿6) y™ i ■ Ci (¿! + Ls) + c2(¿2 + L3+L4 + L6)

B, the portion of the slip surface intersecting the granular fill remains constant, and therefore no increase in the average parameters is experienced. Finally, when the granular fill is sufficiently wide, the slip surface intersects it at its right boundary, as depicted in case 2 in Figure 11. In reality, the shape of the slip surface depends on the geometry of the granular trench, and therefore the results given by the average-parameters method are unrealistic. Hence, special care must be taken when considering this method for design purposes. For case 1 (see Figure 11): 80 %, it was good enough to apply three cycles to obtain the sample liquefaction; however, eight cycles were needed to reach soil liquefaction of the samples having a coefficient B close to 50 %.Reference [10] indicated that an increase in Skempton's pore-pressure coefficient B reduced the soil dilatancy and consequently amplified the phase of contractancy. The authors of [30] reported that an increase of Skempton's coefficient (B) from 89 % to 95 % induced a decrease of the initial stiffness of sandy soil and its shear strength and consequently an increase of the contractancy phase leading to a significant increase in the excess pore-water pressure of the tested granular sandy soils. Moreover, they indicated that for lower 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils loading amplitudes (CSR<0.40) for the undrained cyclic tests, the number of cycles required to cause liquefaction increased appreciably with the decrease of the Skemp-ton's coefficient (B) of Chlef sandy soils. On other hand, the influence of particle size distribution is an important subject when assessing the undrained shear strength (liquefaction resistance) response of soils [3, 5, 6, 7, 8, 15, 31, 32, 33, 34, 35 and 36]. In addition, [5] reported that the undrained shear strength of Chlef sand could be correlated with the mean grain size and coefficient (D50 and Cu). Reference [6] indicated that the grain size distribution in terms of effective diameter (D10), mean grain size (D50), coefficient of uniformity (Cu), effective size ratio (ESR), mean grain size ratio (MGSR) and coefficient of uniformity ratio (Cur) had significant influences on the excess pore-water pressure of silty sand soils. The authors of [34] showed that for samples with the same relative density, the undrained shear strength and the phase transformation deviatoric stress gradually decreases with the increase of the coefficient of uniformity Cu. The authors of [37] observed that the liquefaction resistance of clean sand decreases with a decrease of D50sand and Cmand with the same relative density for the loose samples; however, the undrained shear strength of silty sand soils decreases with an increase of the coefficient of uniformity (Cu). In [7 and 8] it was reported that the gradation and particle shape have a significant influence on the undrained shear strength (liquefaction resistance) of two silty sand soils. Moreover, their test results confirm the existence of simple correlations between the liquefaction resistance and the different grading characteristics (D10 , D30 , D50 , D60 , Cu , D10R , D50R and Cra)of the tested soils. The authors of [8] suggested that the instability stress and steady-state ratios can be correlated to the grading characteristics (D10 , D30 , D50 , D60 , Cu , D10R , D50R and CUR). Indeed, they decrease in a logarithmic and a linear manner with the decrease of grain size (D10, D30 , D50 and D60) and an increase of fines content, respectively. However, they decrease logarithmically with an increase of the coefficient of uniformity for the different graded sand-silt mixtures. It was reported in [3] that the grain size distribution in terms of extreme diameters (maximum diameter "Dmax" and minimum diameter "Dmin") and the mean grain size (D50) had appropriate effects on the liquefaction resistance of the wet deposited sandy samples reconstituted in the laboratory with an initial relative density (Dr=25 %). However, in the published literature, previous studies have not reported the influence of the grain-size distribution on the shear strength of partially saturated sandy soils under consideration. For this purpose the present study is undertaken to evaluate the influences of a newly proposed grain size distri- bution named the relative effective diameter (RED=D10/ Dmax) on the ultimate shear strength of two groups of granular sandy soils named A1, B1 and C1 with the maximum diameter ranging from 1 mm < Dmax < 4 mm and a constant effective diameter D10= 0.25 mm for group 1, A 2 , B2 and C2 with a maximum diameter ranging from 1 mm < Dmax < 4 mm and the same effective diameter D10=0.08 mm for group 2. Wet deposition is the most popular laboratory method to prepare loose sandy-soil samples and it consists of placing sand layers of specified thickness into a mold and tamping each layer with a flat tamper [12]. Therefore, it is selected as a suitable sample depositional technique for the present experimental program. All the samples were reconstituted using the wet-deposition method (w=5 %) at an initial relative density (Dr=25 %), tested under three Skempton pore-pressures values (B=20 %, 50 % and 90 %) and subjected to a constant confining pressure (P'c=100 kPa). 2 EXPERIMENTAL PROGRAM 2.1 Index properties of tested soils For this investigation the granular sandy-soil samples were obtained from liquefied soil deposit areas along the Chlef River where liquefaction cases were recorded during the 1980 El Asnam earthquake. The samples were collected from the banks of the Chlef River and were used in the preliminary tests as well as in the triaxial tests presented in this laboratory research work. The tested materials were classified according to the Unified Soil Classification System (USCS) as poorly graded sandy soils. The index properties of the sandy soils used in this study are summarized in Table 1. Their grain size distribution curves are shown in Figure1. The tested samples were prepared according to their maximum diameter (Dmax) and effective grain size (D10), subdivided into two groups, as indicated in Table 1. The values of the extreme void ratios (emax and emin) of the sandy-soil samples were determined according to standards [38 and 39] for this experimental research. 2.2 Sample preparation and test procedure Numerous reconstitution methods have been reported in published literature reviews for the deposition of granular sandy soils, such as wet deposition, dry funnel pluviation, water sedimentation, etc. [2, 5, 7, 8 and 40]. In addition, several researches have clearly shown the impacts of sample-reconstitution techniques on the undrained shear response of granular sandy soils and they have claimed that the wet-deposition method approximates closely to the in-situ fabric of fluvial sandy 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils Table 1. Index properties of tested materials Sample Gs DmaX(mm) Dmin(mm) Du(mm) Cu Q emax emin Ai 2.674 4 0.0016 0.25 2.080 1.053 0.840 0.635 Group 1 Bi 2.675 2 0.0016 0.25 1.960 1.058 0.880 0.664 Ci 2.675 1 0.0016 0.25 1.920 0.963 0.889 0.666 A2 2.683 4 0.0016 0.08 5.750 2.446 0.772 0.527 Group 2 B2 2.684 2 0.0016 0.08 5.875 2.237 0.804 0.552 C2 2.679 1 0.0016 0.08 5.625 2.336 0.817 0.555 (a) (b) Chief Sand (D10=0.25 mm) -Al -B1 CI AJfcA'A 10 1 0,1 0,01 0,001 Soil Particle Diameter (mm) 100 90 80 70 60 pû m 50 40 30 20 10 0 Il I I I I I_I_ Il I I I I I_L Chief Sand ( D 10=0.08 mm) 10 1 0,1 0,01 0,001 Soil Particle Diameter (mm) Figure 1. Grain size distribution curves of tested materials. (a) Group 1 (D10=0.25 mm), (b) Group 2 (D10=0.08 mm). soils [12, 13, 18 and 41]. In this study all the tested sandy-soil samples were prepared by the wet-deposition method with a constant water content (w=5 %) and then placed in a cylindrical mold with a diameter of D=100 mm and a height of H=200 mm "H/D=2" in successive layers with a constant thickness of 20 mm for each layer (10 layers). A constant number of strokes were applied with a flat tamper to obtain a homogeneous and isotropic soil fabric. Then, the samples were placed in the classic monotonic triaxial compression. After that, the sandy-soil samples were purged by passing carbon dioxide (CO2) for different times (15 min for B=20 %, 25 min for B=50 % and 35 min for B=90 %). In addition, the samples were also saturated with de-aerated and demineralized water. In this experimental investigation, a back pressure of 200 kPa was applied for all the performed tests and the sandy-soil samples were subjected to a constant effective stress of P'c=100 kPa. All the undrained monotonic triaxial tests were carried out at a constant strain rate of 0.225 mm per minute, which was slow enough to allow the pore-pressure change to equalize throughout the sample with the pore pressure measured at the base of the sample. 2.3 Relationship between the void ratio index and the relative effective diameter For the purpose of evaluating the relationship between the extreme void ratio index in terms of maximum void ratio (emax) and minimum void ratio (emin) with the proposed grain size ratio named as the relative effective diameter (RED=D10/Dmax) for the two groups of sandy soils named as A1, B1 and C1 with the maximum diameter 1 mm < Dmax < 4 mm and effective diameter D10=0.25 mm for group 1 and A2 , B2 and C2 with the maximum diameter 1 mm < Dmax < 4 mm and the effective diameter D10=0.08 mm for group 2 under study. It is clear from Figure 2 that the extreme void ratios (emax and emin) display a logarithmic relationship with the relative effective diameter (R2=0.99) for all the tested materials under consideration. Indeed, the maximum and minimum void ratios (emax and emin) increase with the increase of the relative effective diameter of the selected sandy soils for the two values of the effective diameter (D10=0.25 mm and D10=0.08 mm) and the maximum diameter range (1 mm < Dmax < 4 mm). 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils (a) S 0,8 0,6 S 0,4 ■o 0,2 Chief Sand (D10 1 * •=0.25mm) ——♦ ■ emin „ ♦ emax 0,1 0,2 0,3 Relative effective Diameter (RED) £ 0,8 c c 0,6 s = 0,4 ® S 7» 0,2 ( hlefSa nd (D 10=0.0 —--- S mm) i : : ■ 1 -1 1 1 1—-1 ■ 4 lemin L'iriax 0 0,02 0.04 0,06 0,08 0,1 Relative effective Diameter (RED) (b) Figure 2. Extreme voids ratios index versus relative effective diameter. (a) Group 1 (D10=0.25 mm), (b) Group 2 (D10=0.08 mm). 3 RESULTS AND DISCUSSION_ The result of the undrained monotonic compression triaxial tests performed on two groups and each group reconstituted of three sandy-soil samples named as A1, B1 and C1 with the same effective diameter (D10=0.25 mm) and three values of the maximum diameter of Dmax=4 mm for A1, Dmax=2 mm for B1 and Dmax=1 mm for C1 for group 1 and three sandy soils termed as A2, B2 and C2 reconstituted with an effective diameter of D10=0.08 mm with three different maximum diameter values (Dmax=4 mm, 2 mm and 1 mm), respectively, for group 2. All the samples were reconstituted with an initial relative density of Dr=25 % and subjected to three different values of Skempton's pore-pressure parameter (B=20 %, 50 % and 90 %) and a constant confining pressure of P'c=100 kPa. 3.1 Undrained monotonic triaxial compression test results 3.1.1 Group 1 (4, B1 and C1 with D10=0.25 mm) Figures 3,4 and 5 illustrate the undrained monotonic triaxial compression tests performed on three laboratory-reconstituted sandy-soil samples named A1, B1 and C1 with the same effective diameter (D10=0.25 mm). It can be observed from these figures that completed and limited static liquefaction cases were recorded for the different tested samples under different Skempton pore-pressure parameters (B) with a clear impact of the maximum diameter Dmax on the undrained shear strength (static liquefaction resistance) response. Moreover, it is clear that the undrained shear strength (liquefaction resistance) of the samples A1, B1 and C1 increases with the decrease of 160 140 I 120 "cr 100 80 Chief Sand (Dmax=4mm, Dmin=0.0016mm) J 60 40 20 :........ ......... ......... ......... i : 1 = a1 (d10=0.25mm) -x- b=20% —4— b=50% -a- b=90% i : : : : (a) 5 10 15 Axial Strain (%) 20 25 (b) 50 100 Effective Mean Stress, P" (kPa) Figure 3. Undrained monotonic behavior of Chief sandy soils (Aj) (Dmax=4 mm, Dmin=0.0016 mm, D10=0.25 mm, Dr=25 %, P'c=100 kPa). (a) Deviator stress versus axial strain. (b) Devia-tor stress versus Effective mean Stress. 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils (a) (b) Chief Sand (Dmax=1mm, Dmin=0.0016mm) J 140 120 100 80 60 40 20 0 10 15 Axial Strain (%) Chief Sand (Dmax=2mm, Dmin=0.0016mm) -1 i i i i i i i i i i i i i i i i i i i i 111111 _ : B1 (D10=0.25n»n) E -X- B=20% .X : —+— B=50% * : -&- B=90% / - : i j *V„ E i / \ ! 1 / \\ J : / \ ¡/ i 1 .........: i111 | 0 50 Effective 100 P" (kPa) 150 Figure 4. Undrained monotonie behavior of Chief sandy soils (BJ (Dmax=2 mm, Dmin=0.0016 mm, D10=0.25 mm, Dr=25 %, P'c=100 kPa). (a) Deviator stress versus axial strain. (b) Devia-tor stress versus Effective mean Stress. (a) ta 140 120 100 80 60 40 20 0 5 10 15 20 Axial Strain (%) Chief Sand (Dmax=1mm, Dir»n=0.0016mm) ....... .........E ; CI (DICH).25mm) -X- B=20% —+— B=50% —A— B=90% 1 | ; 1 / \ 1 / \ l / \ 1 ......... (b) 0 50 100 150 Effective Mean Stress, P" (kPa) Figure 5. Undrained monotonie behavior of Chief sandy soils (C1) (Dmax=1 mm, Dmin=0.0016 mm, D10=0.25 mm, Dr=25 %, P'c=100 kPa). (a) Deviator stress versus axial strain. (b) Devia-tor stress versus Effeetive mean Stress. Table 2. Summary of undrained monotonie triaxial tests for group 1. Charaeteristies of materials Ai B1 Q B (%) 20 50 90 20 50 90 20 50 90 qu (kPa) 117.85 21.83 17.67 90.54 28.68 24.86 91.40 26.43 22.50 RED (-) 0.0625 0.0625 0.0625 0.125 0.125 0.125 0.25 0.25 0.25 9u (°) 34.48 57.62 57.71 35.18 49.16 58.29 35.10 52.99 57.73 Ib (-) 0.17 0.71 0.75 0.31 0.66 0.64 0.32 0.70 0.69 Soil response All the samples exhibit flow behavior Skempton's pore-pressure parameter in the range 50-90 % and becomes more obvious at a lower Skempton's pore-pressure parameter B=20 %. In addition, the obtained data indicate that the percentage of increasing in the shear strength is 0.09 %, 22 % and 20 % for the range of Skempton's pore-pressure parameter (B=50-90 %) and becomes very significant (87 %, 55 % and 53 %) in the range of Skempton's pore-pressure parameter (B=20-50 %) for the 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils tested sandy-soil samples Aj, Bj and C respectively. This sandy-soils trend can be attributed to the increase in the particle interlocking between the coarse grains of sandy soils due to the decrease of Skempton's pore-pressure parameter inducing a contractive character of the tested samples. Therefore, the higher peak undrained shear strength correspondsto the higher maximum diameter (Dmax=4 mm) and the lower Skempton's pore-pressure parameter (B=20 %). In contrast, the inverse tendency is observed for the higher Skempton's pore-pressure parameters (B=50 % and B=90 %) where they exhibit the lower peak undrained shear strength with the increasing of the maximum diameter from Dmax=1 mm toDmax=4 mm for the tested materials. In addition, all the samples A1, B1 and C indicate that the ultimate shear strength is reached within 10-20 % axial strain (Figures 3a, 4a and 5a). Our findings are in good agreement with the observations of [4, 10 and 24]. The stress path in the (p', q) plane clearly shows the role of the Skempton's pore-pressure parameter and the maximum diameter (B and Dmax) in the decrease of the average effective stress and the maximum devia-toric stress (Figures 3b, 4b and 5b). A summary of the undrained monotonic triaxial tests results for group 1 are summarized in Table 2. 3.1.2 Group 2 (A2, B2 and C2 with D10=0.08 mm) Figures 6, 7 and 8 show the undrained-shear-strength response of three granular sandy soils termed as A2, B2 and C2, reconstituted with an effective diameter of D10=0.08 mm. The test results demonstrate clearly that the completed and limited static liquefaction cases were recorded for the different tested samples under different values of the Skempton's pore-pressure parameter (B=20 %, 50 % and 90 %) with a clear impact of Chief Sand (Dmax=4mm, Dmin=0.0016mm) Chief Sand (Dmax=2mm, DminM).0016mm) (a) f\ : f\ \ : A2 (D10=0.08rrm) -X- B=20% -^- B=50% -&- B=90% H \ : ......... : 5 10 15 20 Axial Strain (%) 25 B2 (D10=0.08rrm) X- B=20% B=50% B=90% (a) 5 10 15 20 Axial Strain (%) (b) 100 80 Chief Sand (Dmax=4mm, Dmin=0.0016mm) 40 20 : '7a : : - J£ \ \ - JF \ : A2(D10=a08mm) X B=20% ♦ B=50% —A— B=90% : : 0 50 Effective 100 i, P- (kPa) 150 Figure 6. Undrained monotonie behavior of Chief sandy soils (A2) (Dmax=4 mm, Dmin=0.0016 mm, D10=0.08 mm, Dr=25 %, P'c=100 kPa). (a) Deviator stress versus axial strain. (b) Devia-tor stress versus Effective mean Stress. (b) 120 100 Chief Sand (Dmax=2mm, Dmin=0.0016mn) ', & 80 1 60 40 20 : B2(D1(N).08nnrn) —X— B=20% : ♦ B=50% —a— B=90% E >< ~z M : 50 100 150 Effective Mean Stress, P* (kPa) Figure 7. Undrained monotonie behavior of Chief sandy soils (B2) (Dmax=2 mm, Dmin=0.0016 mm, D10=0.08 mm, Dr=25 %, P'c=100 kPa). (a) Deviator stress versus axial strain. (b) Devia-tor stress versus Effective mean Stress. 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils (a) C2 (D10=0.08mm) -X- B=20% B=50% B=90% 10 15 Axial Strain (%) (b) C2 (D10=0.08mm) —X— B=20% ♦ B=50% -A- &30% 50 100 150 Effective Mean Stress, P" (kPa) Figure 8. Undrained monotonie behavior of Chief sandy soils (C2) (Dmax=1 mm, Dmin=0.0016 mm, D10=0.08 mm, Dr=25 %, P'c=100 kPa). (a) Deviator stress versus axial strain. (b) Devia-tor stress versus Effective mean Stress. the maximum diameter Dmax on the undrained shear strength behavior. Moreover, the undrained shear strength of the samples A2, B2 and C2 decreases with the increase of the Skemptons pore-pressure parameter. In addition, for the lower Skempton's pore-pressure parameter (B=20 %), the higher maximum diameter (Dmax=4mm) exhibits the lower peak undrained shear strength. The inverse tendency is observed for the higher values of the Skempton's pore-pressure parameter (B=50 % and B=90 %). Therefore, the obtained data indicate that the ultimate shear strength is reached within 4-20 % axial strain for all the sandy-soil samples A2, B2 and C2, see Figures 6a, 7a and 8a, respectively. The obtained results are in good agreement with the results of [4, 10 and 24]. The stress path in the (p\ q) plane shows clearly the role of the Skempton's pore-pressure parameter (B) and the maximum diameter (Dmax) in decreasing the average effective stress and the maximum deviatoric stress (Figures 6b, 7b and 8b). A summary of the undrained monotonic triaxial tests results for group 2 are summarized in Table 3. 3.2 Effect of Skempton's pore pressure on the undrained ultimate shear strength Figure 9 summarizes the effect of Skempton's pore-pressure parameter (B=20 %, B=50 % and B=90 %) on the undrained ultimate shear strength of six sandy-soil samples (A1, B1, C1, A2, B2 and C2). It is clear from Figure 9 that the decrease of the Skempton's pore-pressure parameter leads to a remarkable increase of the undrained ultimate shear strength and becomes more significant for the lower Skempton's pore-water pressure (B=20 %) of the tested sandy-soil samples. The obtained data indicate that the sandy-soil samples with the higher maximum diameter (Dmax=4 mm) and the lower Skempton's pore-pressure parameter (B=20 %) exhibit a higher undrained ultimate shear strength compared to that induced by higher Skempton's pore-pressure parameters (B=50 % and 90 %) for the tested samples of group 1. In contrast, the inverse tendency was observed in the case of group 2, where the sandy-soil samples with the higher maximum diameter (Dmax=4 mm) exhibited a lower undrained ultimate shear strength for the same Skempton's pore-pressure parameter (B=20 %) Table 3. Summary of undrained monotonie triaxial tests for group 2. Charaeteristies of materials A2 B2 C2 B (%) 20 50 90 20 50 90 20 50 90 qu (kPa) 27.57 24.46 24.58 38.33 24.02 23.95 56.92 25.97 23.67 RED (-) 0.02 0.02 0.02 0.04 0.04 0.04 0.08 0.08 0.08 fu (°) 49.65 58.58 57.71 42.78 57.69 57.72 37.99 57.44 57.57 ib (-) 0.71 0.69 0.63 0.62 0.71 0.63 0.46 0.71 0.69 Soil response All the samples exhibit flow behavior 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils 140 I 120 o* j- 100 s ei S 80 60 40 20 ♦Al ■A2 B1 XB2 *C1 ; \ i i C2 : 4 t : / ; l \ l 1 - 0 20 40 60 80 100 Skempton's Pore Pressure Parameter, B (%) Figure 9. Ultimate shear strength versus Skempton's pore-pressure parameter of the sandy soils (Dr=25 %, P'c=100 kPa). compared to the higher Skempton's pore-pressure parameters (B=50 % and 90 %). In addition, the ultimate shear strength of the sandy-soil samples of group 1 is more significant than that of group 2. This behavior indicates that the increasing in the effective diameter leads to a significant increase in the undrained ultimate shear strength of the tested sandy soils under consideration. Moreover, the sandy-soil samples of group 2 with an effective diameter D10=0.08 mm demonstrate clearly that the presence of the low plastic fines between the coarse grains of A2, B2 and C2 makes the sandy soils more compressible, leading to a reduction of the interparticle forces and consequently to a decrease in the undrained ultimate shear strength of the used materials. 3.3 Effect of the maximum diameter on the undrained ultimate shear strength The influence of maximum diameter (Dmax) on the undrained ultimate shear strength (qu) of the two groups of sandy soils is illustrated in Figure (10). It is clear from the bar chart that the maximum diameter (Dmax) exhibits a significant influence on the undrained ultimate shear strength of the tested materials. Indeed, the undrained ultimate shear strength increases with the increase of the maximum diameter for the lower value of the Skempton's pore pressure (B=20 %) in the case of the group 1 samples, while the inverse tendency was observed for the sandy-soil samples of group 2, where the undrained ultimate shear strength decreases with the increase of the maximum diameter for the same Skempton's coefficient (B=20 %). The observed tendency can be attributed to the combined effects of the maximum and effective diameters with (B=20 %) in the increasing and decreasing of the interparticle forces between the coarse grains for group 1 (D10=0.25 mm, the small amount of (a) (b) Figure 10. Ultimate shear strength versus maximum diameter of sandy soils. (Dr=25 %, P'c=100 kPa). (a) Group 1 (D10=0.25 mm). (b) Group 2 (D10=0.08 mm). low plastic fines Fc<3 %) and group 2 (D10=0.08 mm), the presence of low plastic fines Fc=10 % leading to a decrease of the interlocking of the coarse particles and consequently to a significant decrease in the undrained shear strength. In contrast, quite similar observations were recorded for the ultimate shear strength in the case of the intermediate and higher Skempton's coefficients (B=50 % and 90 %) for all the tested sandy-soil samples. 3.4 Effect of the relative effective diameter on the undrained ultimate shear strength The effect of the relative effective diameter (RED=D10/ Dmax) on the undrained ultimate shear strength (qu) of the two groups of sandy soils is shown in Figure (11). The obtained data from the current study indicate that the relative effective diameter (RED) could be correlated 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils ja S a a ce 140 120 100 80 60 40 I 20 Chief Sand (Group"l") -♦ —I_I_I_I_ —I_I_I_I_ ♦ B=20% ■B=50% B=90% —I_I_I_l_ (a) 0 0,1 0,2 0,3 Relative Effective Diameter (RED) (b) 60 6 50 s a" ■5 40 5» a aj S 30 h a v £ 20 10 Chief Sand (Group"2") 0 0,02 0,04 0,06 0,08 0,1 Relative Effective Diameter (RED) Figure 11. Undrained ultimate shear strength versus relative effective diameter (Dr=25 %, P'c=100 kPa). (a) Group 1 (D10=0.25 mm). (b) Group 2 (D10=0.08 mm). with the undrained ultimate shear strength of the tested sandy soils for the tested Skempton's pore-pressure parameter (B=20 %, 50 % and 90 %) values under study. In addition, it is clear from Figure 11a that the undrained ultimate shear strength decreases with an increase of the relative effective diameter for the lower Skempton's pore-pressure parameter (B=20 %) for the samples of group1. However, the inverse tendency is observed for the intermediate and higher Skempton's pore-pressure parameter (B=50 % and 90 %) for the same group, where the ultimate shear strength increases with the increase of the relative effective diameter (RED). For group 2, the tested sandy-soil samples exhibit the opposite trend for all the considered Skempton's pore pressures (B=20 %, 50 % and 90 %) (Figure11b). On the other hand, it can be seen from these plots that the variation of the undrained ultimate shear strength (qu) with the relative effective diameter (RED) of the tested sandy soils for the lower Skempton's pore-pressure parameter (B=20 %) is highly affected rather than that of the intermediate and higher Skempton's pore-pressure parameter (B=50 % and B=90 %) for both groups under consideration.These tendencies confirm that the decrease in the Skempton's pore-pressure parameter induces a significant increase of the interparticle forces between the coarse and fine grains, inducing a clear increase of the undrained ultimate shear strength of the tested materials under study. However, the relative effective diameter (RED) has an effect on the resistance liquefaction. The results of the ultimate shear strength are summarized in Tables 2 and 3, respectively, for group 1 and group 2. 3.5 Effect of the maximum diameter and Skempton's pore-pressure parameter on the brittle-ness index of the sandy soils To quantify the amount of strain softening during undrained loading, [42] proposed a new parameter to identify this behavior named the brittleness index (Ib), which is defined as: IB rhlpf SanH iflrniin"?'^ Figure 12. Brittleness index versus maximum diameter of sandy soils (Dr=25 %, P'c=100 kPa). (a) Group 1 (D10=0.25 mm). (b) Group 2 (D10=0.08 mm). (a) 0,80 ,0,60 a (b) 0,40 0,20 0,00 Chief Sand (D10=0,25 mm) t ' 1 ♦ ♦ ♦B=20% : ♦ ■B=50% i B=90% 0,80 0,70 J g0,60 -o a g 0,50 a a l0,40 n 0,30 0,20 0 0,1 0,2 0,3 Relative Effective Diameter (RED) r Chief Sand (D10=0.08 mm) 1 ♦B=20% ■B=50% A B=90% J_I_I_I_I_I_I_I_I_I_I_I_I_L 0 0,02 0,04 0,06 0,08 0,1 Relative Effective Diameter (RED) Figure 13. Brittleness index versus relative effective diameter of sandy soils (Dr=25 %, P'c=100 kPa). (a) Group 1 (D10=0.25 mm). (b) Group 2 (D10=0.08 mm). 3.6 Relationship between the brittleness index and the relative effective diameter The relationship between the brittleness index (IB) and the relative effective diameter (RED) of six sandy-soil samples (A1, B1 and C1 for group 1 and A2, B2 and C2 for group 2) is discussed in this section. The test results show that the brittleness index could be correlated with the relative effective diameter of the tested materials for all the parameters under consideration. In addition, it is clear from Figure 13a that the brittleness index (Ib) increases with an increase of the relative effective diameter (RED) for the lower Skempton's pore-pressure parameter (B=20 %) of the tested sandy-soil samples of group 1. However, the inverse tendency was observed for the intermediate and higher Skempton's pore-pressure parameter (B=50 % and 90 %), where the brittleness index (Ib) decreases with the increase of the relative effective diameter (RED) of the tested materials (A1, B1 and C1) under study. Moreover, it is clear from Figure 13b that the brittleness index decreases with an increase of the relative effective diameter (RED) of the tested materials under a lower Skempton's pore-pressure parameter (B=20 %). In contrast, the inverse tendency was shown for two other Skempton pore-pressures parameters (B=50 % and 90 %), where the brittleness index increases with an increase of the relative effective diameter for the sandy-soil samples (A2, B2 and C2). 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils 3.7 Relationship between Skempton's pore- pressure parameter and the mobilized ultimate internal friction angle of sandy soils 3.8 Relationship between the mobilized ultimate internal friction angle and the relative effective diameter of the sandy soils The correlation between the Skempton's pore-pressure parameter (B) and the mobilized ultimate internal friction angle (fu) for the two groups is presented in Figure (14). The obtained data indicate that the Skempton's pore-pressure parameter could be correlated with the mobilized ultimate internal friction angle of the tested materials. Indeed, the mobilized ultimate internal friction angle increases in a good polynomial manner an increase of the Skempton's pore-pressure parameter from B=20 % to B=90 % of the A1, B1 and C1 for group 1 and A2, B2 and C2 for group 2. Moreover, it is clear from Figure14 that the higher Skempton's pore-pressure parameter (B=90 %) indicates a similar undrained mobilized ultimate internal friction angle for all the used materials compared to the lower Skempton's pore-pressure parameter (B=20 % and 50 %), where the tendency shows a higher range between the values of the mobilized ultimate internal friction angle of the tested materials under study. This trend confirms that the Skempton's pore-pressure parameter has a significant influence on the undrained shear strength and consequently on the undrained mobilized ultimate internal friction angle of the sandy soils, where it has a remarkable effect on increasing the interparticle forces between the coarse and fine grains, leading to an important increase in the mobilized ultimate internal friction angle of the tested sandy soils. a #o u ■c "3 a t- J =9 ■a a> ■o o M) a < 65 60 55 50 45 40 35 30 y = -0,007x2+ 1,037 R2 = 0,833 x +21,264 2 /1 ► A / ' s x/ ■ at V ♦Al AA2 XB1 XB2 ®C1 ■ C2 0 50 100 Skempton's Pore Pressure Parameter, B (%) The variation of the mobilized ultimate internal friction angle (fu) and the relative effective diameter (RED) of the two groups of sandy-soil samples is presented in Figure 15. The obtained data indicate that the relative (a) : Chief Sac d (B=20%) ..........♦ 0 0,1 0,2 0,3 Relative Effective Diameter (RED) (b) 80 60 5 C m 40 20 - ; Chief S: ind (B=50°/ ») ♦ ♦ 0 0,1 0,2 0,3 Relative Effective Diameter (RED) (c) 60 &50 40 Chief Sa ad (B=90%) ♦ f-f ♦♦ 0 0,1 0,2 0,3 Relative Effective Diameter (RED) Figure 14. Mobilized ultimate internal friction angle versus Skempton's pore-pressure parameter of sandy soils (Dr=25 %, P'c=100 kPa). Figure 15. Mobilized ultimate internal friction angle versus relative effective diameter of sandy-soil samples (Dr=25 %, P'c=100 kPa). (a) B=20 % , (b) B=20 % , (c) B=90 %. 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils effective diameter (RED) could be correlated with the mobilized ultimate internal friction angle (fu) for the materials under study. Indeed, the mobilized ultimate internal friction angle decreases in a polynomial manner with an increase of the relative effective diameter for the lower and intermediate Skempton's pore-pressure parameters (B=20 % and 50 %). However, the influence of the relative effective diameter on the mobilized ultimate internal friction angle is insignificant for the higher Skempton's pore-pressure parameter (B=90 %). The obtained sandy-soils tendency confirms that the decrement of the Skempton's pore-pressure parameter plays a major role in increasing the mobilized ultimate internal friction angle-relative effective diameter response leading to a significant increase in the interparticle forces between the coarse grains of the tested sandy soils under consideration. 4 CONCLUSION This laboratory research work is based on a series of undrained compression tests using static triaxial apparatus for the purpose of evaluating the effects of the relative effective diameter (RED=D10/Dmax) on the mechanical behavior of partially saturated sandy soils. The tested samples were subdivided into two groups: A1, B1, C1 and A2, B2, C2. They were reconstituted with the wet-deposition method (w=5 %) at an initial relative density (Dr=25 %), examined under three different Skempton's pore-pressure parameter values (B=20 %, 50 % and 90 %) and subjected to a constant confining pressure (P'c=100 kPa). The main conclusions of this study are summarized below: 1. The obtained test results show that the Skempton's pore-water-pressure parameter has a significant influence on the undrained-shear-strength response of the wet-deposited sandy-soil samples. Indeed, the ultimate shear strength increases with the decrease of the Skempton's pore-pressure parameter from the higher value (B=90 %) to the lower value (B=20 %), and it becomes more significant for the effective diameter (D10=0.25 mm). Completed and limited static liquefaction response cases were observed for all the tested partially saturated sandy-soil samples, as illustrated in Figures 3b, 4b, 5b, 6b, 7b and 8b. 2. The test results demonstrate clearly that the maximum diameter could be correlated with the ultimate shear strength and the brittleness index of the used materials. Indeed, the increase of the maximum diameter leads to an increase of the ultimate shear strength and a decrease of the brittleness index of the samples of group 1 (A1, B1 and C1) for (B=20 %) and the inverse tendency was observed for group 2, where the ultimate shear strength decreases and the brittleness index increases with an increase of the maximum diameter for the same Skempton's coefficient (B=20 %). In contrast, similar observations were made for the ultimate shear strength in the case of medium and higher Skempton's coefficients (B=50 % and 90 %) for all the tested sandy-soil samples, as shown in Figure (12 and 13). 3. The obtained findings confirm that the relative effective diameter (RED) has a significant influence on the mechanical behavior of sandy soils in terms of the ultimate shear strength (qu). The increase of the relative effective diameter leads to a decrease of the ultimate shear strength for the lower Skempton's pore-pressure parameter (B=20 %) of the tested sandy-soil samples of group 1 and increase it for the intermediate (B=50 %) and higher (B=90 %) Skempton's pore-pressure parameter for the same group and consequently to an increase of the undrained shear strength (liquefaction resistance). Moreover, the inverse trend was observed in the case of group 2 for all the tested Skempton's pore-pressure-parameter values (B=20 %, 50 % and 90 %), as illustrated in Figures 10 and 11. 4. Finally, the relative effective diameter (RED) could be correlated with the brittleness index (¡¿) and the mobilized ultimate internal friction angle (fu) of the two groups and control effectively the undrained shear strengthof the sandy soils tested under three Skempton's pore-pressure-parameter values (B=20 %, 50 % and 90 %) at an initial relative density (Dr=25 %) and subjected to a constant confining pressure (P'c=100 kPa), as presented in Figure (15). Acknowledgments The authors are grateful to Professor Tom Schanz for putting at the disposal of the research teams all the necessary laboratory equipment to achieve the objective of this research project. The writers acknowledge the technician Michael Skubisch, who effectively contributed to the achievement of this experimental program. 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils REFERENCES [1] Kramer, S. L. 1996. Geotechnical Earthquake Engineering Prentice Hall. New York. [2] Belkhatir, M., Arab, A., Delia, N., Schanz, T. 2012. Experimental study of undrained shear strength of silty sand: effect of fines and gradation. Geotech GeolEng 30(5), 1103-1118. DOI: org/10.1007/ s10706-012-9526-1 [3] Hazout, L., Zitouni, Z. E. A., Belkhatir, M., Schanz, T. 2017. Evaluation of Static Liquefaction Characteristics of Saturated Loose Sand Through the Mean Grain Size and Extreme Grain Sizes. Geotechnical and Geological Engineering, pp 1-27. DOI: org/10.1007/s10706-017-0230-z [4] Arab, A., Shahrour, I., Lancelot, L. 2011. A laboratory study of liquefaction of partially saturated sand. J IberGeol 37(1), 29-36. DOI: 10.5209/ rev_jige.2011.v37.n1.2 [5] Belkhatir, M., Arab, A., Missoum, H., Della, N., Schanz, T. 2011. Laboratory study on the liquefaction resistance of sand-silt mixtures: effect of grading characteristics. Granular Matter 13, 599-609. DOI: org/10.1007/s10035-011-0269-0 [6] Belkhatir, M., Schanz, T., Arab, A., Della, N., Kadri, A. 2014. Insight into the effects of gradation on the pore pressure generation of sand-silt mixtures. Geotech Test J 37(5), 922-931. DOI: org/10.1520/ gtj20130051 [7] Cherif Taiba, A., Belkhatir, M., Kadri, A., Mahmoudi, Y., Schanz, T. 2016. Insight into the effect of granulometric characteristics on the static liquefaction susceptibility of silty sand soils. Geotechnical and Geological Engineering 34(1) 367-382. DOI: org/10.1007/s10706-015-9951-z [8] Cherif Taiba, A., Mahmoudi, Y., Belkhatir, M., Kadri, A., Schanz, T. 2017a. Experimental characterization of the undrained instability and steady state of silty sand soils under monotonic loading conditions. International Journal of Geotechnical Engineering, pp 1-17. DOI: org/10.1080/19386362. 2017.1302643 [9] Cherif Taiba, A. 2017b. Etude en laboratoire de la susceptibilité à liquéfaction des sables limoneux: Influence de la taille et forme des grains (Doctoral dissertation, KADRI Abdelkader, BELKHATIR Mostefa). [10] Della, N., Arab, A., Belkhatir, M. 2011. Static liquefaction of sandy soil: an experimental investigation into the effects of saturation and initial state. Acta Mech 218(1-2), 175-186. DOI: org/10.1007/ s00707-010-0410-x [11] Lade, P. V., Yamamuro, J. A. 1997. Effects of non plastic fines on static liquefaction of sands. Cana- dian Geotechnical Journal 34(6), 918-928. DOI: org/10.1139/t97-052 [12] Mahmoudi, Y., Cherif Taiba, A., Belkhatir, M., Schanz, T. 2016. Experimental investigation on undrained shear behavior of over consolidated sand-silt mixtures: effect of sample reconstitution. Geotech Test J 39(3):515-523. DOI:10.1520/ GTJ20140183 [13] Mahmoudi, Y., Cherif Taiba, A., Belkhatir, M., Arab, A., Schanz, T. 2016. Laboratory study on undrained shear behavior of over consolidated sand-silt mixtures: effect of the fines content and stress state. International Journal of Geotechnical Engineering, pp 1-15. DOI:10.1080/19386362.2016 .1252140 [14] Xenaki, V.C., Athanasopoulos, G.A. 2003. Liquefaction resistance of sand-silt mixtures: an experimental investigation of the effect of fies. Soil DynEarthqEng 23(3), 183-194. DOI: org/10.1016/ S0267-7261(02)00210-5 [15] Yilmaz, Y., MollaMahmutoglu, M., Ozaydin, V., Kayabali, K. 2008. Experimental investigation of effect of grading characteristics on the liquefaction resistance of various graded sands. EngGeol J 100, 91-100. DOI: org/10.1016/j.enggeo.2007.12.002 [16] Cherif Taiba, A., Mahmoudi, Y., Belkhatir, M., and Schanz, T. 2018. Experimental Investigation into the Influence of Roundness and Sphericity on the Undrained Shear Response of Silty Sand Soils. Geotechnical Testing Journal. DOI.10.1520/ GTJ20170118 [17] Cherif Taiba, A. Mahmoudi,Y. Hazout, L. Belkhatir, M. Baille, W. 2019. Evaluation of hydraulic conductivity through particle shape and packing density characteristics of sand-silt mixtures. Marine Georesources & Geotechnology. DOI: 10.1080/1064119X.2018.1539891 [18] Mahmoudi, Y., Cherif Taiba, A., Hazout, L., Wiebke, B. Belkhatir, M. 2018. Influence of Soil Fabrics and Stress State on the Undrained Instability of Overconsolidated Binary Granular Assemblies. StudiaGeotechnica et Mechanica.40(2), 96-116. DOI.org/10.2478/sgem-2018-0011 [19] Atigh, E., Byrne, P.M. 2004. Liquefaction flow of submarine slopes under partially undrained conditions: an effective stress approach. Can Geotech J 41, 154-165. DOI: org/10.1139/t03-079 [20] Chaney, R. C. 1978. Saturation effects on the cyclic strength of sands. In From Volume I of Earthquake Engineering and Soil Dynamics—Proceedings of the ASCE Geotechnical Engineering Division Specialty Conference, June 19-21, 1978, Pasadena, California. Sponsored by Geotechnical Engineering Division of ASCE in cooperation with: (No. 56. Acta Geotechnica Slovenica, 2020/1 K. Doumi et al.: Experimental investigation of the influence of relative effective diameter on the ultimate shear strength of partially saturated granular soils Proceeding). [21] Kokusho, T. 2000. Correlation of pore pressure B-value with p-wave velocity and Poisson's ratio imperfectly saturated sand or gravel. Soils Foud 40(4), 95-102. D0I:org/10.3208/sandf.40.4_95 [22] Mathiroban, S., Grozic, J. 2004. A model to predict the undrained behaviour of loose gassy sand. In: Proceedings of 57th Geotechnical conference, session 6G, 16-22. DOI: org/10.11575/ PRISM/22820 [23] Pietruszczak, S., Pande, G., Oulapour, M. 2003. A hypothesis for mitigation of risk of liquefaction. Geotechnique 53(9), 833-838. D0I:org/10.1680/ geot.2003.53.9.833 [24] Sherif, A.M., Tsuchiya, C., Isibashi, I. 1977. Saturation effects on initial soil liquefaction. J. Geotech. Eng. Div. 103(8), 914- 917. [25] Yang, J., Sato, T. 2001. Analytical study of saturation effects on seismic vertical amplification of a soil layer. Geotechnique 51(2), 161-165. DOI: org/10.1680/geot.2001.51.2.161 [26] Yoshimi, Y., Tanaka, K., Tokimatsu, K. 1989. Liquefaction resistance of a partially saturated sand. Soils and foundations, 29(3), 157-162. DOI: org/10.3208/sandf1972.29.3_157 [27] Ishihara, K., Tsuchiya, H., Huang, Y., Kamada, K. 2001. Recent studies on liquefaction resistance of sand effect of saturation. In: Proceedings of 4th conference, recent advances in geotechnical earth engineering, Keynote Lecture. [28] Ishihara, K., Tsukamoto, Y., Kamada, K. 2004. Undrained behavior of near-saturated sand in cyclic and monotonic loading. In: Proceedings of conference, cyclic behavior of soils and liquefaction phenomena, pp 27-39. [29] Martin, G.R., Finn W.D.L., Seed H.D. 1978. Effects of system compliance on liquefaction tests, Journal of the Geotechnical Engineering Division 104 (4), 463-479. [30] Arab, A., Belkhatir, M., & Sadek, M. 2016. Saturation Effect on Behaviour of Sandy Soil Under Monotonic and Cyclic Loading: A Laboratory Investigation. Geotechnical and Geological Engineering, 34(1), 347-358. DOI: org/10.1007/s10706-015-9949-6 [31] Finn, W.D.L., Bransby, P.L., Pickering, D.J. 1970. Effect of strain history on liquefaction of sand. J Soil Mech Found ASCE 96(SM6), 1917-1934. [32] Ishihara, K., Tatsuoka, F., Yasuda, S. 1975. Undrained deformation and liquefaction of sand under cyclic stresses. Soils Found 15(1), 29- 44. DOI:org/10.3208/sandf1972.15.29 [33] Lee, K.L., Fitton, J.A. 1968. Factors affecting the cyclic strength of soil. Vibration effects on earth- quakes on soils and foundations. ASTM STP 450, 71-95. [34] Liu, Y.J., Li, G., Zhen, Y.Y., Dano, C., Hicher, P.Y., Xiao, H.X., Wang, J.H. 2014. Influence of grading on the undrained behavior of granular materials. Symposium about: recent advances in the mechanics of granular and porous media at the European solid mechanics conference in Graz, Austria. DOI: org/10.1016/j.crme.2013.11.001 [35] Miura, S., Toki, S., Tatsuoka, F. 1994. Cyclic undrained triaxial behavior of sand by a cooperative test program in Japan. In: Ebelhar RJ, Drenvich VP, Kutter BL (eds) Dynamic geotechnical testing II, ASTM STP 1213, 246-260. DOI: 10.1520/STP13217S [36] Seed, H. B., Peacock, W. H. 1971. Test procedures for measuring soil liquefaction characteristics. Journal of Soil Mechanics & Foundations Div. [37] Monkul, M. M., Etminan, E.,§enol, A. 2016. Influence of coefficient of uniformity and base sand gradation on static liquefaction of loose sands with silt. Soil Dynamics and Earthquake Engineering, 89, 185-197. DOI: org/10.1016/j. soildyn.2016.08.001 [38] ASTM D 4253-00, 2002. Standard test method for maximum index density and unit weight of soils using a vibratory table. Annual Book of ASTM Standards. American Society for Testing and Materials, West Conshohocken, PA, pp. 1-14, www. astm.org. [39] ASTM D 4254-00, 2002.Standard test method for minimum index density and unit weight of soils and calculation of relative density.Annual Book of ASTM Standards. American Society for Testing and Materials, West Conshohocken, PA, pp. 1-9, www.astm.org. [40] Vaid, Y.P., Sivathayalan, S., Stedman, D. 1999. Influence of specimen reconstituting method on the undrained response of sand. Geotech. Test J 22(3), 187-195. DOI: org/10.1520/GTJ11110J [41] Sze, H.Y., Yang, J. 2014. Failure modes of sand in undrained cyclic loading: impact of sample preparation. Geotech GeoenvironEng 140, 152-169. DOI: org/10.1061/(ASCE)GT.1943-5606.0000971 [42] Bishop, A. W. 1967. Progressive failure: with special reference to mechanism causing it. Proceedings of the geotechnical conference, Oslo 2, 142-150. 56. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils LOAD-SETTLEMENT AND LOAD-SHARING BEHAVIOUR OF A PILED RAFT FOUNDATION RESTING ON LAYERED SOILS ODNOSA OBTEŽBA-POSE-DEK IN OBTEŽBA-DELITEV OBTEŽBE ZA TEMELJNO PLOŠČO NA PILOTIH NA SLOJEVITIH TLEH Plaban Deb (corresponding author) Sujit Kumar Pal National Institute of Technology Agartala, National Institute of Technology Agartala, Civil engineering department Civil engineering department 799046 Tripura, India 799046 Tripura, India E-mail: plaban930@gmail.com E-mail: skpal1963@gmail.com https://doi.Org/10.18690/actageotechslov.17.1.71-86.2020 DOI Keywords piled raft, load-settlement behaviour, load-sharing behaviour, load-improvement ratio, numerical analysis Ključne besede temeljna plošča na pilotih, odnos obremenitev-posedanje, odnos obtežba-delitev obtežbe, razmerje obtežba-izboljša-nje, numerična analiza Abstract To understand the load-settlement and load-sharing behaviour of a piled raft foundation resting on different layered soils, small-scale laboratory model tests were conducted. The investigational programme includes a prototype model test on an unpiled raft, braced by a single pile, 2 x 2 and 3 x 3 pile groups. 3D numerical analyses were also implemented to ensure the experimental model's verification and to explore the settlement criterion of the piled raft foundation. The influences of the number of piles, the diameter of the piles and the raft sizes on the ultimate failure load, the load-improvement ratio and the percentage load carried by the raft at different settlement levels and different types of settlements that influence the foundation design criteria are illustrated. The relative contributions of these different design variables were also found using statistical analyses. The test results show the efficacy of using piles as settlement reducers with the raft. From the studies it is clear that as the number of piles below the raft increases from 1 to 9, the load-improvement ratio increases by 20% to 69%, whereas the load shared by the raft decreases with the addition of piles below the raft. From the numerical analyses it was found that the normalized differential settlement of the rafts can be optimized by introducing a number of piles in the central area of the raft. Izvleček Za lažje razumevanje odnosov obtežba-posedek in porazdelitve obremenitve na temeljni plošči na pilotih, ki leži na različnih slojevitih zemljinah, so bili izvedeni laboratorijski preizkusi v zmanjšanem merilu. Raziskovalni program vključuje preizkus prototipnih modelov temeljne plošče, in temeljnih plošč podprtih z enim pilotom, ter skupinama pilotov 2 x 2 in 3 x 3. Za preverjanje eksperimentalnega modela in preučitev kriterija posedanja temeljne plošče na pilotih so bile izvedene 3D numerične analize. Razloženi so vplivi števila pilotov, premera pilotov in velikosti pilotov na mejno porušno obtežbo, razmerje obtežba-izboljšanje in delež obtežbe, ki jo nosi temeljna plošča pri različnih nivojih posedanja ter različnih tipov posedanja, ki vplivajo na kriterije zasnove temeljev. Relativni prispevki teh različnih projektnih spremenljivk so bili ugotovljeni tudi s pomočjo statističnih analiz. Rezultati preizkusov kažejo na učinkovitost uporabe pilotov za zmanjšanje posedkov temeljnih plošč. Iz študij je razvidno, da se s povečanjem števila pilotov pod temeljno ploščo z 1 na 9 razmerje obtežba-izboljšanje poveča za 20 % do 69 %, medtem ko se obtežba, ki jo prevzame temeljna plošča, zmanjša z dodajanjem pilotov pod temeljno ploščo. Iz numeričnih analiz je bilo ugotovljeno, da je mogoče normalizirani diferencialni posedek temeljne plošče optimizirati z uvedbo števila pilotov v osrednjem območju temeljne plošče. 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils 1 INTRODUCTION The growing need for high-rise structures and the scarcity of the available land area are forcing us to have concerns about a weaker soil base to use for foundation purposes. This is compelling us to have various alternative foundation techniques in various soil types. These foundation systems transfer the superstructure's load safely into the firm soil, confirming the overall strength and serviceability. Piled raft foundations are generally used when the isolated footing covers more than 70 % of the area under a superstructure. These are being used in several countries to sustain different kinds of superstructures, like buildings, bridges or industrial plants in various types of subsoil. Using these foundation systems, differential settlement can be reduced to a great extent as the piles improve the load-carrying capability of the raft. In the piled-raft foundation technique, the raft is directly contacted with the subsoil and hence there is a greater number of load exerts on the raft, thereby making the method very tedious, which also leads to an over-design of the foundation. Although considering the settlement reduction, a piled raft is a commonly used foundation technique in today's world. The idea of using a pile as a settlement reducer was suggested by Burland et al. [1]. Using this concept, an approach has been specified by some researchers, considering piles as settlement reducers (Randoplh [2], Burland [3], de Sanctis et al. [4], Fioravante et al. [5]). The fundamental concept of this foundation approach is to minimize the number of piles by allowing a certain limited number of piles that are enough to reduce the settlement to an acceptable level by transmitting the load from the superstructure through the raft. This methodology permits the piled raft foundation's design to be enhanced and the number of piles under a raft to be substantially reduced. However, a comparatively small number of piles in a piled raft can lead to complications like high bending moments and hence there is a tendency of the raft to offer cracking, and a large amount of axial stress originates in the pile heads (Wong et al. [6]). If the structural load is assumed to be relatively consistently distributed over the whole area of the raft, the unpiled raft tends to deflect in the center. If piles are added underneath the central portion of the raft and are loaded nearly to their probable critical capacity, they will reduce the central settlement and this will help in minimizing the differential settlement. In an orthodox piled raft foundation having a large number of piles, the piles are usually expected to take all the loads after the mobilization of pile, and the raft starts taking the loads up to its structural capacity. For the effective design of a raft with settlement-reducing piles, the capacity of the piles is anticipated to be 80 % mobilized under the external load (Clancy and Randolph [7]). The nature of a piled raft along with the self-supporting pile group and unpiled raft foundations were compared by Cooke [8], who employed some prototypical tests on the piled raft foundation. It was observed that the spacing between the piles and the quantity of piles under the raft were the focal object for understanding the load-sharing response. It was also seen that the central settlement of the raft foundation was larger than those at the edges of the raft. According to Horikoshi and Randolph [9] and Reul and Randolph [10] the strategic placement of the piles under the centre of the raft can effectively decrease any differential settlement under uniform loading conditions. The effects of a variation in the pile and raft geometry to determine the piled raft's stiffness were studied with a centrifuge test on the piled raft system (Conte and Mandolini [11]). To explore the effect of the pile installation and the pile-raft interaction, Lee and Chung [12] performed an experimental test on a piled raft foundation system in a sandy soil. The load-carrying capacity and the load-sharing behaviour of the foundation can be changed depending on the specific settlement level. Bajad and Sahu [13] accomplished a 1-g model test to study the influence of the pile's dimensions and the number of piles on the settlement behaviour and the load-sharing response with different raft thicknesses using various pile configurations such as 4, 9 and 16 piles. To study the effect of the pile's dimension and the pile's arrangement Kumar and Kumar [14] performed a 1-g model test on a piled raft foundation in sandy soil at different relative densities. Based on this experimental model test data, Kumar and Kumar [15] developed an ANN model on Matlab to predict the settlement behaviour of a piled raft foundation. In today's groundwork context, as stated earlier, deep foundations are now being furnished in every soil type, regardless of the profile of the soil and its strength criterion. However, it was found from a review of the literature that only very limited research has been accomplished on the behaviour of a piled raft on different layered soils. Every researcher has restricted their investigations to a single-layer system, whereas most of the soil profiles generally consist of multiple layers. Hence, in this paper, the load-settlement response and the load-sharing behaviour are explored through a prototype laboratory model test as well as through a finite-element analysis on various piled raft foundation systems resting on multiple layered soils. 2 LOAD-SHARING BEHAVIOUR OF A PILED RAFT The load-settlement relationship for a piled raft can be decomposed into the raft and the piles. From Fig. 1 it is obvious that the proportions of the load withstood by 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils the raft and the piles vary as a function of the settlement. The load-sharing behaviour can be described using the load-sharing ratio apr that represents the ratio of the load carried by the piles to the total load imposed on the piled raft, as follows: V = = Qp Qpr Qr+Qr = 1 _Qr_ Qpr (1) where, Qpr = load executed on a piled raft; and Qr and Qp = loads supported by the raft and the piles, respectively. The values of apr were investigated experimentally using centrifuge tests (Horikoshi and Randolph [16], Giretti [17]). Horikoshi and Randolph [16] conducted centrifuge tests and presented the values of apr for flexible foundations with different numbers of piles. It was observed that apr decreases with an increasing load level and increases with the number of piles. Giretti [17] and Comodromos [18] have also presented a similar variation of apr decreasing with an increasing settlement. Several acceptable settlements are specified and used in the design of a foundation. For an improved piled raft design, the variation of apr with the settlement needs to be identified and considered appropriately. The non-linear behaviour of the load responses in a piled raft foundation also needs to be included as it has an effect on the values of apr for a given settlement. Since the most current definitions and propositions of apr are addressed in terms of the stiffness and the geometry of a foundation, the research is much needed to idealize the settlement-reliant, non-linear variation of apr . Piled raft ultimate load B Pile ultimate load ^^^ 7A / i / i / i / i / i / i / raft I Pile capacity reached, / l'ei°re | raft before yielding / yielding | Pile and raft capacity reached -> Settlement Figure 1. Simplified contact piled raft load-settlement curve. 3 EXPERIMENTAL PROGRAMME_ A number of laboratory tests were accomplished on the models of the unpiled raft and the central piled raft with two different sizes of raft and different diameters -f o (a) 12 12 O 1 1 1 O 1 o 1 1 o 1 1 _ 0.25 . \ 7.5 .1. 8.25 O 1 CD t O -© 0- O i © t 0 Ml. 7.5 . L 7.5 .1« (b) Figure 2. Pile configurations (a) for a pile diameter of 20 mm and (b) for a pile diameter of 25 mm (all dimensions are in cm). Table 1. Model test programme. Foundation type Test denotation Dia. of pile (mm) s/d Unpiled Raft R-240 - - R-180 - - PRF-1-240-25 25 - Raft + PRF-1-240-20 20 - 1 central pile PRF-1-180-25 25 - PRF-1-180-20 20 - PRF-4-240-25 25 3 Raft + PRF-4-240-20 20 3 4 central pile PRF-4-180-25 25 3 PRF-4-180-20 20 3 PRF-9-240-25 25 3 Raft + PRF-9-240-20 20 3 9 central pile PRF-9-180-25 25 3 PRF-9-180-20 20 3 R - Raft, PRF - Piled raft foundation, s/d - Spacing to diameter ratio of the piles. The programme of the laboratory model tests on the unpiled raft and the piled-raft foundations are presented in Table 1. The pile configurations for the model raft's dimension of 240 mm are shown in Fig. 2. Similar pile configurations and spacings were also used for the raft's dimension of 180 mm. The dimensions of the model pile and the raft were chosen to ensure that no stress is concentrated at the periphery of the tank. 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils 3.1 Subsoil materials In the present study, three types of subsoil materials were used, i.e., silty-clay soil, silty-sand soil and sandy soil. The engineering and physical properties of the soil samples were evaluated using laboratory tests as per the standard techniques as recommended by the ASTM codes. The summary of the index and engineering characteristics of the subsoil materials is shown in Table 2 and the particle size distribution curves for all the soil samples are shown in Fig. 3. Table 2. Engineering characteristics of subsoil materials. Properties Silty-clay soil Silty-sand soil Sandy soil Grain size: Sand (%) 4.5 58.5 97.3 Silt (%) 43.23 35.7 2.7 Clay (%) 52.27 5.8 - Specific gravity, G 2.71 2.65 2.6 Maximum Dry Density, MDD (kN/m3) 17.05 17.41 15.35 Optimum Moisture Content, OMC (%) 23.5 16.2 13.6 Liquid Limit, LL (%) 45.2 26.32 NP Plastic Limit, PL (%) 23.78 17.12 NP Cohesion, c (kN/m2) 21.22 16.4 Cohe-sionless Angle of internal friction, f (degree) 20.54 27.4 31.2 NP = non-plastic 110 100 90 80 70 h 4> I 60 50 40 30 20 10 1 1 II II II -s-Silty-clay so -♦-Silty-sand sc il >il Sai id r Ol > / y / ✓ / { / 1 I / / I, / J, / y A V / i- smooth surface. The plate had screwed holes from the bottom surface to fix the different groups of circular piles. The outer surface of each pile top (up to 10 mm) was screw threaded to ensure a rigid connection between the pile and the raft. The arrangements of the piles and the raft are shown in Fig.4. A notch was then created in the centre of the raft to support the calibrated proving ring. In choosing the material for the model piles, aluminium was found to be more convenient than steel due to its light weight. Two sizes of pile diameter, i.e., 25 mm and 20 mm outer diameter, with 250-mm-long pipes were used as the model pile. The model piles under the rafts were smooth, hollow and circular, closed-ended, non-displacement piles. The spacings between the piles for the group pile were kept at three times the diameter of the pile. A scaling factor of 1/60 was used to select the model dimension, similar to Park and Lee [19]. The configuration for the pile groups selected in the study were a single pile, a 2 x 2 pile group and a 3 x 3 pile group. 0.01 0.1 1 Particle Size (mm) Figure 3. Particle size distribution curve. 10 Figure 4. Arrangement of piles and raft. 4 EXPERIMENTAL SET-UP AND TEST PROCEDURE 3.2 Model raft and pile materials The model raft was made of steel plates having a square cross-section with dimensions of 180 mm x 180 mm x 10 mm and 240 mm x 240 mm x 10 mm, and having a 4.1 Testing tank and test set-up The testing tank was made up of 12 mm thick translucent Perspex sheet and 10 mm thick steel plates. It had a square cross-section with a base dimension of 70 cm x 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils Figure 5. Experimental test set-up. 70 cm and having a height of 60 cm. Horizontal stiffeners were attached to two sides of the tank, where the translucent Perspex sheet was present, with nuts and bolts to avoid any bulging of the tank throughout the preparation of the soil bed and during the loading process. It is obvious that the dimensions of the test tank are large enough to overcome the scaling effects and the boundary conditions on the piled raft. The vertical load test set-up used for the piled raft is shown in Fig. 5, consisting of a soil tank, a piled raft, a screw-gear wheel, loading frames, a proving ring and two dial gauges. A movable wheel based on screw-gear mechanism was used to apply the vertical load. A vertical shaft made of steel acted as a screw jack that worked on the principle of nut and screw motion, and a manually operated movable wheel was used to control the vertical movement of the shaft. The application of the load through the shaft and wheel was controlled by the action of a ball-bearing system. The whole arrangements of the test set-up were rested on the horizontal beam of a loading frame, attached with the vertical frames, and the frames were bolted to the concrete base for better stability of the test set-up. A proving ring a having capacity of 25 kN with 0.01 kN accuracy was mounted over the steel plate's center notch via a ball bearing for measuring the vertical load and two dial gauges on each side with 0.01 mm sensitivity were used to measure the vertical deformations. 4.2 Soil bed preparation and model pile installation The depth of the soil bed was maintained at more than two times the embedded length of the pile (i.e., 54 cm) to ensure a minor effect of the rigid base response of the piles (Horikoshi and Randolph [20]). The top 6 cm thickness of the soil tank was kept free to avoid any overflow of the soil during the whole compaction process. The test was performed on three-layered soil, keeping an equal depth (i.e., 18 cm) for each layer. The top layer of the soil bed consisted of silty-clay soil; the second layer consisted of silty-sand soil; and sandy soil was kept in the bottom layer. For all the tests the bottom layer soil, i.e., the sandy soil, was used, having a relative density of 70 %. To achieve a uniform density over the soil bed, the 'sand-raining technique' was used. The relative density of the sandy soil is dependent on the height of the free fall of the sand particles and, therefore, a relationship between the relative density and the height of the free fall was obtained, and is presented in Fig. 6. The height of the free fall of the sand particles corresponding to a relative density of 70 % was found to be 270 mm. 50 60 70 Relative density (%) Figure 6. Calibration of height of free fall vs. relative density. 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils For the middle layer, i.e., for the silty-sand soil, it was placed with a bulk density (y) of 20.23 kN/m3 at a water content (w) of 16.2 %. For achieving the required bulk density, the silty-sand layer was divided into three sub layers of 6 cm each. The dry weight of the silty-sand soil was calculated and mixed with water (at w = 16.2 %), and then filled uniformly in the respective layers. A steel plate hammer (with a light compaction energy) was used to compact the silty-sand for each layer. The achieved density was checked by accumulating samples in several containers of known volume kept at several locations of each layer of the silty-sand soil during the soil filling. For all the tests the achieved density was found to vary by only ± 2 %, which was insignificant. The top-soil layer, i.e., the silty-clay soil layer, was prepared with an undrained shear strength for a given consistency. A series of UCS tests were performed on a silty-clay soil sample at different water contents to obtain the undrained shear strength at a given consistency. For all the tests, the water content of the silty-clay soil was maintained at around 34 %. The undrained shear strength and the bulk density corresponding to a 34 % water content was 10 kPa and 18.76 kN/m3, respectively. The arrangement of the installation of pile was started from the inner pile, then the corner pile, and finally the edge piles. In the first stage, the soil was poured up to the tip of the pile and then the pile was employed on the centre line alignment. These arrangements helped in the equal distribution of the load over the whole raft. After that, the sub-soil was again packed up to the necessary height. At the time of this filling process the pile was kept vertical and any inclination of the piles was detected cautiously by a level. This process of pile fabrication was assumed to optimize the stress conditions around the piles that were cast in-situ. After the completion of one test, all the piles and the soil from the tank were removed and a similar method was repeated for the subsequent tests. Prior to the test, the soil was permitted to set at room temperature for about 24-30 hours to confirm the uniform circulation of the moisture content. The vertical settlements were recorded at the end of each load increment. The rate of loading was 0.1 kN/min. For each test, the loading was continued until the settlement of the foundation reached about 25 mm (El-Garhy et al. [21]). 5 FINITE-ELEMENT ANALYSIS OF THE MODEL PILED RAFT The numerical analysis of the piled raft foundation presented in this paper was carried out with the 3D finite-element software package ABAQUS. The dimen- sions of the elements of the numerical model were kept as similar as possible to the experimental model. However, to counteract the complexity of the numerical simulation, the circular pile was replaced by square piles having the same shaft circumference (Reul [22], Sinha and Hanna [23]). An isotropic elastic Hooke's model was used to verify the elements of the pile and the raft. To account for the non-linear behaviour, elastic-perfectly-plastic stress-strain relationships were selected to establish the soil mass, considering the total stress on the model because of the undrained behaviour of the soil mass in the numerical model (Rose et al. [24], Alshenawy et al. [25]). To account for the settlement behaviour of the piled raft, elastic solutions were chosen in most of the studies (Poulos [26], Lee [27], Teh and Wong [28]). These elastic solutions are mostly based on 'Mindlin's analytical point load solution' within an infinite elastic half space. Here, the soil yielding is not considered at the pile-soil interface. However, it was also observed that the elastic solutions overestimated the settlement for the piled raft foundation (Lee et al. [29]). The Mohr-Coulomb failure criterion was selected to relate the material behaviour with the parameters described in Table 3. The loading was simulated into three steps: the geostatic step was used to generate the initial conditions where the geostatic stress field (equal to the gravitational acceleration of 10 m/s2) was applied to the soil mass, prior to the installation of the pile; to keep the model in equilibrium, self-weight was introduced in the whole model in the second step and finally the loading was applied to the top of the raft in the third step. The pore change in the model was not required to be accounted for, as the loading was applied relatively quickly (Rose et al. [24]). All the elements in the model were selected as 8-noded hexahedron brick elements (C3D8R) with a reduced integration so as to lower the Gaussian integral points that would allow the FEA solver to reduce the computational time with a minimal reduction in accuracy. Different mesh densities were utilized in this model, seeded with different aspect ratios so as to minimize the computational efforts. Thicker meshes lead to a more Table 3. Properties introduced in the modelling. Soil Layers Modulus of Elasticity (kPa) Poisson's ratio Density (kN/m3) Silty-clay soil 6x103 0.5 18.76 Silty-sand soil 30x103 0.35 20.23 Sandy soil 25x103 0.38 14.25a Pile 69x106 0.28 27 Raft 12x107 0.3 78.7 a = density corresponding to relative density of 70% 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils Figure 7. Mesh and boundary of 3D FE model. accurate analysis, but it may slow the running process. Therefore, a high concentration of meshes was generated in the vicinity of the pile-soil interface and the thinner meshes were applied in the regions of less importance. The meshing pattern and boundary conditions are described in Fig. 7. 5.1 Soil-structure interaction The interaction between the soil and the related structure plays an important role in simulating the actual loading conditions, and hence it becomes one of the major issues in identifying the accurate soil-structure interaction. When the compressive load is applied only on the piled raft, the pile conveys the shear stresses along their lateral surfaces, but when the piles come into contact with the soil surface, a frictional surface is developed according to the modified Coulomb friction theory. In this paper the Coulomb friction model was used to mimic the soil-pile interface, which related the maximum allowable shear stress (friction) across an interface to the contact pressure between the contacting bodies. An interface frictional coefficient and a limiting shear displacement of 5 mm were used to engender the elasto-plastic slip analysis. The contact surface between the pile and the raft was selected as being perfectly rigid, which means there was no relative motion between the nodes of the pile and the raft. The raft-soil interface was considered as a smooth surface. 5.2 Validation of the numerical model The vital phase of constructing a model using any numerical software is to make it compatible with the reference model in order to ensure that the numerical model is acting as expected. To confirm that the simulated model created in ABAQUS is capable of representing the actual behaviour of the laboratory model test, it is required to run various trials in numerical analyses. The model test outcomes from the actual laboratory investigations can be correlated with those of the results obtained from the replicated tests in the software and thus it can be confirmed whether the numerical model resembles to the actual situation in a realistic way or not. Thus, a series of numerical model tests were carried out for the raft and the piled raft system. The numerical model represents a copy of the experimental set-up with the same model configurations, while the loading conditions are also simulated with the experimental circumstances. Fig. 8 shows a comparison between the laboratory tests and the numerical model. As exemplified in the presented figure, no peak load is detected in the results i.e., with an increasing of the pile settlement, the vertical load increases. It is obvious that the numerical analysis provides acceptable remarks and can be utilized to predict the piled raft's behaviour. It is also observed that the numerical analysis (model scale) is in a close agreement with the results from the experimental tests. 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils Figure 8. Comparison of experimental and numerical results. 'A Poiut Comer -------ft- Figure 9. Reference settlement. 5.3 Settlement behaviour From the numerical model, the settlement was acquired at three different positions, i.e., the central point on the raft, the corner point on the raft and the point at a distance of one-quarter from the corner. From these vertical settlement values, the average settlement (Savg), the differential settlement (Sc_c) and the reference settlement ($rf) were calculated by Eqs. (2) to (4) (Reul and Randolph [10]), $avg = (2$Center + $Corner)/3 (2) where, Savg = average settlement, $Center = Settlement at the raft centre and $Corner = the settlement at the raft corner. The differential settlement and the reference settlement are generally considered as the most crucial ones, because these might be the reason for increasing the internal stress inside any superstructure, which ultimately possesses a negative impact by reducing the service life of the superstructure on which the piled raft foundation is provided. The minimal differential settlement criterion can be achieved by providing the pile group only in the central portion of any flexible raft (Prakoso and Kulhawy [30]). Fig. 9 demonstrates the schematic representation of the differential settlement and the reference settlement. The differential settlement and the reference settlement were calculated using Eqs. (3) and (4), respectively (Prakoso and Kulhawy [30]). $c-c $Center $Corner (3) $ref = ($Center + 2$1/4 + 2$Corner)/5 (4) where, $c-c = differential settlement, $ref = reference settlement, $Center = Settlement at the raft centre, $Corner = Settlement at the raft corner and $1/4 = settlement at a point at a distance of one-quarter from the corner. 6 RESULTS AND DISCUSSION_ The experimental results obtained from the laboratory tests and the outcomes obtained from the numerical simulations are analysed and discussed in this section. For the experimental data acquisition, each test was repeated twice and the average value of the two trials was selected as the final test result. In most cases the measured values for the two trials differed by less than 1 % and in some cases, the result differed by 2-3 %. If the difference in the results was found to be more than 5 %, the test data were discarded and the test was repeated. 6.1 Effect of the size of the unpiled raft and the number of piles on the load-settlement curve The experimental load-settlement curves for the unpiled raft and the piled raft are illustrated in Figs. 10-12. The load-settlement curves for all the tests indicate that the curves do not show any peak behaviour, i.e., with an increase in the settlement value, the load-carrying capability increases. From Fig. 10 it is clear that the load-carrying capability of the unpiled raft can be boosted by increasing the raft size. This phenomenon can be ascribed to the fact that the increased raft size aids in increasing the relative raft stiffness, which helps in sustaining more load. From the test outcomes it can be seen that the capability of the raft of size 240 mm x 240 mm at a settlement of 25 mm is 12.7 % more than the capacity of the raft of size 180 mm x 180 mm. From Figs. 11 and 12 it can be seen for all the cases that the load capacity of the piled raft rises as the amount of pile increases below the raft. It can also be detected that with an improvement in the number of piles below the raft, the yield load level of the piled raft system is also increased. Usually, an increased number of piles can be effective in enhancing the stiffness of any piled raft and hence the capability of piled raft is also increased. A similar trend was found by El-Garhy et al. [21]. From Figs. 11 and 12 it can also be concluded that with an increase of the pile diameter below the rafts, the load-carrying capacity is also increased. This increase might be due to the increase of the surface area and the stiffness of piled raft system as the diameter of the settlement-reducing pile increases. 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils From Fig. 13 it is seen that the piled raft behaviour shows three separate stages when subjected to an external load. In the first stage, the load-settlement relationship is found to be linear, because both the pile and the raft are elastic. In the second stage, once the load increases, the load-settlement relationship converts to more curvilinear, because the settlement rises little more than the first stage. This stage can be called the critical stage of the pile. During the final stage, the load-settlement criterion changes to linear again, although the settlement increases relatively faster, but the load does not increase substantially. In Fig. 13, during stage I, the OA is linear since the load-settlement relationship for the pile and raft are elastic. During stage II, AB is nonlinear because the pile reaches its full capacity. During stage III, BC becomes a straight line because the settlement increases quickly, but load taken by the pile does not increase substantially. This agrees with the observation of Patil et al. [31]. Load (kN) Load (kN) 4 6 Figure 10. Load-settlement curves for the unpiled raft. Load (kN) 10 12 —R-240 -■»- PRF-1-240-20 —9—PRF-1-240-25 -A- PRF-4-240-20 —A— PRF-4-240-25 -«- PRF-9-240-20 PRF-9-240-25 Figure 11. Load-settlement curves of the raft and the central piled raft. 8 •9—R-180 ■»- PRF-1-180-20 -0— PRF-1-180-25 A- PRF-4-180-20 -A— PRF-4-180-25 e- PRF-9-180-20 -Q— PRF-9-180-25 Figure 12. Load-settlement curves of the raft and the central piled raft. Load (kN) ID 15 0 ~ 5 S J- 10 a aj I 15 f 20 25 30 1 i i i i i i i * 1 1 t 1 < > t i r i \IB i Stage I T|L 1 1 \ 1 « Stage II i x » I i i i ! \c i i ! Stage III < 6.2 Figure 13. Load-settlement curve for PRF-9-240-25. Effect of the number of piles on the ultimate failure load For the present study, the ultimate failure load is determined by the tangent intersection method (Ismael [32]). In this methodology a tangent from the initial point and another tangent at the point where the curved part of the load-settlement curve changes to a steep straight line has Figure 14. Evaluation of the ultimate failure piled raft load for R- 240. 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils been drawn and their intersecting point gives the value of the ultimate failure load. The evaluation procedure for the ultimate failure load is shown in Fig. 14. The variation of the ultimate failure load with the number of piles is shown in Fig. 15. From the relation between the ultimate failure loads vs. the number of piles, it can also be anticipated that the noted ultimate failure load increases with an increase in the number of piles in a piled raft. 10 -R-240 with 25 mm dia. pile -R-240 with 20 mm dia. pile 6.3 14 9 Number of piles Figure 15. Ultimate load vs. number of piles curves of central piled raft. Effect of number of piles on the load-improvement ratio (LIR) at different settlement levels The enrichment of the external load-carrying capability of the raft at 10 mm and 25 mm settlements is defined as the load-improvement ratio (LIR). It is expressed as a fraction of the load conveyed by the piled raft to the unpiled raft at settlements of 25 mm and 10 mm. Figs. 16 and 17 demonstrate the changes in the LIR with the number of piles at 25 mm and 10 mm settlements for pile diameters of 25 mm and 20 mm, respectively. From the figures it can be seen that for the same raft sizes, as the number of piles under the raft and the pile diameter increases, the LIR also shows an increment. This is due to the fact that more load might be taken by the piles as the number of piles increases. For example, at a settlement of 25 mm, for a raft size of 240 mm x 240 mm, installing 9 piles having a 25 mm diameter increases the load taken by the raft by 69 %, while installing one pile with the same diameter increases the raft load by 20 %, as compared to the unpiled raft. From Figs. 16 and 17 it can be seen for all the cases that the LIR for a 10 mm settlement is little more than that of the LIR at a 25 mm settlement. A similar kind of variation of the LIR has been reported by El-Garhy et al. [21]. This clarifies the method of load sharing among the raft and the piles, i.e., at the start of the loading on the piled raft, the piles carry a major portion of the load, and with the increase of the settlement, the load is transferred to the raft. It means that, at the early stage of loading, the pile carries more load than the raft. This can be attributed to the pile-soil-raft interaction and enhancing more overburden pressure over the neighbouring soil, inducing a more effective stress and shear resistance. In addition, the piles can also be regarded as reinforcement elements that escalate the shear resistance of the supporting soil, due to which the load-improvement ratio decreases as the settlement level increases. - O- R-180 @ 10 mm Settlement -e— R-180 @ 25 mm settlement - o- R-240 @ 10 mm settlement -A— R-240 @ 25 mm settlement 4 6 Number of piles 8 10 Figure 16. Variation of the load-improvement ratio with the number of piles at 10 mm and 25 mm settlements for a 25 mm diameter pile. - O- R-180® 10 mm settlement -e—R-180® 25 mm settlement - □- R-240® 10 mm settlement -A—R-240® 25 mm settlement 10 0 2 4 6 8 Number of piles Figure 17. Variation of the load-improvement ratio with the number of piles at 10 mm and 25 mm settlements for a 20 mm diameter pile. 6.4 Load sharing between the raft and the pile Poulos [33] stated that the proportion of the load shared by the raft can be expressed as the inverse of the load-improvement ratio (LIR). Fig. 18 displays the variation of the proportions of the loads carried by the raft with the number of piles for the raft models of 240 mm x 240 mm and 180 mm x 180 mm at settlements of 10 mm 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils and 25 mm. The figure illustrates that the proportion of the load carried by raft continues to reduce as the number of piles in a piled raft starts increasing. This is because of the additional resistance offered by the piles by the increase in the pile bearing and the pile skin friction, for a larger number of piles. It can also be observed that the percentage of the total load that is gained by the raft decreases as the diameter of the pile beneath the raft increases. This may be due to the increase of the stiffness and the surface area the of the piled raft system as the diameter of the settlement-reducing pile increases. 0.2 -e— R-240-25@10 mm settlement - e -R-240-25@25 mm settlement -a— R-240-20@10 mm settlement - A -R-240-20@25 mm settlement 4 6 Number of piles 10 Figure 18. Proportion of the load carried by the raft vs. the number of piles. 6.5 Normalized average settlement The average settlement for a piled raft foundation having different pile and raft configurations is obtained from Eq. (5) and the normalized average settlement is then obtained from the ratio of the average settlement of the piled raft to the unpiled raft, i.e., the normalized average settlement, Sr = Savg of the piled raft/Savg of the unpiled raft. The variation of the normalized average settlement corresponding to the raft's central load is plotted in Fig. 19. It is clear from the figure that the increased number of piles increase the capacity of the piled raft and the normalized average settlement gradually decreases. The reason behind this phenomenon is that more piles would carry more load, as compared to the single pile, which would assist in lowering the settlement. From the figure it is also seen that the raft's size and the pile's size also affect the load-carrying capacity and the normalized average settlement of the piled raft. The larger size of the pile and raft showed a minimum value of the average settlement. The 3 x 3 pile group having a pile diameter of 25 mm provides the maximum load sustainability compared to the other configurations and it possesses the least value of the normalized average settlement. Z 0.2 -PRF 1-240-20 -•—PRF 1-240-25 -A- PRF 4-240-20 -PRF 4-240-25 -»-PRF 9-240-20 -e-PRF 9-240-25 6 g 10 12 14 Load (kN) 16 Figure 19. Normalized average settlement vs. raft central load. 6.6 Normalized differential settlement The relationship between the piled raft and the raft's differential displacement or normalized differential settlement (Sn) and the factor of safety (FS) is mapped in Fig. 20. The factor of safety for a raft or piled raft ensures the capacity of the external load without failure. The overall FS can be obtained from the formula used by Prakoso and Kulhawy [30] and it is labelled as, Factor of safety = (Qg + Qr)/Q (5) where, Qg = pile group compression capacity, Qr = raft compression capacity, and Q = applied load. From Fig. 20 it is clear that with the increase of FS in the piled raft system, the normalized differential settlement gradually decreases. A similar trend was also found by Prakoso and Kulhawy [30]. Among the different relationships, an exponential relationship showed a better goodness of fit with a co-relation factor of R2 = 0.79, which indicates that the normalized differential settlement has a substantial effect to ensure the level of safety for any superstructure. It can also be observed from the relationship that, although the normalized differential settlement takes an imperative role, 1.00 0.20 «„ = Z.8e-°'»re • A \ ♦ R! = 0.79 A'"-.. • A •R-180 with 20 mm dia. pile ' AR-180 with 25 mm dia. pile fr'-.. «R-240 ■ R-240 with 20 mm dia. pile with 25 mm dia. pile ♦ '■ Factor of Safety Figure 20. Normalized differential settlement vs. factor of safety. 72. Acta Geotechnica Slovenica, 2020/1 P. Deb and S. K. Pal: Load-settlement and load-sharing behaviour of a piled raft foundation resting on layered soils there are certain other factors that also need to be considered for verifying the factor of safety for any structure. 6.7 Reference settlement The reference settlement generally represents the nominal average displacement in the piled raft considering its settlement at the centre point, corner point and the one-quarter point from the corner. The reference settlement for the different piled raft configuration is obtained using Eq. (4). Fig. 21 shows the relationship between the load-sharing ratio (apr) and the reference settlement. apr is obtained for all the configurations using Eq. (1), as conferred in the preceding context. From the figure it is understood that the load-sharing ratio has a significant effect on the reference settlement for the piled raft foundation. An increase in the load-sharing ratio certifies the stiffness of any foundation, consequently moderating the settlement. This non-linear response of apr along with the reference settlement can be incorporated for the optimized design of the piled-raft foundation. The reference settlement is then normalized to that of the corresponding raft reference settlement, and 0.1 Q.6S \ • II 1.7S C/" A •-.. R1 = 0.93 ■ n.2J0 nllli 10 mm dis. Pitt AH-2J0 Avitli 2Í mm dli. Pile ♦ R-180 20 mm B ->C ->D ->E ->F ->G. The supporting structures in the pit engineering are divided into three segments, as shown in Figure 2. (1) From part a to part b and part b to part c, B=5.2 m (W1=3.7 m, W2=1.5 m). (2) From part c to part d, B=6.7 m (W1=3.7 m, W2=3.0 m). (3) From part d to part a, B=6.2 m (W1=3.7 m, W2=2.5 m). B is the width of the cement retaining wall. W1 and W2 are the widths of the slabs of the cement retaining wall. A wide cement retaining wall and the double row of 0800 cast-in-place bored piles are used as the supporting structures. Acta Geotechnica Slovenica, 2020/1 87. P. Li et al.: Numerical investigation of the influence of spatial effects and supporting structures during pit excavation Table 2. Monitoring projects. Design value Warning value Alarm value Monitoring projects Ultimate displacements value (mm) Rate of change (mm/d) Cumulative change value (mm) Rate of change (mm/d) Cumulative change value (mm) Horizontal and vertical displacements of the supporting structures A~G 35.00 2.00 21.00 2.50 28.00 Others 30.00 2.00 18.00 2.50 24.00 Deep horizontal A~G 35.00 2.00 21.00 2.50 28.00 displacements Others 30.00 2.00 18.00 2.50 24.00 Pipeline displacements - Rigidity / / / 2.00 20.00 Flexibility / / / 3.00 30.00 2.3 On-Site Monitoring Design Considering the engineering monitoring projects and geological conditions, the monitoring settings are shown in Table 2. The support structures of this foundation engineering and the shallow-buried pipeline monitoring point plane layout are shown in Figure 1. Twenty-five monitoring points are arranged on the support structures, 12 monitoring points are arranged along the enclosure structure and 6 monitoring points are arranged along the shallow-buried pipeline to monitor the horizontal and vertical displacements so that we can evaluate the impact of the pit excavation on the surrounding existing structures. The layout principles and the numbers of test points are shown in Table 3. Days/d (a) On-site monitoring horizontal displacements. (b) On-site monitoring vertical displacements. Figure 3. On-site monitoring displacements of the pipeline. Table 3. Monitoring station layout numbers and principles. Numbering Monitoring projects Quantity Layout principle Monitoring point number 1 Horizontal and vertical displacements of the supporting structures 25 Set one monitoring point along the top of the pile at 20-25m QL1~QL25 2 Deep horizontal displacements 12 Set one monitoring point along the perimeter of the enclosure structures at 25-30m CX1~CX12 3 Pipeline displacements 6 Set one monitoring point along the buried pipeline at 20-25m G1~G6 Acta Geotechnica Slovenica, 2020/1 87. P. Li et al.: Numerical investigation of the influence of spatial effects and supporting structures during pit excavation The results of the on-site monitoring are shown in Figure 3. Both the horizontal and vertical displacements of the pipeline are compatible with the construction conditions of the foundation pit and show a synchronous trend. With the excavation of the pit, the horizontal and vertical displacements of the pipeline monitoring points gradually increase with the excavation depth. Then, the horizontal and vertical displacements of each monitoring point gradually decrease and eventually stabilize. The maximum value of the cumulative amount of horizontal displacement reaches 5.43 mm; the maximum value of the vertical displacement reaches 5.19 mm; both of which are far below the alarm value of the cumulative displacement of 20 mm. In addition, a displacement singularity occurs for the horizontal displacement of the monitoring point G6. This singularity can be caused by temporary loading on the ground and measurement errors when monitoring at this point. 3 NUMERICAL ANALYSIS OF THE PIT EXCAVATION 3.1 Geometric models A three-dimensional numerical analysis using ABAQUS is performed. Because of the asymmetry of the foundation pit model and pipeline position and the asymmetry of the pit excavation, it is necessary to carry out a comprehensive three-dimensional modeling of the project. According to the engineering background, the long east edge of the pit is 170 m, the southwest length of the Stone Bridge Waterway is 150 m, and the depth of the pit excavation is 5.85 m. The back boundary of the supporting structure is five times the depth of the excavation, as is the bottom boundary. Additionally, the scope of influence is 20 m outside of the pipeline. As a result, the total length of the final model is 240 m, the total width is 225 m, and the total depth is 30 m. 3.2 Soil properties and constitutive models In the three-dimensional numerical simulation calculation, the Mohr-Coulomb model is selected as the soil model. The bored pile, cement retaining wall and pipelines are all selected as isotropic elastic models. The thickness, compressive modulus, cohesive force and internal friction angle of the different soil layers in the model are evaluated according to the geological survey data measured at the beginning of the project. The physical parameters are shown in Table 1. The physical and mechanical parameters corresponding to each structural material are shown in Table 4. In the Table 4. Physical and mechanical parameters of soil parameters. Supporting structure Y (kN/m3) E(GPa) £ Cast-in-situ bored pile 24.50 30.00 0.20 Concrete retaining wall 18.30 0.20 0.28 NOTES: All the data were obtained from the laboratory tests. y = the unit weight; E = the elastic modulus; £ = the Poisson's ratio pit-supporting structure system of this project, the water curtain of the cement mixing pile and the cement retaining wall with a variable section width are soil-cement materials, while the double row of cast-in-place bored piles are C30 concrete. 3.3 Finite-element meshes and boundary-condition settings The total number of grid divisions in the model is 84205 units, of which the number of soil units is 72377 and the number of nodes is 83776. The number of units in the pipeline is 1136, and the corresponding number of nodes is 1728. The number of support structure units is 10692, and the corresponding number of nodes is 14652. Each material in the model uses a C3D8 hexa-hedral solid element. The boundary conditions are set in the analysis step to limit the horizontal and vertical displacements of the bottom surface of the geometric model (horizontal displacement equals 0, vertical displacement equals 0). While limiting the horizontal displacements around the soil model (x-direction displacement equals 0 or z-direction displacement equals 0), the vertical displacements can be changed freely. The grid division and boundary-limit displacements are shown in Figure 4. I Figure 4. Meshing and boundary limit diagram. Acta Geotechnica Slovenica, 2020/1 87. P. Li et al.: Numerical investigation of the influence of spatial effects and supporting structures during pit excavation Pipeline length / m .3 13 I" cf Á oS VT ¿y,sy 0{ 9771854017001