Seismic Response of Wharf Structures Supported in Liquefiable Soil Andriani Ioanna Panagiotidou Diploma in Civil Engineering (2009), National Technical University of Athens, Department of Civil Engineering Submitted to the Department of Civil and Environmental Engineering in Partial Fulfillment of the Requirements for the Degree of Master of Science in Civil and Environmental Engineering at the MASSACHUSETTS INS fIE OF TECHNOLOGY Massachusetts Institute of Technology JUL 0 8 2013 June 2013 LIBRARIES © 2013 Massachusetts Institute of Technology. All rights reserved. Signature of Author Department of ti)vil and Environmental Engineering May 13, 2013 Certified by Andrew J. Whittle Professor of Civil and Environme tal Engine ring I T sispupe isor Accepted by Heidi M. Nepf Chair, Departmental Committee for Graduate Students Seismic Response of Wharf Structures Supported in Liquefiable Soil Andriani loanna Panagiotidou Submitted to the Department of Civil and Environmental Engineering on May 13, 2013, in Partial Fulfillment of the Requirements for the Degree of Master of Science in Civil and Environmental Engineering Abstract This research analyzes the vulnerability of wharf structures supported on loose granular waterfront fills that are susceptible to liquefaction during seismic events and considers the effectiveness of pre-fabricated Vertical (PV) drain systems in mitigating potential damage. The analyses are based on non-linear finite element simulations using the OpenSees open-source software framework. The analyses make extensive use of an advance macroelement formulation by Varun (PhD, 2010), which captures efficiently the fundamental mechanisms of saturated granular soil behavior. The thesis explains in detail the mechanical components of the macroelement as well as the necessary calibration steps. Numerical simulations of a typical berth of port facilities on the US West coast have been carried out using as earthquake loading the time histories of free field displacements and of excess pore pressures predicted by Vytiniotis (PhD, 2011) at specific locations along the embedded length of piles (for a suite of 56 ground motions). The results show that the primary cause of the structural damage is indeed the lateral spreading of the soil and therefore retrofitting efforts should be targeted in limiting the development of pore pressures that cause the lateral spreading. This argument is then tested by comparing the performance in numerical simulations for the case where full-depth PV drains are installed at locations behind the crest of the slope (i.e. minimally-intrusive mitigation system) and for the case of a fully densified slope for the same suite of ground motions. These results indicate that soil improvement methods assist in reducing structural damage to pile-deck connection locations. The results also show that the densification of the slope is only marginally superior to the less intrusive improvement with PV-drains. Thesis Supervisor: Andrew J. Whittle Title: Professor of Civil and Environmental Engineering Acknowledgements First and foremost I would like to thank my advisor Prof. Andrew J. Whittle for transforming me from a student to a researcher and for so patiently editing this thesis. I would like also to thank Linde Foundation for the fellowship that I was awarded the first year of my studies as well as Alexandros S. Onassis Foundation for its support during my SM years. Special thanks to Dr. Vytiniotis for being always a "deus ex machina" and to Gonzalo, Despina, Yixing and Davoud for all their valuable help and support. I would also like to thank Dr. Varun and Dr. Shafieezadeh for providing useful code. Finally I would like to express my gratitude to my family and my friends who have stood by me despite my many faults. 5 6 Contents 3 Abstract................................................................................................................................................... 1 Introduction.........................................................................................................................................15 Thesis Overview ........................................................................................................................... 17 Literature Review ................................................................................................................................ 22 1.1 2 3 2.1 Soil-Structure Interaction............................................................................................................22 2.2 Soil-Structure Interaction for Pile Foundations...................................................................... 23 2.3 State of the art methods of analyses ..................................................................................... 25 M odel of Soil-Structure Interaction ................................................................................................ 31 Overview of M acro-elem ent Form ulation of the M acroelem ent ........................................... 31 3.1.1 Drained Behavior of M acroelem ent............................................................................... 32 3.1.2 Undrained Behavior of the Macroelem ent ..................................................................... 35 3.1.3 Partially Drained Behavior of the M acroelem ent .......................................................... 40 3.1 3.2 4 5 44 Calibration of the Varun (2010) Macroelem ent for Toyoura Sand ........................................ 62 Seism ic Response of Pole-supported W harf Structures................................................................. 4.1 Problem Description....................................................................................................................62 4.2 Ground M otions and Free Field Analyses .............................................................................. 65 4.3 Results of Untreated Scenario................................................................................................. 66 4.4 Effectiveness of Ground Improvement System s ................................................................... 71 4.5 M odeling of Pile-Deck Connection.......................................................................................... 73 4.6 Conclusions..................................................................................................................................78 Sum mary, Conclusions and Recom mendations................................................................................104 104 5.1 ................................................................................................................................... Sum ma .ry 5.2 Conclusions................................................................................................................................105 5.3 Recom mendations .................................................................................................................... 107 REFERENCES .............................................................................................................................................. 108 APPENDIX A - Validation of OpenSees Response ..................................................................................... 111 Methodology ......................................................................................................................................... 111 Vertically Loaded Piles...........................................................................................................................112 Elastic Springs....................................................................................................................................112 Elastic Perfectly Plastic Springs ......................................................................................................... 7 113 Laterally Loaded Piles............................................................................................................................114 Elastic Springs....................................................................................................................................114 Elastic Perfectly Plastic Springs ......................................................................................................... 114 M acroelem ent Springs ...................................................................................................................... 115 Com parison of the Pushover Response of the three Spring Elem ents ................................................. 116 Dynam ic M odeling of a single pile ........................................................................................................ 116 APPENDIX B -Classification of Dam age for Piled-W harf Structure...........................................................124 8 Table of Figures Figure 1-1 Seismic hazard for USA by USGS (PGA with 2% PE in 50 year), (USGS, 2008). ........ 19 Figure 1-2 Typical soil-foundation-structural system for pile supported wharf (not to scale) 20 (Shafieezadeh, 20 11) .................................................................................................................... Figure 1-3 Shearing of a pile by ground displacement in Kobe earthquake, 1995 (Finn and Fujita, 20 20 0 2)............................................................................................................................................. Figure 1-4 Damage to a pile by 2 m of ground displacement in Niigata earthquake, 1964 (Finn 21 and Fujita, 200 2). .......................................................................................................................... Figure 1-5 Undamaged pile supporting a crane rail in ground which moved more than 1.0 m during the Great Hanshin Earthquake (Finn and Fujita, 2002)................................................. 21 Figure 2-1 (a) Geometry of soil-pile-structure interaction problem; (b) decomposition into kinematic and inertial interaction problems; (c) two-step analysis of inertial interaction 29 (modified from Varun 2010)..................................................................................................... Figure 2-2 Various observed modes of pile failure in liquefiable soils (Tokimatsu et al, 1996)... 30 Figure 2-3 Typical Winkler spring and dashpot model for pile foundation analysis (from 30 Kavvadas & Gazetas, 1993)....................................................................................................... Figure 3-1 Schematic Formulation of the Macroelement (modified from Varun, 2010)...... 49 Figure 3-2 Effective vertical stress (Pa) plot after seven loading cycles showing the formation of local liquefaction zone - approximately five diameters around the pile (from Varun, 2010)...... 50 Figure 3-3 Relationship between excess pore pressure and shear work for undrained cyclic shear tests on sand (after Towhata and Ishihara, 1985).......................................................... 50 Figure 3-4 Envelope of stress points at equal shear work (Towhata and Ishihara, 1985) ........ 51 Figure 3-5 The liquefaction front concept as extent by Varun for pile. The normalized vertical effective stress S=-- V can be calculated as a function of the normalized soil resistance r=p/Ba o --------------------------- ...... ........................................ -.. -----...............................---- 52 Figure 3-6 Liquefaction front parameter (S0 ) correlation with normalized shear work for different values of (a) critical state friction angle (b) pile diameter (c) maximum friction angle 53 (figure from Varun (2010))............................................................................................................ Figure 3-7 (a) Liquefaction front parameter (S0) correlation with normalized shear work as a function of liquefaction resistance parameter X (b) Shear work correlation curves after 53 normalization with parameter w1 (figure from Varun (2010)). ................................................ 9 Figure 3-8 Parameter r as a function of Poisson's ratio and power exponent n (from Varun (20 10 ))........................................................................................................................................... 54 Figure 3-9 Flowchart for calculation of pile response using the macroelement (Varun, 2010).. 55 Figure 3-10 Pile response to small loops of cyclic unloading-reloading after monotonic loading (Varu n, 20 10 )................................................................................................................................ 56 Figure 3-11 p-y curve for the macroelement. We can observe the degradation of the stiffness and the lateral resistance of the macroelement as a result of the cyclic loading and the pore pressu re built up........................................................................................................................... 56 Figure 3-12 Sinusoidal Input Displacement motion and b. Assumed pore pressure build up .... 57 Figure 3-13 Physical model of Berth 60-63 and FE model properties as used by Vytiniotis (PhD 2011). (Figure from Vytiniotis (PhD (2011))............................................................................... 58 Figure 3-14 Young's Modulus distribution along the depth of the pile. The darker grey line corresponds to the upper hydraulic fill, which is modeled with Toyoura Sand of Dr=40 %. The lighter grey line plots the corresponding distribution along depth of Young's modulus fo Toyoura Sand of Dr=80%. After the depth of 18.4 m below the surface, for the calibration of the macroelement we are using the properties of Dr=80 %Toyoura Sand............................. 59 Figure 3-15 p-y curve for the macroelement. Comparison of the medium and the dense behavior normalized over the vertical effective stress. We have used the same initial stiffness as the purpose isto presented the qualititave differences in the backbone curves.................... 60 Figure 3-16 p-y curve for the macroelement. Demonstration of the influence on the overall response of the macroelement of the parameter X.The denser the sand, the lower the value of X,the larger the soil resistance. Moreover it is worth noticing that the looser sand (X =0.15) liquefies to a higher degree. All the other parameters used same as Figure 3-11.................. 61 Figure 4-1 Configuration of the Berth 60-63 in Oakland (Shafieezadeh, PhD 2011)................ 80 Figure 4-2 OpenSees model of piled wharf structure. ............................................................. 81 Figure 4-3 Constitutive law used to model (a)the moment resistance of the pile section and (b) the axial resistance . ...................................................................................................................... 82 Figure 4-4 Moment curvature curve of the reinforced concrete section of the pile as calculated by online available softw are KSU_RC (2011)............................................................................. 82 Figure 4-5 Geometry of PV-Drains array for mitigating seismic risk for Berth 60-63 (Vytiniotis, 20 11) ............................................................................................................................................. 83 Figure 4-6 Response of the unimproved piled wharf to base case ground motion ,nga0779; (a) time history of acceleration of rock base excitation (b) acceleration contours of free field response (from Vytiniotis, 2011) (c) displacement contours of free field response (from Vytiniotis, 2011) (d) horizontal displacement of the deck ........................................................ 84 10 Figure 4-7 Response of the unimproved piled wharf for ground motion nga0779 at t= 5s (a) Structural displacements (60:1) (b) Snapshot at the same time of the contours of displacement of the free field analysis (Vytiniotis (PhD, 2011)) (c) Snapshot at the same time of the contours of excess pore pressure ratio of the free field analysis (Vytiniotis (PhD, 2011)). ..................... 85 Figure 4-8 Response of the unimproved piled wharf for ground motion nga0779 at t= 10s (a) Structural displacements (60:1) (b) Snapshot at the same time of the contours of displacement of the free field analysis (Vytiniotis (PhD, 2011)) (c) Snapshot at the same time of the contours of excess pore pressure ratio of the free field analysis (Vytiniotis (PhD, 2011)). ..................... 86 Figure 4-9 Response of the unimproved piled wharf for ground motion nga0779 at t= 15s (a) Structural displacements (60:1) (b) Snapshot at the same time of the contours of displacement of the free field analysis (Vytiniotis (PhD, 2011)) (c) Snapshot at the same time of the contours of excess pore pressure ratio of the free field analysis (Vytiniotis (PhD, 2011)).. .................... 87 Figure 4-10 Response of the unimproved piled wharf for ground motion nga0779 at t= 20s (a) Structural displacements (60:1) (b) Snapshot at the same time of the contours of displacement of the free field analysis (Vytiniotis (PhD, 2011)) (c) Snapshot at the same time of the contours of excess pore pressure ratio of the free field analysis (Vytiniotis (PhD, 2011)). ..................... 88 Figure 4-11 Performance of the unimproved piled wharf under ground motion nga0779 (a) bending moment envelopes and (b) location of the plastic hinges in piles............................. 89 Figure 4-12 Time histories of deck displacements for unimproved piled wharf structure for the 90 reference suite of 56 ground motions ..................................................................................... Figure 4-13 Summary of computed deck displacements for unimproved piled wharf structure as functions of earthquake intensity parameters for the reference suite of 56 ground motions 91 a)PGA, b) PGV, c) PGD, d) Arias Intensity ................................................................................ Figure 4-14 Response of the unimproved piled wharf structure and classification of structural 92 damage for the reference suite of ground motions ................................................................. Figure 4-15 Examples of structural damage in selected seismic ground motions for the 93 unim proved piled w harf structure............................................................................................ Figure 4-16 Comparison with results of Shafieezadeh (2011): a) Time Histories of Horizontal deck displacements; b) yielded sections as calculated by Shafieezadeh (2011); c) yielded sections as calculated by Panagiotidou (2013).......................................................................... 94 Figure 4-17 Effectiveness of ground improvement scenarios for piled-wharf structure under reference suite of ground motions as functions of selected intensity parameters: a)PGA, b) PGV, 95 c) PG D, d) A rias Intensity ......................................................................................................... Figure 4-18 Effect of ground improvement systems on fill and wharf response under ground 96 motio n nga0 779 ............................................................................................................................ 11 Figure 4-19 Effect of ground improvement systems on fill and wharf response under ground motio n nga0753 ............................................................................................................................ 97 Figure 4-20 Effect of ground improvement systems on deck displacements under ground motio n nga0779 ............................................................................................................................ 98 Figure 4-21 Effect of ground improvement systems on deck displacements under ground motio n nga0753 ............................................................................................................................ 98 Figure 4-22 Section geometry and reinforcement details of the tested pre-stressed pile (from Lehm an et al., 2009) ..................................................................................................................... 99 Figure 4-23 Bilinear hysteretic model with strength limit (from Ibarra et al.2005).................. 99 Figure 4-24 Modified Ibarra Krawinkler (MIK) Deterioration Model (from Lignos et al. 2011). 100 Figure 4-25 The calibration results of the MIK model for pile-deck connection for the simulated experim ent (Figure 4-22)............................................................................................................101 Figure 4-26 Effect of deteriorating pile-deck connection on response of the unimproved piledwharf structure under base case ground motion, nga0779....................................................... 102 Figure 4-27 Effect of deteriorating pile-deck connection on response of the unimproved piledwharf structure under base case ground motion, nga0779: (a)MIK, (b) EPP............................ 103 Figure A-1. Settlement along the length of the pile supported on Elastic Springs. ................... 117 Figure A-2 Settlement along the length of the pile supported on Elastic- Perfectly Plastic Springs. ..................................................................................................................................................... 118 Figure A-3 Horizontal deformation along the length of the pile supported on Elastic Support. 119 Figure A-4 Constitutive Behavior of springs in lateral loading ................................................... 119 Figure A-5 Horizontal deformation along the length of the pile supported on Perfectly Plastic Springs. We can observe a small difference between the numerical and the analytical solution due to the assumption of the analytical solution that the pile is clamped at the depth where the deform ation changes sign........................................................................................................... 120 Figure A-6 p-y curve for the macroelement. We can observe the degradation of the stiffness and the lateral resistance of the macroelement as a result of the cyclic loading and the pore pressure built up. The excitation is sinusoidal with frequency of 0.1 Hz, amplitude of 0.02 m and total number of cycles 6. We are assuming a linear pore pressure built-up. ........................... 121 Figure A-7 a. Horizontal Reactions b. Horizontal Displacements c. Force-Displacement curves at various depths along the pile. The macroelement exhibits a behavior that is bounded but the responses of the elastic and the elastic-perfectly plastic springs. We can see that the reaction distribution is smoother as it is anticipated from the Bouc-Wen type hysteresis model.......... 122 Figure A-8 Schematic model of earthquake loading of a single pile .......................................... 12 123 Table of Tables 48 - - - - - - --.............................. ..... -----.. ---------------- .---------..-........................... 48 --- 79 . -------.. . ----------------------...........................---- Table 3-1 ................................................................. Table 3-2 ................................................................... Table 4-1 ................................................................ Table 4-2 ..............................................................................--..---------------------......................-- -- - - - 79 86 ...... --------...... ------------------......................... Table 4-3 ....................................................................... 87 ........... - -...------------------------.......................... Tab le 4-4 ..................................................................... 13 14 1 Introduction Liquefaction is a term used to describe the underlying mechanisms of strength loss and ground deformations that occur due to the accumulation of excess pore pressures. It is most commonly associated with cyclic shearing during earthquakes and damage is limited over the short timeframe of the loading event. Liquefaction is commonly observed in seismic events at sites with loose, saturated granular materials often causing large settlements or lateral movements within slopes, and causes damage to embedded structures and facilities. Maritime trade accounts for 80-85 % of international trade and port facilities are exposed to a variety of natural hazards including earthquakes, tsunamis, and hurricanes that can lead to significant disruptions in operations and economic losses. The 'Great Hanshin' earthquake (http://www.city.kobe.lg.ip/foreign/english/disaster/index.html) resulted in almost complete destruction of the port facilities resulting in huge short and long-term economic losses for the city of Kobe. Kobe was one of the world's busiest ports prior to the earthquake. Although the facilities were rebuilt within 2 years, the port has never regained its former status as Japan's principal shipping port. More recently, the facilities of the Port au Prince, Haiti seaport were severely damaged in the January 2010 earthquake (http://www.telegraph.co.uk/news /worldnews/centralamericaandthecaribbean/haiti/7032385/Haiti-earthquake-damaged-port-reopens to -aid-ships.html) due to the liquefaction of the underlying soils and were unable to handle aid shipments almost a month after the event. 15 Many US ports are located in areas with significant seismic hazard on both the West (Oakland, Los Angeles, Long Beach, and Seattle) and East coasts (Charleston, SC, and Savannah, GA), as seen in seismic hazard map from USGS, Figure 1-1. There is a constant need to evaluate the seismic risk, and suitable mitigation measures, (and their associated costs) that can be implemented without severe disruption to ongoing port activities. This thesis focuses on particular mitigation scheme that uses an array of PV drains to prevent pore pressure accumulation on the landward side of the wharf (originally proposed by Rathje et al, 2004). This research focuses on the performance of a common type of wharf structure comprising pilesupported deck that typically supports a rail-mounted container crane (also founded on piles Figure 1-2). The seismic vulnerability of these types of structures is primarily due to the lateral spreading and slope failure within the embankment fills, which often comprises loose granular fills (mostly constructed prior to current seismic design codes). Therefore, the piles that are embedded within these fills can be subjected to large lateral loading due to the spreading (i.e. lateral deformations within the soil). Figures Figure 1-3 to Figure 1-5 show examples of pile failures: Figure 1-3 shows a pile-supported warehouse on Port Island (near Kobe City) that failed completely due to shearing associated with permanent lateral ground displacements greater than 1.5 m, while Figure 1-4 shows damage to a reinforced concrete pile under a building in Niigata (1964) earthquake caused by 2 m of ground. In both of these examples the piles were designed primarily for vertical loads and had limited ability to support bending moments and shear caused by strong shaking and lateral spreading. Figure 1-5 shows a crane rail pile that did not sustain any structural damage in the Kobe earthquake even though the 16 ground deformation was approximately 1 m. In this case, the design was adequate for the imposed structural loads and ground movements. This thesis focuses on the dynamic response of a pile-supported wharf structure embedded in a loose granular fill. The performance is evaluated using an uncoupled substructure approach, which involves separate analyses of the free-field response of the soil mass (i.e. without structural elements) and of the wharf structure (piles, deck, and crane). This research builds on prior research by Vytiniotis (2011) and Shafieezadeh (2011). The interaction between the free field and pile-deck models is handled through 'macro-elements' (Varun, 2010) that require the time histories of free-field displacements and pore pressures at locations along each of the piles as input motions. The thesis evaluates structural damage for a reference wharf structure for a suite of selected earthquake ground motions. It further evaluates the effectiveness of current soil improvement methods in limiting the structural damage in wharf structures during seismic loading. 1.1 Thesis Overview Chapter 2 presents a review of the literature analysis of seismic soil-structure interaction for pile-supported wharves. Chapter 3 gives details of the macroelements used to represent pile-soil interaction and the calibration of the input parameters. Chapter 4 presents results of two dimensional analyses of the response of a typical piled-wharf structure under a suite of earthquake ground motions and compare the structural response of 17 the wharf under its existing condition and after installation of a PV drain system. The study focuses on the effectiveness of the PV drains in mitigation damage to the wharf structure. Chapter 5 summarizes the research, provides a set of general conclusions drawn from the outcomes of the research, and outlines future research needs. 18 C' 0O ui IL LA 04 c~IA E Figure 1-1 Seismic hazard for USA by USGS (PGA with 2% PE in 50 year), (USGS, 2008). 19 I Container Crane Seaside Pre-stressed~ batter piles Pre-stressed vertical piles Figure 1-2 Typical soil-foundation-structural system for pile supported wharf (not to scale) (Shafieezadeh, 2011) Figure 1-3 Shearing of a pile by ground displacement in Kobe earthquake, 1995 (Finn and Fujita, 2002). 20 I" Figure 1-4 Damage to a pile by 2 m of ground displacement in Niigata earthquake, 1964 (Finn and Fujita, 2002). Figure 1-5 Undamaged pile supporting a crane rail in ground which moved more than 1.0 m during the Great Hanshin Earthquake (Finn and Fujita, 2002). 21 2 2.1 Literature Review Soil-Structure Interaction The term free field motions refers to the situation where the ground response during an earthquake is not affected by the presence of the structure. For example, a structure founded on stiff rock, the very high stiffness and strength of the founding rock is often sufficient to ensure that motions are well approximated by the free field case. However, the foundations of structures founded on softer soils generally do not conform to the ground response and will alter the motion experienced by the structure (kinematic interactions), while the dynamic response of the structure itself can induce deformations of the surrounding soil (inertial interactions). This process, in which the response of the soil influences the motion of the structure and vice versa, is referred as soil-structure interaction (SSI). There are two general methods to quantify soil SSI effects, namely the direct approach and the substructure approach. The direct approach accounts for the interaction by simultaneously analyzing the soil and the structure and their interfaces in a single step. While it is possible to carry out direct analyses, they are usually very computationally demanding. The substructure approach decomposes the kinematic interaction and inertia interaction problems. The response of the system is the result of the superposition of the responses (1) due to the kinematic interaction effect, involving the response to base rock excitation of the system is shown in Figure 2-1b, which considers a massless super-structure while inertial interactions (Figure 2-1b) considers the response of the pile-soil system to excitation by D'Alembert forces, 22 -Mii, where accelerations, ii,, are obtained from the kinematic interaction. This kinematic-inertial superposition is only theoretically valid for linearly elastic systems. However Gazetas (1984) observes that in the majority of actual cases, pile deformations due to lateral excitation (transmitted from the super-structure) attenuate very rapidly with depth (typically within 10-15 diameters from the ground surface). Therefore, shear strains induced in the soil due to inertial interaction may be significant only near the ground surface. In contrast, vertical S-waves induce free-field shear strains that are more important at depth. Thus, since soil strains are controlled by inertial effects near the ground surface and by kinematic effects at greater depths, the superposition may be approximately valid even if nonlinear soil behavior is expected, during a strong base excitation. Finally the inertial interaction analyses in the frequency domain can be conveniently performed in two steps as originally proposed by Kausel and Roesset (1975) for embedded foundations, (Figure 2-1c). The primary advantage of the substructure approach is its flexibility and computational efficiency. 2.2 Soil-Structure Interaction for Pile Foundations The seismic design of pile foundations in loose granular soils poses several difficulties problems in analysis and design: 1) liquefaction is associated with a large reduction in shear stiffness of the supporting soil and can induce large shear forces and bending moments within the piles leading to severe cracking and formation of plastic hinges at specific locations (Finn & Fujita, 2002). After liquefaction, the residual strength of the soil may be less than the shear stresses needed for slope equilibrium and significant lateral spreading or downslope displacements can 23 also occur (i.e., post-shaking lateral spreading can cause substantial increases in pile cap displacements above those for the non-liquefied case). In addition, the moving soil can exert damaging pressures against the piles, leading to failure. Such failures were prevalent during the 1964 Niigata and 1995 Kobe earthquakes. Lateral spreading is particularly damaging when the piles are embedded in soil profiles including both liquefiable and non-liquefiable (stable) soil layers. The most commonly observed modes of pile failure during liquefaction are illustrated in Figure 2-2 (Tokimatsu et al., 1996). For end bearing piles, damage in the pile heads can be caused by the horizontal forces and/or the overturning moments imposed on the foundation by the superstructure that can cause excessive shear forces on the piles (Figure 2-2b and f). Another potential failure can be initiated by the settlement of the adjacent ground that would lead in an increase in the loads on the piles and cause rigid body vertical settlements on the superstructure (Figure 2-2c). Finally transient ground deformation can apply significant loads to the supporting piles especially when the pile goes through a non-liquefiable layer, in which case the moment demand at the interface of the two layers can exceed the moment capacity of the pile section (Figure 2-2g). Lateral spreading can cause pile bending, and can be accompanied also with simultaneous loss of pile capacity (Figure 2-2d and h). In the case of friction piles, damage is mostly associated with large settlement of the superstructure and or tilting due to bearing capacity failure induced by soil liquefaction (Figure 2-2a and e).. 24 2.3 State of the art methods of analyses It is apparent from the preceding discussion that seismic response of a pile foundationstructure system should be analyzed using nonlinear dynamic numerical analyses (finite element or finite difference methods). However, these types of analyses are still not popular in engineering practice due to their complexity and computational costs. A range of approximate analysis has been developed, including pseudo-static and simplified dynamic analyses. Pseudo-static structural analysis of a single pile typically solves the bending moment and shear force distributions, where soil-structure interactions are represented by a distribution of nonlinear Winkler springs (p-y curves) and the earthquake loading is represented by an equivalent static inertia force applied at the pile cap (F = PGA'm, where PGA is the peak ground acceleration and m the mass of the superstructure). The key limitations of this approach relate to the selection of p-y curves that are derived from static and cyclic lateral load tests, and are not directly related to mechanisms of stress changes in the soil and pore pressure development associated with loading. Moreover, it is should be noted that this method tends to over predict the maximum bending moments because the inertial forces and soil resistance are not always in phase (Brandenberg et al., 2007). Simplified dynamic analyses account for both the stiffness of the soil and radiation of energy away from the pile (radiation damping). Such methods involve an initial calculation of the onedimensional vertical wave propagation, representative of the response of an infinite, level soil domain (Kavvadas and Gazetas, 1993,). One can then simulate numerically the response for a 1- 25 D column of pile elements in which all nodes at the same vertical elevation are fixed to have the same vertical displacement, thus ensuring that the one-dimensional soil column deforms as a shear beam (Figure 2-3). The free-field motions are imposed at the free field end of complex Winkler springs for the assumed soil profile. The complex spring consists of two frequency dependent parts (i) a spring that simulates soil stiffness, and (ii) a dashpot that accounts for radiation damping. Researchers have developed both closed-form, semi-empirical frequency dependent complex springs of pile foundation embedded in elastic soil (Novak, 1974; Dobry et al., 1982; Kaynia and Kausel, 1982; Kavvadas and Gazetas, 1993). These methods are generally formulated in the frequency domain and give the pile response for a single harmonic excitation. Their application to earthquake motion therefore requires the application of Fourier transforms. Similar efforts have led to the development of frequency independent models for elastic soil for example Kagawa and Kraft (1981), Nogami and Konagai (1988), and Tabesh and Poulos (2001). Some Winkler type models have been developed by Kagawa (1992) and Liyanapathirana & Poulos (2005) for liquefying soil. However, those have limited predictive capabilities since they neglect important features of dynamic pile behavior such as soil-pile gapping and displacement hardening that istypically observed for sand of medium density (Varun, 2010). Finally, several attempts have been made to simulate the soil-pile-structure interaction using macroelements. Macroelements are derived by integrating the material behavior over the 26 locally affected volume. The global stress-strain response is the applied as external loading at representative locations along the soil-structure interface. Bounded by limit equilibrium conditions, the macroelements can simulate the coupled effects of soil plasticity and interface nonlinearities, and thus have a substantial advantage over the simplified analytical procedures. Successive, decomposition of the far-field and near-field domain allows efficient frequencydomain methods to be employed in the far field, since analysis of the superstructure supported by macroelements incorporates nonlinear soil-structure interaction effects. However it should be pointed out that the results are highly dependent on the constitutive soil model and loading path used to calibrate the model as the mathematical expression describing their behavior has been mainly derived from curve fitting of the numerical or experimental results. Boulanger et al. (1999) proposed a macroelement formulation for soil-pile interaction. The components of the macroelement included scaled replicas of the API p-y curves for nonliquefiable soils. In order to capture the effects of liquefaction, a dashpot was included to account for radiation damping, and a gap element to account for strain hardening. Boulanger et al. (1999) report reasonable agreement with centrifuge data, however, as Varun (2010) noted, the macroelement formulation is unable to distinguish the response of pile in soils that differ in terms of hydraulic conductivity, liquefaction resistance or dilation angle as the hardening simulation capabilities of the macroelement are independent of soil properties. Moreover, Boulanger et al. (1999) macroelement cannot account for seepage effects that may develop between the far-field and the region around the piles (Gonzalez et al., 2009). Varun (2010) developed a macroelement for soil-structure interaction analyses of piles in liquefiable soil that 27 attempts to capture efficiently the physical mechanisms of saturated granular material without the aforementioned drawbacks. The formulation of this macroelement is described in Chapter 3. 28 M (a) UU~t UO t) U64t) MU.t) . U,(t) UIt) (b) Uct) Ui) (i) Knmatc Interaction (ii) Inrta Interaction A (K) (c) (ii) (i) Figure 2-1 (a) Geometry of soil-pile-structure interaction problem; (b) decomposition into kinematic and inertial interaction problems; (c) two-step analysis of inertial interaction (modified from Varun 2010) 29 Setternent of adjacent Lose of pile cAPacity Loss of p-e capacity lateral preading ground Failure due to overturning moment Failure due to Iteral apreading Fallure due to translent Faillue due to laterel ground deforntion spreading Figure 2-2 Various observed modes of pile failure in liquefiable soils (Tokimatsu et al, 1996) Seismic free-field 0 Seismic pile 0 T Fu(I /UOZ i-1 I EP, , Lp ZO n 747 i 1rm ffr U 9e''"' Vertical S-waves Figure 2-3 Typical Winkler spring and dashpot model for pile foundation analysis (from Kavvadas & Gazetas, 1993) 30 3 3.1 Model of Soil-Structure Interaction Overview of Macroelement Formulation of the Macroelement During intense seismic events, Soil Structure Interaction (SSI) has been to proven to be a significant factor of the overall response of pile foundations (Mylonakis and Gazetas, 2000). To this end, numerous numerical analysis of the soil have been developed to account of this interaction, including finite element, finite difference and dynamic beam on nonlinear Winkler foundation (BNWF) methods. The basic assumption of the latter method, more commonly referred to as the p-y approach, is that the soil-structure interaction can be represented by a distribution of rheological elements that act independently (i.e., effectively de-coupling the shear stress transferred between adjacent soil layers). This assumption greatly simplifies the analyses and reduces computational costs compared to finite element methods that will indeed describe the aforementioned problem to the fullest. As described in Chapter 2 the macroelement approach represents a practical compromise between simplified Winkler methods and full 3D finite element analyses for simulating complex SSI for piles. The current analyses use a macroelement developed by Varun (2010) that can capture efficiently the response of piles in cohesionelss soils subjected to cyclic lateral loading and can account for soil liquefaction, Figure 3-1. The soil resistance around the pile circumference is modeled along using a nonlinear Winkler spring(dp,), while a viscous damper (dpd) represents radiation damping that varies with non-linear material behavior. A gap 31 element is included in the formulation to account for the formation of a gap at the interface between the soil and the pile. Varun conducted 3D finite element parametric investigation of a single pile in liquefiable soil to interpret the controlling parameters. The macroelement was originally developed around a series of 3D finite element simulations using the finite element program Dynaflow (Prevost,1995) and incorporate the multi yield plasticity model of Prevost (1985) and later validated with using full-scale, force vibration test data from a blast-induced liquefaction test bed and centrifuge data for earthquake loading of piles with superstructure (Varun, 2010). 3.1.1 Drained Behavior of Macroelement Varun (2010) initially considered the response of a pile subjected to lateral loading in dry / drained soil conditions using a modified Bouc-Wen type hysteresis model (Bouc, 1971; Wen 1976). The equation for the quasi-static case is: P=P,& (3.1) where p is lateral pressure, p, is the ultimate lateral resistance of the pile section and { is a dimensionless quantity that describes the nonlinear, hysteretic, lateral soil reaction. This parameter iscomputed incrementally by the following expression: d = {1- f({)[b+ sign(x)={ +1 g -sign(du -{ A r(3.2) for x >0 32 where du is the incremental relative displacement between the pile and the free-field at the location of macroelement; u, = p,/K , u, is the yield displacement, and K the initial stiffness; b(= 1-g) and g are input parameters controlling the unloading and reloading stiffness. The case b = g = 0.5 corresponds to the case where the unloading stiffness is equal to the initial stiffness. Finally, f f({) = 1 for is a monotonically increasing function of ( such that J = 1. Varun f = 0 for = 0 and (2010) assumes the following analytical expression: (3.3) f ({) = tanh(a()/tanh(a) where a is the backbone curve parameter that is fitted to the results of the numerical (FE) simulations and depends on the relative density of the surrounding sand. Varun (2010) reports values of a= 2.7 for dense sands, a= 2.8 for medium sands and a= 2.9 for loose sands. He also states that the initial stiffness K of the p - y curve, is calibrated to the reference Young's modulus of the soil (E,), (p/p, vs while b is fitted to the backbone quasi-static behavior u/u,) such that b=0.6 (g=1-b=0.4). The ultimate lateral resistance is also calibrated to the Coulomb friction angle of the sand through the passive earth pressure coefficient K, and the pile diameter, using a weighting of empirical experiment proposed by Broms (1964) and Fleming at al. (1992): (3.4) p= (3.25K, + 0.3K) -Bov where K, = tan 2 (45+#/2) The macroelement includes a dashpot "in series" with the spring (Figure 3-1). This describes radiation damping caused by energy dissipation and re-distribution. There has been a broad 33 discussion in the literature on whether the dashpot should be placed either in parallel or in series with the spring element. Wang et al. (1998) reported that, when the dashpot is placed in a parallel arrangement with the spring, unrealistically large forces may occur when highly nonlinear loading occurs and therefore an upper bound should be applied. To overcome those Boulanger et al. (1999) placed the dashpot element in "parallel" with the elastic stiffness spring and in "series" with the plastic slider element, so that the total response never surpasses the ultimate yield strength of the soil. Varun (2010) argues that the physical interpretation of the parallel arrangement leads to an incorrect consequence, namely that the soil resistance at maximum displacement should be almost the same for all loading frequencies in sinusoidal loading, since at maximum displacement the loading rate is zero, and thus the dashpot does not contribute to the overall resistance. In his parametric investigation, Varun (2010) proved that the soil resistance at maximum displacement decreases as the loading frequency increases, which is something that the series model can capture. Hence he used this arrangement in the formulation of the macroelement. The total resistance is calculated in an incremental fashion from the following equation: dp =dp,+'dp= pyd{s +p,d4 (3.5) The quasi-static resistance, dp, calculated from equation (3.2), while the dynamic resistance is given by: dp =c.du (3.6) where c,. is the radiation damping coefficient. 34 The radiation damping coefficient c, is calculated iteratively using an equivalent linear approach by modifying the linear damping coefficient proposed by Makris and Gazetas (1992) as: Cr = c(1-f [(b+g-sign(du-{ (3.7) c = pV-ao 2 QB where p, is the density of soil, V, is the shear wave velocity in soil, Q is a coefficient that depends on the soil Poisson's ratio and the shape of the foundation as derived by Novak et al. (1978), ao is the normalized frequency of loading, ao = cB/V , B is the diameter of the pile. For the case of transient loading co is equated to the dominant frequency of the loading. Badoni and Makris (1995) showed that 3.1.2 Q =3 is appropriate for shallow depths. Undrained Behavior of the Macroelement Varun (2010) observed the formation of a zone around the pile where pore pressures are considerably different from those in the far-field due to local soil structure interaction. This zone extends for about five diameters around the pile, Figure 3-2. The same phenomenon has been reported in the literature, (Gonzalez et al., 2009; Boulanger et al. 1999) and has been attributed to both dilation effects in the soil and suction acting on the tension side of the pile. The above observations are indeed in agreement with the principle that soil-structure interaction alters the stress, pore pressures and displacements in the near field compared to the conditions in the free-field. 35 In order to incorporate the effects of changes in effective stress at the soil-pile interface Varun (2010) introduced a pore pressure generator that modifies the drained response to account for local pore pressure generation and dissipation. The model developed by Varun (2010) extends the "liquefaction front" concept originally proposed by Iai et al. (1992). Figure 3-3 summarizes measurements of excess pore water pressure from elemental undrained torsional shear tests on sands, as a function of the total amount of shear work done per unit volume of soil (W) as reported by Towhata and Ishihara (1985). In undrained hollow cylindrical shear tests the total work is calculated in an incremental fashion from the following equation: dW = a 'de, +2ah'de +rdvh(. dW =dW + dW The above increment, dW, consists of a strain energy contribution due to volumetric deformations, dW and one due to shearing deformation, dW, . However, since Towhata and Ishihara (1985) performed undrained tests, the total volumetric change is zero and thus it does not have any in the total work, dW = 0. Therefore the total work equals the total shear work done, W = W,. The same authors showed that this behavior is independent of the mode of shearing and they stated that the excess pore pressure build up depends exclusively on the current stress state and the accumulated shear work. Towhata and Ishihara (1985) constructed contour lines for the accumulated shear work in the effecticve stress space (Figure 3-4). lai et al. (1992) used the above experimental results to formulate a semi-empirical procedure to describe the effective stress paths, where: 36 = 1 =(1- S) (3.9) lai et al. (1992) explained that as shear work is accumulated during dynamic loading, the envelope of stress points at equal shear work gradually moves from the initial envelope to the failure line (Figure 3-4). This envelope is called "liquefaction front". Figure 3-5 shows the implementation of the liquefaction front concept for the pile-soil macroelement by Varun (2010). The predominant parameter is the liquefaction front parameter(S 0 ), which can be interpreted as a measure of the cyclic mobility and is defined as function of shear work, and can be graphically explained as the intersection of the envelope of stress points at equal shear work with the x-axis in the normalized stress space (Figure 3-5). Varun (2010) performed a numerical parametric investigation which concluded that So is independent of critical state friction angle, friction angle and pile diameter as shown in Figure 3-6 a, b and c. However So does depend on the liquefaction resistance parameter of soil, X, a dimensionless parameter used is the soil model of Prevost as shown in Figure 3-7 (a). The liquefaction resistance parameter X depends on the relative density and sand type (Popescu, 1995). The liquefaction resistance parameter X is derived by liquefaction strength analysis using laboratory testing of soil samples or via correlations with field test data. Popescu (1995) quotes values of the parameter range, X = 0.08 for dense specimens to 0.15 for loose sands. Varun (2010) found that the liquefaction front parameter correlation with the normalized shear work could be further normalized by a scaling parameter w, which related to the liquefaction 37 resistance parameterX and to the elastic stiffness properties of the soil through a derived parameter, 77 (Figure 3-8): R = n(-n) (1 - 2v) 2 n (I _ V)20-n> (3.10) wi (3.11) where v is the Poisson's ratio and n is the power exponent describing stiffness variation with confining pressure 1. Figure 3-6b shows that the liquefaction front parameter, So, is a unique function of the ratio The liquefaction front parameter, So, can then be calculated from the following w/i. equations: So = exp -- dSO = KSO (-1ogS (3.12) 0 )1x (3.13) where ris fitted to the simulated behavior shown in Figure 3-5 andFigure 3-6, resulting in a constant value, K = 1.4 . Finally Varun (2010) proposed that the liquefaction front excess pore pressure can be used to represent relations between the vertical effective paths S = d r = p/BaO : S2 + (r:5 r3) SO S= (So-S 2 )2+[(r -T 3 )/mj2 (r r) i.e. E=E0(a/0 38 /O to the pile resistance, S2 = So (1-m 2 /3m) r3=2Som 2 /3in where mi is the slope of the failure line, m2 is the slope of the phase transformation line and So isthe liquefaction front parameter (Figure 3-5). Figure 3-5 shows clearly how the undrained pile resistance (p) can increase once loading exceeds the phase transformation state. According to the Varun (2010), the slope of the phase transformation line to depends on the critical state friction angle, s: (3.15) m2 = 3.25 tan2 (450 +,/2) Finally the quantity dw is the normalized incremental plastic shear work, and is calculated as the difference between total incremental shear work (dW) and elastic incremental shear work (dW) normalized by the product of ultimate soil resistance and yield displacement. From the results of the parametric investigation, Varun (2010) concluded that the amount of plastic shear work done when the soil dilates, doesn't contribute significantly to the build-up of excess pore pressure, therefore only the plastic shear work done in the contractive zone (below the phase transformation line, r ! r ) is calculated: dw= 1 dW-dW pdu-p K pu, pYu, r <r (3.16) 0r>r 39 3.1.3 Partially Drained Behavior of the Macroelement Local mechanisms of pore pressure generation contribute to the gradient of excess pore pressures around the pile hence, provide conditions for partial drainage within the soil mass. In order to account for this effect, Varun (2010) assumes that there is a linear pressure gradient between the local field and far/free-field regions in the radial direction, and that Darcy's Law controls the seepage within the sand. Varun (2010) assumed 1-D radial flow(v, = v = 0): v,. = k Ah L kU ff-Q k, Us =- o (Sg - S) kL pg L (3.17) where k is the hydraulic conductivity, Ah is the difference in piezometric head, L the characteristic length and S = 0v is the effective stress ratio next to the pile and Sg the effective stress ratio in the far/free field. Assuming the characteristic length is proportional to pile diameterL = aB, the first derivative of the drainage velocity is: d k ,k kL , (1 -S)=- O(Sfdr L dr L dv,d'dd dr =-S)=L , kS , = a2 2 ,0GS) 1 -S) aB (3.18) Using the mass flux equation and assuming that we have flow only in the radial direction and that the both the soil and the water particles are incompressible -- = -v at at V.vr +an = at v + an = ar at : v - aevoi at ar = where n is the porosity of the soil. For the case of a linear, isotropic elastic soil: v Dr 1 (a OAu Ky at a at => 1 ar K at _u 1 ad Ks at 40 (3.19) where K,' is the elastic bulk modulus of the soil skeleton, and there are no changes in the mean total stress. For non-linear sand stiffness the value of K ,varies with the effective dense level: 11f 1 Eo so _ gn 0) 3(1-2v) (3.20) where the exponent, n, is obtained from 1-D or hydrostatic compression tests on sands. v = sO Ks K'K It is then possible to express the radial gradient of specific discharge: _ o> dv, a " dr K at o aS KS at _ 0 ES" (3.21) S at Equating results in equations 3.18 and 3.21, it is then possible to rewrite change in effective stress ratio due to partial drainage: aS Esok s"(s'-s) 2 2 at 3(1-2v)a B s (Sf - S) -S" (3.22) at where $ is the partial drainage parameter and depends on the permeability, the Young modulus of the soil, the diameter of the pile and the Poisson's ratio. Varun (2010) assumes that the rates of change in effective stress are equivalent to changes in the liquefaction front parameter (i.e. dS = dS0 ) (from equation 3.14). He makes this assumption as drainage produces only contraction of the skeleton (which is valid for states below the phase transformation). The explicit formulation of Equ.17 requires very small time increments in order to be numerically stable as Varun (2010) reports: p dSO (i) dt -S(i) (3.23) (i)] Instead one can use the implicit formulation for this part which is more stable numerically: dSo (i) = 18. dt - S(i)" (S' (i) - S(i + 1)) 41 S(i +1)=S(i)+dS0 (i) dSo 19-dt-S" (Sff -S) (3.24) The drainage parameter @was obtained as a function of the hydraulic conductivity, Poisson's ratio and the pile diameter from fitting in the numerical data from the 3D simulations (Varun, 2010): #J=550 2(1+ v') k- (3.25) 3(1- 2v') B Another very important feature that is incorporated in the macroelement behavior is the formation of a gap between the pile-soil interface. During cyclic loading gapping will occur near the ground surface on the tension side of the pile, ( as reported in i.e. the 1989 Loma Prieta Earthquake by Pender and Pranjoto, 1996). To study this behavior, Varun (2010) performed numerical experiments where the pile is loaded monotonically to a maximum displacement and then subjected to cyclic loading of magnitude of portions of the original displacement, a scenario representative of earthquake motion with a strong unidirectional impulse followed by smaller cycles. Figure 3-10 shows the pile resistance curve during the above loading scenario. The initial response shows kinematic hardening behavior and afterwards force relaxation is observed after each cycle of unloading-loading. Varun (2010) states that this kind of kinematic hardening behavior is indicative of a cohesionelss soil where after the gap becomes infilled during subsequent load cycles. To simulate this effect, Varun (2010) implemented a gap element as an additional component of the macroelement. The gap element is simply an envelope function used as a multiplier to scale the total p-y response predicted by the macroelement depending on the current 42 displacement and the maximum previous displacement on each side of the pile. A hyperbolic function is used and the gap multiplier iscalculated as mg = c + where (1- c) umax (3.26) I is the maximum prior displacement on each side, ng is a power coefficient (avalue of 2 is recommended by Varun(2010)), uref = 5u, is a reference displacement value used for scaling and cd =0.1 -0.2 is the ratio of drag resistance from the sides of the piles to total resistance. The overall mechanic components of the macroelement are shown in Figure 3-1. Figure 3-9 shows the sequence of calculations for the macroelement: 1. The drained response ({(i),pd, (i)) is calculated using Eq.(3.5); 2. The shear work done is calculated incrementally using Eq.(3.14)., and is used to calculate the change in liquefaction front parameter So associated with the undrained pore pressure generation is obtained from Eq.(3.11); 3. The value of So is updated to account for partial dissipation of excess pore pressure (Eq. 3.19). 4. Given the current values of (, So and r (=m4S), the current level of average-effective stress ratio (S)is calculated. 5. The total resistance taking into account also the gapping effect is then calculated as p = mgpPS and the next increment to the shear work done is computed. 43 The following example presents the behavior of the macroelement in a simplified loading scenario that allows a better understanding of its mechanical behavior. Figure 3-11 illustrates the displacement-resistance behavior of the macroelement for 6 cycles of sinusoidal displacement controlled loading, using the soil parameters shown in the legend (detailed discussion of the parameters follows in Section 3.2). The vertical axis corresponds to the normalized lateral resistance p/p, while the horizontal to the lateral displacement of the pile. The macroelement has as an input a sinusoidal displacement with amplitude of 0.02 m and frequency of 0.1 Hz (Figure 3-12a). We assume a linear pore pressure build-up (Figure 3-12b) in order to demonstrate the effect of the phenomenon in the degradation of the soil stiffness. 3.2 Calibration of the Varun (2010) Macroelement for Toyoura Sand The macroelement has been calibrated for a specific (but hypothetical) soil profile corresponding to site conditions for a wharf structure (Berth 30-63, Figure 3-13) that is founded on piles driven in an hydraulic fill (modeled as loose Toyoura sand). The site consists of three basic layers whose properties are listed in Table 3-1. The first layer (top 18.3 m of the soil profile) consists of a hydraulically placed loose sand fill that is susceptible to liquefaction, this is underlined by a 2.6 m thick layer of dense sand (Dr=80%) on top of stiff-to-hard clay. The water table level is located 4.6 m beneath the ground surface. Vytiniotis (2011) modeled the hydraulic fill using the properties of Toyoura sand with Dr=40% (e=0.825), a saturated unit weight, psat=1.85Mgr/m 3, and hydraulic conductivity, k=3x10 4 m/s, while the two underlying layers are assumed to have properties of dense Toyoura sand (Dr=80%, e=0.673, psat=2.05Mgr/m 3, and 44 k=3x10 4m/s). The macroelement parameters that are related to the physical properties are the following: i. Initial stiffness of the p-y curve: K =1.25- E,, where E, is the Young's modulus of the soil at small shear strain. The small strain, elastic shear modulus of Toyoura Sand is reported in Dafalias and Manzari (2004) as G= Gopa, , 1+e based on PatI recommendations from Richart et al. (1970), where pat isthe atmospheric pressure, e is the current void ratio, and Go is a material property that defines the maximum shear modulus, Gmax, at P'=Pat. Dafalias and Manzari (2004) report Go=125 and n = 0.5 for Toyoura Sand and Vytiniotis (2011) assumes v=0.33 for free field simulations of Toyoura sand. Figure 3-14 summarizes values of the initial Young's modulus along the depth of pile H. In the finite element analysis the actual input parameter for each macroelement- embedded pile node is K, =1.25xE,, (Table 3-2). ii. The slope of the failure line: =(3.25K, +0.3K ), m.,=- where B, K, = tan 2 (45 + / ) is the coefficient of passive elarth pressure and <b is the peak friction angle of the soil. The peak friction angle for Toyoura sand Dr=40 % and Dr=80 % was estimated using data from Pestana (Pestana et al (2002) having values of 360 and 430. The critical state friction angle of #,, = 31' 0 for Toyoura sand, (and is comparable to soil properties measured for Berth 60-63 Table 3-1). 45 iii. The slope of the phase transformation line: m2 = 3.25 tan2 45+ s ), where #,, is the critical state friction angle of the soil. iv. Backbone curve parameter, a describing non-linear p-y relation in first loading (depends on relative density of the sand, D,). v. Shear work correlation parameter w, depending on the power exponent of the soil (n), the Poisson's ratio (v) and the liquefaction resistance parameter (x) _n ) (1- 2v) 2n ( 1_ 1)2 n> X vi. Partial drainage parameter (0) depending on the Poisson's ratio (v), hydraulic conductivity (k)and pile diameter (B): #= 2(1+ v') k 5503(1-2v') B Following the recommendations of Varun (2010) we assume a= 2.8 for the medium density sand and a=2.7 for the denser sand. Figure 3-15 compares the macroelement (p-y curve) response for these loose and dense sand cases. For the sake of simplicity of the figure, we are using the same initial stiffness for both of cases, which indeed is not true but our intention is to demonstrate the effect of the others -not so trivial- parameters. As far as it concerns the shear work correlation parameter w,, the only unknown parameter is the dilation parameter X. In principle this parameter can be obtained by matching cyclic laboratory tests with numerical analyses using the Prevost (1985) soil model. The typical range 46 of X = 0.08 to 0.15, with lower values corresponding to denser specimens. In the absence of specific calibration data from Toyoura sand we assume the same values reported by Popescu et al. (1995) (based on calibration for fine Nevada Sand). Figure 3-16 shows the influence of the parameter in the elemental behavior of the macroelement. Finally, we calculated the radiation damping coefficient (c) . At each macroelement depth, we calculate the shear modulus (G) and from that the shear wave velocityV, = - . Since we will be imposing earthquake loading we used an average predominant frequency for the suite of earthquake loading of about 2 Hz. The macroelement input parameters for the pile No 3 (Figure 3-13) are shown in Table 3-2. 47 Table 3-1: Measure in situ properties of Berth 60-63, data from Vytiniotis (PhD 2011) Table 3-2: Macroelement properties for pile 3 Dr=40% Dr=80% No. Depth of Influence K (kPa) (kN m) a mi m2 w1 1 2 0.5 1 10181 21063 32 136 2.8 2.8 16.97 16.97 10.15 10.15 1.074 1.074 0.43 0.43 4.40E+05 6.33E+05 3 1 27994 240 2.8 16.97 10.15 1.074 0.43 7.30E+05 4 5 6 1 1 1 34617 40161 45029 366 493 620 2.8 2.8 2.8 16.97 16.97 16.97 10.15 10.15 10.15 1.074 1.074 1.074 0.43 0.43 0.43 8.12E+05 8.74E+05 9.26E+05 7 8 1 1 49419 53450 747 874 2.8 2.8 16.97 16.97 10.15 1.074 9.70E+05 10.15 1.074 0.43 0.43 9 10 11 12 13 1 1 1 1 1 1000 1127 1254 1381 1507 10.15 10.15 10.15 10.15 1.04E+06 1.08E+06 1.10E+06 1.13E+06 1634 16.97 10.15 10.15 1.074 1.074 1.074 1.074 1.074 1.074 0.43 0.43 0.43 0.43 1 2.8 2.8 2.8 2.8 2.8 2.8 16.97 16.97 16.97 16.97 16.97 14 57197 60714 64037 67197 70214 73107 0.43 0.43 1.16E+06 1.18E+06 15 16 1 1 75890 78575 17 1 81170 1761 1888 2015 2.8 2.8 2.8 16.97 16.97 16.97 10.15 10.15 10.15 1.074 1.074 1.074 0.43 0.43 0.43 1.20E+06 1.22E+06 1.24E+06 18 1 83685 2141 2.8 16.97 10.15 1.074 0.43 1.26E+06 19 1 107739 2268 2.7 25.58 10.15 2.013 0.43 1.43E+06 20 1 110210 2373 2.7 25.58 10.15 2.013 0.43 1.45E+06 21 1 113602 2522 2.7 25.58 10.15 2.013 0.43 1.47E+06 22 1 116423 2648 2.7 25.58 10.15 2.013 0.43 1.49E+06 23 0.5 119177 2775 2.7 25.58 10.15 2.013 0.43 1.51E+06 48 (s) C (Pa's) 1.01E+06 PORE PRESSURE MACROELEMENT ELEMENT ----------------------------------dp, dw Pile response OUTPUT GAP ELEMENT INPUT Free-Field dUff F Uf Displacement of the free field Sf Effective Stress Ratio at the free filed dp, incrementalforceper unit length exerted by the spring dpd incrementalforce per unit length exerted by the dashpot p,,y Force per unit length for drained case dw Incremental shear work in macroelement S(w) Average effective stress ratio in near-field P Force per unit length with gap correction F Force imposed at pile node in FE analysis Figure 3-1 Schematic Formulation of the Macroelement (modified from Varun, 2010). 49 Syy-STRESS kPa] I I 1.13 4.13 -9.38 14.62 19.87 . 25.12 30.36 35.61 * 40.86 l H1 OB 46.10 Figure 3-2 Effective vertical stress o, [kPa] plot after seven loading cycles showing the formation of local liquefaction zone - approximately five diameters around the pile (from Varun, 2010). 3001 w-u relationship u UT = liquefaction . 0 kN/ m 0 WN/m" 2 way cyclic torsion shear test C , x- C , r = 2 at E C', = 294 kNm' a. 200 K0=1, Toyoura sand Sym . Exp. No Tmax kN/m 2 e o 37 77.7 0.813 o 45 60.6 0.809 A 47 64.9 0.818 <1 60 70.6 0.811 V 69 55.4 0.784 c> 77 70.6 0.812 5 10 J .0 V 100 *)0 (A 0 . Average of loading schemes a,b and e L.. h .02 0.05 0.1 0.2 1 0.5 2 Shear work per unit volume w (kJ/m 20 3) Figure 3-3 Relationship between excess pore pressure and she ar work for undrained cyclic shear tests on sand (after Towhata and Ishihara, 1985). 50 Toyouro Sand 1 0L %. 0 V) 1-50 4> -1 300 Effective Confining Siress -0~' (kPo) Figure 3-4 Envelope of stress points at equal shear work (Towhata and Ishihara, 1985) 51 0 Cm1 Failure Line 0 M Liquefactio 4-J ront M2 Phase Transformation Line -- r2 2/3M2 r3 S2 z _ F SO Vertical Effective Stress Ratio, S =Ta /so Figure 3-5 The liquefaction front concept as extent by Varun for pile. The normalized vertical effective stress S=O- 0 oOcan be calculated as a function of the normalized soil resistance r=/)Bo - 52 1 1I 0.8 0.8 -B=0.5 0-6 . B=0.75 -- 0-6 CS 0.4 0.4 0.2 0.2 0 0 0 0.5 1 1.5 2 B=125 A A --B=1.5 (b) 0 2.5 0.5 1 1.5 2 2.5 w w 1 0.8 0 0.6 Co 0.4 0.2 0 0.5 1 1.5 2 2.5 w Figure 3-6 Liquefaction front parameter (SO) correlation with normalized shear work w for different values of (a) critical state friction angle (b) pile diameter (c) maximum friction angle (figure from Varun (2010)). 1 *X=0.15 0.8 X=0.15 0.8 SX=0. I X=0.05 -- ------ 0.6 A X=0.10 0 .6 U) - - . X=0. 0 5 - - - -0.4 0.4 - 0.2 (a) 0 0 __ 1 _ _0__ 2 _ - 0.2 _(b) 3 4 w 0 0.5 1 1.5 w/w 1 Figure 3-7 (a) Liquefaction front parameter (S0 ) correlation with normalized shear work as a function of liquefaction resistance parameter X (b) Shear work correlation curves after normalization with parameter w1 (figure from Varun (2010)). 53 2 1 A n=0.6 0.8 0 n=0.7 0.4 0.2 0 0 0.1 0.3 0.2 0.4 0.5 V Figure 3-8 Parameter r as a function of Poisson's ratio and power exponent n (from Varun (2010)) 54 Total Macroscopic Pile Response IptParametersK, py, 0, C, M2, W1,y Z !relative dslcmnu()n~ Compute dramed response (i) and pai) Compute Plastic shear work (w) and Liquefaction front, So(i) I Input far field excess pwp, 1-Sa(i) S(i)=S.,(I)I Compute eff. stress ratio, S(i) s|S(i)-S (i)|< to No Yes / Output soil resistance, p(i) / Figure 3-9 Flowchart for calculation of pile response using the macroelement (Varun, 2010). 55 1 0.5 IL: 0 -0.5 -1 0 0.025 0.05 0.075 0.1 u (m) Figure 3-10 Pile response to small loops of cyclic unloading-reloading after monotonic loading (Varun, 2010). 0.6 0.4 E 0.2 0 0 Lb 0 -0.2 0 NI F -0.4 z -0.6 -0.8I -0.02 -0.015 -0.01 1 I -0.005 6 0.005 Horizontal Displacement u (m) Figure 3-11 p-y curve for the macroelement. The degradation of stiffness and lateral resistance of the macroelement is a result of the cyclic loading and the accumulation of excess pore pressure. 56 0.02 E (a) 0.01 o E - o 0 C 0 0 -0.01 -0.02 T 0.8 1> .(b) Lfn < 0.60.4- a~>- CU 0.2- v Lu. 0 > 0 0 10 20 40 30 50 60 70 Time, t (sec) Figure 3-12 Sinusoidal Input Displacement motion and b.Assumed pore pressure build up 57 a. Berth Section 4.m - ASea 0DOm- / Hydraulic Fill -13.7m -6.3m - Water Level \ 5 6Facing Stone Sand FilI Old Bay Mud (Stiff Clay) Pile Locations .33,7m - b. FE Model Properties I.TI 4,6m 0,0M 'El -13.7m.-16.3m - LI .33,7m- LI Figure 3-13 Physical model of Berth 60-63 and FE model properties as used by Vytiniotis (PhD 2011). (Figure from Vytiniotis (PhD (2011)) 58 Young's Modulus distribution along the Macroelement nodes in pile No 3 E(MPa) 100 T---- 80 60 40 20 0 0 120 25 0- Toyoura Sand Dr=40% pTovaura Sand Dr-80% E -5 *---*-*---* 4- -- --- 20 CL 0. o 5 -10 - 00 -25 --- - - ----- - -- -- ------ 0 Figure 3-14 Young's Modulus distribution along the depth of the pile. The darker grey line corresponds to the upper hydraulic fill, which is modeled with Toyoura Sand of Dr=40 %. The lighter grey line plots the corresponding distribution along depth of Young's modulus fo Toyoura Sand of Dr=80%. After the depth of 18.4 m below the surface, for the calibration of the macroelement we are using the properties of Dr=80 %Toyoura Sand. 59 15 a. 10 CL E 0 5 * 0 0 -1 0 N -10 -.02 -0.015 -0.01 -0.005 0 0.005 0.01 0.015 Horizontal Displacement u (m) Figure 3-15 Effect of parameters a and w on the response of the macroelement. 60 0.02 0.6 2 04- x=0.15 - x=0.08 -0.015 -0.01 E $ 0.2- 'U 00 C 15 -0.2 -0.4 E 0 -0.6 -0.02 -0.005 t 0.005 0.01 Horizontal Displacement, u (m) 0.015 Figure 3-16 Effect of parameters X on the response of the macroelement. 61 0.02 4 4.1 Seismic Response of Pole-supported Wharf Structures Problem Description In this chapter we examine the 2D dynamic response of Berth 60-63 (port of Oakland), which represents a typical pile-supported wharf structure that was constructed in the 1960s and hence, reflects now outdated seismic design criteria for loose, liquefiable granular fill. The soil profile comprises three basic layers whose properties are listed in Table 3-1: a 18.3 m thick (hydraulically-placed) loose sand fill (average Dr= 40%), that is susceptible to liquefaction, overlying a 2.6 m base of dense sand (average Dr= 80%) and an underlying stiff-to-hard clay. The berth structure consists of 7 rows of vertical and 2 rows of batter pre-stressed reinforced concrete piles supporting a 30 m long, 0.46 m thick reinforced concrete deck (Figure 4-1). A 2-D finite element (FE) model of this wharf has been developed in OpenSees, an object-oriented FE analysis framework (McKenna et al. 2010) as shown in Figure 5-2. The pile-soil model comprises three components at: 1) 2-D beam-column elements representing the behavior of precast and prestressed piles, 2) vertical spring elements describing the vertical pile-soil interaction, and 3) macroelements describing the horizontal response of the pile-soil interaction. Each macroelement has a free field boundary where the corresponding time histories of free field displacements and excess pore pressure ratio are imposed as input motions for the dynamic analyses. 62 Vertical loads from the superstructure are transferred to the underlying soil through the shear resistance along the shafts of the piles and the tip end bearing resistance, while lateral loads are resisted by the bending action in the vertical piles (the batter piles provide both bending and axial resistance). The pile nodes have three degrees-of-freedom (2 translational, 1 rotational), while the non-linear beam-column elements (each with 5 integration points) are capable of simulating the formation of plastic deformations within the element. The properties of the reinforced piles are represented with an aggregated section in which the momentcurvature response is described by an elastic, perfectly plastic constitutive law with elastic axial stiffness, EVA, bending stiffness, EI,, and maximum section yielding moment, M, (Figure 4-3). The piles for Berth 60-63 have width 45.7 cm with a ring of 16 #32 mild steel vertical bars and 10 #13 pre-stressed tendons, covered by 7.5 cm of unconfined concrete. The calculation of the moment capacity of the section was done using an online available software (KSURC; Esmaeily, 2011), assuming that pre-stressed tendons provided a fixed axial load across the section (i.e. increasing the moment capacity). We assume a concrete compressive strength, f 6.0 ksi = 41.4 MPa, elastic modulus: E, [ksi] = 57 -f 4 [psi], and steel with yield strength =f60 ksi =414 MPa and modulus, E, = 29,000 ksi = 200 GPa. 63 The deck structure is responsible for transferring the dead and live loads of the wharf structure to the underlying foundation. The current case considers a cast-in-place concrete deck slab with thickness, 46 cm, constructed of reinforced concrete (with concrete strength fc=41.4 MPa). Due to the large thickness and high rigidity, the deck acts as a diaphragm wall in the horizontal plane. The deck is modeled with linear elastic, beam-column elements, and additional constraints are imposed (by tying the deformations of all the deck nodes to each other) to prevent flexural deformation of the deck. The pile-deck connections are modeled in a simplified way by imposing the same degrees of freedom at the pile head and the connecting pile deck nodes. The soil-pile interaction in the vertical direction is represented by elastic springs that simulate the friction and tip resistance of the soil using zero length elements, which are defined by two nodes at the same location (the pile node and the corresponding "spring" node). The forcedeformation relationship for the element is prescribed in a uniaxial material. The springs were assigned stiffness properties following the change of the soil Young's modulus with depth (Figure 3-14). This simplified modeling of the vertical soil-pile interactions was judged appropriate as the loading was predominantly horizontal and the vertical displacements are very small. In the lateral direction, the soil-pile interaction is modeled using the macroelement developed by Varun (2010), with input parameters calibrated in Chapter 3. The nodes for the macroelements are offset at a very small distance from the corresponding pile nodes (0.1 m), such that the orientation of the macroelement is perpendicular to the axis of the pile. Loading is 64 simulated by imposing time histories of free field displacements and excess pore pressure ratios 2 at the macroelements. 4.2 Ground Motions and Free Field Analyses The NEES-GC team (Shafieezadeh, 2011) selected a suite of 56 ground motions that are typical of firm-site conditions in coastal California using records from the Next-Generation Attenuation of Ground Motions (NGA) project (Chiou et al., 2008), with moment magnitude M=5.5-8.0, located 15-60 km from rupture zone. A subset of 15 ground motions were randomly selected from the NGA database at rupture distances less than 15 km. Vytiniotis (2011) used the selected suite of earthquake ground motions to predict the free field response at the Berth 60-63 site. The mechanical responses of loose sand fills were computed using a coupled deformation-flow finite element analyses with appropriate free field boundary conditions. The DM 2004 (Dafalias & Manzari, 2004) effective stress soil model was used to simulate the mechanical response of sand in cyclic shearing. Vytiniotis (2011) also investigated the effectiveness of installing an array of PV-drains as a method of mitigating seismic risk for Berth 60-63 by using the same suite of free field ground motions. The performance of this system was evaluated through comparisons with a series of analyses where the whole sand fill was densified though appropriate compaction methods. The PV-drains system offers a less , where U0 is the insitu (hydrostatic) pore pressure 2 07o 65 intrusive solution for retrofitting the berth, than conventional compaction methods that would require a complete reconstruction of the piled wharf structure. Figure 4-5 shows the locations of the PV drain system analyzed by Vytiniotis (2011). The current research uses free field analyses to compare the response of: 1) the current unimproved piled wharf structure; 2) the wharf retrofitted with a PV drain system (Figure 4-5); and 3) the idealized retrofit with complete densification of the fill. 4.3 Results of Untreated Scenario We will examine first the response of the structure for the base case analysis of nga0779motion (Loma Prieta Earthquake, 1989). Figure 4-6a shows the horizontal ground acceleration record of the event as used by Vytiniotis (2011) as an excitation for the free field analysis of the site of Berth 60-63. The results of the above analysis are plotted in contours of acceleration and displacement at 4 snapshots (t = 5, 10, 15, 20 secs) Figures 4-6b and 4-6c. The maximum free field displacement that occurs in the site is greater than 0.9 m as shown in Figure 4-6c. Figure 46d shows the time history of the horizontal displacement of the deck with a net permanent displacement Sh = 55 cm. Figures 4-7 to 4-10 show the evolution of the failure mechanisms within the wharf structure subjected to the base case ground motion, nga0779. The figures represent snapshots of the wharf structure and underlying free field fill response (contours of displacement and excess pore pressure ratio) at 5 sec intervals. At t=5secs (Figure 4-7) no significant displacements or 66 pore pressures have developed in the fill and the deck displacement is less than 1 cm. At t=10 sec, there are significant slope displacements (Figure 4-8b) and pore pressures have already developed with the outmost regions of the slope reaching excess pore pressure ratio approachingAu/O = 1.0. Hinges form at the red (landward) pile-deck connections, and there is a net deck displacement 6 h= 22 cm. The formation of hinges at the most landward pile-deck connections is indeed the most common type of failure observed as later revealed from our analyses. At t=15secs, the wharf has failed with a 3-hinge collapse of five rows in piles and large rotations of the seaward batter piles (Figure 4-9a) and the deck displacement 6 h= 20 cm. However, it should be noted that the deck displacement has previously reached a local maximum of 40 cm at t=11 secs (Figure 4-6a) and then decreased to smaller value following the imposed pulses of the motion. The creation of the 3-hinge collapse mechanism occurs at t = 11 secs. At t = 20 secs; the deck displacement reaches 54 m, while there is a significant spread of the region with plastic hinges (Figure 4-10a). At the end of the analysis (t = 25 sec) the residual horizontal displacement is 6 h = 0.55 m. The moment envelope along with the final deformed shape of the wharf for nga0779 is shown in Figure 4-11. Figure 4-12 shows the time histories of horizontal displacement of the deck for the suite of 56 motions, categorized in 8 (accordingly to the NGA record locator number). It is worth noticing that the two motions that give the largest deck displacements are nga0753 and nga0779 (Figure 4-12a) both records from Loma Prieta Earthquake (1989). Figure 4-13 summarizes the correlations between the computed maximum deck displacements and four measures of 67 earthquake intensity (IM) using power law functions, as proposed by Cornell et al. (2002): PGA, PGV , PGD and Arias Intensity Udeck (4.1) =a-IMb where PGA (peak ground acceleration) is the maximum acceleration of the ground motion recorded to the surface; PGV (peak ground velocity) refers to the maximum velocity of the ground motion recorded at the surface; PGD (peak ground displacement) is the maximum displacement of the ground motion recorded at the surface; and finally, the Arias intensity (Ia) is the measure of total energy content of a seismic excitation defined as: Ia = 2g [a(t)]2 dt (4.2) Table 4-2 summarizes the values of a, b and the regression efficient, r2 , achieved in these cases. While the correlations show that deck displacements increases with each of the intensity measures, there is significant scatter in the results precluding simple predictions. The peak ground acceleration correlation involves a lot of scatter, but has a having a clear trend (Figure 4-13a). An inherit limitation in the use of PGA is the fact that it does not reflect the frequency content of the ground motion. PGV represents a better correlation measure, as it incorporates partially the aforementioned effect but captures only some effect of the sequence of strong pulses. The correlation of damage measure with PGV is indeed higher than with PGA (Figure 4-13b). The correlation with PGD has again a clear trend but high scatter (Figure 4-13c). As discussed by Vytiniotis (2011), Arias Intensity provides, perhaps, the most reliable measure of 68 earthquake intensity currently as it includes both the effects of the acceleration, the frequency content of the record and the sequence of the pulses. Figure 4-13d shows the correlation of deck displacement with Arias Intensity. This presents a more reliable correlation than PGA, with two notable outliers nga0753 and nga0739. It is worth commenting that indeed nga0753 is an outlier for all the above correlation; the reason for this is that the large displacements associated with this ground motion are attributed to the large lateral spreading of the slope, a phenomenon than none of the above measures can capture. However, nga0739 is only an outlier for the correlation with Arias Intensity and indeed follows the trend in the other 3 correlations. This can be attributed to the fact that the nga0739 has a high Arias Intensity value but consists of a large number of small acceleration pulses that produce little lateral spreading. Figure 4-14 shows that there is also a very good linear correlation between the maximum deck displacement and the corresponding maximum free field movement (referred to node 49, Figure 4-2), as reported by Vytiniotis (2011) to quantify free field slope damage. Based on an analysis of the wharf structure the seismic performance can be grouped into 4 classes of response: 1. No damage, structural response is elastic; 2. Light damage, where failure occurs only at the pile-deck connections; 3. Moderate damage, where some plastic hinges develop within the piles; and 4. Heavy damage associated with the creation of a plastic collapse mechanism of the supported wharf. 69 Figure 4-15a illustrates the most commonly observed case of light damage where plastic hinges develop at the pile-deck connection. Moderate damage, Figure 4-15b, is associated with the creation of plastic hinges at both the pile-deck connections and at other locations along the embedded piles. The example shows hinges in vertical piles close to the surface of the slope and at the interface between layers of strong contrasting stiffness (Old Bay Mud-dense sand interface, Figure 4-1Figure 4-5) stiffness for the batter piles. Heavy damage occurs with the creation of a failure mechanism involving plastic hinges at three distinct locations along the vertical piles at the pile-deck connection, on the upper part of the loose fill and at the base of the fill, Figure 4-15c. Of the 56 ground motion records, only two caused extensive damage to the structure, 6 produced moderate damage and 6 light damage. The structure remained elastic for the other 42 ground motions with small permanent displacements, as shown in Table 4-1. The current analysis for nga0753 can be compared with the results reported by Shafieezadeh (2011). The main differences in the two analyses is that Shafieezadeh (2011) uses a more advanced modeling of the reinforced section that can capture the coupled flexural and axial response of the pile and models the deteriorating behavior of pile-deck connections as studied by Lehman et al. 2009. In contrast, the current calculations assume an elastic, perfectly-plastic 70 pile response). The results of our analysis show a remarkable agreement in the location of the plastic hinges and also the time history of the deck displacement (Figure 4-16). 4.4 Effectiveness of Ground Improvement Systems This section considers the effectiveness of soil improvement methods by comparing the response of the piled wharf structure under the same suite of ground motions and two ground improvement systems: 1) a system of PV-drains installed behind the crest of the fill slope (Figure 4-5) and 2) complete densification of the fill (Dr= 80%; e=0.673). The first case assumes the same input parameters for the soil and macroelements as the untreated case while the latter uses a revised set of macroelements parameters corresponding to Toyoura sand with Dr= 80% (Section 3.2). Figure 4-17 summarizes the computed deck displacements for the piled-wharf with three ground conditions as functions of the four reference earthquake intensity measures (PGA, PGV, PGD, and IA). The results show that the predicted (i.e. best fit correlations) deck displacements for the PV-drain mitigation system lie in the range predicted for unimproved and compacted fill conditions. Table 4-2 summarizes the power law relations (Equ. 4.1) parameters and regression coefficients for the four intensity measures. The results show improved regression properties for the simulations with improved ground conditions. Table 4-1 compares the overall classification of structural damage for the wharf with the three different ground conditions (untreated fill, PV-drain system, fully densified fill) for the full suite of reference ground 71 motions. In general, soil improvement methods reduce the number of events where permanent structural damage occurs. For the PV-drain system, there is only heavy damage for one case (nga0779), where there is large lateral spreading in the underlying slope; one case of moderate damage and eleven cases of light damage, (i.e. the installation of the PV-drain system is efficient in protecting the wharf structure from the plastic hinges formation in the lower section of the piles and hence, structural retrofit of the wharf can focus on the pile-deck connections). For the fully densified condition, there are no cases of heavy damage, one case of moderate damage and eleven cases of light damage. Although the wharf deck undergoes larger displacement with the PV-drain system than with the fully densified fill (Figure 4-17) the level of structural damage in both scenarios is quite similar. In this respect, the PV-drain system represents an efficient method for mitigating the seismic risk for the wharf structure. However, it is worth noting that the densification method produces less structural damage for the case of nga0779 where PV drains are ineffective against lateral spreading of the slope. Figure 4-18 and Figure 4-20 compare the wharf responses and underlying free field results for the three ground conditions considered for nga0779. The PV-drain system has little effect on the structural damage of the wharf and displacement for this case. This result occurs as most of underlying ground conditions are used by lateral spreading in the fill slope. The complete densified slope undergoes significantly smaller horizontal displacements (Vytiniotis, 2011) and as a result the structural damage that occurs is moderate (Figure 4-18). 72 Figure 4-19 and Figure 4-21 present a similar survey for nga0753. In this case, damage reduces significantly with the use of PV-drain system, and deck displacements are comparable to the full densified fill. In this case, the contour of displacements in the free field analyses show limited lateral spreading of the loose fill. The above analyses are indicative that the PV-drain system is effective in protecting the wharf structure for cases where pore pressures affect response but does offers only marginal improvement for the case of large induced displacements due to inertial forces. It should be also stressed out than neither the "ideal" densification of the slope nor the PV-drains improvement method are proven extremely capable of limiting the failure of the pile-deck connections (from 6 cases of light damage in the case of untreated, 3 did not sustain any damage in the ideally densified case and only 1 did not sustain any damage in the case of PVdrains improvement). 4.5 Modeling of Pile-Deck Connection The prior analyses have used a simplified elastic perfectly-plastic model for the piles and piledeck connections. However, Lehman et al. (2009) have demonstrated in a full-scale test that pile-deck connections degrade under cyclic loading. The piles and the deck in the wharf for this study are typically connected by T-headed dowel bars and the nonlinear behavior of this connection has been calibrated against a full-scale test conducted by Lehman et al. (2009), Figure 4-22. The pile-deck connection was subjected to lateral cyclic loading at a distance of 73 2.54 m from the interface of the reinforced concrete base while maintaining a constant axial load on the pile simulating the gravity load. This section uses a complete representation of the pile-deck connection using a nonlinear connection element with aggregated section properties that represents a 3.8 cm embedment of the pile into the wharf deck (Figure 4-22). The axial response of the connection section is the same as for the piles. However the moment response is modeled using the Modified IbarraKrawinkler (MIK) deterioration model with bilinear hysteretic response (Ibarra et al., 2005). The MIK model can capture most of the sources of deterioration (i.e. cyclic deterioration of several modes and softening of the post-yielding stiffness) and accounts for residual strength after deterioration. The MIK model is based on standard bilinear hysteretic rules with kinematic strain hardening but it is modified in a way that during cyclic loading the maximum strength that can be achieved during unloading is limited by the smallest strength that the has been previously reached. Had this condition not been established, the strength in the loading path could increase in later stages of deterioration (Ibarra et al., 2005), Figure 4-23. The model incorporates an energy-based deterioration parameter that controls four cyclic deterioration modes once the yield point is surpassed in at least one direction: basic strength, post-capping strength, unloading stiffness, and accelerated reloading stiffness deterioration as 74 seen in Figure 4-24. However accelerated reloading stiffness is not active in the bilinear version of MIK. The rates of cyclic deterioration are controlled by a rule developed by Rahnama and Krawinkler (1993) assuming that every deterioration mode has a reference hysteretic energy dissipation capacity Et, which is an inherent property of the deterioration mode and it is independent of the loading history (Ibarra et al. 2005). The reference hysteretic energy dissipation capacity Et depends on Op, the precapping plastic rotation and My, the effective yield strength of the component: E, = A -O, -M, or (4.3) E,, = A, -M, where A, =1, -d, is the reference cumulative rotation capacity for each mode of deterioration. The basic and postcapping strength deterioration) is modeled by translating the two strength bounds (the lines intersecting at the capping point) after every loading cycle in which energy is dissipated toward the origin at the rate, Figure 4-24: M, = (1- #l) -M, (4.4) where the moment Mi is any reference strength value on each strength bound line. According to Lignos and Krawinkler (2011), Mi can be defined as the intersection of the strength bound line with the y-axis and Pi isthe energy based deterioration parameter: $, =(4.5) E,-XE, 75 where Ei is hysteretic energy dissipated in the loading cycle i; IEj is the total energy dissipated in past cycles ;and c is an empirical parameter that controls the rate of deterioration, usually taken as 1.0, a value that we also used in our calibration. The model can also account for different rates of deterioration in the positive and negative direction, controlled by parameters dnp which however, the current simulation does not consider this feature and uses the default dnp= 1.0. The same concepts apply to modeling of unloading stiffness deterioration, i.e., the deteriorated stiffness after cycle i isgiven by: K, =(1-,)-K, (4.6) Tables 4-3 and 4-4 list the parameters used for the backbone response and deterioration of the pile-deck connection, respectively. The parameters modeling the initial backbone response of the pile-deck connection are listed in Table 4-3, where the initial stiffness and the yield capacity are taken the same as the elastic- perfectly plastic case (section 4-1). The MIK bilinear model is used to simulate the deteriorating behavior of our connection and the resulting moment versus drift ratio (lateral displacement over the height of the column) from the experiment by Lehman et al. (2009), which was numerically simulated. Figure 4-25 shows that the MIK model captures reasonably well the measured nonlinear moment-drift response of physical the pile-deck connection test (note the moment is normalized by the yielding moment of the connection). 76 Figure 4-26a to d summarize the comparison between the elastic perfectly plastic (EPP) piledeck connection and the Modified Ibarra Krawinkler (MIK) pile-deck connection for the base case ground motion nga0779. The results show that the maximum deck displacement for MIK is significantly less than for the EPP model, Figure 4-25a. As the pile-deck connection deteriorates, less and less deformation induced by the lateral spreading on the piles can be transmitted to the deck. So even though the embedded part of the piles undergoes the same amount of deformation, the deck exhibits smaller deformations. Figure 4-26b and Figure 4-26c compare the time histories of the rotation of the connection and of the moment in the pile-deck connection for pile 3 (Figure 4-1). For the case of MIK, the connection undergoes much larger rotations than for EPP, as expected from the deteriorating behavior, where the connection loses its capacity after exceeding the ultimate rotation of capacity. It is worth noting that damage spreads along the length of the embedded piles for simulations using MIK model as seen in Figure 4-27. In conclusion, we see that when the pile deck connection modeling captures the deteriorating behavior, previously established in model tests, smaller deck displacements occur, while there is more pronounced structural damage in the piles. It should be pointed out that the location of the damage is indeed in the same locations are previously identified with the EPP analysis. 77 4.6 Conclusions The primary conclusions of our analyses are that: * The deformations of the piled wharf structure are primarily governed by the lateral spreading of the soil. Deck displacements are better correlated to Arias Intensity and PGV than to PGA. " Three levels of structural damage can be identified for the reference suite of the 56 ground motions. For the non-improved fill only two records cause extensive damage to the structure, 6 cause moderate damage and 6 cause light damage. * The proposed PV-drain system is effective in mitigating seismic damage for most of the ground motions considered. However, it is ineffective when there is a large lateral spreading in the loose fill slope. The comparison with the idealized case of fully densified fill provides a convenient meter for assessing the improvement achieved with the PV-drain system. " Analyses using a more advanced modeling of the pile-deck connection, that captures the deteriorating behavior under cyclic loading, generate smaller deck-displacements but more extensive structural damage to the embedded piles. 78 Table 4-1: Overall classification of structural damage for the wharf with the three different ground conditions. Level of Damage None Light Moderate Heavy Unimproved Loose Fill 42 6 6 2 Compacted Fill 45 10 1 0 PV-Drains System 43 11 1 1 Table 4-2: Values of a and b and the regression coefficient for four earthquake intensity measures. PGA PGV PGD Arnas Intensity Treatment NO PV-DRAINS Density 40% 40% a 0.43300 0.36770 b 1.023 1.137 r2 0.6607 0.7273 IDEAL DENSIFICATION NO PV-DRAINS IDEAL DENSIFICATION NO PV-DRAINS IDEAL DENSIFICATION NO PV-DRAINS 80% 40% 40% 80% 40% 40% 80% 40% 40% 0.28020 0.00689 0.03500 0.00564 0.03132 0.01970 0.02232 0.13710 0.10090 0.986 0.907 0.997 0.851 0.732 0.819 0.704 0.641 0.709 0.7591 0.7505 0.8367 0.8328 0.7381 0.859 0.8311 0.8115 0.8793 IDEAL DENSIFICATION 80% 0.09350 0.599 0.8894 Table 4-3: Modified Ibarra Krawinkler model properties for initial backbone 900 Effective yield strength My (kNm) 0.0082 Effective yield rotation ey (rad) Effective stiffness Ke (kNm 2 ) 110,000 901 Capping strength for monotonic loading Me (kNm) Capping rotation for monotonic loading Oc (rad) 0.0083 Pre-capping rotation capacity for monotonic loading E,(rad) 0.0001 Post-capping rotation capacity Oc (rad) Residual Strength ratio K 0.012 0.0 Ultimate rotation capacity 8u (rad) 0.015 79 Table 4-4 Deterioration Parameters for Modified Ibarra Krawinkler As basic strength deterioration k unloading stiffness deterioration 0.65 Aa accelerated reloading stiffness deterioration 1.0 Ad 1.0 dp post-capping strength deterioration exponent for deteriorations (c = 1.0 for no deterioration) rate of cyclic deterioration for positive loading dn rate of cyclic deterioration for negative loading 1.0 cs,k,a,d 1.0 Node C3 Nle C2 Node Ci Figure 4-1 Configuration of the Berth 60-63 in Oakland (Shafieezadeh, PhD 2011) 80 1.0 1.0 * pile-soil node $ pile nodes (free space) * peck nodes free -field end macroelement nodes Figure 4-2 OpenSees model of piled wharf structure. 81 Moment' Load My EcA curvature displacement Figure 4-3 Constitutive law used to model (a)the moment resistance of the pile section and (b) the axial resistance. EI= 110,000 kN i 2 1000 M'=-900 kNm 900.~ 700 z 600 500 40- 0 30200100 0 0.00 0.01 0.02 0.03 0.04 0.05 0.06 007 Curvature (1/m) Figure 4-4 Moment curvature curve of the reinforced concrete section of the pile as calculated by online available software KSU_RC ( Esmaeily, 2011). 82 Drain Locations Model Berth 4.6m -F Fl F O.Om -13.7m - -16.3m - r F, F. p+ 43.7m V V Figure 4-5 Geometry of PV-Drains array for mitigating seismic risk for Berth 60-63 (Vytiniotis, 2011) 83 0 0.8 E 0.5- 0.6 0.4- S0.4 N 0 0 0.2 10 -0.2 -0.4 -0.1 -0.6 -: -0.2- CL S-.8 0 -1 10 15 20 time (s) +4.6m - +4.6m - ±O.Om tO.Om Ss 5 -16.3m 00 25 time (sec) - (b) -16.3m (c) los 15s Accelerations (m/s0) S1. Is 20s Displacements (m) 0A6 O' 20S 0- Figure 4-6 Response of the unimproved piled wharf to base case ground motion ,nga0779; (a) time history of acceleration of rock base excitation (b) acceleration contours of free field response (from Vytiniotis, 2011) (c) displacement contours of free field response (from Vytiniotis, 2011) (d) horizontal displacement of the deck. . .. .... ......... t=5 s 0 ~-w I - mu mimi U- mi Sea Water -5 Hydraulic Fill -10 - / -15- -20 Sand Fill Old Bay Mud (a) -25 -10 -5 0 5 10 15 20 25 30 35 Displacements (m) 10.9 Excess Pore pressure ratio 0.8 0.7 10.5 03 0.2 0.43.0 0.1 10.9 0.7 00.6 0.5 0.3 0.1 F-1 (b) (c) Figure 4-7 Response of the unimproved piled wharf for ground motion nga0779 at t= 5s (a) Structural displacements (50:1) (b) Contours of free-field displacements (c) Contours of freefield excess pore pressure ratio. 85 t=10 s 0 hI31rSearWater I -5 Q Hydraulic Fill -10 _15 I, -20 Send Fill I Old Bay Mud (a) -25 ' -10 - -5 0 5 10 1s 20 25 30 35 Displacements (m) Excess Pore pressure ratio __ 0.9 0.8 0.7 0.6 0.5 0.40. 0.3 0.2 0.1 0.9 0.7 0.5 03 0.1 . (b) (c) Figure 4-8 Response of the unimproved piled wharf for ground motion nga0779 at t= 10s (a) Structural displacements (50:1) (b) Contours of free-field displacements (c) Contours of freefield excess pore pressure ratio. 86 t=15 s , , -~~g-smu - 0 swum smw A Sea Water N -5 Hydraulic Fill -10 1/ -15 -20 *- Sand Fill Old Bay Mud (a) -25 -10 -5 0 5 10 15 20 25 30 35 Displacements (m) Excess Pore pressure ratio r7 090.9 00.7 0.5 0.5 00.3 0.2 01 0.1 0.4 0.3 (b) (c) Figure 4-9 Response of the unimproved piled wharf for ground motion nga0779 at t= 15s (a) Structural displacements (50:1) (b) Contours of free-field displacements (c) Contours of freefield excess pore pressure ratio. 87 t=20 s mnil - 0 A KK Q -5 -15 r Old Bay Mud (a) -251 -10 zzz F !L I Sand Fill -5 Sea Water _ ea Water r r -10 Hydraulic Fill -20- m ml wui siniuu ~s.w 0 5 10 15 20 25 30 35 Displacements (m) Excess Pore pressure ratio F_ 0. 0.8 0.7 0.6 0.2 01 0.9 0.7 0.1 -1 (b) (c) Figure 4-10 Response of the unimproved piled wharf for ground motion nga0779 at t= 20s (a) Structural displacements (50:1) (b) Contours of free-field displacements (c) Contours of freefield excess pore pressure ratio. 88 (a) 0 ( ( 'I 'I A S" W~t~ L*V4I F -5' Hydruc Fm -10 My~O00 kNm -151- -20 ____II on6 -10 Say LII|ILIZZ\I r F-, ( ued -5 0 5 10 is 20 15 20 V 25 30 35 25 30 35 40 (b) -s -10V -20 -25 ki -10 -- -s - ------ 0 5 --- - 10 ---- s 40 Figure 4-11 Response of the unimproved piled wharf under ground motion nga0779 (a) bending moment envelopes and (b) location of the plastic hinges in piles. 89 0. (a) (b) 0.4 0.2 --- --- 0 ..E-0.2 o 0.6 0) 0.4, W 0. 0 0.6 -i - -0.2 (e) (d) U nga0982 0 0 (g) 5 10 is 20 25 (h) - I0-4 Case: Unimproved piled wharf 0.2 Elastic Perfectly Plastic Piles Ground Motions: 56 records (NGA) -0.2 0 S 10 15 20 25 0 6 10 16 20 25 time (sec) Figure 4-12 Time histories of deck displacements for unimproved piled wharf structure for the reference suite of 56 ground motions . . . . ...... .. .............. 0.7 CD 0.7 0.6 - ) nga0753 (b) nga0753 nga0779- . 0.6 nga0779- 0 CD 0.5 0 0.5 0 C 4) E ) 0.4 0.4 CL, 0 0.3 C 0 N 0 o0.23 r 2=0.6607A A A N o 0.2 -A o 0.2 0 M E -A E E0.1 0.1 0.2 0.3 0.4 0.5 0.7 0.6 0.8 0.9 1 - - CU AAA r 2=0.7505 AOL A E0.1 0 10 20 30 40 PGA (g) 50 60 70 80 90 100 PGV (cm/s) 0.7 S0.6 (D 0 0 a, (D0.5 0 (D CL 0.4 C 0 o0.3 N X 0 M E E E E0.1 CU 0 5 10 15 20 25 30 35 40 45 0 1 2 PGD (cm) 3 4 5 6 7 8 9 Arias Intensity (rms) Figure 4-13 Summary of computed deck displacements for unimproved piled wharf structure as functions of earthquake intensity parameters for the reference suite of 56 ground motions a)PGA, b) PGV, c) PGD, d) Arias Intensity ................ . ........... .......................................... : ::...... ::............ ......... 0.9 E ~0.8E * 9 0 Fit No damage Light Damage Moderate Damage X Heavy Damage CU 0. 7 - (D 0 4 0 0.6N or 2 0.5C R2=0.969 CU 0 02 -5 00 02 .406 . 060. . Maximum Horizontal Displacement of the Soil (node 49)(m) Figure 4-14 Response of the unimproved piled wharf structure and classification of structural damage for the reference suite of ground motions 92 Old Bay Mud 10 -4 J 5 0 10 15 20 25 5 30 0 (a) Light damage, nga08O2 6 0 10 20 1 25 30 (b) Moderate damage, nga0982 0 A K I -5 rr r Hydraulic FIN 1 r -10 -15 -20 SandFINf11 Sea Water Level Q 7 ~lW Old Bay Mud -25 -10 -5 - 0 5 10 15 20 25 30 35 40 (c) Heavy damage, nga0779 Figure 4-15 Examples of structural damage in selected seismic ground motions for the unimproved piled wharf structure 93 Yielded Section b) 0.9 0.8 E W E a, a) 0.7 0.6 0.5 W VU 0.4 a. 0.3 -Shafieezadeh (2011) 0,2 0 -Panagiotidou (2013) 0.1 0 -0.1 5 10 15 20 25 30 time (sec) Figure 4-16 Comparison with results of Shafieezadeh (2011): a) Time Histories of Horizontal deck displacements; b) yielded sections as calculated by Shafieezadeh (2011); c) yielded sections as calculated by Panagiotidou (2013) . ....... .. . E 3 0.7Symbols A p0.6 0 0.5 -0 E 0.4 Case Unimproved PV-drain mitigation system Compacted fill 0.3 5 0- 0.2 N b E 0.1 E x 0 PGV (m/s) PGA (g) E ~0.7 (.0 W, U 0.6 0 r 0.5 (D E (D 80.4 O) 0 0.3 N 0.2 0 E 0.1 _E .x 0 (U 20 25 PGD (cm) 30 35 40 450 1 2 3 4 5 6 Arlashtensity (n/s) Figure 4-17 Effectiveness of ground improvement scenarios for piled-wharf structure under reference suite of ground motions as functions of selected intensity parameters: a)PGA, b) PGV, c) PGD, d) Arias Intensity ...... ............................... ...... UntreatedcaseDisplacements -0 -IC Untreatedcase- 40- Excess Pore 0 (a) Unimproved Soil TreatedcaseDisplacements -%%,Treated caseExcess Pore 1.I'~6~Pressure Ratio 4 -0 0 6 1* 15 so 20 30 56 40 (b) PV-drains mitigation system * Q-%j---------in-- Ezc~ Pore Prnw Rd.~ *S S------S NS- A s~ *~ Q / K contrm'blouml0t WOvift re .1 I~ -20 K -10 1 0 5s 55 0 (c) Compacted fill Figure 4-18 Effect of ground improvement systems on fill and wharf response under ground motion nga0779 96 UntreatedcaseExcess Pore PressureRaio (a) Unimproved Soil Treated caseDisplacements so I J TreatedcaseExcess Pore Pressure Rado 'M' 4$ ai 13 5 20 2 40 n (b) PV-drains mitigation system Sm.PorersPrmaa RanSm. Ie WAINe twe Codiourbva 01 I J -1 0isO18 . 2 2d 30 36 (c) Compacted fill Figure 4-19 Effect of ground improvement systems on fill and wharf response under ground motion nga0753 97 40 Horizontal Displacement of the Deck nga0779 0.3- 0.20.10 | -0.10 S-0.20 -0.3-0.4 -015 - - 0 15 time (sec) Figure 4-20 Effect of ground improvement systems on deck displacements under ground motion nga0779 Horizontal Displacement of the Deck nga0753 W a) F3 5 10 15 20 25 30 time (sec) Figure 4-21 Effect of ground improvement systems on deck displacements under ground motion nga0753 98 AXIAL LOAD 61cm HORIZONTAL LOAD I- 0 A-A' E r' #32 T-HEADED REINFORCEMENT BAR _W11 WIRE SPACING VARIES #36 HOOPS A #25 HOOPS P 2-35m 4 Figure 4-22 Section geometry and reinforcement details of the tested pre-stressed pile (from Lehman et al., 2009) 3 Strength Limit " ........... 6 8co- FcO Envelope Curve Fy4 5 Figure 4-23 Bilinear hysteretic model with strength limit (from Ibarra et al.2005) 99 Uniod. Y Stiff. Det.y u r r y Post -Capping Strength Det Basic Strength Det. Mc Chord Rotation 8 o Effective yield strength and rotation (My and 6y) Effective stiffness K,=M,/6, Capping strength and associated rotation for monotonic loading (M, and 0c) Pre-capping rotation capacity for monotonic loading 6 p Post-capping rotation capacity 6,4 o Residual strength Mr o Ultimate rotation capacity 6. o o o o = KM Figure 4-24 Modified Ibarra Krawinkler (MIK) Deterioration Model (from Lignos et al. 2011) 100 -- 1.5 MIK ModelTest # 9 Leh man et al. (2009) 1 E 0 rr~ NV -M E -0.15 Iff Is -0.1 0.1 0.15 I-. 0 z Drift Ratio f -1.5 Figure 4-25 The calibration results of the MIK model for pile-deck connection for the simulated experiment (Figure 4-22 101 0.6 -- 0.5 0.12 - (a) 0.1 = 0.1 - 0.4 0.3 0.08 0.2 - 0.06 MIKMI 0.1 -- o - 0.04 EPP 0- 0.02 0 N 0* - - -0.2 x -0.3 -0.02 5 10 Time, t (secs) E 20 25 0.08 0.1 600 -EPP40(d (C) - -MIK 200 .-. 15 Time, t (secs) 600 400 -- EPP 200 0 Z 5 15 20 25 -0.02 -200 -400 - -400 z -600 - - -800 -- -600 0.02 0.04 0.06 - EPP -I -MIK 800 -1000 --- 1000Time, t (secs) Rotation at Connection (rad) Figure 4-26 Effect of deteriorating pile-deck connection on response of the unimproved piled-wharf structure under base case ground motion, nga0779 .... ....... 0 (a) Sea Water Level Hydraulic Fill -10- -15- -20- Sand Fill Old Bay Mud -25 2; 1 -5 0 5 10 15 20 25 30 35 40 Figure 4-27 Effect of deteriorating pile-deck connection on response of the unimproved piled-wharf structure under base case ground motion, nga0779: (a) MIK, (b) EPP ...... ... ... 5 5.1 Summary, Conclusions and Recommendations Summary This thesis focuses on the study of the dynamic response of a piled-wharf structure supported within a loose, liquefiable sandy fill soil. The method of analysis that has been used is the uncoupled substructure approach that involves separate analyses for the response of the soil mass "free-field" and for the wharf structure (piles, deck, and crane). The soil-structure interaction between the two models is handled through macro-elements that require the time histories of displacements and pore pressure of the free field response in the locations of the piles supporting the wharf as input motions (Varun, 2010). This research supplements the prior studies by Vytiniotis (2011) on free field behavior and Shafieezadeh (2011) on piled-wharf response. The current study evaluates the effectiveness of ground improvement methods (densification, PV-drains) in terms of limiting structural damage and permanent deformations. To study the seismic response of the wharf structure, we developed a finite element model in OpenSees that can predict the permanent deformations of the structure and the along the piles during seismic loading. The earthquake loading is imposed at specific locations along the embedded length of piles at the free nodes of the macroelements in terms of time histories of free field displacements and excess pore pressures. The soil-pile interaction phenomena are incorporated through the use of the macroelement developed by Varun (2010) that is calibrated using standard engineering soil properties for the site of interest. A suite of 56 ground motions that are typical of firm-site conditions in coastal California was used to evaluate 104 the dynamic response of the wharf of interest. We then compared the effectiveness of 2 ground improvement scenarios for piled-wharf structure under reference suite of ground motions, a PV-drains mitigation system and an ideally compacted fill. Finally, we implemented a more advanced model of the pile-deck connection that can capture the deteriorating behavior exhibited by the connection. 5.2 Conclusions The use of the sophisticated macroelement in the analyses provides significant advantages over the full 3D finite element analyses for simulating complex SSI for piles especially in terms of modeling complexity and computational time. The macroelement captures efficiently the fundamentals mechanics of saturated granular soil-pile interaction. Moreover, the macroelement is easily calibrated using standard parameters obtained from laboratory tests and other available correlations. However, further research is required to identify the validity of the macroelement predictions using experimental data. Moreover, the results of the predictions of the analyses with the macroelement should be compared with 3D analyses using other constitutive models than the ones used in the original analyses Our major conclusion is that the permanent deformations of the structure are primarily governed by the lateral spreading of the soil and therefore retrofitting methods should be targeted in decreasing the soil deformations. We can also support the fact that due to the high nonlinear nature of the phenomenon, no single measure of earthquake intensity can be used as a definite predictive tool for the anticipated damage on the structure. However, reasonably 105 good correlations were achieved with Arias Intensity and PGV. Furthermore, we identified 4 ranges of response which were highly associated with the lateral deformations of the slope : (a) zero damage, structure remains elastic; (b) light damage, where failure occurs only at the piledeck connections; (c) moderate damage, where some failure occurs also inside the soil mass; (d) heavy damage associated with the creation of failure mechanisms. In total for the untreated fill, from the 56 motions, only two caused extensive damage to the structure, 6 of them caused moderate damage and 6 caused light damage. The analyses also show that a system of PV-drains can be used to mitigate permanent deformations of the deck and reduce structural damage to the wharf. In fact, the results indicate a general reduction of the level of damage within the piles. Hence, further improvements can be obtained by retrofitting the more susceptible pile-deck connections. The results of our analyses indicated that the PV-drains mitigation system results are comparable to that of the "ideal" full densification of the slope. Finally, we implemented a more advanced modeling of the pile deck connection that can capture the deteriorating behavior exhibited by the connection. The results of this analysis indicated that the computed deck displacements are reaching smaller values than in the elasticperfectly plastic case. Nevertheless, the structural damage that occurs in the embedded part of the piles is indeed more extended. 106 5.3 Recommendations Future research efforts should be concentrated on development and validation of constitutive laws that can predict more reliably the cyclic response of saturated sandy soils. The lateral spreading is the main cause of the structural damage in the wharf structures; therefore better free-filed prediction capabilities are needed. Other important topics include strategies for structural retrofitting of the wharf and low intrusive improvement methods such as colloidal silica. Future studies should be also directed at modeling newer wharf designs and assess their vulnerability, thus testing the design suggestions given by current design codes. Finally, research efforts should be targeted in validating the developed tools of analysis with 3D numerical analyses and experimental data. 107 REFERENCES Bouc, R. (1971). "Mathematical Model For Hysteresis." Acustica, 24(1), 16-&. Boulanger, R. W., Curras, C. J., Kutter, B. L., Wilson, D. W., and Abghari, A. (1999). "Seismic soilpile-structure interaction experiments and analyses." Journal of Geotechnical and Geoenvironmental Engineering, 125(9), 750-759. Brandenberg, S. J., Boulanger, R. W., Kutter, B. L., and Chang, D. (2007). "Liquefaction-induced softening of load transfer between pile groups and laterally spreading crusts." Journal of Geotechnical and Geoenvironmental Engineering, 133(1), 91-103. Chiou, B. S. J., and Youngs, R. R. (2008). "An NGA model for the average horizontal component of peak ground motion and response spectra." Earthquake Spectra, 24, 173. Cornell, C. A., Jalayer, F., Hamburger, R. 0., and Foutch, D. A. (2002). "Probabilistic basis for 2000 SAC Federal Emergency Management Agency steel moment frame guidelines." Journal of Structural Engineering-ASCE, 128(4), 526-533. Dafalias, Y. F., and Manzari, M. T. (2004). "Simple plasticity sand model accounting for fabric change effects." Journal of Engineering Mechanics-ASCE, 130(6), 622-634. Dobry, R., Vicente, E., Orourke, M., and Roesset, J. (1982). "Horizontal Stiffness and Damping of Single Piles." Journal of the Geotechnical Engineering Division-ASCE, 108(3), 439-459. Esmaeily, A. (2011). "KSU_RC Analysis of Reinforced <http://www.ce.ksu.edu/faculty/esmaeily/KSU RC.htm>. Concrete Members." Finn, W. D. L., and Fujita, N. (2002). "Piles in liquefiable soils: seismic analysis and design issues." Soil Dynamics and Earthquake Engineering, 22(9-12), 731-742. Gazetas, G. (1984). "Seismic response of end-bearing single piles." International Journal of Soil Dynamics and Earthquake Engineering, 3(2), 82-93. Gonzalez, L., Abdoun, T., and Dobry, R. (2009). "Effect of Soil Permeability on Centrifuge Modeling of Pile Response to Lateral Spreading." Journal of Geotechnical and Geoenvironmental Engineering, 135(1), 62-73. Ibarra, L. F., Medina, R. A., and Krawinkler, H. (2005). "Hysteretic models that incorporate strength and stiffness deterioration." Earthquake Engineering & Structural Dynamics, 34(12), 1489-1511. Kagawa, T., and Kraft,(Kobe 2009) L. (1981). "Lateral Pile Response During Earthquakes." Journal of the Geotechnical Engineering Division-ASCE, 107(12), 1713-1731. 108 Kausel, E., and Roesset, J. (1975). "Dynamic Stiffness of Circular Foundations." Journal of the Engineering Mechanics Division-ASCE, 101(6), 771-785. Kavvadas, M., and Gazetas, G. (1993). "Kinematic Seismic Response and Bending of Free-Head Piles in Layered Soil." Geotechnique, 43(2), 207-222. Kobe, C. o. (2009). "The Great Hanshin-Awaji Earthquake Statistics and Restoration Progres ", <http://www.city.kobe.lg.jp/safety/hanshinawaji/revival/promote/img/january.2009.pd f>. (5/12/2013, 2009). Lehman, D. E. (2009). "Seismic performance of pile-wharf connections, in Proceedings of TCLEE 2009: Lifeline Earthquake Engineering in a Multihazard Environment." E. Brackmann, A. Jellin, and C. W. Roeder, eds.Oakland, 865-877. Leonard, T. (2010). "Haiti earthquake: damaged port reopens to aid ships." <http://www.telegraph.co.uk/news/worldnews/centralamericaandthecaribbean/haiti/7 032385/Haiti-earthquake-damaged-port-reopens-to-aid-ships.htm>. (5/12/1023, 2013). Lignos, D. G., and Krawinkler, H. (2011). "Deterioration Modeling of Steel Components in Support of (Leonard 2010)Collapse Prediction of Steel Moment Frames under Earthquake Loading." Journal of Structural Engineering-ASCE, 137(11), 1291-1302. Liyanapathirana, D. S., and Poulos, H. G. (2005). "Seismic lateral response of piles in liquefying soil." Journal of Geotechnical and Geoenvironmental Engineering, 131(12), 1466-1479. Makris, N., and Badoni, D. (1995). "Seismic Response Of Pile Groups Under Oblique-Shear And Rayleigh-Waves." Earthquake Engineering & Structural Dynamics, 24(4), 517-532. Makris, N., and Gazetas, G. (1992). "Dynamic Pile Soil Pile Interaction .2. Lateral And Seismic Response." Earthquake Engineering & Structural Dynamics, 21(2), 145-162. McKenna, F., Scott, M. H., and Fenves, G. L. (2010). "Nonlinear Finite-Element Analysis Software Architecture Using Object Composition." Journal of Computing in Civil Engineering, 24(1), 95-107. Nogami, T., and Konagai, K. (1988). "Time Domain Flexural Response Of Dynamically Loaded Single Piles." Journal of Engineering Mechanics-ASCE, 114(9), 1512-1525. Novak, M. (1974). "Dynamic Stiffness and Damping of Piles." Canadian Geotechnical Journal, 11(4), 574-598. Pestana, J. M., Whittle, A. J., and Salvati, L. A. (2002). "Evaluation of a constitutive model for clays and sands: Part I - sand behaviour." international Journal for Numerical and Analytical Methods in Geomechanics, 26(11), 1097-1121. 109 Popescu, R., and Prevost, J. H. (1995). "Comparison Between Velacs Numerical Class-A Predictions And Centrifuge Experimental Soil Test-Results." Soil Dynamics and Earthquake Engineering, 14(2), 79-92. Prevost, J. H., Abdelghaffar, A. M., and Elgamal, A. W. M. (1985). "Nonlinear Hysteretic Dynamic-Response Of Soil Systems." Journal of Engineering Mechanics-ASCE, 111(5), 696-713. Prevost, J. 1995. DYNAFLOW: A nonlinear transient finite element analysis program. ,:Department of Civil Engineering and Operations Research, Princeton University, Princeton,NJ. Rahnama, M., and Krawinkler, H. (1993) (1993). "Effect of soft soils and hysteresismodels on seismic design spectra." The John A. Blume Earthquake Engineering Center, Stanford University, Stanford,CA. . Rathje, E., and Bray, J.(2000). "Nonlinear Coupled Seismic Sliding Analysis of Earth Structures.", Journal of Geotechnical Geoenvironmental Engineering, 1002-1014. Shafieezaedeh, A. (2011). "Seismic vulnerability assessment of wharf structures." Georgia Institute of Technology, Atlanta, Ga. Tabesh, A., and Poulos, H. G.(2001). "Pseudostatic approach for seismic analysis of single piles." Journal of Geotechnical and Geoenvironmental Engineering, 127(9), 757-765. Tokimatsu, K., Mizuno, H., and Kakurai, M. (1996). "Building Damage Associated with Geotechnical Problems." Soils and Foundations, 36(Special), 219-234. Varun (2010). "A non-linear dynamic macro-element for soil structure interaction analyses of piles in liquefiable soils." Georgia Institue of Technology, Altanta, GA. Verruijt, A. (1995). Computational Geomechanics, Springer. Vytiniots, A. (2012). "Contributions to the Analysis and Mitigation of Liquefaction in Loose Sand Slopes." MIT, Cambridge, MA. Wang, S., Kutter, B. L., Chacko, M. J., Wilson, D. W., Boulanger, R. W., and Abghari, A. (1998). "Nonlinear Seismic Soil-Pile Structure Interaction." Earthquake Spectra, 14(2), 377-396. Wen, Y. K. (1976). "Method For Random Vibration Of Hysteretic Systems." Journal of the Engineering Mechanics Division-ASCE, 102(2), 249-263. 110 APPENDIX A - Validation of OpenSees Response Methodology One important task of this research was to validate the OpenSees framework in vertical and lateral loading of single piles supported on springs. This was done by comparing analytical solutions of the response of a single pile embedded in one soil layer under laterally and vertically loading with the finite element model. In the finite element model, the pile is simulated by beam elements and a series of springs are used to represent the soil. The soil springs are generated and are assigned separate uniaxial material objects in the vertical and 4and lateral directions. The validation of the response of the pile embedded in one soil involved two different constitutive behaviors of the soil: 1) vertically loaded pile with elastic and elastic, perfectly-plastic materials and, 2) laterally loaded pile with elastic and perfectly-plastic materials. The embedded pile length is 20 m and the pile with diameter, D=1.0 m. The mesh is defined by the number of the elements. The spring nodes are created with three dimensions and three translational degrees-of-freedom. The soil springs are generated using zero-length elements to which separate uniaxial material are assigned to represent the force-displacement relationship in the lateral and vertical directions, therefore two sets of nodes are created, sharing the same set of locations. One set of spring nodes, the fixed-nodes, are initially fixed in all three degreesof-freedom. The other set of nodes, the slave nodes, are initially fixed in only two degrees-offreedom, and are later given equal degrees-of-freedom with the pile nodes. The constitutive behavior of the springs is defined such that the springs oriented in the lateral direction 111 represent p-y springs, and the vertically-oriented springs represent t-z and Q-z springs for the pile shaft and tip, respectively. The pile nodes are created with three dimensions and six degrees-of-freedom (3 translational, 3 rotational). With the exception of the uppermost pile head node, the pile nodes are fixed against translation in the y-direction and rotations about the x- and z- axes. The pile head node, where the load is applied, has no rotational fixity, thus simulating a free-head pile. The pile is given elastic behavior for simplicity. Displacement Beam Column elements with elastic section were used to model the pile, with a modulus of elasticity, E = 25 GPa and shear modulus, G = 9.6 GPa (Poisson ration v = 0.3). The use of displacement based elements can allow simulating the spread of plasticity along the element in future research. Vertically Loaded Piles Elastic Springs Vertical-oriented springs were assigned uniaxial Elastic material behavior with elastic spring constant of the shaft k,= 20 MN/m 3 and elastic spring constant of the tip kp=35 MN/m 3 . In order to verify our numerical solution, we compared it with the analytical solution proposed by Veruijt (2010) for pile supported on continuous elastic springs along the length of the pile and at the tip. In Figure A-1, we can see that we are achieving identical response with numerical and analytical solutions. The results presented are for vertical load, P=10 MN. The analytical fpr vertical displacement and the normal force in the pile are: ph cosh[(L- z)/ h] W=EEA n (L sinh /h) (Ah(1) 112 N = -P sinh[(L -z) / h] (A.2) sinh (L / h) where h is the characteristic length h = EA/k, O, k, has the character of a subgrade modulus (r = kw) and 0 is the circumference of the pile. Elastic Perfectly Plastic Springs Vertical-oriented springs were assigned uniaxial Elastic Perfectly Plastic material behavior with elastic spring constant of the shaft k= 20 MN/m 3 MN/m 3 and elastic spring constant of the tip kp=35 and quake (displacement where the perfectly plastic behavior begins) wo=0.01D. For this particular case, we derived the analytical solution, for pile supported on elastic-perfectly plastic springs along the length of the pile and at the tip.: E +k h M=E E-k~h C Uei exp 2Lh (P-roOd)h B= M (P-ro0d)h (1-M) (1-M)EA (1-M)EA =Cexp -) xhdj+Bexp eh Ne =-EA[-exp h U A00 2EA h h P EA (A.3) h exp (_h (M+1)(P--ro0d)h (1-M) EA N,,= -(P+r 0Ox) where ro = (k, wo) is the maximum shear stress, d isthe plastic region Figure A-2 shows very good agreement between numerical and analytical solutions. The results presented are for four different levels of loading starting for small loads where the system exhibits purely elastic behavior, to higher loads where there is only a region where the soil 113 exhibits plastic behavior and to finally to the ultimate load. Laterally Loaded Piles For the horizontal case, the embedded length of the pile was 50m in order to secure a perfect match with the analytical solutions derived by Veruijt (2010) for the case of infinite long pile: Figures A-3 and A-5 show that our assumption was correct and indeed only the upper 20 meter of the pile undergo lateral displacement. Elastic Springs Horizontally-oriented springs were assigned uniaxial Elastic material behavior with elastic spring constant, k= 20 MN/m 3 . We reproduced the analytical solution proposed by Verruijt (2008) for pile supported on elastic springs along the length of the pile: u - PA 3 exp(-z/ 2)cos(z/) 2EI (A.4) where A4_ 4E1 k is the characteristic length of the pile. Figure A-3 shows that we are achieving a perfect match between numerical and analytical solution. The results presented are for P=10 MN. Elastic Perfectly Plastic Springs Horizontally-oriented springs were assigned uniaxial Elastic Perfectly Plastic material behavior. Since we verify our analysis using the analytical solution produced by Brum (reported by Veruijt, 2010), the springs have to be simulated as Perfectly Plastic (Figure A-4).: (K, -K,)7'D(8h2+9hz+3) 360EI Z) 3 ( 114 +K)cD(h+z) 12EI 3 where h is determined by solving (Equ. A.6) for a given level of loading P P= (K, -K.)jDh2+( K-+ (A.6) JK)cDh These springs were modeled by springs with very high stiffness and very small yielding strain. From Figure A-5, we present results for P=3000 kN. We can observe a small difference between the numerical and the analytical solution, this is because of the Blum's assumption that the pile is clamped at the depth where the deformation changes sign. In Brum's assumption the soil pressure at that depth is replaced by a point load in order to maintain equilibrium. It is worth noticing that in this case the strength of the springs depends on the depth of the soil. Macroelement Springs Horizontally-oriented springs were modeled with the Macroelement elements created by Varun (2010). Figure A-6 presents the displacement vs. resistance curves of the macroelements for 6 cycles of consecutive loading using the soil parameters mentioned in the figure. The vertical axis corresponds to the normalized lateral resistance while the horizontal to the imposed displacement at the element. The macroelement has as an input a sinusoidal displacement with amplitude of 0.02 m and frequency of 0.1 Hz, and assumes a linear pore pressure accumulation in order to demonstrate the effect of the phenomenon in the degradation of the soil stiffness. The properties of macroelement necessary for this analysis, which were homogeneous along the depth of the model, are shown in the same figure. 115 Comparison of the Pushover Response of the three Spring Elements Figure A-7 shows the lateral response of an elastic pile embedded in 3 different scenarios (elastic springs, elastic-perfectly plastic springs and macroelements). The finite element mesh is the same as used for the vertical cases as the pile length is 20 m. The macroelement soil springs are generated using special elements PYMacro2D elements (implemented in OpenSees by Shafieezadeh, 2011). In order to be able to compare the results, the same modulus of elasticity has been used for all three scenarios (E = 20 MPa) and for the two latter scenarios the same ultimate resistance at each depth (friction angle 30 degrees). The results presented are for P=10 MN. We can notice that the macroelement exhibits a more realistic and smooth behavior and it is bounded by the two extreme idealized cases. The reaction of the soil springs is pressure depended for the elastic- perfectly plastic spring and the macroelement. Dynamic Modeling of a single pile Figure A-8 presents schematically the model that has been used for the preliminarily dynamic analysis of a single pile. The main difference in the finite element mesh is that for the seismic analysis, a mass element is added at the free node of the pile. The analysis is driven by time histories from free field analysis (displacements and excess pore pressure ratio) at the free nodes of the macroelements. This particular example imposes displacements reordered from the field analysis (Vytiniotis, 2011) using as an outcrop motion from Parkfield earthquake (2004, M= 6.19). In Figure A-9, the recorded displacement, velocities and accelerations at the mass are presented, along with the excitation. 116 0 2 4 6 81Verruijt (2010) Numerical Solution r- 10-o 1214. 16- P (kN) 10000 ks (kN/m 3 ) 20000 kp (kN/m 3 ) 35000 - 18 20 6 0.0011 LPmax (M) 6.5 7 7.5 8 8.5 9 settlement (m) 9.5 Figure A-1. Settlement along the length of the pile supported on Elastic Springs. 117 10 10.5 11 x 10 3 E E 10- lo~11 12-.--- NumericalSolution AnalyticalSolution,Elastic Region AnalyticalSolution,Plastic Region 14- - Numerical Solution Analytical Solution,Elastic Region Analytical Solution,Plastic Region 14- Large Plastic Region ,0 18 20 0.008 a*NmrclSuto 12 12- 0.009 0.01 0,011 0.012 settlement(m) 0.013 0.014 Failure 1618- 0.015 20 0.014 0.015 0.016 0.017 0.018 0.019 settlement(m) 0.02 0.021 0.022 Figure A-2 Settlement along the length of the pile supported on Elastic- Perfectly Plastic Springs. 118 P A .4 510 15 ks 20 - Veirutt(1 7 Analytical Solution - 2530- P (kN) 10000 35 ks (kN/m 3 ) 20000 40- 6 max 0.24 (M) . 45 I 50 -4 -0.8 -0.6 -0.4 I- I 0.2 0 .0.2 horizontal displacement (m) 0.4 0.6 0.8 1 Figure A-3 Horizontal deformation along the length of the pile supported on Elastic Support. h KpO ' +2 c Kp passive 0.5 Ka O' -2 c Ka active Horizontal Displacement Figure A-4 Perfectly plastic soil response 119 c 25- 3035404550 -0.05 0 I 0.05 I I 0.15 0.1 horizontal displacement (m) I 0.2 0.25 0.3 Figure A-5 Horizontal deformation along the length of the pile supported on perfectly plastic springs. 120 0.6 0.4>) a) 0.20 CU) 0- 4- 0 CD, CU -0.2 -0.4 a) N 0 z -0.6- -0.8 -0.02 -0.015 -0.01 -0.005 0 0.005 0.01 0.015 0.02 Horizontal Displacement, u (m) Figure A-6 p-y curve for the macroelement. The excitation is sinusoidal with frequency of 0.1 Hz, amplitude of 0.02 m and total number of cycles 6. This case assumes a linear pore pressure build-up. 121 2 4 6 8 Macroelement Elastic- Perfectly Plastic E 10 .c Elastic -e- =t 12 14 16 18 -600 -500 -400 -300 -200 -100 0 100 200 -0.05 0 0.05 0.1 reaction (kN) 0.15 0.2 0.25 0.3 horizontal displacement (m) 200 - - 100- - Depth=15 m Depth=10 m Depth=7.5 m Depth=5 m Depth=2.5 m Depth=0.5 m 02 .2-100- -200- -300- -400 -0.02 ' 0 ' 0.02 L 0.04 L L 0.1 0.08 0.06 displacement (m) 0.12 0.14 0.16 0.18 Figure A-7 a. Horizontal Reactions b. Horizontal Displacements c. Force-Displacement curves at various depths along the pile. The macroelement exhibits a behavior that is bounded but the responses of the elastic and the elastic-perfectly plastic springs. 122 I I1 444Figure A-8 Schematic model of earthquake loading of a single pile 43- .2 2- jN -2 0 time(s) 5 10 15 time (s) 20 25 30 0 E1 4) 10 15 time (s) 20 25 15 time (s) 30 30 Figure A-9 a Outcrop Horizontal Acceleration used in the free field analysis to produce the free field displacement, b. Recorder Horizontal Displacement at the mass center, c. Recorder Horizontal Velocity at the mass center, d. Recorder Horizontal Acceleration at the mass center. 123 APPENDIX B -Classification of Damage for Piled-Wharf Structure Dwaae Level Record Nube Numiber 33 145 150 Earthquake me AUmasproed Intent Loose Fil Deck Displacemnent im) Compacted Fill PV-Drains System Unimproved Loose Fit Compacted FiH Dins Di 0 0 0.077 0.067 0.061 0.047 0.037 0.153 0.131 (I)Sse 0.45 0.29 Parkfield 0.22 Lake __ __ 0.40 Lake 0 0.19 _ __ __ _ _ 0.77 __ _ _ _ _ _ _ __ _ _ __ _ __ __ _ _ 0.030 _ _ _ 0.136 _ 448 n 0.34 0.68 0.126 0.099 0.101 451 n 0.97 2.89 0.234 0.171 0.203 472 n 0.07 0.10 0.038 0.029 0.026 0.14 0.17 0.025 0.017 0.013 0.032 0.024 0.018 0.034 0.024 632 648 649 669 676 Whittier Narrows01 Whittier Narrows01 Whittier Narrows01 Whittier Narrows01 Whittier Narrows1 01 0.15 0.13 0.15 0 0 0 0.21 0.20 _ 0.035 0.26 0.028 0.043 0.21 II - 0 - - - -- - - - - _ .,. 0.026 0.030 1 1 I I 0.021 _ _ _ 0.16 0.017 0 _ - - - - -- - ...................................... ............... .. 0_ 00 .. ........... . ...... . . ... ........ 1 _ 1 llamage Level Record AisPV- SeEathquake ame NNber Whe Narrows01 684 739 __ _ _ Loma Prieta 0.03 0.02 0.24 __ _ Unmproved Loose FAd 0.012 4.97 _ _ _ _ _ 753 0.50 3.24 X 8.37 X 0.78 791 Pnieta _________en 0.07 _ _ 0.10 01ve Loma Unimproved Compacted Fin Loose Fil PV-Drains System U System 0.009 0.006 _ 0.09 _________ao Compacted FI 0 751 779 U, Ias (mis) PGA (g) _ _ _ _ _ _ _ X _ _ _ __ _ _ _ _ _ ___ 0.169 _ _ _ _ _ 0.138 _ 0.041 0.035 0.032 0.584 0.226 0.285 0.551 0.373 Deck___ 0.08 0.134 _ _ _ 0.506 _ (m)_ _ _ 0.035 0.029 0.031 _____ 802 Loma 0.38 1.45 0.209 0.160 0.178 810 ora 0.46 2.66 0.167 0.129 0.137 897 Landers 0.07 0.12 0.037 0.036 0.042 0.20 0.43 0.036 0.027 0.021 0.06 0.06 0.022 0-017 0.012 982 0.76 3.24 0.334 0.225 0.275 983 0.76 3.24 0.334 0.225 0.275 954 969 Nodig- ...... I -..1--.111 .......... - .1 ... .......... ..... ...... -. 0&z*0 z9z0 1.#0*0 f£.00' SZ0oo £900o 990,0 Z9O*0 9L0 0 __ ___ .. ... .... .. .... ...... .... .................................................................. 0' 0 0t' 0 9zo __ __ ___ _ ___ ___ ___ __ _ _ V91 ale 91.6Lt a'CO 1.0 t 097 tEO 09' NO %*0 ___ 0900o 990 ___ 190*0 lPOO 0V00o 9 6*0 8WO LL O £lLO 'o0 tLI tLEZO SLVO 961-0 Vt'Oo 90*0 Lt.00 11.0* 600*0 9900 9900o __ 1.0 IpoN -06pu 0 LS~ 01.0 L9A)1.SEO C~%I 5pUtpiON ______-e -A 0Z' POumw1M SUWW-Ad -Ad pmPedWO3 JOAal L- 1.0 1.01. 1.0U0 1£0o to.01. 0V PGeMidwRi- IA~18Owe (W)wqwqOe*u pea to £owl. ASewM (6 JO)ObUe 8u ja Sewaet.q a6eujea poj I Damae Level ReNuc Nmber r~J Ea ake Name Arias PGA (g) Intensity (MIS) Unimped Fig Compacted Fill PV-Drains System Deck Di~4acenent (m~ Unimproved Loose Fif Compacted Fill PV- Sse 2374 ahi-2 0.02 0.01 0.006 0.005 0.004 2393 Tahi-2 0.03 0.01 0.007 0.005 0.004 2397 Chi-Chi, Taiwan-2 0.02 0.01 0.008 0.006 0.004 2399 TCi-Chi2 0.05 0.02 0.008 0.006 0.005 2490 Taiwan-3 0.08 0.09 0.032 0.023 0.024 2498 Chi-Chi, Taiwan-3 0.08 0.17 0.081 0.061 0.063 2658 Tai-n3 0.61 0.81 0.055 0.046 0.040 2716 Chi-Chi, Tain-4 0.03 0.02 0.008 0.006 0.005 2804 Chi-Chi, Tauan-04 0.02 0.00 0.003 0.002 0.000 2867 Chi-Chi, 0.02 0.0 0.01 0.011 0.009 0.007 0.022 0.015 0.015 _____TaiWan-04 2871 281Taiwan-04 Chi-Chi, 2883 3008 Ch-Chi TChi-nh, _____ 0.05 _____ 0.05 ____ 0.03 0.03 0.016 0.014 0.012 0.06 0.05 0.010 0007 0.005 .... .......