You are on page 1of 303

Analysis and Performance of the Excavation for the Chicago-State Subway Renovation Project and its Effects on Adjacent

Structures
by Richard J. Finno Michele Calvello Sebastian L. Bryson Department of Civil Engineering Northwestern University September 2002

Table of Contents
TECHNICAL REPORT DOCUMENTATION AND DISCLAIMER ACKNOWLEDGEMENT INTRODUCTION TECHNICAL BACKGROUND 2.1 BEHAVIOR OF STIFF EXCAVATION SUPPORT SYSTEMS IN SOFT TO MEDIUM CLAY 2.1.1 General Deflection Behavior of Excavation Support Systems 2.1.2 Influence of Soil Conditions 2.1.3 Influence of System Stiffness and Installation Techniques 2.1.4 Wall Installation Effects 2.1.5 Settlement Behind Excavation Support Walls 2.1.5.1 Peck (1969) 2.1.5.2 Clough and O'Rourke (1990) 2.1.5.3 Hsieh and Ou (1998) 2.2 BUILDING RESPONSE DUE TO EXCAVATION-RELATED DEFORMATIONS 2.3 PREVIOUS STUDIES TO DEFINE LIMITING CRITERIA 2.3.1 Limiting Criteria Based on Self-Weight Settlement Only 2.3.2 Limiting Criteria Based on Excavation-Induced Distortions 2.3.3 Summary CHICAGO AVENUE AND STATE STREET SUBWAY RENOVATION PROJECT OVERVIEW

3.1 GENERAL PROJECT DESCRIPTION 3.2 SUBSURFACE CONDTIONS 3.2.1 Geology and Site Stratigraphy 3.2.2 Laboratory and Field Testing 3.2.3 Engineering Properties of the Soil 3.3 ADJACENT STRUCTURES 3.4 EXCAVATION SUPPORT SYSTEM DESIGN METHODOLOGY 3.5 EXCAVATION SUPPORT SYSTEM DESCRIPTION 3.6 FIELD INSTRUMENTATION 3.6.1 Ground Instrumentation 3.6.2 Structural Support Instrumentation 3.6.3 Adjacent Structure Instrumentation 3.7 SUMMARY PERFORMANCE AND OBSERVED BEHAVIOR OF FIELD INSTRUMENTATION 4.1 CONSTRUCTION SEQUENCE AND PROCEDURES 4.1.1 Construction Along the East Side of State Street 4.1.2 Construction Along the West Side of State Street 4.1.3 Construction Along Chicago Avenue 4.1.4 Reduction in the Bending Stiffness of the Wall 4.2 OVERVIEW OF FIELD INSTRUMENTATION RESPONSES 4.2.1 Ground and Building Movements 4.2.1.1 Responses along East Side of State Street

4.2.1.2 Responses along West Side of State Street 4.2.1.3 Responses along Chicago Avenue 4.2.2 Pore Water Pressure Response 4.3 DETAILED OBSERVATIONS OF LATERAL MOVEMENTS 4.3.1 Stage 1-Wall Installation 4.3.1.1 East Side of State Street 4.3.1.2 West Side of State Street 4.3.1.3 Chicago Avenue 4.3.2 Stage 2-Excavation and Support System Installation 4.3.2.1 East Side of State Street 4.3.2.2 West Side of State Street 4.3.2.3 Chicago Avenue 4.3.3 Stage 3- Renovation and Backfill 4.3.3.1 East side of State Street 4.3.3.2 West Side of State Street 4.3.3.3 Chicago Avenue 4.3.4 Summary of Lateral Deformations and Discussion 4.3.4.1 Incremental Lateral Soil Movements 4.3.4.2 Lateral Deformation Vectors 4.3.4.3 Comparison of Settlements and Lateral Movements 4.4 LOADS IN CROSS-LOT BRACES AND TIEBACKS 4.5 SUMMARY

EXCAVATION-INDUCED RESPONSE OF THE FRANCES XAVIER WARDE SCHOOL 5.1 SETTLEMENT RESPONSE 5.1.1 Settlement Contours 5.1.2 Zones of Sagging and Hogging 5.2 DISTORTIONS 5.2.1 Computing Distortions from Settlement and Inclinometer Data 5.2.2 Development of Distortions During Construction 5.2.3 Comparison of Deflection Ratios and Distortions 5.2.4 Distribution of Distortions 5.3 TILT OF FOUNDATION WALLS 5.4 EXCAVATION-RELATED DAMAGE 5.4.1 Damage on the First Floor Level 5.4.2 Damage on the Second Floor Level 5.4.3 Damage on the Third Floor Level 5.4.4 Damage to Exterior Walls 5.4.5 Damage Summary 5.5 CRACK GAUGE DATA 5.6 ANALYSIS OF DAMAGE 5.6.1 Crack Types Observed 5.6.2 Initiation of Selected Cracks 5.6.2.1 Onset of Interior Cracking 5.6.2.2 Onset of Exterior Cracking 5.6.3 Summary of Distortions

5.7 SUMMARY DEEP EXCAVATIONS: THE OBSERVATIONAL METHOD AND INVERSE ANALYSIS 6.1 INTRODUCTION 6.2 MODEL CALIBRATION BY INVERSE ANALYSIS 6.2.1 An inverse analysis algorithm: UCODE 6.2.2 Model Fit statistics 6.2.3 Input parameters statistics 6.2.4 Observation weighting 6.3 UPDATE DESIGN PREDICTIONS USING MONITORING DATA BY INVERSE ANALYSIS 6.4 PROCEDURE VALIDATION: THE CHICAGO AND STATE CASE STUDY 6.4.1 Finite Element simulation of the problem 6.4.1.1 Calculation phases 6.4.1.2 Hardening-Soil model initial calibration 6.4.2 Inverse analysis setup 6.4.2.1 Observations and weighting 6.4.2.2 Parameterization 6.4.3 Results 6.4.3.1 Optimization based on observation from construction stage 1 6.4.3.2 Model fit for all construction stages 6.4.3.3 Best-fit parameters 6.4.3.4 Model statistics for all construction stages

6.4.3.5 Comments on the calibrated model 6.5 SUMMARY SUMMARY AND CONCLUSIONS REFERENCES APPENDIX A: CONSTRUCTION ACTIVITY, TIME HISTORY AND STRUCTURAL PLAN Figure A-1- Chicago Avenue and State Street Subway Renovation Project Structural Staging Plan Table A-1- Construction Activities Time History

APPENDIX B: FIELD INTRUMENTATION DATA Table B-1- Survey Data Table B-2- Pore Water Pressure Table B-3- Strut Loads Table B-4A- Tieback Anchor Loads at Inclinometer 2 Location Table B-4B- Tieback Anchor Loads at Inclinometer 1 Location Table B-5A- Inclinometer 1 Data: A-Axis Table B-5B- Inclinometer 2 Data: A-Axis Table B-5C- Inclinometer 4 Data: A-Axis Table B-5D- Inclinometer 5 Data: A-Axis Table B-5E- Inclinometer 5 Data: B-Axis

ACKNOWLEDGEMENT

Many people and a number of organizations contributed to the success of the ChicagoState project. Special thanks are due Messrs. Kevin Hayes and Robert Staiton of Baker Engineering, the construction managers of the project, for their outstanding efforts providing access to the site under difficult conditions. Dr. Andy Longinow and Mr. Nicholas Hyatt of Wiss, Janney Elstner Associates Inc. were responsible for the monitoring of the school and provided structural expertise throughout the project. The help and interest of Mr. Tom Ambry, Ms. Cindy Williams and Mr. Manoher Chawla of the Department of Transportation of the City of Chicago made the work much easier. The support of Mr. C. Gregory Veith, the manager of facilities and construction for the Archdiocese of Chicago, is appreciated. His willingness to accept the proposed solution was key to the success of the project. Mr. Dan McCarthy of CTI instrumented the support elements and was very helpful during all aspects of instrument installation and data collection. Aldridge Drilling installed the secant pile wall and TCDI installed the tiebacks. The help of Messrs. Steven Scherer, Kyle Camper and Jeff Hill of TCDI is greatly appreciated. Dr. Safdar Gill of Ground Engineering designed the wall system for the Chicago DOT. Walsh Construction was the general contractor. Collins Engineers designed the wall for Walsh. The authors also thank Gilles Marchadier, Dan Priest, Benoit Paineau, Jill Roboski, Kristi Kawamura, and Amy Rechenmacher, graduate students at Northwestern University, who generously gave of their time to help with data collection throughout the length of the project. Frank Voss of Northwestern University helped with the preparation of this final report.

The authors acknowledge the financial support for a portion of this work provided by the Infrastructure Technology Institute of Northwestern University. The continuing support of Mr. David Schulz, its director, is greatly appreciated. Support was also provided by funds from a National Science Foundation grant CMS-0115213. The authors also thank Dr. Richard Fragaszy, program manager of Geotechnical and GeoHazardsl Systems, for his support of this work.

CHAPTER 1

INTRODUCTION Infrastructure rehabilitation has become commonplace in many urban areas across the country. Space for construction activities is usually limited because of the proximity of adjacent structures; this limited space necessitates the use of excavation support systems. A major concern for projects involving deep excavation is the impact of constructionrelated ground movements on adjacent buildings and utilities. The degree of stiffness required for an excavation support system is determined from the movements associated with its construction and their effects on the adjacent structures. It is necessary then to quantify the tolerable deformations for each affected structure. However, determining tolerable deformation by purely analytical means is extremely difficult. The amount and severity of excavation-related damage depends on the building type and configuration, support system type and performance, ground conditions, and construction activities. Consequently, a semi-empirical approach based on tolerable deformations is usually employed to design excavation support systems. The damage to a structure is related empirically to the magnitude and distribution of the deformations expected during the project. The stiffness of the excavation support system is chosen based on the limiting criterion developed from this relation. Limiting criteria have been developed from observing the excavation-related ground deformations and building responses associated with different types of support systems in various soil conditions. A necessary component of this semi-empirical approach is complete and comprehensive documentation of the response of buildings adjacent to excavation support systems,

including a complete record of the ground conditions and building responses before, during, and after the construction activity. This information contributes to the understanding of the interactions between the soil, excavation support system, and adjacent structures. Predicting the magnitude and distribution of these movements is critically important to the design process, as these predictions are used to estimate the tolerance of the structures to the induced deformations. However, there are a number of uncertainties associated with these predictions, such as soil properties variations, support system details, construction sequence, and surcharge loads. In practice, the observational

method is commonly used to compare predicted performance with observed responses. This method requires much engineering judgment, and it is very difficult to quantitatively judge how well the work is proceeding in terms of the predicted responses, especially given the time constraints associated with construction. Ad hoc numerical tools are needed to compare observations and predictions. This report presents the observations made during the excavation and construction the Chicago Avenue and State Street Subway renovation project in Chicago, IL. It also summarizes a numerical procedure that objectively updates design predictions of deformations for supported excavations in clay. The monitoring data from the ChicagoState excavation are used as observations in an inverse analysis that calibrates the numerical model of the excavation. The report shows how the updated predictions

made at an early stage of construction can be used to accurately predict the responses at the later stages of the project.

The Chicago Department of Transportation (CDOT) began constructing the subway improvements in June 1999. The renovation project includes excavating 12.2 m of soft to medium clay to expose the existing subway station and tunnels and to allow for capital improvements. The improvements included adding a mezzanine level for office and vendor spaces and making the station handicap accessible. Space limitations presented significant challenges to the project. The project was located in close proximity to the Frances Xavier Warde School and the Holy Name Cathedral. The Warde School was founded on shallow foundations that were located within 1.3 m of the face of the excavation. The Holy Name Cathedral was located approximately 15 m from a corner of the excavation. Options for providing lateral support of the excavation support system was further restricted due to the presence of the subway station and twin subway tubes. The temporary support walls were 2.2 m from the outer wall of the subway tubes Baker Engineering, the construction manager for the project, contracted with Wiss, Janney, Elstner Associates, Inc. (WJE) to monitor ground movements and structural responses associated with the excavation and the subsequent renovations. WJEs primary responsibility was to assess the potential for structural damage to the adjacent buildings. Northwestern University was subcontracted by WJE to perform the portion of the contract involving monitoring the field performance of the support system and predicting the ground movements that resulted from the excavation activities. Data obtained from the monitoring activities included lateral soil deformations, pore water pressures, building movements, and support loads. Data were obtained daily during wall installation and excavation, and at least on a weekly basis after the excavation had

reached its final depth. Building movements were optically surveyed weekly during construction. Chapter 2 of this document presents technical background concerning excavations in soft clay that employ stiff excavation support systems. This chapter discusses the general deflection behavior of braced walls in soft clays and the deformation response of the soil resulting from the wall deflections. It shows the influence of the soil conditions, the system stiffness, and the installation techniques of the support system on the deflections of the support wall. It reviews methods to empirically evaluate the settlement distribution behind excavation support walls in soft clays and summarizes the observed responses of buildings situated adjacent to an excavation. Lastly, this chapter provides a literature review and discussion of limiting deformation criteria established to minimize excavation-related damage to structures. Chapter 3 presents the subsurface conditions of the Chicago Avenue and State Street project site and describes the excavation support system and field instrumentation. It provides descriptions of the structures adjacent to the excavation. It also summarizes the design methodology used to select the excavation support system. Chapter 4 presents the observations and performance of the field instrumentation during the project. It summarizes in detail the project construction sequence and presents the lateral ground movements and building settlements measured during the project. This chapter also presents the pore pressure measurements and the loads in the structural supports observed during excavation.

Chapter 5 presents the observed response of the building adjacent to excavation. These observations include building settlements and distortions as well as observations of excavation-related damage to the structure. This damage is compared to observed distortions. Chapter 6 summarizes the inverse analysis used to calibrate the finite element model of the excavation and procedures needed to update design predictions. This procedure is validated by applying it to the detailed observations made at the Chicago-State project. The soil was represented as by the Hardening Soil model, a multi-yield surface, elastoplastic effective stress model. The results of the inverse analysis are presented for optimizations made at an early stage of construction as well as at later stages. Model statistics that quantify the improvement of the predictions over the original, nonoptimized values are given. Chapter 7 summarizes this work and presents conclusions.

CHAPTER 2

TECHNICAL BACKGROUND

2.1 BEHAVIOR OF STIFF EXCAVATION SUPPORT SYSTEMS IN SOFT TO MEDIUM CLAY Deformations of an excavation support system and the adjacent ground are influenced by a number of factors including support system stiffness, method of support system installation, and soil conditions. When average to good workmanship is employed and the clays are relatively soft, the resulting deformations are most influenced by the support system stiffness, and thus, is the key design parameter used to control ground movements. The ability to explicitly consider the support system stiffness is important to producing predicted behavior consistent with the observed ground and wall response. A stiff braced support system is typically used when the deformations of ground adjacent to the excavation must be limited, particularly when excavating through soft clays. The stiffness of an excavation support system is a function of the flexural rigidity of the wall element, the structural stiffness of the support elements, and the type of connections between the wall and supports, and the vertical and horizontal spacing of the supports. The overall stiffness of the support system is typically expressed in terms of an effective stiffness of the system. Mana and Clough (1981) gave the effective wall stiffness, S, as:

(2.1)

where EI is the wall flexural stiffness per horizontal unit of length (E is the modulus of elasticity of the wall element and I is the moment of inertia per length of wall), H is the average vertical spacing between supports, and is the total unit weight of the soil behind the wall. Koutsoftas et al. (2000) and Clough at el. (1989) defined effective system stiffness similar to the definition in (2.1), except the unit weight of soil is replaced with the unit weight of water, dimensionalize the equation. Walls that are considered stiff on the basis of the rigidity of the wall element include secant and tangent pile walls and structural slurry walls (often referred to as diaphragm walls in the literature). Walls that are considered flexible on the basis of the rigidity of the wall element include steel sheet pile walls and soldier pile and lagging walls.
w.

The unit weight is introduced to non-

2.1.1 General Deflection Behavior of Excavation Support System The behavior of excavation support systems can be expressed in terms of the ground surface settlements and lateral wall deformations. These movements are a function the flexural rigidity of the wall component, the stiffness of the supports, the earth pressures and water loads, the general soil and groundwater conditions, and the construction procedures. Lateral ground deformations associated with excavation support systems are a response to the wall deflection. For relatively small deformations, the profile of lateral ground movements behind the support system tend to match the deflected shape of the wall near the wall. Braced excavations are typically performed in three stages:

(i) Wall installation (ii) Cycles of excavation and lateral support installation (iii) Removal of the supports and backfill. General ground movements profiles, not considering the effects of wall installation, as observed for a typical excavation are illustrated in Figure 2-1.

Figure 2-1.Typical Profiles of Movement for Braced and Tieback Walls (Clough and ORourke, 1990) These profiles would most likely be obtained from inclinometer and settlement measurements. The figure shows that during the initial excavation, before the installation of the first level of lateral support, the wall deforms as a cantilever. Settlement of the adjacent soil tends to decrease with distance from the edge of the excavation. Settlements during this stage may be represented by a triangular distribution of displacements. When the excavation advances to deeper elevations, upper wall

movements are restrained by the installation of new supports. Deep inward movements of the wall occur. The deflected shape of the wall shows a bulge in the deeper portion of the excavation. If deep inward movements are the predominant form of wall deformation, as in the case with deep excavations in soft to medium clay, then the settlements tend to be bounded by a trapezoidal displacement profile. If cantilever movements predominate, as can occur for excavations in sands and stiff to very hard clay, then settlements tend to follow a triangular pattern. Further inward bulging of the wall occur as the bottom supports are removed. The combination of cantilever and deep inward components results in the cumulative wall and ground surface displacements shown in Figure 2-1. Additional cantilever-type deformation at the top of the wall results when the upper supports are removed. In soft to medium clays, small amounts of deformation may be observed until backfill is complete. There are typically no additional movements in stiff clays during the backfill operations. The patterns of movements shown in Figure 2-1 have been justified by theory and from observed behavior. Theoretical and experimental studies by Milligan (1985) have shown that incremental deformations of the wall will generate deformations consistent with those for the cantilever and deep inward movements delineated in Figure 2-1. ORourke et al. (1976), ORourke (1981), and Finno et al. (1989) have presented cases of observed behavior supporting the patterns of movements shown in Figure 2-1. However, this figure only describes the general deflection behavior of the wall in response to the excavation. The soil conditions, wall installation methods, and the

effective stiffness of the excavation support system are specific factors that influence the magnitude of movements of the support system.

2.1.2 Influence of Soil Conditions The behavior of an excavation support system in clay is greatly influenced by the undrained shear strength of the clay. Clough and ORourke (1990) concluded that the average horizontal and vertical movements of support systems in stiff clays were roughly 0.2 percent and 0.15 percent of the total excavated depth, respectively. Their findings agree with guidance established in Canadian Foundation Engineering Manual (1985), which states that lateral movements of temporary support systems in stiff clay are less than 0.2 percent of the excavation depth. This compares to guidance established in NAVFAC DM-7.2 (1982) that suggests in stiff fissured clays lateral movements may reach 0.5 percent of the total excavated depth or higher depending on quality of construction. Clough and ORourke (1990) also conducted a finite element parametric study on stiff clays. Their analyses showed that parameters such as wall stiffness and support spacing have only a small influence on the predicted movements in these soils because in most circumstances these soils are stiff enough to minimize the need for stiff support elements. They found soil modulus and coefficient of lateral earth pressure have a more significant impact on the ground movements. Their results suggested that in a stiff soil, variations in soil stiffness have a more profound effect on wall behavior than system stiffness.

Clough and ORourke (1990) noted that basal stability is typically not an issue in stiff clays. In soft clays however, a major portion of movement occurs below excavation bottom as a result of basal instability. Lateral movement may be in the range of 0.5 percent to 2 percent of excavation depth, depending on the factor of safety against bottom instability and the stiffness of the support system. Higher movements are associated with smaller factors of safety against basal heave. Peck (1969) and, Clough and Reed (1984) showed that the movements of an excavation support system become large when the magnitude of the stability number, N, exceeds the bearing capacity factor for failure of the base of the excavation. The stability number is defined as:

(2.2)

where is the unit weight of the soil above the excavation bottom, H is the depth of the excavation, and Su is the undrained shear strength of the clay beneath the excavation. Clough and Reed (1984) concluded that the increase of movements were a result of plastic yielding of the soft clay at and beneath the bottom of the excavation. Clough et al. (1989) presented the design curves in Figure 2-2.

Figure 2-2. Design Curves for Maximum Lateral Movement in Soft to Medium Clay (Clough et al., 1989) The figure allows the user to estimate lateral movements in clay in terms of effective systems stiffness and the factor of safety against basal heave. The factor of safety against basal heave used in the figure is that given by Terzaghi (1943). For wide excavations (H/B<1), the factor of safety against basal heave is given as:

(2.3)

For wide excavations where there is a strong stratum near the base of the excavation, the factor of safety is given as:

(2.4)

where Sub and Suu is the undrained shear strength below and above the bottom of the excavation, respectively, Nc is the bearing capacity factor at the bottom of the excavation, H is the height of the excavation, B is the width of the excavation, is the unit weight of the soil, and D is the distance from the bottom of the excavation to a relatively hard stratum. Figure 2-2 was created from parametric studies performed using results of finite element analyses. For the analysis, sheetpile and slurry walls with varying effective stiffness were modeled. The figure illustrates the influence of basal stability on movements. In particular, the figure shows that for a given wall stiffness, a lower factor of safety against basal heave results in higher movements caused by the excavation. Clough et al. (1989) suggested that the figure could be used to estimate maximum lateral wall movement in circumstances where movements are primarily due to the excavation process.

2.1.3 Influence of System Stiffness and Installation Techniques

Figure 2-2 also shows that in soft clays, where the factor of safety against basal heave is low, increasing the stiffness of the support system helps to reduce movements. A stiffer support system can be obtained by reducing the vertical spacing between the supports. This assertion agrees with observations made by Goldberg et al. (1976), which showed that both vertical and horizontal support spacings are an important factor in increasing the support stiffness. They concluded that closely-spaced horizontal supports provided as significant contribution to the effective stiffness of the support system as closelyspaced vertical supports. The overall stiffness of a support system is also influence by the stiffness of the supports themselves. The support components include either cross-lot braces or tiebacks, and walers, which are structural members that distribute the load from the wall to the brace. The theoretical stiffness of the supports can be defined in terms of its axial stiffness, KA:

(2.5) where A is the cross sectional area of support, E is the elastic modulus of the support, and L is the unsupported length of the support. Cross-lot braces are compression members, which are subjected to axial loading and elastic bending. Movement of the wall or preloading of the member is required before the stiffness of these members is engaged. ORourke (1981) concluded that their actual stiffness is significantly affected by the nature of the connections to the wall, the use of preloading, and the elastic deformation of the brace. Conversely, tiebacks are preloaded in tension. The stiffness of these members is engaged when the tiebacks are stressed higher than the design load

and then unloaded to a lock-off load, typically 70 percent to 80 percent of the design load. Their theoretical stiffness is close to their actual stiffness. It is noted that although the stiffness of the support components contribute to the overall support system stiffness, support stiffness is not as important a factor to the system stiffness as either the stiffness of the wall component or the spacing of the supports. Clough and ORourke (1990) used finite element analyses to show that within the normal range of system parameters, variations in the stiffness of the cross-lot braces and tiebacks accounted for about 20 percent of the overall combined wall and support stiffness. The effective stiffness of the support system can be improved by preloading the supports. Preloading reduces the slack in the support system that otherwise would have to be taken up by movements of the wall. For a compression member like a brace, this increases the effective stiffness of the brace. Also, preloading reduces the shear stress levels in the soil that are induced by the excavation process. The reduction of shear stresses allows the soil to follow an unloading-reloading response instead of the softer primary loading response. However, quantifying the effects of preloading is difficult using only observed data. Mana and Clough (1981) performed additional finite element analyses to determine the influence of preloading. They found that the use of preloads in the struts reduced movements, although there is a diminishing returns effect at higher preloads. Very high preloads may, in fact, be counter productive since local inward movements at support levels can damage adjacent utilities by inducing horizontal strains.

2.1.4 Wall Installation Effects

Often when estimating movements for a wall it is common to envision the wall in place, and consider only what occurs after that point. However, ground movements are caused by factors other than excavation-induced stress relief. One principal source of movements is related to the construction of the wall itself. DAppolonia (1971) showed that poor construction techniques could also account for large movements of the ground adjacent to insitu walls. He defined insitu walls as secant piles walls, tangent piles walls, structural slurry walls, and soldier pile walls that are augured into place. The quality of construction for an insitu wall project depends upon many factors, including the contractors experience with the subsurface conditions at the site and with the insitu wall system being used. ORourke (1981) also found that significant surface settlement can occur as a result of installing insitu walls. He noted that settlement in soft clays and sands occurred as a result of ground loss when excavating the trench for a diaphragm wall or when drilling shafts for secant and tangent pile walls. In soft clays, soil squeeze appeared to contribute to the surface settlement, depending on the amount of time the trench or shaft remains open before placing the concrete. ORourke (1981) presented case histories where between 50 percent and 70 percent of the total settlements observed were associated with the construction of the insitu wall. Koutsoftas et al. (2000) presented a case history that involved installation of both a soldier pile and tremie concrete (SPTC) wall and a conventional diaphragm wall. The SPTC wall was constructed by first installing wide-flange steel sections in pre-drilled shafts spaced at 3.7 m intervals. The spaced between the steel sections was then excavated using techniques similar to those used to install the conventional diaphragm wall. They initially estimated that in soft clays, settlement caused by the wall

installation could extend to a distance equaled to approximately 1.0 to 2.0 times the depth of the wall. They also estimated that the maximum settlement would occur directly behind the wall with a maximum value equaled to about 0.12 percent of the depth of the wall. Similar to the findings of ORourke (1981), Koutsoftas et al. (2000) found that the actual observed surface settlement associated with the wall installation was a function of the length of time the drilled shafts and trenches remained unsupported. They observed surface settlements equaled to approximately 0.2 percent of the wall depth at locations where drilled shafts were augured through fill without casing. However, lateral wall deflections and the consequent surface settlements were relatively small during the diaphragm wall installation because a positive (net outward) differential slurry pressure was maintained prior to placing the concrete. The pore water pressures are also significantly impacted by the installation of the support wall. Support walls consisting of driven or pre-augured H-piles typically experience a significant increase in pore water pressure during the driving and auguring process. However, the pore pressures tend to dissipate to near hydrostatic levels shortly after installations are complete (Koutsoftas et al., 2000 and Poh and Wong, 1998). Insitu walls that remain open for sometime prior to placing the concrete tend to act as sinks. Groundwater levels often decrease and do not return to pre-construction levels. These drops in the groundwater level increase the insitu effective stress of the soil, leading to increased ground surface settlements.

2.1.5 Settlement Behind Excavation Support Walls

2.1.5.1 Peck (1969)

The first rational basis for estimating ground movements adjacent to excavations was presented by Peck (1969). He compiled ground surface settlement data measured adjacent to temporary braced sheetpile and soldier pile walls, and summarized the data in a chart. Figure 2-3 presents Pecks (1969) chart.

Figure 2-3. Summary of Soil Settlements Behind Insitu Walls (after Peck, 1969)

The chart presents normalized values of ground settlement versus the distance from the excavation. Both axes are normalized using the final depth of the excavation. Peck (1969) grouped the data on the chart into three categories. The categories were developed on the basis of the soil conditions and the level of workmanship employed when constructing the wall. Peck (1969) also recognized that portions of the ground surface deformation patterns might be due to basal instability in the soft and medium clays. Category I includes excavations in sands, stiff clays, and soft clays of small thickness. Category II includes excavations in very soft to soft clays extending a small distance below the bottom of the excavation or with a stability number, Nb, less than 6 or 7. Category III includes excavations in very soft to soft clays that extend to a significant depth below the bottom of the excavation, and with stability numbers greater than the critical stability number for basal heave. In the figure, is the unit weight of the soil above the excavation, H is the final depth of the excavation, and Cb is the undrained shear strength of the soil beneath the excavation. The remaining variables are defined in the figure. It can be seen from Figure 2-3, that for Category I soils the maximum surface settlement is limited to 1 percent of the final excavation depth. The maximum surface settlement of Category II soils is 2 percent of the final excavation depth. However, the extent of the influence extends 2 to 4 times the depth of the excavation.

2.1.5.2 Clough and ORourke (1990) Clough and ORourke (1990) observed that a relatively well-defined grouping of excavation-induced settlement data was evident when the settlements were plotted as fractions of maximum settlement. They presented dimensionless settlement profiles in

Figure 2-4 as a basis for estimating vertical movement patterns adjacent to excavations. Separate profiles were developed for sand, stiff to very hard clays, and soft to medium clays. With knowledge of the maximum settlement, the dimensionless diagrams in Figure 2-4 can be used to obtain an estimate of the actual surface settlement. The figure shows that the settlement influence zone is 3H for excavations in stiff to very hard clays and 2H for excavations in sands and soft to medium clays.

Figure 2-4. Settlement Profiles Recommended for Estimating the Settlement Distribution adjacent to Excavations (Clough and O'Rourke, 1990) Figure 2-4 shows that a trapezoidal envelope bounds the settlement distribution in soft clays. Inside the envelope two zones of movement could be identified. The zone in which the maximum settlement occurred was at 0 = d/H = 0.75 (d is the distance from

the excavation, H is the final height of the excavation). At 0.75 < d/H = 2.0, there was a transition zone in which settlements decreased from maximum to negligible values. In using the diagrams presented by Clough and ORourke (1990), it should be recognized that they pertain to settlements caused during the excavation and bracing stages of construction. Movements associated with other activities, such as dewatering, deep foundation removal or construction, and wall installation, must be estimated separately. Excavations in stiff to very hard clays show variable behavior, with heave possible for some conditions. For stiff to very hard clays, the dimensionless diagram in Figure 2-4 should be used as a conservative estimate, provided that the wall is stable and not affected by poor construction techniques

2.1.5.3 Hsieh and Ou (1998) Hsieh and Ou (1998) suggested that there were two types of settlement profiles caused by excavations: (i) spandrel type, in which maximum settlement occurs very close to the wall; and (ii) concave type, in which maximum settlement occurs at a distance away from the support wall. The spandrel type of settlement profile occurs if a large amount of wall deflection occurs at the first stage of excavation when cantilever conditions exist and the wall deflection is relatively small due to subsequent excavation. After the initial stages of excavation, additional cantilever wall deflection is restrained by installation of support as the excavation proceeds to deeper elevations. The concave settlement profile reflects the ground settlement profile that develops when the movements are more deepseated.

Hsieh and Ou (1998) presented the relationship shown in Figure 2-5 for a spandrel-type condition. The data are presented as normalized settlement, v/
vm,

where

vm

is the

maximum ground surface settlement, versus the square root of the distance-from-theedge-of-the-excavation divided by the-excavation-depth (d/He). This relationship was based on 10 case histories from Taipei, Taiwan. The "mean" estimate curve shown in the figure was derived based on the results of regression analysis.

Figure 2-5. Proposed Method for Predicting Spandrel Settlement Hsieh and Ou, 1998)

Hsieh and Ou (1998) developed the curve in Figure 2-6 for the concaved settlement profile from case histories compiled by Clough and O'Rourke (1990) and obtained from additional sites in Taipei. Hsieh and Ou (1998) concluded that the distance from the wall to the point where the maximum ground surface settlement occurred was approximately equal to half the excavated depth. Assuming the maximum lateral wall deflection occurs near the excavation bottom, the distance where the maximum ground surface settlement occurs can be taken as half the final excavation depth (He/2). Using case histories, the settlement at the wall was established as 0.5
vm.

The point marked by

d/He=2 corresponds to the extent of the primary influence zone, which was defined by Hsieh and Ou (1998) as being equaled to approximately two excavation depths (2He). The case histories also showed that settlement was practically negligible at a distance from the wall equaled to four excavation depths (4He) and was thus used as the farthest most point on the curve. For simplicity, a linear relationship was assumed between each

turning point.

Figure 2-6. Proposed Method for Predicting Concave Settlement Profile (Hsieh and Ou, 1998) 2.2 BUILDING RESPONSE DUE TO EXCAVATION-RELATED DEFORMATIONS The response of buildings adjacent to deep excavations refers to the translation and rotation of individual structural members and to the translation and rotation of the structure itself as a rigid body in reaction to lateral ground movements and surface settlement. These translations and rotations result in direct tensile strains, bending

strains, and diagonal tensile strains in the structural and non-structural members of the buildings. For buildings adjacent to deep excavations, the severity of the responses are dependent upon the stiffness of the excavation support system, the installation procedures of the system, the soil conditions, the excavation procedures, the type of building, the distance of the building from the excavation, the orientation of the building with respect to the excavation, and the size of the building with respect to the excavation. A purely theoretical approach to estimating building response to excavation-related deformations is not possible due to the variability of the many factors that contribute to the response. Consequently, building response is estimated and evaluated on the basis of empirical observations and simplified structural approximations. The goal of estimating and evaluating building response is to provide limiting criteria that will safeguard the structure against unacceptable damage. Burland and Wroth (1974) presented definitions that describe types of ground movements and building responses that result from ground settlement. These definitions are presented in Figure 2-7a to Figure 2-7c. Boscardin and Cording (1989) added definitions describing the ground movement and building response associated with excavations.

Figure 2-7. Building and Ground Response to Vertical and Lateral Deformations (after Burland and Wroth, 1975) These definitions are present in Figure 2-7d

Descriptions of the terms given in Figure 2-7 are as follows: 1. Settlement, Relative Settlement, and RotationThe symbol in Figure 2-7a denotes downward displacement. The symbol
h

implies upward displacement, which is termed and is used to denote differential

heave. Relative settlement is given by the symbol

settlement or differential heave. As can be seen in the figure, the differential settlement is the difference between two settlement points of interest. The symbol denotes rotation and is angle formed from the differential settlement, between two points divided by the distance, l, between them. Rotation is typically used to describe the slope of the settlement trough. 2. Relative Deflection and Deflection RatioThe term shown in Figure 2-7b is the maximum displacement relative to the straight line connecting two reference points a distance L apart. Relative deflection that produces an upward concavity is termed relative sag. Relative deflection that produces a downward concavity is termed relative hog. The term /L is the deflection ratio and is an approximate measure of curvature of the settlement curve. The deflection ratio is often correlated with bending related distortions in a structure. 3. Tilt and Angular DistortionThe rigid body rotation of the entire superstructure is termed tilt and is denoted as in Figure 2-7c. Angular distortion is denoted as . It is often referred to as relative rotation in the earlier literature. Angular distortion is the rotation given in Figure 2-7a minus the rigid body tilt. Angular distortion is a measure of the shearing distortion of a structure. Of note, Burland et al. (1977) suggested that accounting for tilt in frame buildings on separate footings might be quite inappropriate.

This also agreed with Leonards (1975) who observed that, for frame structures on isolated footings, it was unlikely that each individual footing would rotate through the same angle as the overall structure. Therefore, tilting would contribute to the stresses and strains in the frame and should be included in the distortion calculations for these types of structures. 4. Horizontal Displacement and Horizontal StrainFigure 2-7d gives horizontal displacement and strain as
h,

and h, respectively. These two parameters are typically

associated with excavation-related movements and thus describe the direct lateral movement component of the building. Building response to excavation-related ground movements differs from the building response to ground movements caused by application of the weight of the building, i.e., self-weight settlement. Excavations generate horizontal and vertical ground movements. Thus, excavation-related deformations will induce some direct tensile strains in structures. This is not to say that the horizontal strains in a building adjacent to an excavation equal the associated horizontal ground strains. Horizontal buildings strains do not equal the associated horizontal ground strains in structures where there is significant horizontal stiffness. The stiffness is a result of tensile reinforcement in the foundation system and walls, and rigid floor systems. The substructures of most modern buildings consist of reinforced concrete bearing walls or reinforced strip footings and grade beams. In addition, the floors of most modern buildings are laterally stiff relative to the other structural members. 2.3 PREVIOUS STUDIES TO DEFINE LIMITING CRITERIA

2.3.1 Limiting Criteria Based on Self-Weight Settlement Only In the initial studies of building response to ground movements, researchers studied damage caused by differential settlement due to the self-weight of the building only. Table 2-1 summarizes the results of empirical studies that relate building damage to these ground movements. Skempton and MacDonald (1956) reviewed case histories of 98 buildings and observed the onset of damage at various magnitudes of total and differential settlement. The buildings included both steel and reinforced concrete frame structures and structures with load bearing walls. They observed that most of the damage appeared to be in response to distortional deformations. Thus, they selected angular distortion, , as the critical index for building response and established the limiting angular distortion as the distortion at the initiation of visible cracking in a structure. They concluded the following: (i) Cracking of panels in frame buildings or walls in load bearing wall structures was likely to occur if exceeded 1/300, and, (ii) Structural damage to columns and beams was likely if exceeded 1/150. TABLE 2-1. SUMMARY OF EMPIRICAL LIMITING CRITERIA

Damage Description Safe limit against cracking Cracking of panels in frame buildings or walls in load bearing wall structures

Limiting Distortion, 1/500 1/300

Source Skempton and MacDonald (1956) Skempton and MacDonald (1956)

Structural damage in columns and beams Cracking in load bearing walls or continuous brick cladding Cracking of infilled frames Cracking in beams and columns of frame structures For steel and reinforced concrete frame structures (cracking of infill) For end rows of columns with brick cladding For structures where auxiliary strain does not arise during non-uniform settlement of foundations Tilt of rigid structures (smokestacks, towers, silos, etc.) Slope of crane way, as well as tracks for bridge crane truck Danger to machinery sensitive to settlement Danger for frames with diagonals Safe limit for buildings where cracking is not permissible First cracking in panel walls is to be expected Difficulties with overhead cranes are to be expected Tilting of high, rigid buildings becomes visible Considerable cracking in panel walls and brick walls Safe limit for flexible walls, L/H > 4 Structural damage of general buildings is to be feared Safe limit for hogging of unreinforced load-bearing

1/150 1/1000 1/500 1/250 1/500 1/1000

Skempton and MacDonald (1956) Meyerhof (1956) Meyerhof (1956) Meyerhof (1956) Polshin and Tokar (1957) Polshin and Tokar (1957)

1/200

Polshin and Tokar (1957)

1/250 1/300 1/750 1/600 1/500 1/300 1/300 1/250 1/150 1/150 1/150 1/2000

Polshin and Tokar (1957) Polshin and Tokar (1957) Bjerrum (1963) Bjerrum (1963) Bjerrum (1963 Bjerrum (1963) Bjerrum (1963) Bjerrum (1963) Bjerrum (1963) Bjerrum (1963) Bjerrum (1963) Meyerhof (1982)

walls Safe limit for sagging of unreinforced load-bearing walls 1/1000 Meyerhof (1982)

Limiting to less than 1/500 would provide a factor of safety against cracking. In a discussion of the Skempton and MacDonald (1956) paper, Meyerhof (1956) suggested more stringent criteria and made a distinction between load bearing wall structures and frame structures. Meyerhofs (1956) recommendations to preclude damage were to limit angular distortion to 1/1000 for load bearing walls, 1/500 for panel walls of brick and similar unit masonry (infill frames), and 1/250 for beams and columns of frames. Meyerhof (1953) performed laboratory experiments on full-size brick bearing walls and infill frame walls and reported observations of tensile stress, deflection ratio, and angular distortion at the onset of cracking. Although he suggested some permissible values, Polshin and Tokar (1957) are credited for introducing deflection ratio, /L, as an index to establish limiting criteria. Polshin and Tokar (1957) correlated the deflection ratio to the onset of damage in structures with varying length to height (L/H) ratios. They observed that cracks in masonry bearing walls typically occurred after the tensile capacity of the material had been exceeded and concluded that the maximum allowable deflection ratio was a function of the development of a critical value of tensile strain in the wall. For brick walls, the critical tensile strain was observed to be 0.05 percent. Polshin and Tokar (1957) also used the slope of the settlement trough (rotation) as an index to relate damage due to settlement in frame structures. Later, Bjerrum (1963) presented data relating angular distortion to building performance based on additional

data and the Skempton and MacDonald (1956) data. These data provided the framework wherein damage severity could be categorized as a function of angular distortions. Angular distortion and relative deflection have been used to define limiting conditions in these empirical studies. It is noted that the table includes later recommendations from Meyerhof (1982), which provide safe limits for both the sagging and hogging modes of deflection for the load bearing wall structures. The termed hogging refers to concave downward deflection profiles and sagging refers to concave upward deflection profiles. Burland and Wroth (1974) established limiting criteria by applying the Polshin and Tokar (1957) concept that visible cracking begins once a critical value of tensile strain is reached. Burland and Wroth (1974) assumed that a building could be represented approximately as an elastic deep beam. Figure 2-8 shows the deep beam approximation applied to a building. Using beam-bending theory and assuming a centrally loaded beam, Burland and Wroth (1974) developed limiting deflection ratios of masonry and brick walls of varying L/H ratios and building stiffness. The limiting deflection ratios corresponded to critical tensile strains resulting from bending and shearing deformations. Burland and Wroth (1974) considered the effects of a sagging deflection profile on a building by assuming the neutral axis of the deep beam model was located at the center of the building. The effects of a hogging deflection profile were considered by assuming the neutral axis was located at the bottom of the building.

Figure 2-8. Elastic Deep Beam Approximation (after Burland and Wroth, 1975) For load bearing walls with little to no tensile reinforcement undergoing a sagging deflection profile, Burland and Wroth (1974) presented the following equation that relates deflection ratio, /L, to the maximum bending strain, b(max):

(2.6) For load bearing walls with little to no tensile reinforcement undergoing a hogging deflection profile, Burland and Wroth(1974) presented the following relation

(2.7)

For framed buildings and reinforced load bearing walls, Burland and Wroth (1974) presented the following equation that relates deflection ratio to the maximum diagonal strain,
D(max):

(2.8) Equation 2.8 assumes the neutral axis of the deep beam model is located at the bottom of the building. This assumption was made in an effort to approximate real structures where the foundation and soil would offer considerable restraint. They concluded that structures with relatively low stiffness in shear or a significant degree of tensile restraint, cracking due to diagonal tensile strain occurring at the neutral axis would be the limiting factor. Burland and Wroth (1974) found that the critical tensile strains ranged from 0.5 percent to 0.1 percent for brick walls and masonry. They used 0.075 percent, the average of that range, as input for the strain terms in Equation 2.2 to 2.8. From these equations, they

established limiting criteria for cracking in brick walls and brick infill frames for various L/H ratios. Note that self-weight movements of buildings generally result in sagging profiles, whereas excavation-induced movements can result in sagging, hogging, or both, of an adjacent building. Consideration of the type of settlement profile typically leads to more restrictive limits in a structure, especially for a building that hogs.

2.3.2 Limiting Criteria Based on Excavation-Induced Distortions Boscardin and Cording (1989) extended the concepts of Burland and Wroth (1974) to develop limiting deformation criteria for buildings adjacent to excavations by explicitly considering the effects of horizontal strains associated with excavation-induced movements. Figure 2-9 presents a summary of these efforts.

Figure 2-9. Relationship of Damage to Angular Distortion and Horizontal Extension Strain (after Boscardin and Cording, 1989) The figure presents a plot of horizontal strain versus angular distortion. The plot is divided into categories of potential damage to buildings. The categories are established by theoretical considerations of structural response to deformation, field observations of building damage, and measurements of horizontal and vertical differential displacements associated with the damage. Also included in Figure 2-9 are a limited number of case histories involving damage. The cases were used to verify the validity of using the figure as a basis for limiting criteria. The categories of damage in the figure

are based on the damage classification criteria presented by Burland et al. (1977), as given in Table 2-2. In Figure 2-9, each curve represents a given value of critical tensile strain. Polshin and Tokar (1957) initially defined critical tensile strain as the strain at the onset of visible cracking, which is considered the start of observable damage. Burland and Wroth (1974) modified that definition by expressing critical tensile strain as either maximum bending strain or maximum diagonal tensile strain. Boscardin and Cording (1989) extended the Burland and Wroth (1974) definition by defining the critical tensile strain as the sum of the maximum extreme fiber strain TABLE 2-2. CLASSIFICATION OF VISIBLE DAMAGE (AFTER BURLAND ET AL., 1977

Damage Category Negligible Hairline Crack Very Slight

Description of Damage

Approximate Crack Width < 0.1 mm 1 mm

Fine cracks which can easily be treated during normal decoration. Perhaps isolated slight fracture in building. Cracks in exterior brickwork visible on close inspection Cracks that can be easily filled. Redecoration probably required. Several slight fractures showing inside building. Cracks are visible externally. Some repointing may be required for watertightness. Doors and windows may stick slightly.

Slight

5 mm

Cracks may require cutting out and patching. Suitable linings can mask recurrent cracks. Repointing of external Moderate brickwork and possibly a small amount of brickwork to be replaced. Doors and windows sticking. Service pipes may be fracture. Weathertightness often impaired. Severe Extensive repair involving removal and replacement of sections of wall, especially over doors and windows. Windows and door frames distorted, floor slopes noticeably. Walls lean or bulge noticeably, some loss of

5 mm to 15 mm or several cracks > 3 mm 15 mm to 25 mm, depends on number of cracks

bearing in beams. Utility service disrupted. Very Severe Major repair required involving partial or complete reconstruction. Beams lose bearing; walls lean badly and require shoring. Windows broken by distortion. Danger of instability. Usually > 25 mm, depends on number of cracks

and the horizontal strain. From Figure 2-9, it is noted that when angular distortions are equaled to zero, the horizontal strains equal the tensile strains. When the horizontal strain equals zero, the tensile strain equals the diagonal tension strains. The critical tensile strains corresponding to the boundaries of the zone labeled very slight were taken as 0.0005 and 0.00075. This is the threshold for cracking to first become noticeable. The upper bound of the zone in which damage is considered slight was established at a critical tensile strain of 0.0015. This corresponds to an angular distortion of 1/300 for a horizontal strain of zero. This value of angular distortion is the threshold for first cracking in panel wall and load-bearing walls for structures settling under their own weight as reported by Skempton and MacDonald (1956). The upper bound of the zone in which the damage is considered moderate to severe was established at a critical tensile strain of 0.0030, which corresponds to an angular distortion of 1/150 for a horizontal strain of zero. An angular distortion of 1/150 corresponds to the threshold angular distortion for severe cracking and structural damage in structures settling under their own weight, also given by Skempton and MacDonald (1956). Boscardin and Cording (1989) concluded: 1. Buildings sited adjacent to excavations are generally less tolerant to excavationinduced differential settlements than similar structures settling under their own weight.

They attributed this to the lateral strains that develop in response to most excavations. These strains add to the strains imposed by the vertical movements associated with the excavation. 2. Angular distortion is an appropriate parameter to correlate with observed response. 3. If reasonable estimates of critical tensile strain and L/H can be made for a structure adjacent to the proposed excavation, then the limiting deflection ratio and angular distortion for that structure can be estimated and compared to the estimated ground movement. This will allow the engineer to assess the potential for damage and suitability of possible remedial measures. 4. The magnitude of angular distortion associated with critical tensile strains and the effect of the horizontal strains is a function of the building type and the lateral stiffness of a structure. Frame-type structures, depending on geometry and number of stories, can often resist some ground movements better than masonry bearing-wall buildings. Conversely, a frame structure would be affected more by horizontal ground strains than a structure with reinforced concrete walls supported by continuous footings or with stiff floor systems. Boone (1996) proposed an approach, summarized in Figure 2-10, to assess damage that considers flexural and shear stiffness of building sections, the nature of the ground movement profile, location of the building within this profile, degree of slip between the ground and the foundations, and building configuration. He used cumulative crack width as an indicator of damage severity, and defined severity in terms of tensile strains from bending, elongation of the ground and direct lateral extension. The basis of the equations in Figure 2-10 is the assumption that the building wall deforms as a simply

supported, uniformly loaded, deep beam. On that basis, component tensile strains and shear strains can be derived for a wall, using the relative rotations at the ends. These component strains are then used to obtain principal strains. Crack widths are estimated from the principle strains and compared to damage severity levels Boone (1996) developed from 20 case histories.

Figure 2-10. Estimation of Crack Width and Building Damage (after Boone, 1996)

The following procedures are employed for Boones (1996) approach: 1. Calculate settlement and horizontal movement of an infill wall at the end of the wall closest to the excavation (S1 and h1, respectively) and at the end of the wall furthest from the excavation (S1 and h1, respectively). Also, calculate the slope (g) of the settlement trough beneath the infill wall. In the absence of measured data, these values can be obtained from the maximum settlement at the excavation edge (Smax), the distance to the point where settlement/lateral movement is zero (Dmax), the maximum horizontal movement (hmax), and the distance each end is from the excavation (D1 and D2). 2. Measure or calculate the rigid body tilt (t) of the infill wall and obtain the rotation slope (v), which is equivalent to angular distortion 3. Determine the proportion of deformation due to moment (v(M) and the proportion of deformation due to shear (vV). Note, substitute (vMax(m)) and (vMax(v)) for an infill wall that is relatively flexible in shear (i.e. the presence of openings). 4. Calculate the radius of bending (Rm), the bending strain at the top of the wall ( M), the lateral extension strain
le

), and the shear strain ( ).

5. Calculate the cumulative maximum tensile strain along the top of the infill wall ( t) and the principal tensile strain ( p) using the previously calculated strain components. 6. Cumulative tension crack width (Ct) and Cumulative diagonal crack width (Cp) are calculated and plotted in the graph given in Figure 2-10. From this information, potential damage severity is estimated for the infill walls.

As a point of clarification, critical tensile strains represent the tensile strain at which cracking becomes evident. The critical tensile strains often given in the literature (Boscardin and Cording, 1989; Burland and Wroth, 1974; and Polshin and Tokar, 1957) are given in terms of bending and diagonal strains for buildings settling due to selfweight only (Burland and Wroth, 1974) and given in terms of either bending strains plus direct horizontal strain or diagonal strain plus direct horizontal strain for buildings adjacent to open excavations (Boscardin and Cording, 1989). Critical cracking strain has also been given in terms of shear strain (Bozozuk, 1962, Mainstone and Weeks, 1970; and Mainstone, 1971) and in terms of principal tensile strains (Boone et al., 1999 and Mainstone, 1974). Principal tensile strain only equals the maximum bending strain at the edge of the tensile fiber and only equals diagonal strain at the neutral axis or for shear only conditions. Burland and Wroth (1974) concluded that the critical tensile strain for brickwork and other masonry infill frames using the maximum bending and direct tension strains ranged from 0.0005 to 0.001 with the recommended value being 0.00075. Bozozuk (1962) reported that critical shear strains for concrete and masonry structures ranged from 0.001 to 0.002. Mainstone (1974) gave the range of critical principle tensile strains for full-scale frames with brick infills as 0.00015 to 0.0003. Thus, a range of critical tensile strain criteria has been reported in the literature.

2.3.3 Summary The implication from the previous work related to excavation-induced damage to buildings is that limiting criteria based on self-weight settlement are inadequate for precluding damage resulting from excavation-related movements. One possible explanation for this inadequacy is that tensile strains develop in structural members as a

result of self-weight settlement. These tensile strains become locked in and become the pre-existing strain condition at the beginning of excavation. Excavation-related movements result in additional tensile strains. The excavation-related tensile strains add to the pre-existing tensile strains to cause damage in adjacent buildings at much lower distortions than may be found in literature (i.e. Skempton and MacDonald, 1956; Meyerhof, 1956; Polshin and Tokar, 1957; Bjerrum, 1963; and Burland and Wroth, 1974). Boscardin and Cording (1989) and Boone (1996) have both recognized that excavation-related limiting criteria is a function of building type and orientation with respect to the excavation, type of support system, excavation techniques, and soil conditions.

CHAPTER 3

CHICAGO AVENUE AND STATE STREET SUBWAY RENOVATION PROJECT OVERVIEW 3.1 GENERAL PROJECT DESCRIPTION This chapter presents support system performance observations and building response data collected during the excavation and construction the Chicago Avenue and State Street Subway renovation project in Chicago, IL. The Chicago Department of Transportation (CDOT) authorized the renovation and expansion of the existing subway station. The renovations included adding a mezzanine section for office and vendor spaces and making the station handicap accessible. The station renovations began in June 1999 and were completed by using the cut and cover technique of construction. The construction included excavating 12.2 m of soft to medium clay to expose the existing subway station and tunnels and to allow for capital improvements. CDOT contracted with Wiss, Janney, Elstner Associates, Inc. (WJE) to monitor ground and structural movements associated with the excavation and the subsequent renovations, to monitor the vertical movements of the adjacent school, and to assess the potential for structural damage to the adjacent buildings. The adjacent buildings of most concern were the Frances Xavier Warde School and the Holy Name Cathedral. The structural response of the Warde School to the excavation was of particular interest because of its close proximity to the excavation. Northwestern University was subcontracted by WJE to perform the portion of the contract involving monitoring and predicting the ground movements that resulted from the excavation activities.

Figure 3-1 shows the Chicago and State subway renovation project site as viewed from two different directions. Figure 3-2 shows a plan view of the Chicago and State subway renovation project site including instrument locations, temporary wall types and strut locations. The temporary wall support along State Street consisted of tieback anchors and cross-lot braces.

Figure 3-1. Chicago Avenue and State Street Subway Renovation Project

Figure 3-2. Project Site Plan Monitoring and performance data collected during the construction included lateral soil deformations, pore water pressures, vertical building movements and loads in several cross-lot braces and tiebacks. Inclinometers measured lateral movements at five locations around the excavation. Pore pressures were measured using pneumatic piezometers at two locations on opposite sides of the excavation. Data from the inclinometers and piezometers were obtained daily during wall installation and excavation, and at least on a weekly basis after the excavation had reached its final depth. Strut loads were determined from measurements of strain in the struts for three of

the struts. Loads in the tieback anchors were determined using load cells. The load cells were installed on four anchors, two on either side of the excavation. Strain gauge and load cell data were obtained daily during excavation, and weekly after the excavation was completed. Inclination and elongation of the foundation wall were measured using tiltmeters and extensometers, respectively. Data from these devices were obtained daily by remote access during construction. Lastly, building movements were surveyed optically once per week during the project.

3.2 SUBSURFACE CONDITIONS

3.2.1 Geology and Site Stratigraphy The soils in the Chicago area are primarily glacially derived sediments. Four major geologic events contributed to the formation of the subsurface conditions in this area. The geologic events included: (i) Marine sediment depositionthese deposits formed the bedrock in the vast majority of the Chicago area; (ii) A long period of erosion, which occurred prior to the Pleistocene glaciation; (iii) Successive advances and retreats of a continental glacier during the Pleistocene epochthe majority of the area soils were deposited during the Wisconsin stage of this epoch; and (iv) The variation of elevation of the glacial Lake Chicago, which included water levels from 18 m above to 30 m below the current levels of Lake Michigan (about 177 m above mean sea level). Much of the subsoil in the Chicago area was deposited during advances and retreats of the Wisconsin Stage glacier. These subsoils consisted of debris that was aggregated and deposited directly by glacier ice without disaggregation by other agents, such as melt water. The advance and retreat process was marked by terminal moraines, which created

readily identifiable strata. The geologic names of the more prominent strata in order of deposition are the Valparaiso, Tinley, Park Ridge, Deerfield, Blodgett and Highland Park. The two older strata, Valparaiso and Tinley, are comprised of hard clays and silts. The Park Ridge, Deerfield, Blodgett and Highland Park strata are comprised primarily of silty clays. These younger strata tend to be softer and more compressible than the older strata. The varying stiffness of the different strata is attributed to the changes in the weight of the overlying glaciers during deposition. The Park Ridge, Deerfield, and Blodgett strata define the Lake Border morainic system, which partially bounds the Lake Chicago lake plain. The Chicago and State project site is located within the lake plain. The Blodgett, Deerfield and Park Ridge strata are ice margin deposits. Deposition of these strata occurred under water, when Lake Chicago was at its highest elevation of approximately 195 m above mean sea level. Each morainic stratum can be identified by their sediment properties that resulted from deposition in a particular sub-environment. The depositional environment determined the fundamental geotechnical properties of the glacial deposits including texture, particle shape, fabric, density and stress history. Chung and Finno (1992) found that the Blodgett, Deerfield, and Park Ridge strata are distinguished by water content and undrained shear strength and that the geotechnical parameters reflect the variability of the sediment properties. Figure 3-3 shows the geologic subsurface conditions for the Chicago and State project site. The figure presents the geologic units and the generalized site stratigraphy. The elevations in Figure 3-3 are given in terms of Chicago City Datum (CCD). An elevation

of 0 m CCD corresponds to the mean average level of Lake Michigan. The geologic strata encountered at the Chicago and State project site include:

Figure 3-3. Geologic Subsurface Conditions (Roboski, 2001)

The Blodgett stratum is supraglacial in origin, which implies deposition

from the upper surface of the ice. A relatively wide range of water contents and liquid limits characterizes supraglacial deposits. The Blodgett stratum is subdivided into two sub-strata based on the undrained shear strength. The Upper Blodgett stratum consists of a desiccated crust and underlying soft clays with

undrained shear strengths that increase with depth. The desiccated clay crust was formed as a result of a drop in the water level after deposition. The Lower Blodgett stratum consists soft clays with slightly higher undrained shear strengths that continue to increase with depth. The average thickness of the Upper Blodgett stratum is 3.4 m while that of the Lower Blodgett stratum is 2.4 m.

The Deerfield stratum consists of medium stiff clay and is characterized by uniform water contents. The Deerfield stratum is about 3.7 m thick. The excavation bottoms out in the Deerfield stratum at elevation -7.9 m CCD. The Park Ridge stratum is a stiff to very stiff clay with lower water contents than the Deerfield stratum. The Park Ridge stratum is about 4.6 m thick. The Deerfield and Park Ridge strata are either basal melt-out tills or waterlain paratills. The Tinley stratum, a lodgement till, underlies the ice margin deposits and consists of very stiff to hard clays and silts. The hard soils encountered below elevation -18.3 m CCD are known locally as hardpan. The lodgement tills are often characterized as being very dense to hard and relatively incompressible soils.

With the exception of the clay crust in the Upper Blodgett stratum, the Lake Border deposits are lightly overconsolidated as a result of lowered groundwater levels after deposition and aging. From Figure 3-3, it is seen that a fill deposit overlies the glacial clays. The fill is mostly medium dense sand, but also contains occasional construction debris. The fill deposit is approximately 4.3 m thick.

3.2.2 Laboratory and Field Testing Several researchers had developed stratigraphic correlations for the geologic strata in the Chicago area based on index tests and undrained shear strengths (Peck and Reed,

1954; Otto, 1963; and Chung and Finno, 1992). A profile showing the variation of undrained shear strength and index properties with depth is presented in Figure 3-4.

Figure 3-4. Subsurface Profile The field vane data, unconfined compression data, pocket penetrometer data, and much of the index properties data were obtained from samples of the Blodgett, Deerfield, and

Park Ridge strata at or adjacent to the Chicago and State project site. The triaxial test data and some additional index testing data were obtained from test specimens of the geologic strata found at the Chicago and State site. Kawamura (1999) and Roboski (2001) performed extensive laboratory testing on specimens from the Blodgett, Deerfield, and Park Ridge strata including index property, consolidation, and triaxial tests. The index properties included natural water content, Atterberg limits, and specific gravity at various depths. The triaxial testing included drained triaxial compression (TXC) and reduced triaxial extension (RTXE) testing and undrained-consolidated triaxial compression (TXC CIU) testing. Table 3-1 presents the summary of the index tests. Table 3-2 summarizes the consolidation properties. The summary of the drained and undrained triaxial tests is given in Table 3-3. TABLE 3-1. SUMMARY OF AVERAGE INDEX PROPERTIES (AFTER ROBOSKI, 2001)

Stratum Upper Blodgett Deerfield Park Ridge Definitions:

Depth(m) 5.5-8.8 11.3-15 15-19.5

wn (%) 32.9 21.1 16.5

LL (%) 42.0 30.5 24

PL (%) 22.2 16.6 14.3

SG 2.63 2.72 2.76

wn: Natural moisture content LL: Liquid limit PL: Plastic limit SG: Specific gravity

TABLE 3-2. SUMMARY OF CONSOLIDATION PROPERTIES (AFTER ROBOSKI, 2001)

Stratum Upper Blodgett Upper Blodgett Upper Blodgett Lower Blodgett Lower Blodgett Deerfield Deerfield Deerfield Park Ridge Park Ridge Park Ridge Definitions:

Sample Depth (m) Cc S-5 I-2A I-2b S-11 I-4 S-13 I-1 I-2 S-20 I-2a I-2B 5.3-6.1 5.3-6.1 5.3-6.1 10.7-11.4 10.5 12.2-13 12.2-13 12.2-13 18.3-19.1 18.3-19.1 18.3-19.1 0.25 0.28 0.25 0.17 0.23 0.15

Cr 0.04 0.03 0.02 0.03

vo

'(kPa)

'(kPa) OCR eo 1.00 1.06 1.10 1.02 1.0 1.01 1.0 1.2 1.03 1.0 1.0 0.79 0.79 0.79 0.59 0.73 0.59 0.58 0.59 0.43 0.45 0.45

85.2 103.5 129.3 149.4 128.6 200.2 169.5 169.5

85.2 109.5 114.3 132.3 123.8 151.3 123.8 152.4 206.9 100.0 166.7

0.028 103.5 0.038 128.6

0.186 0.031 128.6 0.181 0.02 0.12 0.03 0.107 0.02 0.106 0.02

Cc: Compression index sp: Pre-consolidation pressure Cr: Recompression index OCR: Overconsolidation ratio svo': Effective overburden pressure eo: Initial void ratio TABLE 3-3. SUMMARY OF TRIAXIAL TEST DATA (AFTER ROBOSKI, 2001)

Drained Triaxial Test Data

Stratum Upper Blodgett Upper Blodgett Upper

Depth Sample (m) S-5 S-5 S-4 5.3-6.1 5.3-6.1 4.6-5.3

Winitial Wfinal ( 1TXC/RTXE vc '(KPa) (%) (%) 3)f(KPa) TXC TXC RTXE 200 400 85 32.71 25.52 280.4 34.01 23.06 465.0 32.1 31.3 -68

'o

24.3 23.9 41.8

Blodgett Lower Blodgett Lower Blodgett Lower Blodgett Deerfield Deerfield Deerfield Deerfield Park Ridge Park Ridge Park Ridge Park Ridge S-11 S-11 S-10 S-13 S-13 S-13 S-15 S-18 S-18 S-18 S-19 S-6 10.711.4 10.711.4 TXC TXC 220 400 130 175 350 450 164 200 350 450 193 200 ~22 ~22 23.2 21.03 21.02 21.11 375.0 670.0 27 27 30.7 29.8 28.0 28.6 41 32.3 31.3 30.5 35 32.6

9.9-10.7 RTXE 12.2-13 TXC 12.2-13 TXC 12.2-13 TXC 14.515.2 16.817.4 16.817.4 16.817.4 7.6 RTXE TXC TXC TXC

19.95 -88.0 17.94 341.7 16.47 617.5 16.29 827.0 -130.0

21.39 20.4

15.23 14.08 482.0 15.23 12.89 758.0 15.23 13.67 928.0 20.2 17.54 -141.0 467.0

17.5-18 RTXE TXC

13.06 9.77

Undrained Triaxial Test Data

Stratum Upper Blodgett Lower Blodgett Deerfield Park Ridge

Sample Dept(m) S-5 5.3-6.1 S-11 S-15 S-20 10.711.4 14.515.2 18.319.1 0.34 0.37 0.45 0.54 100 130 168 204

Wintial(%) Wfianl(%) 34.29 23.65 20.4 29.21 19.43 20.3 18.68 67.6 96.0 151.0 221.0

3.2.3 Engineering Properties of the Soil It is apparent from Figure 3-4 that the natural water content of the Deerfield stratum is

relatively constant. The range of values is from roughly 20 percent to about 22 percent. The natural water content of the Blodgett stratum is more erratic and higher than the Deerfield. These values range from about 18 percent to about 32 percent. The erratic behavior of the Blodgett stratum water content values reflects the fact that these sediments resulted from supraglacial deposition. The natural water content continues to decrease with depth below the Deerfield, such that the water content of the Park Ridge stratum varies between 16 percent and 18 percent. Figure 3-4 shows that the natural water contents of the Blodgett stratum are near the liquid limits, suggesting that the layer is normally to lightly overconsolidated. Conversely, the natural water contents of the Deerfield and Park Ridge are closer to the plastic limits, suggesting these layers are slightly more overconsolidated than the Blodgett. From the index properties, the glacial clays at the Chicago and State project site were classified as lean clay (CL) using the United Soil Classification System (USCS). In general, the index properties and the consolidation results presented in Table 3-1 and Table 3-2 reflect that the strata become less compressible as the depth increases. The results presented in Table 3-3 indicate the shear strength increases with depth. The specific trends obtained from the three tables are as follows:

The initial void ratio (eo) decreased with depth from about 0.79 to

about 0.43.

The specific gravity (SG) increased slightly with depth from 2.63 to

2.76.

The maximum preconsolidation stress ( p') increased with depth from

about 85 kPa at 6 m to about 200 kPa at 18 m.

The compression index (Cc) decreased with depth from about 0.25 to

about 0.106.

The TXC drained friction angle ( ') increased with depth from about 24

degrees to about 32 degrees.

The RTXE drained friction angle ( ') decreased with depth from about

42 degrees to about 35 degrees.

The normalized undrained shear strength (Su/

' vc )

increased with depth

from 0.34 to 0.54. It is observed that there are some occasional variances in the overall trend of stiffer and less compressible strata with increasing depth. Roboski (2001) noted that this was possibly due to sample disturbance.

3.3 ADJACENT STRUCTURES The existing structures adjacent to the project site influenced many of the design decisions relating to the selection of the excavation support system. The potential for damaging these structures was estimated based on the proximity of these structures to the excavation, the building type, the foundation system, and the expected magnitude and distribution of the ground movements. The Frances Xavier Warde School was of primary concern due to its proximity to the excavation. The school was built in the late sixties and is a 3-story reinforced concrete frame structure. When viewed in plan, the building has an L shape. The floor system at each level consists of a reinforced concrete pan-joist system, supported by reinforced concrete beams. The beams are supported on concrete columns at interior locations and

masonry bearing walls and columns around the perimeter. The bearing walls rest on a reinforced concrete basement wall. The basement wall is nominally 2.8 m tall and 400 mm thick and is supported by an approximately 1.2 m wide continuous footing. The interior columns are supported on reinforced concrete spread footings that are nominally 760 mm thick and vary in size from 4 m by 4 m to 5 m by 5 m. The average depth bottom of the footings is approximately 3.7 m below ground surface. The building faade on the north, south, and west elevations of the building consists of windows and coarse limestone with a concrete masonry unit (CMU) backup wall. The east elevation of the building consist of a Chicago common brick veneer wall and a CMU backup wall. The school was located approximately 2 m from the excavation. The proximity of the school relative to the excavation is shown in Figure 3-2. Also adjacent to the excavation, is the 125-year-old Holy Name Cathedral. In plan view, the cathedral is a cruciform. The structure is primarily a masonry bearing wall system with the exception of the central spire, which is supported on drill piers that extend to bedrock. The faade of the cathedral consists of limestone masonry and ornate masonry features. The steep roofs are cover with slate shingles. The cathedral was located approximately 15 m southeast of the excavation. Although of initial concern, measurements recorded throughout the project indicated that the excavation-related deformations at the cathedral were insignificant. The Chicago and State Street subway tunnel and station were the belowground structures adjacent to the excavation. The subway station and tunnel were constructed between 1939 and 1941. Excavation was performed using the liner-plate tunneling method. The tunnel consists of twin subway tubes and passenger platforms and is

symmetrical about its centerline. The tunnel travels in the north and south directions. Each tube is approximately 5 m wide and 6 m tall in the interior and each passenger platform is 2 m wide and 5 m tall in the interior. The bottom elevation of the tunnel is 9 m CCD. The tunnel was located approximately 4.5 m west of the Warde School. The existing subway tunnel increased the overall stability of the excavation because of its mass and stiffness.

3.4 EXCAVATION SUPPORT SYSTEM DESIGN METHODOLOGY The principal focus of the excavation support system design was to protect the Warde School from excavation-related damage. Given the proximity of the school to the excavation, a major design issue was the movement associated with construction. The continuous wall footing, along the west side of the Warde School, was about 1.5 m from the centerline of the secant pile wall. Because of this close proximity, it would have been extremely difficult to perform the excavation without causing any damage to the school. Thus during the design phase of the project, the owner of the school and the designers agreed that the support system would prevent structural damage to the Warde School while permitting minor architectural damage. The intent was that the architectural damage would be repaired at the end of the project. It became apparent that a stiff excavation support system was needed given the soft clays at the site and the requirement of minimizing movements. A complicating factor faced by the designers was that the CDOT contractually dictates that all temporary support is the responsibility of the contractor. Consequently, the designers were reluctant to specify excavation procedures and design details in the bid documents. Rather, the bid documents contained a performance specification, a

conceptual design, a lateral earth pressure loading diagram, and several deformation limits that, if exceeded, would trigger specific responses from the contractor. The deformation limits were defined by: 1. Computing maximum horizontal wall deformation using the charts developed by Clough et al. (1989). This method relates system stiffness (defined as EI/h4
w

where EI

is the bending stiffness of the wall, h is the average spacing between support levels and
w

is the unit weight of water) and factor of safety against basal heave to the maximum

horizontal wall movement. This value was reduced to account for the 3-dimensional nature of the excavation along State Street using recommendations of Ou et al. (1996); 2. Assuming the maximum horizontal movement equals the maximum settlement and computing the vertical settlement distribution behind the wall by the procedure recommended by Hsieh and Ou (1998), and; 3. Using the expected settlement profile as input to a finite element model of the Warde School to evaluate the extent of damage to the building. The potential extent of damage was defined and evaluated in terms of the computed tensile stresses. The potential damage obtained from the finite element simulation was compared to the range of damage deemed acceptable by the owner. If the potential damage exceeded the acceptable range, a stiffer wall support system was used and the evaluation process was re-started at step 1. The process was iterated until the estimated level of damage was within an acceptable range. Deformation action limits and corresponding actions were established for the project based on the results of this process and the condition survey

of the building. These limits and corresponding actions are given in Table 3-4. The movements shown in the table could be either lateral movement or settlement. A number of aspects of this project were not specifically considered when predicting the maximum horizontal movement, and therefore created uncertainties in the prediction. These aspects and their effect on the predicted movements are summarized in Table 3-5.

Table 3-4. Project Performance Specifications Related to Movements

Horizontal or Action required Vertical Movement 19 mm 25 mm 32 mm Contractor and owner notified Contractor and owner notified and site meeting followed up by written report and recommendations Contractor and owner notified, immediate inspection of structure, and contractor with owner's approval to initiate probable remedial measures

Table 3-5. Project Details and Expected Effects on Predicted Movements Project Detail Presence of tunnel in center of evacation Expected Effects on Predicted Movement Reduce predicted movements because of stabilizing weight in center of excavation Reduce predicted movements because of the smaller load to support, i.e., school weighs less than excavated soil Increase predicted movements because this source of movement not considered Increase predicted movements because this source of movement not considered

Effects of construction of Warde Scholl

Installing tiebacks through soft clay under Warde School

Secant pile wall installation; details of installation procedure not specified to contractor

Three-dimensional geometry resulting from simultaneous excavation along Chicago Ave. When the Increase predicted movements in movement predictions were made, the Chicago Ave. northwest corner of school excavation was not part of the main excavation Effects of tunnel installation on in situ soil stresses, i.e., Unknown no known conditions prior to start of excavation While the effect of some of these aspects could have been evaluated with a finite element analyses, time constraints precluded such analyses. Furthermore, even if such analyses had been conducted, significant uncertainty would have remained. No guidance could be found from local practice because this was the first time this type of support system was used in the soft Chicago clays. These uncertainties and the lack of local experience were the reason why the instrumentation program was implemented.

3.5 EXCAVATION SUPPORT SYSTEM DESCRIPTION The primary excavation support system used for the project consisted of a secant pile wall with three levels of support. Figure 3-5 presents a section view of the excavation support system. The plan view showing the excavation support system was presented previously in Figure 3-2. The wall consisted of overlapping 915-mm-diameter drilled shafts installed along the east and west sides of the excavation. Each shaft nominally overlapped adjacent shafts by 150 mm. A W24x55 section was placed in alternating shafts to provide additional stiffness to the wall.

Figure 3-5. Section View of Excavation Support System The top level of supports consisted of cross-lot pipe struts, which were shimmed against the upper waler. The steel pipe struts had a diameter of 610 mm and a nominal wall thickness of 17 mm. The center-to-center horizontal spacing between the struts was approximately 6.1 m. Tieback anchors were used for the second and third levels of support. The regroutable tiebacks were bundled steel tendons. The steel tendons consisted of four to five, 15-mm, 1860 MPa steel strands. The 150-mm-diameter tieback anchors were installed at a 1.5-m center-to-center spacing. They were installed at a 45degree angle and had bond lengths of 9.1 m to 10.7 m. Unbonded lengths were at least 9.1 m. The bond zone was located in the stiff clay below elevation -12.5 m CCD. The bond length was estimated using an allowable interface strength of 125 kPa.

The portion of the excavation north of the Warde School was braced with two levels of struts. The upper level struts consisted of HP10x42 sections. The lower level consisted of HP12x53 sections. A soldier pile and lagging wall was placed along the north and south ends of the excavation. The east-west cross-section through the excavation given in Figure 3-5 shows why the combined support was required at this site. The Warde School had a 3-m deep basement that precluded using tiebacks for the first level. Also, there was the need to place the first brace near the top of the wall to minimize movements. The presence of the tunnel precluded using cross-lot supports for the second and third levels. Tiebacks in the lower level were installed with the drilling equipment placed atop the tunnel, 3 m away from the face of the wall.

3.6 FIELD INSTRUMENTATION The field instrumentation was divided into three groups; ground instrumentation, structural support instrumentation and instrumentation for the adjacent structure. The ground instrumentation included inclinometers and piezometers. The structural support instrumentation included load cells and strain gauges. Optical survey points, string pot extensometers, tiltmeters, and crack gauges were used to monitor the response of the Frances Xavier Warde School to the construction activities. The ground instrumentation was installed and initialized prior to any construction activities at the site. Data from the ground instrumentation were obtained daily during the wall installation and excavation activities and at least weekly during the station renovation and backfill activities. The excavation support system instrumentation was installed as soon as the particular

structural element was placed. Data were collected from these instruments on the same schedule as that of the ground instrumentation. The adjacent structure instrumentation was also installed and initialized prior to any construction activities. Optical surveys were made at least weekly during the project. The tiltmeters and extensometers were queried several times per day by remote access throughout the project. The crack gauge data were collect weekly after the onset of damage was observed.

3.6.1 Ground Instrumentation Lateral movements of the soil behind the secant pile wall were recorded using five inclinometers located around the excavation site. The locations of the five inclinometers relative to the site are shown in Figure 3-2. Lateral ground movements were monitored around the Warde School with Inclinometer 1 and Inclinometers 2 placed along the east side of the excavation and Inclinometer 5 placed around the north side of the school. The free field lateral movements were measured with Inclinometer 4 placed on the west side of the excavation. Inclinometer 3 was located on the southwest corner of the school. This inclinometer was installed to monitor deformations that might affect the cathedral. Inclinometer 1, Inclinometer 2, and Inclinometer 3 were all installed to depth of 22.5 m (elevation -18.3 m CCD). Inclinometer 5 and Inclinometer 4 were both installed to depth of 24.4 m (elevation -20.1 m CCD). For all inclinometers, the primary direction of inclination was towards the adjacent excavation and the secondary direction was 90 degrees to the primary direction. According to the manufacturer, the accuracy of the inclinometer measurements was 6 mm per 25 m of casing.

Figure 3-6. Subsurface Location of Piezometers Pore pressures were measured using pneumatic piezometers at two locations. The piezometers were placed at opposite sides of the excavation, adjacent to Inclinometer 1 and Inclinometer 4. The piezometers were installed in nested pairs. At both locations,

the two piezometers were separated vertically by about 3 m. Figure 3-6 presents the subsurface locations of the piezometers. The piezometers adjacent to Inclinometer 1 were installed at nominal elevations of -5.8 m CCD and -8.8 m CCD within the soft to medium clay. The piezometers adjacent to Inclinometer 4 were installed at nominal elevations of -6.4 m CCD and -9.4 m CCD. Note the elevation of the bottom of the excavation was at -8 m CCD.

3.6.2 Structural Support Instrumentation Strut loads were determined from measurements of strain in the struts. Vibrating wire strain gages were attached to Struts 3, 4, and 5. Three strain gauges were attached to each strut at 120-degree intervals around the circumference. The strain gauges on Struts 3 and 4 were placed at a distance from the east end of approximately 2/5 the length of the strut. The gauges on Strut 5 were placed in the middle of the strut. Temperature corrections were applied to all strain measurements. Loads in the tieback anchors were determined using hollow-core load cells mounted directly onto the tieback anchors. Four load cells were installed at two separate locations along the east wall. For both locations, one load cell was install on an upper level tieback and one load cell was installed on a lower level tieback. The locations of the load cells corresponded to Strut 3 and Strut 4. The locations of Struts 3 and 4 relative to the excavation site are shown in Figure 3-2. The load cells measured the forces in the tieback anchors directly.

3.6.3 Adjacent Structure Instrumentation

Figure 3-7 denotes the locations of the instrumentation used to monitor the response of the Warde School.

Figure 3-7. Instrumentation Location Plan Settlement was monitored by optical survey points established on the interior columns, on the roof, and along several locations on the exterior walls of the Warde School. The interior column survey points were all placed in the basement level. There was also

several survey points located on top of the secant pile wall. The survey was conducted to an accuracy of about 3 mm. Tiltmeters were installed along the perimeter walls in the basement level of the Warde School to measured inclination of the foundation wall directly. The tiltmeters were mounted on baseplates that were attached to the wall and were typically placed about 1.5 m above the basement floor. Tiltmeters T1, T2, T3, and T6 measured tilt in the north-south direction. Tiltmeters T4, T5, T7, T8, and T9 measured tilt in the east-west direction. The figure also shows that extensometers were mounted along the wall in the basement level. These devices were string pot potentiometers mounted on baseplates and measured length changes of the foundation wall directly. Extensometers S5 and S6 were horizontal to the wall and Extensometers S1 and S2 and, S3 and S4 were placed diagonally along the wall. Measurements obtained from the extensometers were not corrected for temperature effects and were exposed to variations in the temperature during the duration of the project. Consequently, accurate estimations of cumulative elongation or shortening of the basement foundation walls could not be obtained. Although there were occasional spikes in the data, estimates of the incremental changes in length of the basement wall were negligible. The extensometer data was ultimately not use for any analyses because the uncertainty in the quality of the data. A crack gauge was used to monitor a diagonal wall crack located on the south wall of a third level classroom. The classroom was located at about the center of the building, just

north of the C5 location (refer to Figure 3-7). The crack gauge monitored the change in width of the crack in both the horizontal and vertical directions

3.7 SUMMARY The excavation along State Street was approximately 40 m long and 24 m wide and was advanced to an average final depth of 12.2 m (elevation -7.9 m CCD). The excavation along Chicago Avenue was approximately 24 m long and 7 m wide and was advanced to a depth of 8.2 m (elevation -4 m CCD). The secant pile wall was approximately 18.3 m deep and was constructed by drilling overlapping 915-mm diameter shafts. W24x55 sections were placed in alternating shafts. Center-to-center spacing of each shaft was 750 mm. The pipe struts were installed without preload at a depth of 0.6 m below ground surface. The tiebacks were installed at a 45-degree angle, down to the stiff and hard clay. Only about one-half of the anchors on either side were regrouted. The minimum bond length for the east side tiebacks was 9.1 m. The minimum bond length for the west side tiebacks was 10.7 m. No tiebacks were used for the support system along Chicago Avenue and the secant pile wall system was only used for about half the south wall. Soldier piles and lagging were used for the rest of the wall at this location. Two levels of cross-lot struts were installed without preload along Chicago Avenue at depths of 0.9 m and 4.3 m. The soils at the Chicago and State project site are primarily lightly overconsolidated glacial clays. The clay layers are distinguished by water content and undrained shear strength, with the strata becoming stiffer and stronger as depth increases. The Frances Xavier Warde School was located approximately 2 m from the excavation and was

instrumented with settlement points located on interior basement columns and along the exterior foundation walls. Lateral movements of the soil behind the secant pile wall were recorded using two inclinometers located on the west side of the school, one inclinometer located on the north side of the school, one inclinometer located on the west side of the excavation and one inclinometer located at the southeast corner of the excavation. Pore pressures were measured using pneumatic piezometers placed at opposite sides of the excavation. Strut loads were determined from vibrating wire strain gages attached to Struts 3, 4, and 5. Loads in tiebacks were determined from load cells mounted on both upper and lower level tiebacks at the Strut 3 and Strut 4 locations.

CHAPTER 4

PERFORMANCE AND OBSERVED BEHAVIOR OF FIELD INSTRUMENTATION The excavation for the Chicago and State subway renovation project consisted of two areas of excavation. The primary excavation area was west of the Warde School, along State Street. This portion of the project involved constructing a mezzanine section atop the subway tunnel and making the subway handicap accessible by adding escalators and elevators. Figure 4-1 presents the subway renovation structural plan and includes the numbering scheme for the secant piles. The figure also shows the locations of the inclinometers with respect to the excavation.

Figure 4-1. Chicago Avenue and State Street Subway Renovation Structural Plan The excavation area west of the Warde School was approximately 40 m long and about 24 m wide. At the center of the west side excavation area, the excavation was advanced to the top of the subway tunnel. The top of the subway tunnel was about 6.2 m below the ground surface (elevation -1.9 m CCD). Along the sides of the west excavation area, on either side of the tunnel, the excavation was advanced to a depth of 12.2 m (elevation -7.9 m CCD) everywhere along the length of the tunnel except at the locations of the

escalators. At the escalator locations, the excavation was advanced to a depth of 13.7 m (elevation -9.4 m CCD). The secondary excavation was located on the north side of the Warde School, along Chicago Avenue. The purpose of this excavation activity was to add new stairs and an escalator to the existing station. The secondary excavation area was approximately 24 m long and 7 m wide. The excavation was advanced to a depth of 8.2 m (elevation -4 m CCD) at this location. Although the excavation activities were relatively uniform as the project progressed from the south end to north end along State Street, different deformation behavior of the ground was noted between the east and west sides of the primary excavation area. To provide a complete picture of the observed responses, representative observations are presented for the east and west sides of the State Street excavation and the south side of the Chicago Avenue excavation.

4.1 CONSTRUCTION SEQUENCE AND PROCEDURES Construction Day 0 corresponds to the beginning of the secant pile wall installation at the project site. The construction days correspond to calendar days and advance sequentially until the end of construction. The construction days include weekends and holidays. The construction activities at the site are separated into three distinct stages; (i) Stage 1Wall Installation, (ii) Stage 2Support System Installation and Excavation, and (iii) Stage 3Station Renovation and Backfill. The actual beginning and duration of each stage differs depending on the location within the excavation. However, general

groupings of construction activities are made based on construction sequence along the (i) east side of the primary excavation area, (ii) west side of the primary excavation area, and (iii) the south side of the secondary excavation area. The time history of the significant excavation and construction activities that occurred at the site is presented in Table 4-1. The overall construction activity time history for the project and a structural plan sheet for the station renovations are given as Appendix A.Table 4-1. Days of Significant Excavation and Construction Activity

Construction Activity

Day Day - INCL 4 INCL 1 Stage 1: Wall Installation 2 11 74 81 87 102 105 116 137140 172177 225310 258 15 18 74 81 91 98 110 156

Day - INCL 5

Beginning of secant pile wall installation End of secant pile wall installation End of strut installation Excavate below upper level tieback/second level struts Tension upper level tiebacks/Install second level struts Excavate below lower level tieback Tension the lower level tiebacks Excavate to final grade Reduce the bending stiffness of the secant pile wall Place concrete for escalator bottom slabs Place backfill Remove struts
1

70 79 128 1961 2061 N/A N/A 208

Stage 2: Excavation and Support

Stage 3: Station Renovation and Backfill 117-134 156-163 225-310 258 199 210 248-310 235

Note: Tiebacks were not used along the north side of the school. Day corresponds to second level struts Tiebacks

4.1.1 Construction Along the East Side of State Street The secant pile wall was installed along the entire east side of the Warde School between Day 0 and Day 30. However, secant piles directly adjacent to Inclinometer 1 and Inclinometer 2 were installed between Day 0 and Day 11. The locations of the inclinometers relative to the excavation are shown in Figure 4-1. The secant pile wall was approximately 18.3 m deep and was constructed by drilling overlapping 915-mm diameter shafts. Center-to-center spacing of each shaft was 750 mm. Construction of the wall began by first drilling primary shafts. These shafts were located 1.5 m apart and were constructed by first auguring down to the clay crust and installing temporary steel casing. The steel casing was slightly oversized and provided support for the sand fill during shaft construction. The remainder of each shaft was typically drilled uncased. Once the primary shafts were drilled, W24x55 sections were placed in the holes and a neat concrete grout was end dumped from concrete trucks. The grout typically was placed within 2 hours of completing the hole, but occasionally longer intervals resulted from delays in grout delivery. Secondary shafts were installed between the primary shafts by auguring through green grout to provide the 150-mm overlap. Green grout is the condition of the grout at which it has hardened enough to provide adequate bending stiffness, yet weak enough so that the edges of the primary shafts can be removed using conventional drilled shaft equipment. Typically, the secondary shafts were installed within 24 hours of installing the primary shafts. The design 28-day compressive strength of the grout was 6.9 MPa. Significant excavation in a given area did not begin until at least 28 days after the wall in that area had been installed.

The excavation activities along the east side of State Street occurred between Day 28 and Day 163. Figure 4-2 shows the profile of the east wall excavation face. The pipe struts were installed without preload at a depth of 0.6 m below ground surface. Pipe Struts 1 and 2 were installed on Day 56 and Day 58, respectively, when the maximum depth of the excavation was approximately 3.7 m beneath the struts.

Figure 4-2. Profile of East Wall Excavation Face The maximum excavated grade at the strut locations was an approximate depth of 4 m during the installation of Strut 3, Strut 4, and Strut 5. Strut 6 was installed when the excavated grade at that location reached approximately 4.5 m. Both levels of tiebacks were installed with excavated grade no more than 0.6 m below the tieback level. The maximum unsupported height of the wall was 4.5 m and occurred after the installation

of Strut 6 (Day 79) and prior to stressing the upper level tiebacks (Day 81). The excavation was advanced to the bottom depth of 12.2 m at locations adjacent to Inclinometer 1 and Inclinometer 2 on Day 116. The entire east side was advanced to the bottom depth of 12.2 m on Day 149. The pit for Escalator #4 was excavated to a completed depth of 13.7 m on Day 163. The pit for Elevator #3 was also excavated to a final depth 13.7 m and was completed on Day 165. The tiebacks were installed by coring through the secant pile wall and drilling down to the stiff and hard clay (elevation -10.7 m CCD). Each tieback was approximately 30.5 m long. The tieback holes were cased with temporary 178-mm diameter steel casing in areas where the tiebacks passed through the softer clays. The upper level tiebacks were cased for the first 10.7 m of the hole and the lower level tiebacks were cased for the first 4.6 m. The steel casing was advanced using an external water flush. The minimum bond length for the east side tiebacks was 9.1 m. Once installed, each tieback anchor was tested to 1.33 times the design load and the load was locked off at 80 percent of the design load. Along the east side, the design load for the upper and lower level tiebacks were 278 kN/m and 351 kN/m, respectively. Experience gained through performance and proof testing showed the required loads could be attained without regrouting, and consequently only about one-half the anchors were actually regrouted. The base slab between Stair #7 and Escalator #4 was poured between Day 172 and Day 177. Other activities of note along the east side of State Street included pouring the remainder of the pit slab between Escalator #4 and Elevator #3, pouring the base slab for the mezzanine, and pouring the exterior walls of the mezzanine. These activities occurred between Day 178 and Day 224. A summary of the activities that occurred

during this time can be found in Table 4-1. Backfilling along the east side of State Street began on Day 225. Struts 1 and 2 were removed on Day 247. Strut 3 was removed on Day 249 and Struts 4, 5, and 6 were removed on Day 258.

4.1.2 Construction Along the West Side of State Street The secant pile wall along the west side of State Street was installed between Day 0 and Day 37. An abandoned drift tunnel was encountered on Day 9 during the installation of the south end of the west wall. Wall installation activities at location of the drift tunnel were stopped and were not resumed until Day 24. The portion of the west secant pile wall adjacent to Inclinometer 4 was installed between Day 15 and Day 18. The portion of the west secant pile wall involving the drift tunnel was installed between Day 24 and Day 37. The days and the sequences in which all the secant piles were installed along the west side are given in Table 4-1. The construction procedures used for the secant pile wall along the west side of State Street were the same as those used for the east wall. However, the total installation time for a single shaft was typically longer for the west wall than for the east wall. It took as long as 5 hours after drilling a shaft for grout to be placed at some locations along the west wall. Although some minor excavation activities began on the south end of the west wall on Day 19, the excavation along the west side did not begin in earnest until Day 59. The excavation activities along the west side of the primary excavation area were not as consistent as those observed along the east side. Figure 4-3 presents the profile of the west wall excavation face. The average depth of the excavated grade was about 2.6 m

below ground surface at the time Struts 1, 2, and 3 were installed. Strut 4 was installed on Day 60. At this time, the depth of excavation was 2.6 m at the Strut 4 location, but the maximum depth of the excavation along half of the west excavation face was 5.8 m. From Day 60 to about Day 67, soil from various other excavation activities was stockpiled in the northwestern corner of the primary excavation area. About 1/3 of the west side was backfilled to a height of approximately 1.2 m above the original grade. The stockpile was removed on Day 74 and excavation activities along the west side were resumed. Strut 5 and Strut 6 were installed on Day 74 and Day 79, respectively. The unsupported heights of the west secant pile wall were limited to a maximum 4 m during excavation, which occurred at the Strut 5 location.

Figure 4-3. Profile of West Wall Excavation Face

The excavated grade was advanced to depths of approximately 6.1 m and 9.1 m during the installation of the upper and lower level tiebacks, respectively. The upper level tiebacks were located at a depth of 4.6 m and the lower level tiebacks were located at a depth of 8.5 m. The excavated grade was not more than 1.5 m below the tieback levels during the installation. The excavation was advanced to a depth of 11.9 m at the locations near Inclinometer 4 on Day 117. The excavation was advanced to the 11.9-m depth across the entire bottom of the west side of State Street on Day 137 and remained at that depth until Day 156. The excavation was completed to a final depth of 12.2 m on Day 156. Also at this time, the pit for the west side escalator (Escalator #1) was excavated to a completed depth of 13.7 m. The installation procedures used for the tiebacks along the west wall were very similar to those used for the east wall tiebacks. One difference was that the minimum bond length for the west side tiebacks was 10.7 m. Also, the design loads differed from the east side; the upper level tiebacks were designed for 397 kN/m and the lower level tiebacks were designed for 429 kN. The base slab between Stair # 6 and Escalator #1 was poured between Day 156 and Day 163. All subway station renovations along the west side of State Street were completed on Day 224. Backfill activities along the west side commenced on Day 225.

4.1.3 Construction Along Chicago Avenue

The construction procedures along Chicago Avenue on the north side of the Warde School differed from those along State Street. There were no tiebacks used at this location and the secant pile wall system was only used for about half the south wall. Soldier piles and lagging were used for the rest of the wall at this location. A plan view of this configuration is shown in Figure 4-1. The soldier piles were HP10x42 sections placed in 510-mm diameter drilled holes. The center-to-center spacing of the soldier piles was 1.8 m. The combination of secant pile wall and soldier pile and lagging support systems were used because the excavation along the north side of the Warde School was not as deep nor as wide as the excavation along the west side of the school. The presence of the existing subway station provided some additional basal stability to the north side excavation.

Figure 4-4. Profile of West Wall Excavation Face * Elev of excavated surface on Day No.

Figure 4-4 shows the profile of the excavation along the north wall. The excavation at this location occurred between Day 125 and Day 196. Two levels of cross-lot struts were installed without preload at depths of 0.9 m and 4.3 m below ground surface. The upper level struts consisted of HP10x42 sections. The lower struts were HP12x53 sections. The upper level struts were installed on Day 128 and Day 133. The excavation was advanced to a maximum depth of 2.1 m on Day 130 and remained at that depth until Day 137. The excavation was advanced to approximately 8.2 m on Day 191 to accommodate the pit for the north side escalator (Escalator #3). The area west of the escalator pit was lowered to a final depth of 7 m and the area east of the pit was completed a final depth of 6.1 m between Day 191 and Day 196. The lower level struts were installed on Day 206.

4.1.4 Reduction in the Bending Stiffness of the Wall The excavation along State Street was completed and demolition of portions of the existing subway tunnel walls began during the period between Day 110 and Day 140. Prior to backfilling, portions of the secant pile wall had to be chipped away so that the required clear span between the east and west mezzanine walls could be attained. The secant pile walls between Strut 1 and Strut 6 were chipped to the flange of the W24x55 sections for their entire exposed length. The east wall was chipped to the flange between Days 110 and 140. The west wall was chipped to the flange between Days 117 and 134. The majority of the chipping activities along the north wall occurred on Day 199. Removal of the hardened grout from the wall resulted in a reduction in bending stiffness of the wall.

The station renovations were completed from Day 172 to Day 255. After the major structural modifications to the station were completed, the excavation was backfilled. The backfill activities began on about Day 225.

4.2 OVERVIEW OF FIELD INSTRUMENTATION RESPONSES Ground movement, building settlement, and pore water pressure observations are presented and related to the excavation and construction activities over the extent of the project. These observations provide summaries of the performance of the excavation support system and the response of the Warde School during the project. The instrumentation data collected during the course of the project are presented as Appendix B. The optical survey data is given in Table B-1. Table B-2 presents the pore water pressure data. The strain gauge data for the struts are given in Table B-3 and Table B-4 presents the load cell data for selected tiebacks. The inclinometer data is presented in Table B-5. It is noted that the only the A-axis data is presented for Inclinometers 1, 2, 4, and 5. The B-axis data for these inclinometers showed that the ground movement parallel to the plane of the excavation was relatively small at the inclinometer locations. Therefore, this data was not included. Also, the deformations recorded at Inclinometer 3 during the project were negligible and thus not used in any of the analyses. As a result, the Inclinometer 3 data are not included in Appendix B.

4.2.1 Ground and Building Movements 4.2.1.1 Responses along East Side of State Street Lateral ground movements resulting from the excavation and construction activities along the east side of State Street are evaluated using the Inclinometer 1 and

Inclinometer 2 responses. These lateral movements are compared to settlement data obtained from points on the Warde School adjacent to the inclinometer locations. Although Inclinometer 2 was also located along the east side of the State Street, its data were locally affected by installation of a tieback. As a result, the lateral movements observed for Inclinometer 2 near elevation -3.4 m CCD were not completely representative of the lateral response of the clay at that elevation. Summaries of the movements that developed throughout construction at these locations are presented in Figure 4-5 and Figure 4-6. These figures show deformations plotted against time for the duration of the project. The horizontal movements were measured for both inclinometers at depths corresponding to the maximum lateral deformation: (i) within the soft clay (elevation -3.7 m CCD), (ii) at the bottom of the excavation (elevation -7.9 m CCD), and (iii) in the stiff clay below the bottom of the excavation (elevation 11.0 m CCD).

Figure 4-5. Lateral Deformations at Inclinometer 1 During Construction

Figure 4-6. Lateral Deformations at Inclinometer 2 During Construction The Inclinometer 1 and 2 deformations in Figure 4-5 and Figure 4-6 are the lateral movements perpendicular to the secant pile wall in the east-west direction. Inclinometer 1 and Inclinometer 2 were located approximately 2.2 m and 1.7 m west of the school, respectively. The settlement shown in Figure 4-5 was obtained from optical survey data at settlement point W10 on the west exterior wall of the Warde School near Inclinometer 1 (see Figure 3-2). The settlement shown in Figure 4-6 was obtained from survey data at settlement point W8, near Inclinometer 2 (see Figure 3-2). Settlement point W10 was located approximately 2.2 m east of Inclinometer 1 and settlement point

W8 was located about 1.7 m east of Inclinometer 2. Also plotted in Figures 4-5 and 4-6 are the major construction activities at the Inclinometer 1 and Inclinometer 2 locations. It is apparent from both figures that the settlement of the building follows the development of lateral movements in the soil. The lateral movements within the soft clay (elevation -3.7 m CCD) and the settlements at W8 and W10 were virtually identical until Day 145 when excavation was complete and the secant pile wall had been chipped to the flange at both inclinometer locations. Thereafter, the settlements became slightly greater than the maximum lateral movements. Of the approximately 40 mm of lateral deformation measured at Inclinometer 1, 10 mm occurred during wall installation, 18 mm developed as the soil was excavated and the lateral support installed, and 12 mm occurred during tunnel demolition and station renovation as a result of creep and the reduction of wall stiffness. The maximum amount of lateral deformation at Inclinometer 2 was about 10 mm less (30 mm total) than was observed at Inclinometer 1. Similar to what was observed at Inclinometer 1, 10 mm of deformation occurred during wall installation. About 12 mm of deformation at Inclinometer 2 was attributed to installation of the support system and excavating to the final grade. From Figure 4-6, it is noted that there is a sharp dip in the elevation -3.7 m CCD curve at Day 74. This dip is a result of installing the upper level tieback at that location. This response is presumably a very local effect and, hence if the tieback did not pass so close to Inclinometer 2, then more deformation would have been recorded in the soft clay. The last 8 mm of deformation occurred in response to the reduction of wall stiffness and creep.

Figures 4-5 and 4-6 show that lateral movements and settlement first occurred as the secant pile wall was installed. The movements increased very slightly during the installation of the struts and the upper level tiebacks. However, both figures indicate a rapid increase in movements once the excavation was advanced below the stiff clay crust at elevation -1.2 m CCD. Both figures show a dramatic reduction in the rate of deformation once the excavation is completed, between Day 114 and Day 116. However, both figures show an increase in the rate of movement between Day 150 and Day 175. This increase was caused by the reduction in wall stiffness caused by chipping away the face of the secant pile wall. Placing the base slab between Day 172 and Day 177 at both the Inclinometer 1 and the Inclinometer 2 locations reduced the rate of lateral movements. This was because the slab acted as a diaphragm once it abutted the existing tunnel. Note that the movements associated with stress relief from the excavation were less than one-half the total movement. This is significant in that the movements associated with stress relief during excavation and support placement are many times the only ones considered when predicting the movements associated with deep excavations. 4.2.1.2 Responses along West Side of State Street Lateral ground movements resulting from the excavation and construction activities along the west side of the State Street are evaluated using the responses observed at Inclinometer 4. Summaries of the movements that developed throughout construction at Inclinometer 4 are presented in Figure 4-7. Similar to Figure 4-5 and 4-6, deformations are plotted versus time and the horizontal movements are measured at depths corresponding to the maximum lateral deformation: (i) within the soft clay (elevation 4.4 m CCD), (ii) at the bottom of the excavation (elevation -8.1 m CCD), and (iii) in the

stiff clay below the bottom of the excavation (elevation -11.1 m CCD). The excavation and construction activities shown above the deformation curves in Figure 4-7 are applicable to the area adjacent to Inclinometer 4. The figure shows that the values between Day 109 and Day 123 were estimated. The original Inclinometer 4 was destroyed while installing a sanitary sewer line on the west side of State Street.

Figure 4-7. Lateral Deformations at Inclinometer 4 During Construction A new inclinometer was installed approximately 1 m north of the original location. Based on the deformation trends observed at Inclinometers 1 and 2 at this same time, it was assumed that the increase in lateral movements at Inclinometer 4 were relatively

linear between Days 109 and 123. Consequently, the deformations shown in Figure 4-7 by the dashed lines were extended to Day 123 at the slope observed prior to the inclinometer being damaged. The deformations after Day 123 were obtained by adding measured deformations from the new inclinometer to the estimated cumulative value at the end of Day 123. It is observed from Figure 4-7 that the general lateral deformation trends on the west side of the excavation are very similar to those on the east side. However, the magnitude of deformation on the west side of State Street was approximately 30 percent greater than on the east side. An estimated 57 mm of lateral deformation was measured within the soft clay layer at Inclinometer 4. Of this total deformation, approximately 17 mm occurred during the wall installation. This deformation represents 34 percent of the total lateral deformation measured within the soft clay layer. An additional 36 mm (59 percent of the total) of lateral deformation was measured during the excavation and support installation stage and the remaining 4 mm (7 percent of the total) of deformation occurred during the station renovation. Thus, just a little more than half of the lateral deformations measured on the west side are in response to excavation-induced stress relief. The percentages of movements caused by excavation-induced stress relief are similar to those observed on the east side, where 45 percent of the total lateral deformation developed in the soft clay as a result of this factor. From Figure 4-7, it is noted that the majority of lateral deformations resulting from stress relief did not occur until the excavation was advanced about midway into the soft

clay layer at elevation -1.8 m CCD. This is apparent from observing that the lateral deformation measured at elevation -4.4 m CCD only increased from 17 mm to about 20 mm from Day 25 to Day 92. The corresponding change in the excavated depth during this time was from elevation +4 m CCD to elevation -1.8 m CCD. However, as the excavation was advanced from that point to the final elevation of -8.2 m CCD (Day 93 to Day 156), the lateral deformation increased from 20 mm to an estimated 53 mm. Along the east side, the significant increase in lateral movements occurred once the excavation was advanced below elevation -1.2 m CCD. Figure 4-7 shows that the rate of deformation appeared to slow between Day 120 and Day 130, but then began to increase after Day 130 at the previous rate. The decrease in the deformation rate was in response to the excavation being stalled at elevation -7.6 m CCD. The increase in the rate of deformation corresponds to advancing the excavation to its deepest point (elevation -8.2 m CCD) beginning on Day 137 and chipping the secant pile wall between Day 117 and Day 134. The excavation was completed at Inclinometer 4 location on Day 156 and the pit slab for Escalator #1 was completed on Day 163. Only minor creep movements (approximately 4.4 mm) were observed after the pit slab was completed.

4.2.1.3 Responses along Chicago Avenue Figure 4-8 presents the lateral deformations measured at Inclinometer 5 located between the school and the excavation along Chicago Avenue. These deformations represent the lateral movements at the north side of the Warde School. The figure shows deformations in both the east-west (B-axis) and north-south (A-axis) directions. The

depths at which the lateral movements were measured corresponded to: (i) the depth of maximum lateral deformation for Inclinometer 5 (elevation -2.6 m CCD), this also corresponds to the bottom of the excavation on the north side of the Warde School; (ii) the bottom of the Blodgett layer (elevation -6.8 m CCD); and (iii) the bottom of the Deerfield layer (elevation -9.9 m CCD). Also shown in Figure 4-8 are the settlement data obtained from settlement point W13. This settlement point was adjacent to the Inclinometer 5 location. Two excavation profiles are given in Figure 4-8, one represents the construction record at Chicago Avenue at the Inclinometer 5 location, and the other represents the construction record at the northwest corner of the school.

Figure 4-8. Lateral Deformations at Inclinometer 5 During Construction

Although the north-south direction was the primary direction of measurement, it is apparent that the east-west direction represented a significant component of the movement. The total lateral movement is represented by the resultant of the A- and Baxis movements and is also shown on the figure. Figure 4-8 shows that the lateral deformations in the east-west direction were consistently greater than those in northsouth direction throughout the project. More importantly, the resultant of the two components of the lateral movement tracks the development of the surface settlement much better than either of the individual components. The north secant pile wall was installed between Day 70 and Day 79. In response to this activity, the maximum lateral deformation perpendicular (north-south component) to the secant pile wall was about 4 mm, but the maximum lateral deformation parallel (eastwest component) to the secant pile wall was about 10 mm. Presumably, the increase in the east-west component on Day 79 was in response to the excavation at the northwest corner being lowered beneath the clay crust. The figure shows that the excavated grade at the northwest corner of the east secant pile wall was advanced to approximately elevation -1.2 m CCD on Day 79. The excavation activities on the north side of the Warde School occurred between Day 125 and Day 208. During this period, the north-south component of the maximum lateral deformation increased from about 4 mm to about 14 mm and the east-west component of the maximum lateral deformation increased from 10 mm to 20 mm. During this time, the excavated grade at the northwest corner was advance from elevation -6.4 m CCD to elevation -8 m CCD on Day 150. The excavated grade at the Inclinometer 5 location was not advanced below the clay crust until about Day 170. It appears that, at least until Day 170, the east-

west component of lateral movement was primarily in response to the excavation along State Street. There were no additional lateral deformations observed after the pit slab was pour. The pit slab acted as a continuous support across the bottom of the excavation. The optical survey data showed that from Day 70 to Day 79 the settlement at W13 increased from 6 mm to 11 mm. By comparing this settlement with the resultant lateral deformations at elevation -28 m CCD, which also increased to the 11 mm during this time, one can see that the settlements were caused by the combined effects of the installation of the north secant pile wall and the excavation along State Street. Figure 48 shows that the settlement rate increased after Day 170, as a result of excavating below the clay crust along Chicago Avenue. The settlement increased to a maximum of about 28 mm on Day 208. These settlement responses indicate that the deformational behavior of the northwest corner of the building was a function of the 3-D excavation geometry at that location.

4.2.2 Pore Water Pressure Response Pneumatic piezometers were installed at elevations -5.8 m CCD and -8.8 m CCD adjacent to Inclinometer 1 and at elevations -6.4 m CCD and -9.45 m CCD adjacent to Inclinometer 4. These piezometer locations were designated Piezometer 1 and Piezometer 4 to correspond to the inclinometer locations Piezometer 1 was located 2 m from the secant pile wall whereas Piezometer 4 was located 3 m from the wall. Data were collected at both piezometer locations throughout the project. However, the piezometers at the Piezometer 1 location were destroyed during installation of the upper level tiebacks. Figure 4-9 presents the pore water pressure responses at the Piezometer 1

location, up to the day the piezometers were destroyed. The pore water pressure responses at the Piezometer 4 location are presented in Figure 4-10. The figure uses two colors for each piezometer curve. This is because the original piezometers at this location were destroyed as a result of the west side construction activities. New piezometers were installed at the same depths of the original piezometers, a few feet north of the original location. The first color in each curve represents the pore water pressure data from the original piezometers and the second color is the replacement piezometers. Note that there was some overlap for the piezometer at elevation -9.45 m CCD.

Figure 4-9. Pore Pressure Response at Piezometer 1 During Construction

Similar responses were observed at both the Piezometer 1 and Piezometer 4 locations. The data obtained before construction began indicate that essentially hydrostatic conditions initially existed in the clays. The elevation of the piezometric surface at the Piezometer 4 location was approximately +0.75 m CCD and the piezometric surface at the Piezometer 1 location was approximately +0.6 m CCD. The piezometric surface is slightly above the top of the stiff clay crust layer and is apparently sloping towards Lake Michigan.

Figure 4-10. Pore Pressure Response at Piezometer 4 During Construction Figures 4-9 and 4-10 show that large drops in water pressure occurred as the secant pile wall was installed. The water heads dropped as much as 5.5 m at the Piezometer 1

location and about 4.5 m at the Piezometer 4 location. Essentially, the open shafts of the secant piles near the piezometer locations acted as temporary sinks. Because grout was placed within each of the holes soon after excavation, the water levels at Piezometer 4 recovered most of their initial values within about 38 days. The excavation had not proceeded below the clay crust at this time (Figure 4-10). The water levels at Piezometer 1 had recovered about 82 percent of their initial values before the excavation was advanced below the clay crust on Day 60. Afterwards, the water levels dropped in response to the excavation until the piezometers were destroyed on Day 87. From this point, the Piezometer 4 data are used to illustrate the pore pressure responses to the subsequent construction activities. The pore pressures at Piezometer 4 were relatively steady from about Day 38 to Day 90. Thereafter, the excavation along the west side was lowered below elevation -1.8 m CCD, whereupon the pore pressures began to drop. This time corresponds to the time where the clay crust was excavated and significant movements began to develop during excavation. Another transient response was observed when the second level tiebacks were stressed. Water heads increased as much as 5 m, but dissipated within several days. Given that the anchors passed within 2 m of both piezometers and that the pressures rapidly dissipated, the pore pressure responses caused by stressing the anchors were localized. By Day 150, the water levels had dropped about 3 m from their initial values at both piezometers. This drop increased the effective stress in the soft clay by 30 kPa, apparently enough for small consolidation settlements to develop. If one assumes that the pore pressures on both sides of the excavation responded to the construction activities in a similar manner, the data beyond Day 150, shown in Figure 4-5, can be

interpreted in light of these pore pressure changes. At settlement point W10, movements larger than observed in the inclinometer data at elevation -3.7 m CCD began to develop. These movements were presumably a result of the small consolidation settlement arising from the drop in water pressure. The maximum excavation-related drop in the water level was observed on Day 203, prior to backfilling. At this time, the heads in the piezometers were 7.1 m at elevation -9.45 m CCD and 4.5 m at elevation -6.4 m CCD. After Day 203, the pressures began to recover slightly. However, after backfill was completed on Day 354, water levels remained about 2.4 m below the initial values. The water levels should increase slowly over time depending on the effectiveness of the waterproofing for the renovated station.

4.3 DETAILED OBSERVATIONS OF LATERAL MOVEMENTS Detailed observations of lateral movements are given for locations corresponding to the excavation activities along the east and west sides of State Street and along Chicago Avenue. Inclinometer 1 is representative of the east side of State Street, because the lateral responses at Inclinometers 1 and 2 were very similar (see Section 4.2.1.1). Inclinometer 4 was used to represent the west side of State Street and Inclinometer 5 is representative of the activities along Chicago Avenue. The detailed observations include inclinometer responses to significant excavation and construction activity. Activities common to all three locations are given in Table 4-1. The table shows the activities and the days of occurrences at the three locations. Note that the activities are listed in groups of the major construction stages of the project. The

subsequent sections discuss the lateral movements observed at the Inclinometer 1, 4, and 5 locations during each construction stage.

4.3.1 Stage 1 Wall Installation Table 4-1 indicates that the secant pile wall installation activities at the representative inclinometer locations along State Street (Inclinometer 1 and Inclinometer 4) were completed between Day 2 and Day 18. The table also shows that the secant pile wall was installed along Chicago Avenue between Day 70 and Day 79. However, these dates only include the installation of the secant piles that influenced the responses of the inclinometers. Using a trail-and-error method of comparing secant pile installation with inclinometer response, it was ascertained that the inclinometer responded to installation activities when secant piles were being installed at a center-to-center spacing of 3.8 m. Referring to Table A-1 in Appendix A, it is seen that the entire east secant pile wall was installed between Day 0 and Day 30. The complete west secant pile wall was installed between Day 9 and Day 32. The dates given in Table 4-1 for the Inclinometer 5 location include installation of all the secant piles along Chicago Avenue.

4.3.1.1 East Side of State Street

Figure 4-11. Lateral Soil Movements and Settlement Response at Inclinometer 1 Location: Wall Installation Figure 4-11 presents the lateral responses for Inclinometer 1 and the settlement response of the Warde School at a section perpendicular to the east secant pile wall through Inclinometer 1. The inclinometer was located 2.3 m from the centerline of the secant pile wall. This section is typical of the response in the middle of the east secant pile wall and shows the effects of construction on the Warde School. Lateral soil movements developed as the wall was installed with the maximum movement occurring within the soft clay. Note that the stability number (Peck, 1969) of the clay at the depth of maximum movement was 7.8. This high stability number would suggest that inward movements would occur due to stress relief as a shaft was drilled without support (i.e.,

with no drilling mud or casing). The lateral movements extend to the depth of the secant pile wall. The building settled as much as the soil displaced laterally, with the settlement extending as far behind the excavation as the secant pile extended below the bottom of the schools foundation. These patterns of movement are expected because the movements into the shafts of the secant pile wall are the result of undrained deformations in saturated clay. Figure 4-11 shows that the maximum lateral deformation at the end of the secant pile wall installation (Day 11) was approximately 10 mm. This initial 10 mm of deformation corresponded to about 26 percent of the total lateral deformation observed along the east secant pile wall. The figure also indicates that below the soft clay layer, the inflection points of the deformation curves correspond approximately to the layer interfaces. This trend is likely the result of the differences in shear strength and stiffness among the layers. Peck and Reed (1954) indicated that the undrained shear strength of individual strata in the downtown area of Chicago generally increases in steps from one stratum to the next.

4.3.1.2 West Side of State Street

Figure 4-12. Lateral Soil Movements at Inclinometer 4 Location: Wall Installation Figure 4-12 presents the lateral soil movements during wall installation at the Inclinometer 4 location. The inclinometer was located approximately 3 m from the centerline of the west secant pile wall. The figure shows that the maximum deformation

of 17 mm observed at the end of the secant pile wall installation was greater than what was observed at Inclinometer 1. The point of maximum deformation occurred in the medium clay layer at elevation -5 m CCD. The larger lateral movements at Inclinometer 4 reflect the shafts of the west wall secant piles being open for longer times prior to placing the concrete. Along the east secant pile wall, steel casing was advanced well into the soft clay when constructing shafts near the Inclinometer 1 location and concrete was typically placed within 2 hours of drilling the shafts. However, along the west secant pile wall the steel casing was only advanced to the stiff clay layer and, in some instances, the steel casing was not used at all. In addition, concrete placement for the west secant pile wall sometimes took as long as 5 hours after drilling the shaft. Another factor contributing to the larger lateral movements at Inclinometer 4 are the higher insitu soil stresses on the west side of the excavation. Construction of the Warde School decreased the insitu stress state along the east side of the project site as a result of the net unloading associated with the excavation of the basement and the reloading associated with the construction of the building. Thus, for the same undrained shear strength, the stability number is higher on the west side of the excavation than on the east side. As a result, soil squeeze into the open, uncased shaft and the subsequent ground loss will occur faster and be greater on the west side of the excavation than on the east side. No settlement data were recorded on the west side of the excavation. 4.3.1.3 Chicago Avenue Figure 4-13 gives the lateral soil response for Inclinometer 5. The inclinometer was located 700 mm south of the centerline of the north secant pile wall and represents the lateral soil response along the north side of the Warde School. The figure gives the

lateral movements in the direction towards Chicago Avenue (4-13a) and the movements in the direction towards State Street (4-13b).

Figure 4-13. Lateral Soil Movements at Inclinometer 5 Location: Wall InstallationIt should be noted that the inclinometer was damaged several times during the installation

of the secant pile wall along Chicago Avenue (refer to Table A-1 in Appendix A). Consequently, the upper 3.5 m to 4 m of the data reflects movements caused by impacts against the inclinometer casing and not lateral deformations of the soil resulting from wall installation. More reliable lateral deformation data were observed below elevations corresponding to the middle of the stiff clay crust (approximately elevation +0.6 m CCD). As was the case in other inclinometers, it is apparent that maximum wall installation-related deformations occurred within the soft clay layer. However, the magnitude of maximum deformation in the direction perpendicular to face of the excavation was about 40 percent of that observed at the Inclinometer 1 location during wall installation along State Street. Figure 4-13a gives the maximum deformation in response to wall installation for Inclinometer 5 as approximately 4 mm, as compared to 10 mm at Inclinometer 1. Figure 4-13b shows that the maximum lateral deformation at Inclinometer 5 in the direction toward State Street was just slightly less than 10 mm on Day 79, of which 6 mm occurred as a result of installing the wall between Day 70 and 79. The decidedly 2-D movement in response to wall installation was caused by the fact that Inclinometer 5 was located within 0.8 m of the wall, and thus was affected locally in two directions by the drilling operations. As was noted for the other inclinometer locations, inflection points along the depth of the curves were observed at the approximate layer interfaces. The bottom of the secant pile wall served as the lower bound for the lateral movements.

4.3.2 Stage 2 Excavation and Support System Installation Stage 2 covers the period of excavation and support system installation. The struts along State Street were installed between Day 60 and Day 74. All upper level tiebacks for

both the east and west secant pile walls were completed as of Day 106 and all lower level tiebacks were completed as of Day 114. The struts along Chicago Avenue were installed between Day 128 and Day 206. No tiebacks were used on the north side of the Warde School. Dates given in Table 4-1 for strut and tieback installation are for the struts and tiebacks in the vicinity of the inclinometers only. Refer to Appendix A for a complete record of the support system installation.

4.3.2.1 East Side of State Street Figure 4-14 presents the Stage 2 lateral deformations observed along the east secant pile wall. The grade was excavated to elevation +2.7 m CCD on Day 59. Negligible movement occurred from the end of wall installation to when the excavation was lowered to elevation +2.7 m CCD. Significant lateral movements began to develop as the excavation was advanced through the clay crust. This large increase in movements was due to the reduction in the basal heave factor of safety. Clough et al. (1989) defined the factor of safety against basal heave for relatively wide excavations (H/B < 1) as

(4.1) where H is the depth of the excavation, B is the effective width of the excavation, Nc is the bearing capacity factor, Suu is the undrained shear strength above the bottom of the excavation, Sub is the undrained shear strength below the bottom of the excavation, and is the total unit weight of the soil above the bottom of the excavation. Note, the presence of the tunnel restricts the width of the failure zone. Thus, to account for the

limited failure zone, the top of the tunnel is used as a rigid base. Table 4-2 gives the basal heave parameters prior to the excavation advancing below the stiff clay crust and after the excavation has advanced into the soft clay. The undrained shear strength values above the bottom of the excavation were obtained from the weighted averages of the layers within the depth H. The undrained shear strength values below the bottom of the excavation are the average values within the failure zone. From Table 4-2, it can be seen that the factor of safety decreased by about 53 percent after the excavation was advanced below the clay crust

Figure 4-14. Lateral Soil Movements and Settlement Response at Inclinometer 1 Location: Excavation and Support Installation

Table 4-2 Estimates of Factor of Safety Against Basal Heave

Basal Heave Parameters H B L (length of excavation Nc Suu Sub Factor of Safety (FS)

Excavation at Clay Crust Layer Excavation at Soft Clay Layer (bgs) (bgs) 5.5 m 2.6 m 35 m 7.1 19.63 kN/m3 30 kPa 15 kPa 5.8 35 m 7.3 18.94 kN/m3 23 kPa 15 kPa 2.7 6.5 m 2.6 m

Note: (1) bgs = Below ground surface Figure 4-14 shows that the lateral movements extend to the depth of the secant pile wall. The building settled as much as the soil displaced laterally, with the settlement extending as far behind the excavation as the secant pile wall extended below the bottom of the schools foundation. The equality of lateral movements and settlements are expected because the movements that occur are essentially undrained. Figure 4-14 shows that the incremental deformations increased slightly more than 2 mm as the excavation was advanced below the first tieback level (elevation -1.4 m CCD) on Day 81, and an additional 3 mm before stressing the first level tieback on Day 87. The increment of movement between excavating below the first tieback level on Day 81 and excavating below the second tieback level (elevation -5.2 m CCD) on Day 102 was

about 9 mm, resulting in a cumulative deformation of 21 mm. The upper portion of the wall moved about 2 mm toward the soil both times the tiebacks were stressed. As a result, the rate deformation was temporarily halted at each instance the tiebacks were stressed. This can be seen from the response of the soft clay layer in Figure 4-5. After the tiebacks were stressed, the secant pile wall moved incrementally toward the excavation in response to excavation-induced stress relief. The excavation was lowered to the final elevation of -7.9 m CCD on Day 116. When the excavation reached final grade, the maximum lateral movement increased to 28 mm. The school settled as the secant pile wall moved laterally. The maximum settlement of the school at the end of excavation was equal to this maximum lateral movement of 28 mm. Note that the extent of the settlement trough did not increase during this time.

Figure 4-15. Lateral Soil Movements at Inclinometer 4 Location: Excavation and Support Installation

4.3.2.2 West Side of State Street

The observed and estimated Stage 2 lateral deformations for Inclinometer 4 are presented in Figure 4-15. As was discussed in Section 4.2.1.2, the original Inclinometer 4 was destroyed on Day 109. Values reported after that date were estimated based on incremental data from the new inclinometer and observed trends of increase in lateral movements at Inclinometers 1 and 2. The lateral movements at Inclinometer 4 during this stage were about twice those observed at Inclinometer 1. The excavation along west secant pile wall was advanced below the strut level (elevation +2.7 m CCD) on Day 59 with no significant lateral movements. Significant lateral movements again developed as the excavation was advanced through the clay crust. Two millimeters of movement occurred as a result of advancing the excavation below the first level tieback to elevation -1.8 m CCD on Day 81. Note that the shape of the deformation curves indicated cantilever movement of the wall until the first level tieback was installed on Day 81. This large cantilever-type movement is in contrast to that observed on the east side of the excavation, where very little cantilever movement was observed. Apparently, load from the west side of the excavation was being transferred to the east side through the struts, resulting in larger cantilever-type movements on the west side. The increment of movement between excavating below the first tieback level on Day 81 and excavating below the second tieback level (elevation -4.9 m CCD) on Day 98 was about 12 mm. This is compared to the 9 mm of incremental movement at Inclinometer 1 for the same excavation activity, even though a greater height of the wall was unsupported along the east side than was observed on the west side prior to stressing the second level tieback. The maximum lateral deformation increased to 43 mm on Day 123, which corresponded to the excavation along the west side being advanced to elevation -

7.6 m CCD. The stress relief-induced movements continued to increase until the excavation reached final grade at elevation -8.2 m CCD on Day 156. The maximum lateral deformation was 57 mm on Day 156. From the shape of the lateral response curves shown in Figure 4-15, it is observed that deep inward movement of the secant pile wall characterized the incremental lateral deformation when the excavation was lowered from the second tieback level to final grade. The secant pile wall was chipped from elevation -4.26 m CCD to elevation -7.62 m CCD between Day 117 and Day 134. This process resulted in the wall becoming more flexible and ultimately contributed to the increase in lateral movements in the soft clay layer. 4.3.2.3 Chicago Avenue The lateral deformation measured at Inclinometer 5 for the Stage 2 construction activities are presented in Figure 4-16. The lateral movements toward Chicago Avenue are given in Figure 4-16a while those toward State Street are shown in Figure 4-16b. The depth of the excavation along the north side of the Warde School was about half that along State Street. Consequently, the maximum lateral movements along north secant pile wall are smaller than those observed on either side of State Street. Cumulative lateral movement at Inclinometer 5 was 3 mm in the north-south direction and 11 mm in the east-west direction on Day 109. At this time, the north side excavation was extended to below the upper strut level at elevation +2.1 m CCD. An additional 2 mm of movement was observed in the north-south direction between the initial excavation and installing the upper level strut on Day 128. No movement was observed in the east-west direction during this time. The excavation along the north side

was first advanced into the soft clay layer (elevation -1.5 m CCD) on Day 170. This resulted in a maximum incremental movement toward Chicago Avenue of 2 mm and maximum incremental movement towards State Street of 4 mm, clearly showing the effects of the excavation on State Street.

Figure 4-16. Lateral Soil Movements at Inclinometer 5 Location: Excavation and Support Installation

Figure 4-16 shows that after Day 170 the incremental movements were primarily directed towards Chicago Avenue as a result of excavating from -1.5 m CCD to -4 m CCD on Day 208. The increment of lateral deformation in the soft clay between Day 170 and Day 208 was 7 mm toward Chicago Avenue and 1 mm toward State Street. There were no additional excavation activities along the east side of State Street after the pit for Escalator #3 was completed on Day 163. Thus, deformations observed after Day 163 were primarily caused by excavation activities along Chicago Avenue, although small amounts of creep movements may have developed as well. Note that at the Inclinometers 1 and 2 locations creep movements were observed after the excavation activities had been completed. It was observed that the rate of creep at these two inclinometers produced an increment of maximum movement in the soft clay layer of about 2 mm for a given 38-day period. This maximum increment of movement compares to the 1-mm increment toward State Street observed at Inclinometer 5 between Day 170 and Day 208.

4.3.3 Stage 3 Renovation and Backfill The last major stage in the Chicago Avenue and State Street subway renovation project involved completing the capital improvements and backfilling the excavation. Table 4-1 indicates that this stage began with chipping the secant pile wall adjacent to the inclinometer locations. However, chipping of the secant pile wall occurred prior to completing the excavation at the Inclinometer 4 and Inclinometer 5 locations.

4.3.3.1 East Side of State Street

The lateral deformations measured at Inclinometer 1 during this stage of the construction are presented in Figure 4-17. Very small creep movements developed after the excavation was advanced to final grade. This can be seen from the incremental movement between Day 116 and Day 137.

Figure 4-17. Lateral Soil Movements and Settlement Response at Inclinometer 1 Location: Station Renovation and Backfill To accommodate the exterior walls of the new mezzanine section, the grout on the face of the secant pile wall was chipped to the face of the flange from elevations -4.3 m to -

7.6 m CCD between Days 137 and 140. This chipping activity resulted in a reduction of the bending stiffness of the wall. Consequently, more movements were observed between Days 137 and 176, when the bottom slab for Escalator #4 was poured, than were observed during the remainder of this stage of construction. Reducing the stiffness of the wall resulted in approximately 5 mm of incremental movement within the softer clays. After Day 176, approximately 3 mm of small creep movements continued until the excavation was completely backfilled. Note that the tiebacks were not cut, and remain stressed in place. Removal of the top level of bracing on Day 258 resulted in small cantilever movements near the top of the wall, as is evident from the Day 266 curve in Figure 4-17.

4.3.3.2 West Side of State Street

Figure 4-18. Lateral Soil Movements at Inclinometer 4 Location: Station Renovation and Backfill

The estimated lateral deformations measured at Inclinometer 4 for Stage 3 are presented in Figure 4-18. The base slab for Escalator #1 was completed on Day 163 and the backfill was placed between Day 225 and Day 310. A total of 3 mm of incremental

creep movement was observed between the time when the escalator bottom slab was placed and backfill was completed. The west secant pile wall was also chipped to the flange along its exposed length However, unlike the east secant pile wall, the west secant pile wall was chipped prior to advancing the west side excavation to final grade. The chipping activities occurred between Day 117 and Day 134 and the excavation along the west side was completed on Day 156. Hence, the incremental lateral wall movements given in Figure 4-18 are a result of creep and strut removal.

4.3.3.3 Chicago Avenue Figure 4-19 presents lateral deformations measured at Inclinometer 5 during Stage 3. It is apparent from the figure that little movement occurred along the north side of the Warde School after the excavation was advanced to final grade on Day 208. Very minor creep movements in the direction of State Street were observed after the base slab for Escalator #3 was poured (Day 210). There were no additional lateral movements observed in the direction of Chicago Avenue after the base slab was pour. The base slab effectively acted as an additional support level across the bottom of the excavation.

4.3.4 Summary of Lateral Deformations and Discussion

4.3.4.1 Incremental Lateral Soil Movements

Figure 4-19. Lateral Soil Movements at Inclinometer 5 Location: Station Renovation and Backfill Figure 4-20 shows a comparison of the lateral increments of movements that occurred adjacent to the excavation at Inclinometers 1, 2, 4, and 5. The incremental movements

are shown for the three stage of construction: wall installation (20a), excavation to final grade (20b), and station renovation and backfill (20c).

Figure 4-20. Summary of Incremental Lateral Soil Movements Inclinometers 1 and 2 along the east side of the excavation exhibited very similar responses, with the exception of the kink in the data at elevation -3.4 m CCD at Inclinometer 2. This kink was the result of installing a first level tieback next to the inclinometer. Inclinometer 4 along the west side of the excavation showed incremental movements that were about 50 percent larger than those on the east side. Incremental movements along the north side at Inclinometer 5 were about 50 percent less than those on the east side.

During wall installation, 17 mm of lateral movement developed at the Inclinometer 4 location, which was about twice that at the other locations. The magnitude of this movement depended on the amount of time the shafts remained open after it was drilled and before it was filled with grout. Near the location of Inclinometer 4, the shafts remained open for as long as 5 hours, as compared to times of no more than 2 hours along the east side of the excavation. During excavation, the increment of movement was about twice as much along the west wall compared to that along the east wall and about four times as much as that along the north wall. The large difference in incremental movements between Inclinometer 4 and Inclinometer 5 was primarily due to the different depths of excavation and the geometry of the excavation adjacent to the inclinometer. However, along the east side the excavation the depth and geometry of the excavation and the construction sequences were similar to that along the west side. The larger movements observed during the wall installation at the Inclinometer 5 location apparently impacted the subsequent movements. As the soil deformed more into the open shaft, the soil mobilized its shear strength, leaving less available resistance for stresses imposed by subsequent excavation. Cunningham and Fernandez (1972) observed similar responses. They reported larger movements during excavation at locations where larger movements occurred as large diameter caissons were installed near a slurry wall before the excavation began. The presence of surcharge loads on the west side, in the form of construction equipment and materials, could have also contributed to the larger movements. No such loads were imposed on the east side of the excavation because of the lack of space between the secant pile wall and the school. Another contributing cause to the larger incremental movements during excavation could be the higher insitu stresses in the soil on that side of the excavation. There were

no buildings on the west side, whereas the 3story Warde School with its 3-m deep basement imposed less stress on the wall than the soil on the west side. The smaller creep movements along the west secant pile wall can be attributed to the base slab being placed more quickly at that location than along the east wall. The creep movements along the north wall were negligible because the excavation was extended only partially into the soft clay layer.

4.3.4.2 Lateral Deformation Vectors Figures 4-13, 4-16, and 4-19 indicated that although the north-south direction (towards Chicago Avenue) was the primary direction of measurement for Inclinometer 5, a significant component of the movement was in the east-west direction (towards State Street). This suggests that the lateral movements observed at Inclinometer 5 were heavily influenced by excavation activities performed along the west side of the school. Prior to excavating to final grade, the magnitudes of movements in the east-west direction were greater than those observed in the north-south direction. The implication is that the movements resulting from the excavation-induced stress relief at the corner were a function of the 3-D geometry of the excavation. Thus, in order to compare the lateral deformational response of Inclinometer 5 with those of Inclinometer 1 and Inclinometer 2, the deformations at Inclinometer 5 are presented in term of the deformation resultant. The movements towards Chicago Avenue serve as the northern component of the resultant and the movements towards State Street serve as the western component. Figure 4-21 presents a plan view of the lateral deformation vectors measured at Inclinometers 5, 1, and 2 to show the effects of opening the excavation

along Chicago Avenue at the same time as the State Street excavation. The deformations are given at depths within the soft clay stratum corresponding to the maximum observed lateral movements in the inclinometers. For Inclinometer 1, the depth of maximum movement was 7.9 m below ground surface. The depth of maximum movement for Inclinometer 2 was 8.6 m and the depth for Inclinometer 5 was 6.9 m. The construction days used in the figure correspond to the days of significant construction activity presented in the previous sections.

Figure 4-21. Lateral Deformation Vectors in the Soft Clay

Figure 4-21 shows that the ground movement was nearly perpendicular to the State Street excavation at the Inclinometer 1 and Inclinometer 2 locations, indicating that the movements were in response to the activities along State Street. The slight inclination observed in both inclinometers on Day 11 is in response to secant pile installation activities. Between Days 11 and 81, Inclinometers 1 and 2 respond to the excavation advancing towards the north. Note that movements between Day 137 and Day 176 were directed towards the south at Inclinometer 1 and towards the north at Inclinometer 2. These incremental movements were caused by excavating the escalator base slab pit located between the two inclinometers (refer to Figure 4-1). The direction of movement observed between Day 266 and Day 353 for both inclinometers was in response to the backfill activities, which proceeded from south to north. The lateral ground movements at Inclinometer 5 were primarily oriented towards State Street until day 109 indicating that the ground movements to this point developed mostly as a result of activities along the State Street excavation. The small component of northward movement between Day 0 and Day 79 developed as the secant pile wall was installed along Chicago Avenue. The excavation along State Street was completed at the Inclinometer 1 location on Day 116. At this time, the lateral movements began responding more to the north excavation activities. This resulted in the deformation vector moving towards the northwest.

4.3.4.3 Comparison of Settlements and Lateral Movements Figure 4-22 shows a comparison of lateral movements measured at Inclinometers 1 and 2 and settlement data for sections taken perpendicular to the east secant pile wall through the inclinometer locations. Because the settlement data represents soil movements at elevation +0.6 m CCD, the inclinometer data were plotted horizontally by

rotating its axis 90 degrees about this elevation, at the respective inclinometer locations. Good agreement is observed between the two data sets at both locations. Slight differences developed after backfill was completed on Day 355, as a result of small consolidation settlements arising from the lowered water table (see Figure 4-10). The agreement between the settlement behind the excavation support wall and the lateral movements shows that there was essentially no volume change in the saturated clays until after the excavation had reached final grade. The volume changes that occurred after Day 117 were a consequence of the excavation-induced drop in the groundwater level. The agreement between the inclinometer and settlement data suggests that settlement behind the support wall can be reliably estimated from inclinometer data when considering excavations through saturated clay.

Figure 4-22. Comparison Between Settlement and Lateral Movements This fact is extremely beneficial when designing an excavation support system to minimize excavation-related damage to adjacent structures. Current state-of-the-art procedures use finite element models to design excavation support systems. Finite element models used to predict wall deflections provide relatively good accuracy if a

representative soil model is used and construction procedures are accurately represented in the analysis. However, prediction of the ground surface settlement induced by excavation is not as good as that of the wall deflection. The finite element models tend to underpredict the vertical movements at the wall and overpredict the vertical movements away from the wall (Finno and Harahap, 1991), and hence any direct predictions of distortions under a building will not be accurate. However, based on the observations of the agreement between inclinometer and settlement data, computed lateral deformations can be used to predict the settlement distribution behind the support wall, and hence to predict expected distortions beneath a building. This approach is valid only for cases involving no volume changes during excavation, as is typical for excavations through saturated soft to medium clay.

4.4 LOADS IN CROSS-LOT BRACES AND TIEBACKS The loads in the struts and tiebacks located adjacent to the Inclinometer 1 location are given in Figure 4-23. These loads included strain gauge data from Strut 4 and Strut 5 and load cell data from the upper and lower tiebacks below Inclinometer 1. The loads in the struts and tiebacks that corresponded to the Inclinometer 2 location are given in Figure 4-24. These loads included strain gauge data from Strut 3 and Strut 4 and load cell data from the upper and lower tiebacks below Inclinometer 2. In the figure, the loads are given in units of force per unit width of the tributary area of the support. The loads presented for the tiebacks were obtained dividing the load recorded by the load cell by the center-to-center spacing of the tiebacks (1.5 m). The loads per width of wall given for the struts were obtained by dividing the axial load derived from the strain gauge readings by the center-to-center spacing of the struts (6.1 m). The axial loads

were calculated by first plotting the observed strains with respect to their distances from the neutral axis, which yielded the strain distribution in the strut. To remove the effects of bending due to self-weight of the strut, the axial strain at the neutral axis was multiplied by the cross-sectional area and the modulus of elasticity of the strut to obtain axial load.

Figure 4-23. Loads in Support System Adjacent to Inclinometer 1 Location

Figure 4-24. Loads in Support System Adjacent to Inclinometer 2 Location The struts were not preloaded prior to recording the strain measurements. Consequently, the strut loads begin at zero and gradually increased to between 115 kN per m and 130 kN per m in response to the excavation. The design load for all struts was 120 kN per m. After the excavation reached its full depth at both locations (Day 114 at Inclinometer 2 and Day 116 at Inclinometer 1), fluctuations of 25 kN per m were observed. The fluctuations observed in the strut loads are primarily due to fluctuations in the temperature. The upper and lower level tiebacks on the east side were locked-off at 220 kN per m and 280 kN per m, respectively. These lock-off loads were equivalent to 80 percent of the design loads. Both Figure 4-23 and Figure 4-24 show that when the lower level of tiebacks were

installed and prestressed, the loads in the upper level tiebacks decreased by about 20 kN per m. During the same time, the loads in the lower tiebacks increased by a similar amount. Thereafter, the tieback loads essentially remained constant with variations of no more than 10 kN per m. This response suggests that the anchors embedded in the stiff to hard clays did not creep significantly over the 170-day period when data were obtained.

4.5 SUMMARY The construction at the site was separated into three stages; wall installation, support system installation and excavation, and station renovation and backfill. Lateral soil movements of the secant pile wall developed as the wall was installed with the maximum movement occurring within the soft clay. The lateral movements extended to the depth of the secant pile wall. The lateral movements associated with stress relief from the excavation were on average about one-half the total movement. The significance of this observation is that often only the movements associated with stress relief during excavation and support placement are considered when predicting the movements associated with supported excavations. Some of the lateral movements were attributed to a reduction in bending stiffness of the wall, which was the result of chipping the secant pile walls to the flange of the W24x55 sections for their entire exposed length. It was observed that larger lateral movements occurred at the location of Inclinometer 4 on the west wall compared to the locations of Inclinometer 1 and 2 on the east wall. This was due to the presence of surcharge loads on the west side, in the form of

construction equipment and materials and the higher insitu stresses in the soil because of the absence of an overcompensated structure on that side of the excavation. The larger movements on the west side were also a function of the large movements that occurred during wall installation. At the Inclinometer 5 location, the lateral movements were divided into a north-south component (toward Chicago Avenue) and an east-west component (toward State Street). The resultant of the two components of the lateral movement tracks the development of the surface settlement much better than either of the individual components and suggests that the deformational behavior of the northwest corner was a function of the 3-D excavation geometry at that location. It was observed that Warde School settled as much as the soil displaced laterally, with the settlement extending as far behind the excavation as the secant pile wall extended below the bottom of the schools foundation. The adjacent settlements were virtually identical to the lateral movements within the soft clay until after the excavation was complete and the secant pile wall had been chipped to the flange. Thereafter, the settlements became slightly greater than the maximum lateral movements. The struts were not preloaded prior to recording the strain measurements. Consequently, the strut loads begin at zero and gradually increased to between 115 kN per m and 130 kN per m in response to the excavation. The tieback anchors were embedded in the stiff to hard clays. Loads in these elements essentially held their lock-off loads and did not creep significantly over the 170-day period when data were obtained.

CHAPTER 5

EXCAVATION-INDUCED RESPONSE OF THE FRANCES XAVIER WARDE SCHOOL The primary responses of the Warde School to the adjacent excavation were settlement of the shallow spread and continuous wall footings and tilting of the foundation walls. The optical survey data indicated that no lateral movement occurred at the ground level of the building. Consequently, slip must have occurred between the bottom of foundations and the ground. The building settled vertically and its walls tilted, but it did not translate in response to the excavationinduced ground movements. These primary responses produced distortions in the infill walls of the school

5.1 SETTLEMENT RESPONSE Figure 5-1 presents the settlement data at points along the west side and north side of the Warde School. Settlement points W10 and W8 are located along the west side and are adjacent to the Inclinometer 1 and Inclinometer 2 locations, respectively. Settlement point W13 is located along the north side of the school and is adjacent to the Inclinometer 5. A general overview of the settlement responses to the excavation can be seen from comparing increases in settlement to excavation activities. The figure shows that the building response matched the soil response. Settlement points W8 and W13 showed approximately 8 mm to 9 mm of movement, respectively, in response to installation of the secant pile wall. After the wall installation, the movements were relatively small until the excavation was advanced below the clay crust on Day 73. The

excavation along the west side of the Warde School was completed on Day 116, after which time the movement developed more slowly as a result of creep.

Figure 5-1. Settlement versus Time at Settlement Points Adjacent to Inclinometer Locations

The movements increased at the settlement locations once the chipping of the east secant pile wall was completed on Day 140. Figure 4-2 also shows that settlement point W13 was clearly affected by the excavation along State Street. The excavation along Chicago Avenue was not advanced into the soft clay layer until Day 170. From the figure, it can be seen that approximately half the total amount of settlement observed at W13 had already occurred at that time. From Day 170 to Day 208, which was the period of excavation along Chicago Avenue, the settlement at point W13 increased from 22 mm to 29 mm. There were no further settlements at settlement point W13 after Day 208.

5.1.1 Settlement Contours

Figure 5-2. End of Stage Settlement Response of the Warde School

Settlement contours were developed for the Warde School using settlement data from points located on the exterior walls and interior columns of the school. The location of these points can be seen in the inset of the Warde School located in Figure 5-1. In the inset, W is used to denote wall points, C is used for column points and R is used for roof points. Figure 5-2 presents the settlement response of the Warde School at the end of wall installation (5-2a), after the excavation had reached final grade (5-2b), and after backfill was completed (5-2c). By comparing the contour of zero settlement for each of the construction stages, it is observed that the extent of the settlements behind the support wall developed mainly while installing the 19.3 m deep secant pile wall. Note that the small settlements that occurred in the northeast portion of the school developed as a result of relocating utilities along Chicago Avenue. The extent of movements can be defined in terms of the ratio of distance from wall to depth of wall of approximately 1. Several of the methods currently used to estimate settlement distribution behind excavation support walls (Hsieh and Ou, 1998; Clough and ORourke, 1990; and Peck, 1969) normalize the extent of settlement with the depth of excavation. However, Figure 5-2 shows that while the magnitudes of movements increased significantly during excavation and support stage, the extent of the movements did not. Thus, for the case where significant movements develop during wall installation, perhaps a more reasonable approach is to normalize the extent of settlement by wall depth as opposed to the excavation depth. Additional case studies should be evaluated to see how general is this response. In further evaluating the settlement response of the school, it is helpful to evaluate the settlement response on days corresponding to (i) Day 73 first observation of interior

cracking; (ii) Day 131 representing conditions when the State Street excavation was to final grade, and (iii) Day 365 a time when all movements had stopped. Figures 5-3, 5-4, and 5-5 presents 2D and 3D settlement contours on Day 73, Day 131, and Day 365, respectively. These figures show that the general patterns of the settlement contours were established as of Day 73 and remained similar throughout the project. The 3D portions of the figures show the settlement response at the foundation level of the building.

Figure 5-3. Settlement Contours of the Warde School on Day 73 From Figure 5-3, it is seen that the settlement contours were slightly asymmetrical. This is due to the location of the excavation with respect to the Warde School. The approximate wall location given on the figure shows that the southern limits of the

excavation along State Street only extended to the midway point of the south portion of the school. Also, the contour patterns observed on Day 73 were solely in response to excavation along State Street because no significant excavation had began on Chicago Avenue at this time. As a result of the location of the excavation and the governing excavation activities, the contours formed a bowl shaped pattern that was centered north of the buildings center. The bowl tilts fairly uniformly towards the west and is approximately symmetrical between columns C2 and C6 (see inset of Figure 5-1). The back end of the bowl was located approximately 10 m to 15 m from the western face of the school. This distance corresponds to the location of the north-south hallway of the school. The inset of Figure 5-1 shows that the east wall of the north-south hallway consists of columns C2, C4, C6, and C8, and the west wall consists of columns C1, C3, C5, and C7. The west wall of the hallway was located at about the 4 mm contour in Figure 5-3. Overall the north side of the school showed a uniform taper on Day 73, progressing from about 1 mm on the east end of the school to about 6 mm on the west end. During this same period, the settlement at the central portion of the schools west face was approximately equal to 10 mm. The north and south segments of the west wall

both sloped toward the middle, whereas the entire wall tilted slightly to the north.

Figure 5-4. Settlement Contours of the Warde School on Day 131 Figure 5-4 gives the settlement contours at the end of excavation along State Street on Day 131. The figure shows that the bowl shaped pattern became asymmetrical in response to increased settlement towards the north end of the school. The center of the settlement trough moved further north and the school began showing a pronounced tilt toward the northwest. It is observed that although the general shape of the contours was no longer symmetrical, the eastern extent of the trough area remained relatively similar to what was observed for Day 73. The contours gradually slope toward the west up to

the north-south hallway location and afterwards become dramatically steeper. On Figure 5-4, the west wall of the north-south hallway was located at the 10 mm contour. The excavation along Chicago Avenue was advanced below the first level bracing (elevation +2.1 m CCD) on Day 131. Figure 5-4 shows that a distinct break point in the slope along the north wall formed approximately 21 m from the west wall of the school. This break point appears to correspond to the zone of influence of the excavation activities along Chicago Avenue, which extended approximately 11 m north of the secant pile wall. The lowest point on the contours on Day 131 was located at settlement point W10 (see Figure 5-1 inset). Although this was the location of the pit for Escalator #4, excavation activities for the escalator did not start until Day 162. The north and south segments along the west side of the school both showed increased total settlement on Day 131, but the differential settlement along the north segment was roughly 4 mm as compared to approximately 8 mm of differential settlement along the south segment. Figure 5-5 shows that the 3-D effects became more pronounced as the excavation along Chicago Avenue was further advanced and completed. Settlement of the north wall of the school increased, and the northwest corner of the school settled approximately three times more than the southwest corner, thus producing greater overall building tilt towards the northwest. Again the distinction between the settlement contours gradually sloping towards the west and the point whereupon the contours steeply slope towards the west is the north-south hallway. In the figure, the west wall of the north-south hallway was located at the 14 mm contour.

All movement had stopped as of Day 365. Thus, Figure 5-5 shows the final settlement contours. The settlement trough had deepened, but did not widen in extent. An inflection point was observed along the south segment of the west wall. This inflection point approximately coincides with the southern extent of the excavation activities.

Figure 5-5. Settlement Contours of the Warde School on Day 365 The settlement profile in Figure 5-5 suggests that the southeast corner of the excavation had the effect of reducing the total settlement by an amount proportionate to the distance from the corner. Several researchers (Lee et al., 1998; Ou and Shiau, 1998; and Ou et al., 1996) have investigated corner effects on the behavior of diaphragm walls. However, their work primarily pertained to corner effects on the lateral movements of the walls. No work was found in the literature pertaining to corner effects on the settlement response of building adjacent to support walls. Although settlement can be

implied from lateral movements for no-volume-change conditions, there are several complicating factors, which preclude applying the lateral deformation results directly to building responses at corners of excavations. These factors include building type and configuration, and building orientation with respect to the excavation, depth of excavation, and extent of excavation. Additional research that considers these factors is required.

Figure 5-6. Deflection Mode of Building in Response to Deformation (after Mair et al., 1996)

5.1.2 Zones of Sagging and Hogging

The contours presented in Figures 5-3, 5-4 and 5-5 show that the school experienced both hogging and sagging modes of deformation. These terms are defined in Figure 5-6. Sagging implies that the settlement profile is concave upward, whereas hogging indicates that the settlement profile is concave downward. The figure shows the building response idealized as an elastic deep beam in bending. The sagging mode produces tension along the bottom fiber and compression along the top fiber. The hogging mode produces tension along the top fiber and compression along the bottom. However, actual buildings differ from beam theory in that the surrounding soil at the foundation level provides lateral restraint in the sagging mode. This restraint can be idealized by assuming the neutral axis is at the bottom fiber, effectively limiting tension in the beam to direct tensile strains with no bending-induced tension. The hogging mode of deformation is not the reciprocal of the sagging mode. There is no external lateral restraint at the roof level for most buildings. Consequently, the neutral axis remains at

the centroid of the building/beam model in the hogging deformation.

Figure 5-7. Regions of Sagging and Hogging The upper half of the beam is in tension and the tensile strain resulting from bending increases with distance from the neutral axis. The definitions of the deflection ratio (sagging) and the deflection ratio (hogging) are included in Figure 5-6. These ratios are simply the deflection ratios in the sag and hog zones, respectively. The deflection ratio (sagging) is a measure of curvature of a member whose deformed shape is concaved up. The deflection ratio (hogging) is a measure of curvature of a member whose deformed shape is concaved down.

Figure 5-7 shows areas of the building where hogging and sagging occurred. These areas were defined based on the settlement profiles shown in Figure 5-3 and Figure 5-5. Figure 5-7a shows the sagging and hogging zones on Day 73 when damage was first observed in the Warde School. Figure 5-7b show the hogging and sagging zones on Day 116, which was when excavation was completed on the east side of State Street. Figure 5-7c gives these zones based on the post construction survey on Day 365. The figures show that the sagging mode of deformation developed within the central portion of the west side of the school. Hogging occurred elsewhere. The north-south extent of the sagging zone remained relatively constant between Day 73 and Day 365. After the excavation had bottomed out on Day 116, movements continued toward the excavation as a result of reducing the bending stiffness of the wall and creep of the soft clay. During this time, the east-west length of the sagging zone gradually increased until the maximum east-west extent of the sagging zone was approximately equal to the depth of the secant pile wall. To illustrate further the deformation patterns throughout the building, several settlement profiles are plotted in Figure 5-8 and Figure 5-9. Figure 5-8 shows east-west settlement profiles taken along the north wall in an area of hogging (5-8a) and through the Inclinometer 1 location in the region of sagging (5-8b). Sagging was limited to an average distance of about 12 m from the west edge of the building, the location of the west wall of the north-south hallway. The deflection ratios were computed from the profiles taken through the interior of the building, in the east-west direction (5-8b). The sag zone included settlement points W11, W10, and C1. The hog zone started at point C1 and included points C2 and W19. Table 5-1 shows the deflection ratios for each of

the construction days given in Figure 5-8b. The deflection ratios (hogging) were negligible until some time near Day 108 and did not approach the initial deflection ratio (sagging) value until after the escalator base slab was poured.

Figure 5-8. East-West Settlement Profiles Figure 5-9 shows north-south settlement profiles taken along the west wall of the building (5-9a) and along the west most interior column line (5-9b). The figure also shows that the interior of the building experience a sagging mode of deformation, while the ends experienced hogging. Hogging was particularly pronounced along the north wall where the effects of the excavation along Chicago Avenue had the most impact on the school. The school transitioned from the sagging zone to the hogging zone near settlement point C1. This point is the northern end of the interior column lines.

Table 5-2 presents the deflection ratios computed from the data given in Figure 5-9b. The sag zone included settlement points C3, C5, C6, and W6. The hog zone included settlement points W13, C1, and C3. From the table and from Figure 5-9b it is seen that the interior hog along

Figure 5-9. North-South Settlement Profiles Table 5-1. Maximum Deflection Ratios: East-West Direction

Day 11 73 108

Deflection Rate (Sag) 0.29 x 10


-3

Deflection Rate (Hog) 0.03 x 10-3 0.06 x 10-3 0.23 x 10-3

0.32 x 10-3 0.63 x 10-3

116 177 365

0.68 x 10-3 0.81 x 10-3 0.79 x 10-3

0.23 x 10-3 0.31 x 10-3 0.29 x 10-3

the north-south plane was negligible throughout the project. Also, the interior northsouth sag was about half that of the interior east-west sag. Figure 5-9b shows that the overall tilt of the building progressively increased towards the north. Hogging was the primary deformational mode along the north side of the Warde School. Yet, the deflection ratio (hog) values along the north side were an order of magnitude less than the deflection ratio (sag) values along the west side. It can therefore be concluded that the sagging zone was the most critical deformational mode for the Warde School. Table 5-2. Maximum Deflection Ratios: North-South Direction

Day 73 108 131 245 312

Deflection Rate (Sag) 0.12 x 10


-3

Deflection Rate (Hog) 0.048 x 10-3 0.036 x 10-3 0.069 x 10-3 0.073 x 10-3 0.073 x 10-3

0.21 x 10-3 0.25 x 10-3 0.3 x 10-3 0.3 x 10-3

5.2 DISTORTIONS Distortion is defined herein as the differential deformation between two points, , divided by the distance them, . This definition of distortion is the measure of the shearing strain of a member. Angular distortion, , can be obtained from the distortion by subtracting the rigid body tilt, , from the measured settlement. Thus, angular distortion is a measure of the rotation of a member at the support relative to the rigid body tilt. It is noted that the rigid body tilt of the Warde School was 1/15,000 when final

grade was reached on Day 116. At the end of the project, the rigid body tilt was 1/3900. Given these small values, there is no significant difference between distortion and angular distortion for this building. This is to be expected for a building that is large enough such that only parts of it are affected by movements induced by excavationrelated activities. Therefore, for practical purposes the distortions and angular distortions were the same for this project.

5.2.1 Computing Distortions from Settlement and Inclinometer Data

Figure 5-10. Distortion from Settlement and Inclinometer Data

The distortion data presented herein was computed from settlement and lateral deformation data. Figure 5-10 presents the procedure used to compute distortions. The distortions based on settlements were computed between distances 4.5 m and 12 m from the center of the secant pile wall (marked by L2 in the figure). Distortions based on inclinometer data were computed at the same relative location after the inclinometer data was rotated according to the procedure described in Section 4.3.4.3. Figure 4-24 in Chapter 4 showed that there was good agreement between the settlement and the inclinometer data, at least at the Inclinometer 1 location. This suggests that inclinometer data can be used to evaluate the distortions induced in adjacent structures when the excavation is made through saturated clays. The implicit assumption is that there is no volume change induced in the soil. This condition occurs in many excavations through soft clay. Agreement between distortions computed with inclinometers data and distortions computed with settlement data cannot be expected if volume changes occur in the soil beneath the structure. For example, if the Warde School was founded on footings a meter or two below ground surface, and hence on several meters of granular fill, the volume changes within the fill would alter the distribution of settlements from what would be indicated from the inclinometer data. Similar disparities would be observed if significant consolidation settlements occurred. Hence, when excavating through soft clay, inclinometer data can be used to estimate distortions under an adjacent building, at least when the movements are small, as would be the case when a stiff system is used to support a well-constructed excavation. Large movements would result in localized strains within the soft clay and agreement between

lateral movements measured at the wall and ground surface settlement points would be poor (Finno et al, 1989). The agreement between the inclinometer and settlement data, when the movements are small, is particularly useful in assessing potential damage to structures that an abundance of settlement data are rarely available for projects because of the cost and, in many instances, the adversarial relation between the building owner and constructor. In many projects, inclinometer data are available and can be used to estimate distortions when no volume changes occur within the soil adjacent to the wall.

5.2.2 Development of Distortions During Construction Distortion versus time was plotted along the west side and north side of the Warde School to illustrate the agreement between distortions computed with inclinometer data and settlement data and to evaluate the distortion response of the building at specific locations. Figure 5-11 plots distortion with respect to time along the west side of the school, in the plane perpendicular to the excavation face. Also shown is a summary of construction at that location. The settlement-derived distortions were computed from the differential settlement between two settlement points (W10 and C1, W9 and C3, and W8 and C5) divided by the distance between the two points. The distortions from Inclinometer 1 and Inclinometer 2 were computed as the difference in lateral deformation between depths of 9 m and 17 m, divided by a length of 8 m. Note that all distortions presented in this figure are located in the sagging zone (Figure 5-7). The figure shows that the distortions computed from the Inclinometer 1 data matched those computed from settlement points W10 to C1 very well. However, the

Inclinometer 2 distortions also matched the W10 to C1 distortions. It was expected that the Inclinometer 2 distortions would more closely agree with the distortions of the adjacent settlement points, W8 to C5. These distortions most likely did not agree because of the geometry of the excavation relative to the location of the building. The corner of the excavation has the effect of reducing the building settlement. This can be seen in the 3D contours of Figures 5-3, 5-4, and 5-5. In each of these figures, the settlement on the south end of the site is much less than the settlement on the north end. Consequently, settlement points W8 and C5 experienced less settlement than settlement points W10 and C1. This resulted in the W8 to C5 distortion being less than the W10 to C1 distortion. The data suggest that the distortions of the Warde School are a function of building size and orientation with respect to the excavation.

Figure 5-11. Distortion versus Time Perpendicular to the West Side of the Warde School The effect of the corner on the distortions of the school is further demonstrated in Figure 5-11 by comparing the distortions computed from the survey data. The magnitudes of the distortions are shown to be functions of the distance from the southeast corner of the excavation. The largest distortions were at the W10 to C1 location, which was about 23.8 m from the corner of the excavation. The distortions were smaller at the W9 to C3 location, which was approximately 15.2 m from the corner of the excavation and smaller still at the W8 to C5 location, which was approximately 7 m from the corner. Figure 5-11 shows that the distortions were as much as approximately 0.001 (1/1000) in response to installing the secant pile wall. The distortion essentially remained constant until the excavation was advanced through the stiff clay crust to a depth of about 5.5 m. The first interior crack was observed on Day 73 when the distortion based on Inclinometer 1 data was 0.0011 (1/920). Thereafter, the distortion increased as the excavation was lowered until final excavated grade was reached between Day 108 and Day 116. The figure shows that installing the struts and tiebacks had little effect on the distortion rate. The first crack in the marble faade in the entranceway foyer was observed on Day 108. The distortion at the W9 to C3 location was 0.0018 (1/555) on Day 108. This is roughly the location of the marble entrance foyer. After completion of the excavation on Day 116, the distortion continued to increase, but at a slower rate than during excavation. These continuing distortions were caused by the reduction of the wall stiffness and by creep of the clay during the subway station renovation and backfill

activities. By Day 177, the distortions became approximately constant. The maximum distortion value computed from the inclinometer data was approximately 0.0033 (1/300). The maximum value calculated from the settlement data (W10 to C1 location) was slightly less at 0.00317 (1/315). Figure 5-12 presents the distortion versus time data for the north and south segments of the exterior west wall of the school. These distortions are in the north-south direction along the west elevation. Distortion was computed for the north segment using settlement points W12 and W11, and distortion in the south segment were computed using settlement points W7 and W8. These distortions represent the hogging zones on the west side of the school.

Figure 5-12. Distortion versus Time along the Exterior West Wall of the Warde

SchoolDistortions at the north and south segment locations were impacted greatly by the end effects of the excavation. The southern limit of the excavation was approximately at the settlement point W7 location, but the excavation extended north and east past the northwest corner point, W12 (see inset of Figure 5-12). As a result of the excavation end effects, the total excavation-induced settlements were greater at the north segment than at the south segment, but the differential settlements were greater at the south segment than at the north segment. This is also seen in the 2-D contours of Figures 5-3, 5-4, and 5-5. The excavation end effects ultimately resulted in the excavation-induced distortions between settlement points W7 and W8 being larger than those between settlement points W12 and W11. The distortion for the north segment was approximately 0.00026 (1/3900) and approximately 0.00034 (1/2900) in the south segment in response to installing the east secant pile wall. It is observed that the north segment showed some variability in the initial data. This variability was due to utility relocation activities at the northwest corner of the school, which had only a temporary effect on the distortional response. The distortions for the north and south segments remained fairly constant and at similar levels until Day 70 because the excavation had not proceeded below the soft clay layer. During this period the excavation end effects had no impact on the distortions. However, as the excavation was advanced into the soft clay between Day 70 and Day 73, excavation end effects have a significant impart on the distortions. The distortions for the south segment increased to 0.00054 (1/1850) whereas the distortion for the north segment began to decrease. The first cracks in the south segment of the exterior west wall were observed on Day 127, at which time the distortion in the south segment

reached 0.00108 (1/925). Cracks in the north segment were first observed on Day 129 and the distortion reached 0.00062 (1/1625). It is noticed that although the cracks occurred at about the same time in the north and south segments, cracking was observed in the north segment at a distortion almost half of that observed in the south segment. The north segment may have cracked at lower distortions than observed for the south segment because of the presence of direct tensile strains caused by the northward components of movements of the school in response to excavation activities along Chicago Avenue. These direct tensile strains are additive to the shear-induced strains created by the distortion of the foundation. The distortions in the northwest corner of the Warde School are presented in the northsouth direction in Figure 5-13 and in the east west-direction in Figure 5-14. Distortion in the north-south direction was computed from the north-south Inclinometer 5 data and from the differential settlement data between settlement points W13 and C1. The eastwest distortions were computed from the east-west Inclinometer 5 data and the differential settlement data from settlement points W14 and W13, and W13 and W12. The Inclinometer 5 distortions were computed as the difference in lateral deformation between depths of 8 m and 16 m, divided by a length of 8 m.

Figure 5-13. Distortion versus Time Perpendicular to the North Wall of the Warde School

Figure 5-14. Distortion versus Time Parallel to the North Wall of the Warde School Figure 5-13 shows that the initial Inclinometer 5 north-south distortions were almost exclusively in response to the installation of the secant pile wall along Chicago Avenue from Day 70 to Day 79. The distortions show a sharp increase during the installation of the wall and then began to flatten after the wall was completed. The Inclinometer 5 north-south distortions were approximately 0.0008 (1/1290) on Day 79. The distortions gradually increased, in response to the excavation along State Street, until the excavation along Chicago Avenue was advanced into the soft clay layer (Day 170). Afterwards, the distortions were primarily a result of the excavation activities along Chicago Avenue. The distortions computed from point W13 to C1 follow the same basic trend. The maximum distortion computed from the Inclinometer 5 north-south

data was 0.0016 (1/625) and the maximum distortion attained at the W13 to C1 location was 0.00143 (1/700). These distortion levels were observed during the period corresponding to completion of the station renovations and beginning the backfill activities along the west side. No further north-south distortions were observed after backfill began along State Street (Day 225), which further demonstrated the influence of the excavation activities on these distortions. Figure 5-14 shows that between Day 70 and Day 79 the east-west distortions of Inclinometer 5 developed in response to the installation of the secant pile wall along Chicago Avenue. Recall that Inclinometer 5 was located quite close to the secant pile wall and its movements were affected by this relative position. It is likely that the eastwest movements at this time did not represent building movements. After Day 116, the inclinometer distortions began to increase primarily in response to the open excavation along State Street. The increase in the inclinometer distortion is attributed to the State Street excavation because between Day 116 and Day 170 the excavation along Chicago Avenue had not been advanced beneath the clay crust. Further evidence that these distortions were mainly caused by the State Street excavation activities is provided by the fact that increases in the inclinometer distortions stopped once backfilling began along State Street on Day 225. Distortions computed from points W12, W13, and W14 follow similar trends, except that they were not influenced as much by installing the secant pile wall. Distortions at W14 to W13 did not increase until approximately Day 73 whereas W12 to W13 distortions did not increase until after Day 86. These rapid increases occurred after installation of the secant pile wall and more accurately reflect the response of the

building to the construction activities than the east-west data from Inclinometer 5. From Figure 5-14, it is seen that the settlement-derived distortions increased from approximately 0.00021 (1/4700) to as much as 0.00085 (1/1175) as the State Street excavation was advanced from the from the top of the soft clay layer to the final grade. The W12 to W13 distortions essentially paralleled the inclinometer distortions after excavation was completed along State Street (Day 116), but were approximately 90 percent its magnitude. In any case, the data shown in Figures 5-11 through 5-14 show that inclinometer data can be used to reliably compute distortion in buildings, when excavation are made through saturated clay.

5.2.3 Comparison of Deflection Ratios and Distortions Table 5-3 presents a comparison of deflection ratios and distortions computed in the east-west direction. The deflection ratios used for the table are those given in Table 5-1. Note that the first set of distortions given in the table is not the maximum distortions for the sagging and hogging zone reported in the previous section. The distortions in the sagging zone were computed from settlement points C1 and W11, and the distortions in the hogging zone were computed from settlement points C1 and W19. Boscardin and Cording (1989) found that angular distortions were typically 2 to 3 times the deflection ratios. The data given for the sag zone indicates this ratio varies between 2 and 4, while the data in the hog zone shows that the angular distortions were approximately 3 to 7 times the deflection ratios. Tablalong the west side. No further north-south distortions were observed after backfill began along State Street (Day 225), which further demonstrated the influence of the

excavation activities on these distortions. Figure 5-14 shows that between Day 70 and Day 79 the east-west distortions of Inclinometer 5 developed in response to the installation of the secant pile wall along Chicago Avenue. Recall that Inclinometer 5 was located quite close to the secant pile wall and its movements were affected by this relative position. It is likely that the east-west movements at this time did not represent building movements. After Day 116, the inclinometer distortions began to increase primarily in response to the open excavation along State Street. The increase in the inclinometer distortion is attributed to the State Street excavation because between Day 116 and Day 170 the excavation along Chicago Avenue had not been advanced beneath the clay crust. Further evidence that these distortions were mainly caused by the State Street excavation activities is provided by the fact that increases in the inclinometer distortions stopped once backfilling began along State Street on Day 225. Distortions computed from points W12, W13, and W14 follow similar trends, except that they were not influenced as much by installing the secant pile wall. Distortions at W14 to W13 did not increase until approximately Day 73 whereas W12 to W13 distortions did not increase until after Day 86. These rapid increases occurred after installation of the secant pile wall and more accurately reflect the response of the building to the construction activities than the east-west data from Inclinometer 5. From Figure 5-14, it is seen that the settlement-derived distortions increased from approximately 0.00021 (1/4700) to as much as 0.00085 (1/1175) as the State Street excavation was advanced from the from the top of the soft clay layer to the final grade. The W12 to W13 distortions essentially paralleled the inclinometer distortions after excavation was completed along State Street (Day 116), but were approximately 90 percent its magnitude. In any case, the data shown in Figures 5-11 through 5-14 show that

inclinometer data can be used to reliably compute distortion in buildings, when excavation are made through saturated clay. 5.2.3 Comparison of Deflection Ratios and Distortions Table 5-3 presents a comparison of deflection ratios and distortions computed in the east-west direction. The deflection ratios used for the table are those given in Table 5-1. Note that the first set of distortions given in the table is not the maximum distortions for the sagging and hogging zone reported in the previous section. The distortions in the sagging zone were computed from settlement points C1 and W11, and the distortions in the hogging zone were computed from settlement points C1 and W19. Boscardin and Cording (1989) found that angular distortions were typically 2 to 3 times the deflection ratios. The data given for the sag zone indicates this ratio varies between 2 and 4, while the data in the hog zone shows that the angular distortions were approximately 3 to 7 times the deflection ratios. Table 5-3. Comparison of Deflection Ratio and Distortions

Sag Day (C1-W11) 11 73 108 116 177 365 Day 0.67 x 10-3 0.67 x 10-3 1.58 x 10-3 1.64 x 10-3 2.03 x 10-3 2.06 x 10-3 (C1-W11)/ ( /L) 2.3 2.1 2.5 2.4 2.5 2.6 Sag

Hog (C1W19
3

(C1W19/L) /( /L) 3.1

0.08 x 10-

0.2 x 10-3 3.2


3

0.66 x 100.67 x 101.04 x 101.02 x 10-

2.9 2.9 3.3 3.5 Hog

(C1-W10) 11 73 108 116 177 365 1.05 x 10


-3

(C1-W10) ( /L) 3.7 3.4 3.8 3.7 3.8 4


3

(C1-C2)/( (C1-C2) /L) 0.16 x 100.33 x 101.32 x 101.32 x 101.81 x 101.20 x 106 5.5 5.8 5.8 5.8 7.1

1.09 x 10-3 2.42 x 10-3 2.54 x 10-3 3.11 x 10-3 3.17 x 10-3

5.2.4 Distribution of Distortions

Figure 5-15. Distortion Contours Figure 5-15 presents distortion contour maps developed from all the building foundation settlement data. The contours are given for Day 73 (the onset of damage), Day 116 (end of excavation along State Street), and Day 365 (the post-construction damage survey). These contour data will not necessarily be the same as the contour data

given in Section 5.1.1 because these data include interpolations of settlement data in areas were no settlement was measured. The actual measured settlement and the interpolated settlement are used to produce discrete points of distortions. The distortions between these discrete points are also interpolated to produce the contour lines. In the previous sections, the distortion data were produced from the two nearest settlement points in the area of question. It is apparent from the figure that the highest levels of distortion corresponded to the region of the school that experienced a sagging mode of deformation. The boundary between the hogging and sagging regions has been superimposed onto the distortion maps. The area of higher distortion is parallel to the north-south hallway and is located between the west wall of the hallway and the west wall of the school. In general, the distortions at the onset of damage (5-15a) were higher on the south side of the school than were on the north side because the excavated depth was greater at the south end of the excavation at this time. The general pattern of distortions at the end of excavation along State Street (5-15b) was similar to that observed for Day 73. However, the maximum distortions moved further north, between survey points W10 and W9 (see inset of Figure 5-12) and the distortions on the north side of the school were larger than those on the south side. The redistribution in the distortions reflected the progress of the excavation, and the area of peak distortion corresponded to the area of maximum settlement. The peak distortion at this time was about 0.0019. The distortion contours on Day 365 had a similar pattern as the contours for Days 73 and 116. Similar to the Day 116 contours, the peak distortion moved further north. The peak distortion was

located adjacent to survey point W10 and was approximately 0.0031. Distortions along the north wall of the school partially reflect the excavation along Chicago Avenue.

5.3 TILT OF FOUNDATION WALLS Tiltmeters were installed along the perimeter basement walls of the Warde School to measure inclination of the foundation wall directly. The responses of the tiltmeters to the excavation and construction activities are given in Figure 5-16. The tiltmeters were mounted on base plates that were attached to the wall, and measured tilt in the directions noted in the building inset of Figure 5-16. The devices were typically placed about 1.5 m above the basement floor.

Figure 5-16. Tilt Response of the Warde School Along the west wall of the school, T7 and T8 indicated that the central portion of the wall tilted to the east in response to the settlement pattern shown by points W11 and W10 in Figure 5-8b. This pattern is consistent with the sagging deformation mode in the

central portion of the school. The tilt at the south end of the west wall indicated by T9 was negligible, because the excavation along State Street did not extend that far south. Along the north wall, T3 indicated that the wall rotated to the north in response to the settlement pattern shown by W13 and C1 in Figure 5-9b. This pattern is consistent with the hogging deformation mode in the north portion of the school. Tilt at the east end of the north wall indicated by T1 was negligible, because it was far from the excavation along Chicago Avenue. The small tilts recorded between Days 60 and 90 at this location were related to utility relocation. The simultaneous excavations along State Street and Chicago Avenue affected the patterns of wall rotations in the northwest corner of the school. The north-south components of tilt recorded in T3 and T6 clearly show that tensile strains were induced in the structure at the northwest corner. These direct tensile strains contributed to the cracking on the north segment of the west wall, as suggested in Section 5.2.2. Tiltmeter T3 shows the north wall of the school rotating towards the north, but Tiltmeter T6 shows the north segment of the west wall rotating towards the south. The east-west tilt recorded by T4 and T5 reflect the hogging deformations illustrated in Figure 5-8a near points W12 and W13. In general, the tiltmeter data agreed with the trends in the settlement data.

5.4 EXCAVATION-RELATED DAMAGE Excavation-related damage within the Warde School was primarily in the form of cracking of interior infill walls and exterior masonry walls. A summary of the significant construction activity along State Street and building damage is presented in Table 5-4. The construction activity in the table is given for the area adjacent to the

Inclinometer 1 location. As indicated in the table, no damage was observed in the Warde School during the secant pile wall installation activities. Damage was first observed on Day 73 when the cross-lot struts were in place and the excavation had extended into the soft clay layer at 5.5 m below ground surface. Cracking was first observed in rooms along the west side of the school on all three floors. The initial damage mostly consisted of diagonal hairline cracks about 300 to 500 mm long in nonloading bearing walls. It was observed that the occurrences of cracking were greater on the second and third levels than on the first level. Table 5-4. Summary of Construction Activity and Building Damage

Stage Day 1 2 0 to 11

Construction Activity Secant pile wall installation

Building Damage

60 to Install cross-lot struts 74 73 Interior cracks observed; hairline cracks in infill Excavate below first tieback walls concentrated in second and third floors; level second floor door replaned Cracks in mortar and limestone faade in exterior north wall at the west end of the school; existing cracks extend, maximum width is 1 mm Install first level tiebacks Tension first level tiebacks Install second level tiebacks New cracks observed in first floor wall panels; existing cracks widen and extend Tension second level tiebacks Cracks observed in marble faade of entranceway foyer; hairline cracks in north foundation wall in the west corner of the school; existing cracks widen and extend Cracks observed in first floor wall panels

78 79 87 98 99 105

108

110 116 127 and 129 137 to 140 142 151 172 to 177 207 225 to 310 258

Chip face of secant pile wall to flange from EL 4.3 m CCD to EL -4.3 m CCD Excavate to final grade Existing cracks widen and extend on all Step cracks observed in the south and north segments of the west exterior wall Chip face of secant pile wall to flange from EL -4.3 m CCD to EL -7.6 m CCD Cracks observed in floor tiles of cafeteria; second floor door replaned Movements exceeded 32 mm; diagonal cracks observed in first floor wall in cafeteria; existing cracks widen and extend Place concrete for escalator pit slab New crack observed in marble faade in entranceway foyer Place backfill Remove struts

The maximum lateral deformation on Day 73 was 12 mm. As the excavation was lowered to a depth of 12 m on Day 108 and the maximum lateral movement increased to 22 mm, cracks developed in the marble faade in the entranceway foyer on State Street. Also, hairline cracks were observed cracks in the basement, along the west end of the north foundation wall. The general excavation along the west side of the school was completed on Day 116. At this time, previously observed cracks widen and extended on all levels. During station renovation and backfill, only a few instances of new damage were observed; existing cracks generally became larger during this time. Step cracks were observed in the mortar of the south and north segments of the exterior west wall on Days 127 and 129, respectively.

5.4.1 Damage on the First Floor Level

Figure 5-17 Damage Locations and Typical Cracks on the First Level Figure 5-17 presents a plan view of the first level and identifies the areas where damage was observed. The figure shows the locations of cracks at Day 73 and Day 365. These days reflect the onset of cracking and the post-construction damage survey dates, respectively. The damage location plan in the figure shows that the majority of the damage on the first floor developed in east-west trending walls. The first observation of cracking damage on the first level developed in the south wall of the assistant principals secretarys office (5-17a). The largest crack in this wall was a diagonal crack

radiating upwards and towards the west at a 45-degree angle. Also, a horizontal crack was located near the bottom of the wall. On Day 73, the diagonal crack was approximately 610 mm long and had a maximum width of about 2 mm to 3 mm. The horizontal crack was about 405 mm long and had a maximum width of 1 mm to 2 mm. By Day 365, the widths of these diagonal and horizontal cracks had increased to about 6 mm and 4 mm, respectively. A vertical crack developed in Room 103 (5-17b) during the excavation activities. The vertical crack was approximately 610 mm long and emanated from the top corner of the connection between the beam and the column.

Figure 5-18 Damage Locations and Typical Cracks in the Entranceway Foyer

Figure 5-18 presents additional first floor damage that occurred to the marble faade on both the north and south walls of the entranceway foyer. The cracks in the north wall (518a) were typically diagonal cracks at the corners of the marble panels and at the corners of fixtures inset into the wall. Most of these cracks were inclined at about 45 degrees and radiated upwards towards the west. This pattern of cracking indicates that the north wall cracks were caused by shear distortions. The damage along the south wall (5-18b) consisted of long horizontal cracks that tended to follow the preferential pathways created by grain variations in the marble panels. The horizontal cracks all emanated from the ends of the panels and radiated inwards. The south wall cracks were most likely the result of out-of-plane bending, or possibly out-of-plane shear, because the upper ends of the cracked panels were not flush with lower ends. The initial cracks in the foyer were detected in the north and south walls on Days 108 and 109, respectively. Cracks in the north wall tended to develop in the upper half of the wall, while cracks in the south wall tended to develop in the lower half of the wall. The marble panels for both walls became misaligned in response to the excavation-related movements. A crack, 3 mm wide and 102 mm long, developed on Day 207 in the marble base near the floor along the south end of the foyer, well after the excavation along State Street had been completed (Day 116) and the Escalator #4 pit slab had been poured (Day 177).

Figure 5-19. Evaluation of Damage on the First Level of the Warde School Figure 5-19 presents an overlay of the distortion contours, the hogging and sagging regions, and the crack locations for the first level. The evaluations are given for Day 73 (5-19a) and Day 365 (5-19b). It is apparent from the figure that cracking occurred in areas of highest distortion, with most of the cracking occurring in the sagging zone. The distortion for the south wall of the secretarys office on Day 73 was 0.00090 (1/1100). No damage was observed in the north partition and infill walls of the Assistant Principals office. Figure 5-19b shows that the distortions were slightly less at the south foyer wall than they were at the north wall, and would help explain why cracks were observed in the north wall of the entranceway foyer before the south wall.

5.4.2 Damage on the Second Floor Level

The damage location plan and characteristic cracks for the second floor level are presented in Figure 5-20. The figure shows that more damage was noted on the second level than on the first level. Furthermore, the damage first observed on Day 73 occurred almost exclusively between the west wall of the school and the west wall of the northsouth hallway.

Figure 5-20. Damage Locations and Typical Cracks on Second Level The initial damage on the second level typically consisted of horizontal cracks emanating from doorways within the upper portions of the walls. The damage that developed on the second level later in the project were vertical and diagonal cracks. In the three northern classrooms on the second floor (Room 210, Room 208, and Room

207), the damage consisted of horizontal cracks in the south walls near the doors (see Figure 5-20a), vertical cracks in the north walls, and diagonal cracks in the east walls. In the rooms on either side of the library (Room 206 and Room 202), the cracks consisted of vertical cracks in the corners of the east walls, horizontal cracks at the beam-wall interface along the tops of the north walls and vertical cracks also along the tops of the north walls. Some vertical cracks were observed emanating from the corners of doors on the second level. These cracks were at locations where doorframes had racked. In some instances, the racking of the doorframes was bad enough to require replaning of the doors. The partition wall in the library apparently shifted to the east, which caused the floor tiles near the door to bubble up. These tiles had to be removed and resized. Also, several horizontal cracks were observed in the north walls of the east-west hallway.

Figure 5-21. Evaluation of Damage on the Second Level of the Warde School

Figure 5-21 presents an overall evaluation of the damage on the second floor level of the school. Figure 5-21a shows that the damage first observed on Day 73 occurred primarily within the central interior of the school where sagging had developed in response to the excavation along State Street. Initial damage observed in the hogging zones occurred at lower distortion levels than were observed in the sagging zones. Two of the incidents of cracking outside the sagging zone occurred near the boundary of the sag and hog region. The damage in the northeast corner of Room 208 (see Figure 5-20 damage plan) was a vertical hairline crack in the plaster coating of the column. The first observed damage in the east-west trending walls in the library developed in a partition wall, which spanned an area of high distortion. From the contour on Day 73, the distortion at the foundation level below the partition wall was about 0.0009 (1/1100). Cracking developed in the far north infill wall of the library within a few days after Day 73. The distortion contours in Figure 5-21b show the region of highest distortions remained located between the west wall of the north-south hallway and the west wall of the school. Consequently, most of the damage observed in the central portion of the school developed within this area. Less damage developed east of the north-south hallway in the central portion of the school. Figure 5-21b shows that damage developed in the south wall of Room 208 and along the north wall of the east-west hallway in the corresponding location in the form of horizontal cracks in the upper portions of the walls. 5.4.3 Damage on the Third Floor Level The damage location plan and characteristic cracks for the third floor level are presented in Figure 5-22. The figure shows that the damage was as extensive on the third level as

that observed on the second level. Also, similar to the second floor level, the damage first observed on Day 73 occurred primarily within the central interior of the school. Comparing Figures 5-20 and 5-22 it is seen that the distribution of damage on the second and third levels were similar. However, the incidents of first observed damage On Day 73 were less on the third level than on the second level.

Figure 5-22. Damage Locations and Typical Cracks on Third Level The damage observed on the third level were primarily diagonal cracks (Figure 5-22). The third level cracks also tended to be wider and longer than the cracks observed on the first and second levels. On Day 78, five days after the first interior cracks were

observed, typical crack widths on the third level were approximately 0.75 mm. The crack shown in Figure 5-22 was 2-mm wide on Day 78. These widths are compared to a typical width of 0.2 mm on the first and second levels during this same time. Throughout subsequent construction, most cracks on the third level lengthened, but widened only slightly. Cracking damage in the east-west trending walls were characterized by large diagonal cracks that often occurred in the middle of the walls and radiated upwards at 45 degrees towards the west. Diagonal cracks that were smaller in width were also observed emanating from the corners of electrical outlets and in-wall cabinets in the east-west trending walls. Corner diagonal cracks were observed in the south walls of rooms north of the north-south hallway and in the north walls of the rooms south of the north-south hallway. Vertical corners cracks also developed at the interface between the south load bearing wall and the infill wall. Vertical cracks on the third level were typically located in the north-south trending walls on the eastern end of the school (the portion from the north-south hallway to the eastern face). In particular, vertical cracks developed at the top of the westwall in Room 301 and at the top of the east wall of the north-south hallway, between the Boardroom and Room 301.

Figure 5-23. Evaluation of Damage on the Third Level of the Warde School Figure 5-23 summarizes the damage to the third level of the school. All damage first observed on Day 73 on the third level occurred within sagging zone. More damage occurred at the southeast end of the school at the third level than was observed on the first and second levels. This is especially the case along the east wall of the north-south hallway. Because this area was in the hogging zone, bending stresses induced in the structure from the hogging mode were higher in the upper levels.

5.4.4 Damage to Exterior Walls The damage location plan and the typical cracks observed along the exterior of the Warde School are given in Figure 5-24. The hogging deformation mode predominated the north and south segments of the exterior west wall and the north exterior wall. The

crack along the north wall of the school (5-24a) was vertical crack located approximately 2.7 m above ground and approximately 600 mm from the west end of the wall. The crack went through both mortar and stone.

Figure 5-24. Damage Locations and Typical Cracks on the Exterior Walls The cracks on the north (5-24b) and south segments of the west wall were diagonal shear step cracks through the mortar. The step cracks were inclined at about 45 degrees and radiated towards the center of the building.

5.4.5 Damage Summary

The damage observed in the Warde School can be characterized as negligible to slight according to the damage severity classification presented by Burland et al. (1977). Their classifications are as follows:

Negligible damage as hairline cracks with widths less than 0.1 mm, Very slight damage as fine cracks easily treated during normal

redecoration and cracks in exterior brickwork visible on close inspection with widths less than 1 mm, and

Slight damage as cracks that can easily be filled with redecoration probably required, exterior cracks visible with some repointing possibly required for water tightness and a maximum crack with less than 5 mm. Also, doors and windows may stick.

The severity of damage was within estimates made during design of the excavation support system. Most of the damage occurred along the west side of the school in the area where distortions were largest, and when the maximum distortions increased from 1/1000 at the end of wall installation to 1/400 at the end of excavation. Further increases in distortion to 1/315 during station renovation and backfill caused little new damage. No damage to structural elements of the school was observed throughout the project. It must be realize that self-weight deformations induce strains in a structure. The additional distortions caused by the excavation-induced movements results in additional strains that, if imposed on an undeformed structure, may not cause damage. But when these strains are imposed on a structure that has already deformed in response to the applied building loads, these excavation-induced strains may result in damage at apparent distortion levels smaller than applicable to self-weight loading. The damage

reported herein was caused by a combination of the two sources of movements, but, as in the case of all reported damage to buildings adjacent to an excavation, the measured performance of the building only reflected the effects of the excavation-induced movements. Hence the impact of the self-weight induced strains on the onset of cracking remains uncertain. It is noted that the response of larger multi-story buildings to hogging and sagging deformations is more complicated than may be allowed by simple beam theory. This is because these structures respond to deformations as a function of the type of connections between the floor and the columns, the rigidity of the floor slab, and the location and rigidity of the interior shear walls. A finite element model is required to account for the many different variables that make up a building system.

Figure 5-25. Crack Gauge Data

5.5 CRACK GAUGE DATA The diagonal wall crack located on the south wall of the third floor faculty room was instrumented with a crack gauge (Figure 5-22). This wall crack was first observed on Day 73. The crack gauge was applied to the crack on Day 124 and measured change in horizontal and vertical width. The data from the crack gauge is presented in Figure 5-25. The figure compares the crack response to relevant excavation activity and distortions computed from data from Inclinometer 2, settlement points W9 and C5, and Tiltmeter T7. It is noted that the crack opening data presented in the figure is the resultant of the cumulative change in the horizontal and vertical widths of the crack, after the gauge was applied. The final crack width given in the figure was not necessarily the total crack

width. The width of the crack varied between 0.5 mm and 1 mm when the gauge was first set across the crack. Figure 5-25 shows that the crack opening increased from 0.1 mm to 1.0 mm between Day 130 and Day 144. However, the increase in distortion during this period was only from 0.00235 (1/425) to 0.0025 (1/400) and the depth of excavation adjacent to the crack location remained constant at a depth of 12.2 m. The east secant pile wall was chipped to the flange between Days 137 and 140. The chipping activity resulted in a reduction in the bending stiffness of the wall, which led to increased movements. The crack opening increased to approximately 1.3 mm from Day 144 to Day 161. This increase was a result of advancing the excavation for the escalator pit to the final depth of 13.2 m. Between Day 161 and Day 190, the cracking opening increased to a width of approximately 3.9 mm. Afterwards, the amount of increase in the crack opening became substantially less. Subway station renovations along State Street were completed between Day 190 and Day 224. The crack opening remained constant after the backfill activities began on Day 225. The total cumulative change in the width of the opening was 4.3 mm. The total final width of the crack opening was estimated to be approximately 4.8 mm to 5.3 mm.

5.6 ANALYSIS OF DAMAGE

5.6.1 Crack Types Observed Cracking damage to the perimeter concrete and masonry bearing walls and the interior infill walls in the Warde School occurred as a result of tensile strains exceeding the tensile capacity of the material. Shear and bending deformations resulted from

differential vertical displacement of building and combined to exceed the structures capacity to deform without cracking. Based on inclinometer results shown in Figures 4-14 and 4-16, horizontal strains developed in the ground in response to the excavation. However, the magnitude of these horizontal ground strains that get transferred to the building as horizontal building strains depends on the type of structure affected by the movements. The Warde School rests on a load-bearing wall supported by a continuous reinforced strip footing around the building perimeter. In addition, there is sufficient tensile reinforcement at the connections between the floors and columns such that the lateral ground strains would not be transferred to the superstructure. Thus, no direct horizontal ground strains were included in the analysis of the building response. This position is supported by optical survey data that showed the horizontal deformations of the settlement points on the walls and columns of the school were negligible.

Figure 5-26. Crack Types and Orientations Figure 5-26 summarizes the typical crack types and orientations encountered in the Warde School. Although there were some variations observed, the crack types and orientations presented in the figure represent the majority of cracking damage encountered in the school. The figure shows that vertical tension cracks were encountered in most infill walls. Vertical tension cracks are caused by the horizontal tension that develops in a wall in response to bending. In the hogging deformation mode, tensile strain develops in the upper fiber. Thus, vertical cracks are located near the top of a hogging wall (5-26a). Conversely, tensile strain develops along the bottom fiber in the sagging deformation mode, which is why vertical cracks are located at the bottom of a sagging wall (5-26b). In both sagging and hogging modes of deformation, vertical

cracks extend from the tension fiber toward the neutral axis. The length of the vertical cracks is a function of the radius of curvature of the deformation and the flexural rigidity of the wall (MacGregor, 1997). Horizontal cracks in infill walls presumably occurred in response to out-of-plane bending. The width and length of these cracks are governed by the out-of-plane flexural strength of the infill wall. Vertical and horizontal stress concentration cracks were observed emanating from the corners of several doors and windows in wall undergoing sagging. These were tension cracks caused by bending deformations. It was observed that many infill walls initially experienced diagonal cracking. This behavior agreed with the conclusions of Burland and Wroth (1974) who found that for walls with relatively low stiffness in shear or a significant degree of tensile restraint, such as may be observed in infill walls of a frame structure, diagonal tensile strain will be the limiting factor. Diagonal shear cracks are caused by diagonal tensile strain and they appear as inclined cracks. In walls with high puncture ratios (area of openings/area of full wall), these cracks emanate mostly from the corners of the openings (5-26b) (Dulacska, 1992). In walls with low puncture ratios, inclined shear crack typically occur in the middle of the wall (5-26a). Inclined shear cracks in a sagging zone radiate outwards and upwards from the region of maximum curvature of the foundation. Vertical shear cracks occur in corners, near the interface between the wall and column. These cracks are caused by high vertical shear stresses transmitted along the wallcolumn interface as the column moves relative to the wall (5-26a). The reinforced exterior load bearing walls were relatively stiff in direct tension and their length to width ratio was less than 1. As was discussed in Section 2.3, Boscardin

and Cording (1989) and Burland and Wroth (1974) have shown that diagonal strain is most critical for a short beam that is more flexible in shear than in direct tension. This explains the appearance of the step mortar cracks in the north and south segments of the west exterior walls (5-26d).

Figure 5-27. Distortion versus Time at West Side of the Warde School 5.6.2 Initiation of Selected Cracks Crack damage was compared to tiltmeter- and settlement-derived distortions at selected locations where there was sufficient enough data such that the progression of distortions could be related to the initiation of cracking. Distortions computed from tiltmeter data were developed by taking the tangent of the tilt angle. For interior cracking, only settlement-derived distortions because there were no tilmeters located on any interior

columns. For exterior cracking, the tiltmeter distortions were then compared to settlement-derived distortions at locations corresponding to tiltmeter locations.

5.6.2.1 Onset of Interior Cracking Figure 5-27 shows the settlement-derived distortions computed in the east-west direction between settlement points W10 and C1, W9 and C3, and W8 and C5. These distortions show the distortional response in the sagging zone along the west side of the school and also show the onset of interior cracking. The first interior cracks were observed primarily in the sagging zone, on all three levels, on Day 73. At this time, the W10 to C1 distortion was 0.00109 (1/920), the W9 to C3 distortion was 0.00085 (1/1180), and the W8 to C5 distortion was 0.00077 (1/1300). From this point, the distortion increased at a rate proportional to the rate of excavation until the east secant pile wall was chipped between Day 110 and Day 140. The first crack in the marble faade of the entranceway foyer was observed on Day 108. The entranceway foyer is situated approximately at the W9 to C3 location. The distortion at this location on Day 108 was roughly 0.002 (1/500). After this date, new hairline cracks were observed throughout the building and previously noted cracks widened and lengthened. A new crack was observed at the base of the south wall in the entranceway foyer on Day 207. The W9 to C3 distortion at this time was 0.0029 (1/350).

5.6.2.2 Onset of Exterior Cracking Figure 5-28 presents the distortion data along the north segment of the west wall of the school. The tiltmeter-derived distortion data at this location are obtained from Tiltmeter T6, which measured the tilt of the foundation wall in the north-south direction. These

distortion data are compared to distortions obtained from settlement points W12 to W11. It is apparent from the figure that there was good agreement between the tiltmeter distortions and the settlement-derived distortions.

Figure 5-28. Distortion versus Time at North Segment of West Exterior Wall The T6 distortion was approximately 0.0002 (1/5000) on Day 73, and increased to 0.00067 (1/1500) on Day 108. Shortly thereafter, the southward tilt and consequently, the distortions gradually deceased as the north end of the wall settled in response to the excavation along Chicago Avenue. The T6 distortion was approximately 0.00059 (1/1690) on Day 129. This was the day damage first appeared in the north segment of

the west wall. Less distortion developed in this wall because point W12 incrementally settled more than point W11 (Figure 5-9) as a result of it being closer to the Chicago Avenue excavation than point W11. Note that this wall was not damaged when subjected to the larger distortions of 0.00067 (1/1500) on Day 108. After Day 108, the wall at this location was subjected to tensile strains, as indicated by the direction of tilt in T3 and T6 (Figure 5-16). Increased tilt of the north wall of the school towards the north and the increased tilt of the north segment of the west wall towards the south resulted in horizontal tensile strains being induced this wall segment. The combination of the induced horizontal tensile strains and the bending tensile strains caused the north segment of the west wall to crack at smaller distortions to which the wall had been subjected previously

Figure 5-29. Distortion versus Time at North Exterior Wall The distortion along the north side of the Warde School is presented in Figure 5-29. Tiltmeter T4 was used to represent the east-west tilt of the north wall and was compared to distortion derived from settlement points W12 to W13. From the figure, it is seen that the trends of the two data are similar, but the tiltmeter-derived distortion is less than the settlement-revived distortion. The top of the north basement wall tilted towards the west in response to the excavation along State Street and, the north wall experienced hogging deformations. The first crack in the west corner of the north basement foundation wall was observed on Day 108 as the T4 distortion increased to approximately 0.00031 (1/3250). The settlement-derived distortion at this time was 0.00056 (1/1800). The distortions continued to increase to values as high as 0.00116 (1/860). During this time,

no significant changes were observed in the crack. The crack in the foundation wall could have possibly resulted from racking strains induced in the wall due to simultaneous movements towards the Chicago Avenue and the State Street excavations.

5.6.3 Summary of Distortions Table 5-5 summarizes the distortions at various locations within the school. Distortions in the sagging zone are taken from the location between survey points W10 and C1, where the maximum values were observed. Distortions in the hogging zones were computed from the Tiltmeter T-6 data obtained along the north segment of the west wall, representative of and exterior walls, and at a representative location in the interior of the building, between survey points C2 and W19. The distortions in the sagging zone were significantly larger than that in the hogging zones throughout construction. While, in general, damage occurs at lower distortions when a structure hogs rather than sags, the proximity of the school to the excavation dictated that much larger distortions developed in the sagging zone, and, consequently, these distortions governed the onset of cracking. Thus, the location of an affected building relative to the excavation is an important factor in determining the level of distortions at which damage will develop. It must be realize that self-weight deformations induce strains in a structure. The additional distortions caused by the excavation-induced movements results in additional strains that, when imposed on a structure that has already deformed in response to the applied building loads, may result in damage at apparent distortion levels smaller than applicable to self-weight loading. The damage reported herein was caused by a combination of the two sources of movements, but the measured performance of the

building only reflected the effects of the excavation-induced movements. Thus, the impact of the self-weight induced strains on the onset of cracking remains uncertain. A finite element model is required to evaluate the impact of the self-weight induced strains in a building system. Table 5-5 Angular Distortion Summary

Stage 1

Day 11 73 78

Sagging Zone (1) Comment 1/1000 No damage 1/920 First Interior Cracks 1/870 Crack in marble 1/500 facade of entranceway foyer 1/400 1/390 Cracks extend and widen on all levels

Hogging Zone (2) 1/6050 1/5000 1/500 1/1500 1/1600 1/1600 (3) 1/23590 1/4720 1/4100 Crack through mortar and stone of north wall Comment

108 116 127

Cracks in west corner 1/3370 of north foundation wall 1/3370 Step crack in mortar 1/2400 along south segment of west wall Step crack in mortar 1/2360 along north segment of west wall 1/1970 1/1310

129

1/380 "Trigger" 1/335 deformation limit exceeded

1/1690

151 310 Notes:

1/1700

1/315 Movement completed 1/3000

(1) Angular Distortions computed in interior o school between survey points W10 and C1 (2) Angular Distortion computed along north segment of west wall at Tiltmeter T6 (3) Angular Distortion computed in interior of school between survey points C2 and W19

5.7 SUMMARY The primary responses of the Warde School to the adjacent excavation were settlement of the shallow spread and continuous wall footings and tilting of the foundation walls. These primary responses produced distortions in the infill walls of the school. The optical survey data indicated that no lateral movement occurred at the ground level of the building. Thus, no direct horizontal strain component was included in the analysis of the building response. The extent of the settlement behind the support wall was defined mainly when the secant pile wall was installed. Although the magnitudes of movements increased significantly during the excavation and support stage, the extent of these movements did not. Tiltmeters placed along the west wall of the school indicated that the central portion of the wall tilted to the east, while the north segment of the west wall tilted towards the south. Tiltmeters on the north wall indicated that the wall rotated to the north whereas the northwest corner of the school tilted towards the west. This pattern of deformation at the northwest corner of the school was caused by the 3-D geometry of the excavation at that location. The sagging mode of deformation developed within the central portion of the west side of the school while hogging occurred elsewhere. The highest levels of distortion corresponded to the region that experienced sagging. Excavation-related damage within the Warde School was primarily in the form of cracking of interior infill walls and exterior masonry walls, and misalignment of doorways and windows. No structural damage was observed during the project. Crack

damage occurred mostly during the excavation and support installation stage. During station renovation and backfill, only a few instances of new damage were observed; existing cracks generally became larger during this time.

CHAPTER 6

DEEP EXCAVATIONS: THE OBSERVATIONAL METHOD AND INVERSE ANALYSIS Predictions of the magnitude and distribution of ground movements are used to estimate the tolerance of structures and utilities to the deformations associated with construction of deep supported excavations. Many factors affect movements associated with excavations, including soil properties (soil type, presence of water), support system properties (wall stiffness, support stiffness, preload) and construction activities or workmanship (construction sequence, installation of support, surcharge loads). In practice, when designers are faced with an excavation where ground movements are a critical issue, they can base their estimate of movements on semi-empirical methods based in part on past performance data or on results of finite element analyses. The main limitation of the first approach is the variety of construction techniques of the case studies used in developing them. These activities can contribute significantly to the movements reported. Therefore design approaches developed from these data should be considered biased towards average construction practices. The only way to explicitly include the effects of the construction activities in the analysis is to perform numerical simulations of the problem. Indeed, if the exact construction procedure is known, a finite element analysis conceptually allows an engineer to model all aspects of excavation that cause stress change in soil: wall installation, dewatering, cycles of excavation, bracing and brace removal, and preloading of anchors. Finite element predictions, however, contain uncertainties related

to soil properties, support system details and construction procedures. If one wants to predict and evaluate the overall performance of a design, a procedure that incorporates an evaluation of the results of the analyses must be defined. The procedure to accomplish this task is usually referred to as the observational method (see section 2.2.3). Morgenstern (1995), in his Casagrande Lecture, emphasizes the importance of the observational method and stresses the need to have plans to cope with possible eventualities. In practice, however, it is very difficult to quantitatively judge how well the work is proceeding especially considering the time constraints associated with construction. Ad-hoc mathematical tools are needed to compare observations and predictions. To improve the state-of-the-practice of controlling ground movements associated with supported excavations, this chapter presents a procedure that objectively updates design predictions of deformations for supported excavations in clay. Monitoring data are used as observations in an inverse analysis that calibrates the numerical model of the excavation and thus supports, in an objective way, the engineering judgments made during the construction of the excavation system. The inverse analysis methodology was developed and tested using data from a 39 ft deep excavation through soft clays in Chicago (Finno et al. 2002).

6.2 MODEL CALIBRATION BY INVERSE ANALYSIS In model calibration, various parts of the model are changed so that the measured values are matched by equivalent computed values until, hopefully, the resulting calibrated model accurately represents the main aspects of the actual system. Despite their

apparent utility, inverse models are used for this purpose much less than one would expect. In practice, numerical models typically are calibrated using trial-and-error methods because, perhaps, of the difficulties of implementing an inverse analysis, the complexity of the simulated systems, and/or the engineers perception that automated estimation of model parameters without engineering judgment is impossible. One outcome of this work is to show that these concerns are, in most cases, unjustified and inverse modeling represents a valuable tool for geotechnical engineers.

Figure 6-1 Schematic of inverse analysis procedure

With an inverse modeling approach, a given model is calibrated by interactively changing model input values until the simulated output values match the observed data (i.e. observations). Figure 6-1 shows a schematic of an inverse analysis procedure. The input parameters are initially estimated by conventional means (e.g. using available laboratory and field test results). Next a numerical simulation of the problem is run and the results are stored in a file (generally in ASCII text format). The simulated results are then compared to the field observations and a regression analysis is performed to minimize an objective function. The objective function quantifies the fit between computed results and observations and its minimization is reached by updating the set of input parameters needed to perform the numerical simulation. If the model fit is not optimal, the procedure is repeated until the model is optimized.

6.2.1 An inverse analysis algorithm: UCODE In the work described herein model calibration by inverse analysis was conducted using UCODE (Poeter and Hill 1998), a computer code designed to allow inverse modeling posed as a parameter estimation problem. UCODE has been developed for ground-water models, but it can be effectively used in geotechnical modeling because it works with any application software that can be executed in batch mode. Its model-independency allows the chosen numerical code to be used as a closed box in which modifications only involve model input values. This is an important feature of UCODE, in that it allows one to develop a procedure that can be easily employed in practice and in which the engineer will not be asked to use a particular finite element code or inversion algorithm.

In UCODE the weighted least-squares objective function S(b) is expressed by:

(6.1) where b is a vector containing values of the number of parameters to be estimated; y is the vector of the observations being matched by the regression; y'(b) is the vector of the computed values which correspond to observations; is the weight matrix; and e is the vectors of residuals. Non-linear regression is an iterative process. The modified Gauss-Newton method used by UCODE to update the input parameters is expressed as:

(6.2)

(6.3) where dr is the vector used to update the parameter estimates b; r is the parameter estimation iteration number; Xr is the sensitivity matrix (Xij= yi/bj) evaluated at parameter estimate br; C is a diagonal scaling matrix with elements cjj equal to 1/XT X)jj;I is the identity matrix; mr is a parameter used to improve regression performance; and
r

is a damping parameter.

6.2.2 Model fit statistics Different quantities can be used to evaluate the model fit. A commonly used indicator of the overall magnitude of the weighted residuals is the model error variance, s2, which equals:

(6.4) where S(b) is the objective function; ND is the number of observations; and NP is the number of estimated parameters. The value of the objective function (Eq. 2.1) is also used to indicate model fit informally, because its variation indicates by how much an optimized model improves with respect to the initial simulation of a problem. The objective function changes can be expressed through a new statistic, the fit improvement (FI), which indicates by what percentage the optimized results improved compared to the initial fit between experimental data and computed results. The fit improvement is defined as:

(6.5) where S(b)initial is the initial value of the objective function; and S(b) optimized is the value of the objective function for the optimized set of parameters.

6.2.3. Input parameters statistics The relative importance of the input parameters being simultaneously estimated can be defined using parameter statistics, including the sensitivity of the predictions to changes in parameters values, the variance-covariance matrix, confidence intervals and coefficients of variation.

Different quantities can be used to evaluate the sensitivity of the predictions to parameters changes. One percent sensitivities, dssij, scaled sensitivities, ssij, and composite scaled sensitivities, cssj, can be used for the purpose. These sensitivities are defined in Eq. (6.6), (6.7) and (6.8), respectively.

(6.6)

(6.7)

(6.8) where y'i is the ith simulated value; yi/bj is the sensitivity of the ith simulated value with respect to the jth parameter; bj is the jth estimated parameter; observation One percent scaled sensitivities represent the amount that the simulated value would change if the parameter value increased by one percent. Scaled sensitivities are dimensionless quantities that can be used to compare the importance of different observations to the estimation of a single parameter or the importance of different parameters to the calculation of a simulated value. Composite scaled sensitivities
jj

is the weight of the ith

indicate the total amount of information provided by the observations for the estimation of one parameter. The reliability and correlation of parameter estimates can be analyzed by using the variance-covariance matrix, V(b'), for the final estimated parameters, b', calculated as:

(Figure 6.9) where s2 is the error variance; X is the sensitivity matrix; and is the weight matrix. The diagonal elements of matrix V(b') equal the parameter variances, the off-diagonal elements equal the parameter covariances. Parameter variances are most useful when used to calculate two other statistics: confidence intervals for parameter values and coefficients of variation. Parameter covariances can be used to calculate correlation coefficients. Coefficients of variation, covi, are equal to:

(6.10) where is the standard deviation of parameter b1

Correlation coefficients are calculated by:

(6.11)

where cor(i,j) indicate the correlation between the ith and jth parameter; cov(i,j) equal the off-diagonal elements of V(b'); and var(i) and var(j) refer to the diagonal elements of V(b'). The coefficients of variation provide dimensionless numbers with which the relative accuracy of different parameter estimates can be compared. Correlation coefficients close to 1.0 and 1.0 are indicative of parameters that cannot be uniquely estimated with the observations used in the regression.

6.2.4 Observations weighting The weights assigned to the observations are an important part of the regression analysis because they influence the value of the objective function, and thus the regression results. UCODE uses a diagonal weight matrix. Weighting is used to reduce the influence of observations that are less accurate and increase the influence of observations that are more accurate. For problems with more than one kind of observation, weighting also produces weighted residuals that have the same units, so that they can be squared and summed. In UCODE the weight of every observation, equal to the inverse of its error variance,
2 ii,

is

(6.12) Users assign the weight of an observation by specifying a value for its variance, standard deviation or coefficients of variation. Assigning appropriate weight values to the observations can be problematic. For regression methods to produce parameter estimates with the smallest possible variance Hill (1998) suggests that weighting needs

to be proportional to the inverse of the variance of the measurement errors. At the end of the regression analysis, the value of the model error variance, s2 (Eq. 2.4), can be used to evaluate the consistency between the model fit and the measurement errors, as expressed by the observations weights. Values larger than 1.0 indicate that the model fits the data less well than would be accounted for by expected measurement errors.

6.3 UPDATE DESIGN PREDICTIONS USING MONITORING DATA BY INVERSE ANALYSIS This section shows how inverse analysis based on field monitoring data can be used to objectively update the predicted performance of supported excavation systems. Movements of the soil surrounding an excavation, measured to evaluate how well the actual construction process is proceeding in relation to the predicted behavior, can be recorded by inclinometers, which measure lateral deformations at various depths at discrete locations around the construction site, and survey points, which record ground movements and/or displacements of structures adjacent to the excavation. With an inverse analysis procedure (see Chapter 2), these recorded displacements can be used to control the construction process and update predictions of movements at early stages of construction. Any time a new set of construction monitoring data are available, the finite element model of an excavation can be recalibrated to provide the best fit to the field observations. Inverse analysis algorithms allow the simultaneous calibration of multiple input parameters. However, identifying the important parameters to include in the inverse analysis can be problematic. Indeed, it is not possible to use the regression analysis to

estimate every parameter of every soil model used in the simulation. The number and type of input parameters that one can expect to estimate simultaneously depend from many factors, among which:

Soil models used. The characteristics of the soil models and the number

and type of observations used in the simulation determine the input parameters that are expected to be successfully calibrated. Some model parameters may be correlated to one another and thus not likely to be estimated simultaneously.

Aspects of the simulated system represented by estimated parameters.

In many instances, supported excavations generally generate only small deformations in the soil surrounding the excavation. In these instances, stiffness parameters are expected to be more important than failure parameters in defining the behavior of the soil mass. Sensitivity analyses can be used to determine the input parameters of a soil model that are most relevant to the computed system response.

Available observations. The number of observation points used in the

inverse analysis is related to the maximum number of parameters that one can expect to estimate by regression analysis. Their spatial distribution influences the number of soil layers whose parameters can be calibrated.

Finite element implementation. Computational time may constitute an important variable for very complex simulations. The number of finite element runs at any given iteration and the number of iterations needed for the convergence of the regression analysis are proportional to the number of estimated parameters, NP.

Figure 6-2 shows a procedural flowchart that can be used for the identification of the soil parameters to optimize by inverse analysis. As subsequently described, the total

number of input parameters can be reduced, in four steps, to the number of parameters that are likely to be successfully optimized by inverse analysis. Step 1: Models input parameters Models uncorrelated parameters. The soil model chosen to simulate the soil behavior determines the total number of input parameters to estimate (e.g., the H-S model has 10 input parameters). The number of parameters that can be estimated by inverse analysis depends from the characteristics of the model and from the type of observations available. Parameter correlation coefficients (Eq. 6.11) can be used to evaluate which parameters are correlated and are, therefore, not likely to be estimated simultaneously by inverse analysis.

Figure 6-2 Flowchart for the identification of soil parameters to optimize by inverse analysis

Step 2: Models uncorrelated parameters Models relevant parameters. The parameters that most affect the computed results are determined by the stress conditions in the soil around the excavation. Composite scaled sensitivity values (Eq. 2.9) can provide valuable information on the relative importance of the different input parameters of a given model. Step 3: Models relevant parameters Total relevant parameters. The number of soil layers to calibrate and the type of soil model used to simulate the layers determines the total number of relevant parameters of the simulation. A new sensitivity analysis may be necessary to check for correlations between parameters relative to different layers. Step 4: Total relevant parameters Parameters to optimize. The total number of observations available and computational time considerations may prompt a final reduction of the number of parameters to optimize simultaneously. Figure 6-2 showed the key role that sensitivity analyses have in determining the parameters that are important for the finite element simulation of an excavation. Once the parameters to optimize have been chosen, sensitivity results continue to play an integral part in the regression analysis. Indeed, a sensitivity matrix is computed at every regression iteration. This is necessary because the simulation of an excavation system by finite element methods is a highly non-linear problem. Thus, the sensitivity of the results to changes in parameter values is not constant but depends on the particular values at which the sensitivity matrix is computed. The design chart (Clough et al. 1989) given in Figure 6-3 will be used to explain the importance of this approach. The graph is generally used to design retention systems for

supported excavations in soft to medium clays. The curves show how the ratio between the maximum horizontal movement of the wall and the height of excavation ( H/H) is a function of the factor of safety against basal heave (FSBS)and of the retaining system stiffness (EI/h4 w), a combination of wall stiffness and strut spacing.

Figure 6-3 Design Curves for Maximum Lateral Movement in Soft to Medium Clay [after Clough et al. (1989)]

The chart can be considered a model of the excavation problem, where FSBS and EI/h4 are the input parameters and
H

/H is a measure of the design performance. The tangent

of the design curves expresses the sensitivity of the movements with respect to the system stiffness, the distance between the curves is related to the sensitivity of the movements with respect to FSBS. The graph shows that, if the system is stiff or the factor of safety against basal heave is high, the performance is less sensitive to either parameter (i.e. a small value of the tangent to the curve) than would be the case if the system is flexible or has a low FSBS (i.e. a higher value of the tangent to the curve).

6.4 PROCEDURE VALIDATION: THE CHICAGO & STATE CASE STUDY The proposed methodology was developed and tested using data from a project in downtown Chicago (Finno et al. 2002), the excavation/renovation of the Chicago & State CTA subway station.

6.4.1 Finite element simulation of the problem The finite element software PLAXIS was used to compute the response of the soil around the excavation. Figure 6-4 shows a schematic of the PLAXIS input. Details about the definition of the finite element problem, the calculation phases and the model parameters used in the simulation described herein can be found in Appendix C.

Figure 6-4 Schematic of PLAXIS input The problem was simulated in plane-strain conditions. The soil stratigraphy was assumed to be uniform across the site (see Figure 4.4). The soil layers considered were 8: a fill layer overlaying a clay crust, a compressible clay deposit (in which 4 distinct clay layers were modeled) and two relative incompressible stiff silty clay strata. All

elevations in the figure refer to the Chicago City Datum (CCD). Note that the figure, for display purposes, does not show the side boundaries of the mesh (600 ft x 94 ft), which was extended beyond the zone of influence of the settlements induced by the excavation (Hsien and Ou 1998 and Caspe 1966). The finite element mesh boundary conditions were set using horizontal fixities, for the left and right boundaries, and total fixities, for the bottom boundary.

6.4.1.1 Calculation phases The tunnel tubes and the school adjacent to the excavation were explicitly included in the finite element simulation of the problem to take into account the effect of their construction on the soil surrounding the excavation. Table 6-1 shows the PLAXIS calculation phases of the simulation described herein. The second column of the table shows the calculation phase number, the third column explains the purpose of the calculation phase, the fourth column indicates the calculation type, and the last column specifies the loading input condition. A plastic calculation indicates that an elastoplastic deformation analysis is carried out in either fully drained or fully undrained conditions. For the simulation described herein, plastic calculations are always associated with staged construction loading conditions, which indicate changes in the geometric configuration of the FE mesh, and clay layers are always assumed to be in undrained conditions. Consolidation calculations are used to analyze the development and dissipation of excess pore pressures in the water-saturated soil layers as a function of time. An ultimate time (loading input condition) is specified to terminate a consolidation calculation. Note than in PLAXIS it is not possible to perform a staged construction calculation with simultaneous consolidation. More details about calculation

types and loading input conditions can be found in the PLAXIS manual (Brinkgreve and Vermeer, 1998).

Table 6-1 PLAXIS calculation phases

6.4.1.2 Hardening-Soil model initial calibration The soil model used to characterize the clays in the PLAXIS simulation of the excavation is the Hardening-Soil model (Schanz et al. 1999). Table 6-2 shows the initial values of the H-S model parameters for the five clay layers that will be calibrated by inverse analysis. Layers 1 to 5 refer to the Upper Blodgett, Lower Blodgett, Deerfield, Park Ridge and Tinley layers, respectively. The model parameters of the soil layers that were not calibrated by inverse analysis can be found in Appendix C.

Table 6-2 Initial values of H-S parameters for the 5 clay layers calibrated by inverse analysis

The initial estimates of the input parameters for layers 1 to 4 are based on the results of the triaxial compression tests(Calvello 2002). Because little laboratory data exists for

the layer 5 soil, the initial values of the parameters for layer 5 are based on the following considerations: (i) layer 5 failure parameters are assumed to have the same values of layer 4 failure parameters, and (ii) layer 5 stiffness modules are assumed to be 1.5 times larger than layer 4 stiffness modules. For all layers, the value of parameter E50refassumed to be equal to 70% Eurref The H-S stiffness parameters are defined with respect to a reference pressure (pref=100 stress units). Thus, it is difficult to relate the values of E50ref, Eoedref, and Eurref to typical geotechnical estimates of stiffness moduli. The following equations define the stress dependent stiffness moduli used in the H-S model:

(6.13)

(6.14)

(6.15) where E50 is the secant Young modulus, Eoed is the oedometric modulus, Eur is the unload-reload elastic modulus, 1'is the major principal stress, stress, is the friction angle and c is the cohesion. Figures 6-5, 6-6 and 6-7 show the variation with the vertical stress of E50, Eoedoed and Eurur. The curves were computed, using the initial values of parameters c, and m (see Table 6-2), according to Equations 5.1, 5.2 and 5.3, respectively. The vertical and
3 '

is the minor principal

horizontal directions were assumed to be principal directions (i.e. v'= 1'and


v '

' ' h = 3 =k0

).

Figure 6-5 Variation of stiffness parameter E50 with vertical stress v' (initial parameter estimates)

Figure 6-6 Variation of stiffness parameter Eoed with vertical stress v' (initial parameter estimates)

Figure 6-7 Variation of stiffness parameter Eurwith vertical stress v'(initial parameter estimates) To compare the stiffnesses of different layers, one has to consider the effective stresses of the soil. Table 6-3 shows the initial vertical effective stress in the middle of the five soil layers. These values can be used to compute, using Eq. 6.13 or Figure 6-5, the

values of the secant stiffness modulus at 50% shear strength, E50, at the beginning of the simulation.

Table 6-3 Initial vertical effective stress of clay layers Table 6-4 shows, for all five layers, the initial E50refref values, the computed E50 values, the estimated undrained shear strength Su and the ratio between E50 and Su. Note that the Su values were estimated from field vane results and correlations based on water content data (Chung and Finno, 1992).

Table 6-4 Initial E50 / Su ratios of clay layers Note that E50 represents a drained modulus. Nonetheless the ratio E50/SuScan be used to judge the relative inherent stiffness of the various soil layers in undrained conditions. The initial /Su ratios used in this simulation show that: (i) the Blodgett layers (1 and 2) are assumed to have about the same relative stiffness, and (ii) the other layers (3, 4 and

5) become, relative to their undrained shear strength, progressively more deformable with depth. Typical values of Eu/Su ratios are often presented in literature to evaluate the stressstrain undrained response of clays. Lambe and Whitman (1969) report Eu/Su values of about 500 and 1000 for soft and stiff clays, respectively. E50/Su values are not generally quoted in literature. However, E50/Su ratios can be related to typical Eu/Su ratios if the initial undrained stiffness modulus Eu is converted into an equivalent E50. The following three steps describe a way of computing E50 from a given value of Eu. Step 1 (EuGin) The following relationship between elastic moduli can be used to convert Eu (initial undrained stiffness modulus) into an equivalent Gin (initial shear stiffness modulus):

(6.16) Step 2 (Gin G50) The value of the shear stiffness modulus of clays, G, decreases with increasing shear strains: G50<Gin<G0 The maximum stiffness, G0, only occurs at extremely small strains ( sh<0.001%). The initial undrained stiffness modulus Eu is generally computed at higher strain levels (
sh=0.05-0.1%).

Thus, the initial shear stiffness modulus, Gin, is smaller than G0. Based

on published results (Viggiani and Atkinson, 1995) the value of G0 is assumed to be 0.5-0.75 times G0and G50 is assumed to be 0.25-0.50 times Gin (i.e. G50 = 0.15-0.35 G0). Step 3 (G50 E50) The following relationship between elastic modules can be used to convert G50 (50% shear stiffness modulus) into an equivalent E50 (50% secant stiffness modulus):

(6.18) Finno and Chung (1992) reported Eu/Su values of 400-600 for normally consolidated compressible Chicago clays (Blodgett and Deerfield layers) sheared in triaxial compression. Following the procedure outlined previously, an equivalent E50/Su ratio can be computed. Assuming Eu/Su=500, G50=Gin/3 and =0.2, the ratio E50/S equals 133. This value is slightly higher than the initial values used in this simulation (see Table 64), suggesting that the initial estimates of the H-S parameters defining the soil stiffness may be conservative.

6.4.2 Inverse analysis set-up The optimization algorithm UCODE was used to calibrate, by inverse analysis, the PLAXIS finite element simulation of the excavation. A schematic of the interaction between PLAXIS and UCODE was presented in Figure 4-7. Examples of input and output files of the inverse problem analyzed in this section can be found in Appendix C.

6.4.2.1 Observations and weighting Table 6-5 shows the construction stages for which the model predictions are updated. Lateral movements of the soil behind the secant pile wall were recorded using five

inclinometers. The excavation, however, was modeled in plane strain conditions. Thus, only two of them (incl. 1 on the east and incl. 4 on the west) were used to compare field data and computed displacements. The measured settlements were not used as observations because the finite element predictions of the ground settlement induced by excavation are generally not as good as those of the horizontal movements of the soil.

Table 6-5 Excavation stages considered for updating model predictions

Figure 6-8 Schematic of retaining system and observation points used from inclinometer readings (e.g., stage 1 data) Figure 6-8 shows the observation points retrieved from the field readings of inclinometers 1 and 4 (the data in the plot refer to stage 1). The soil profile and a schematic of the support system are also shown in the figure. Inclinometer readings were taken in the field every two feet. Not every reading, however, could be used as an observation for the inverse analysis because the finite element displacements were computed only at the intersection between the finite element mesh and the inclinometer location. Thus, 13 observation points were used for the east side and 11 observation points for the west side. The inverse of the variance of the measurement errors was used to assign weights to the

observations (i.e. inclinometer data). Table 6-6 shows the values of the observation points used for the five construction stages and their measurement errors. See Appendix C for details about the inclinometer probe used to monitor the movements at the excavation site, its accuracy and the computed measurement errors. The measurement error of the horizontal displacement inclinometer data is not constant. Its value increases as one moves further away from the bottom of the casing because the inclinometer probe measures tilt and not displacements, thus errors become larger as one gets closer to the ground surface. Note that inclinometer data are available, for the east side, at all 5 construction stages considered. On the west side, however, the inclinometer was damaged by construction activities after stage 3. That is why the west side inclinometer readings are not shown, in subsequent figures, for the last two stages of construction.

Table 6-6 Values of observations on the east side and west side and their measurement errors

6.4.2.2 Parameterization The soil layers calibrated by inverse analysis are the upper Blodgett, lower Blodgett, Deerfield, Park Ridge and Tinley strata. In the analysis described herein they are referred to as layer 1, 2, 3, 4 and 5, respectively. All the layers are modeled using the Hardening-Soil model. The initial estimates of the H-S input parameters were presented in section 6.4.1.2.. The input parameters optimized by inverse analysis were chosen following the procedure described in section 6.3 Note that the first two steps of the procedure (see Figure 6-2) refer to the selection of the model parameters (e.g., H-S model) that are relevant to the problem under study, the last two steps refer to the selection of the total number of simulation parameters (e.g., 5 soil layers calibrated simultaneously) to optimize by inverse analysis. The H-S model features 10 input parameters. The characteristics of the model determine the number of uncorrelated parameters that one can expect to successfully optimize by inverse analysis. The H-S parameters that can be effectively estimated from laboratory data using an automated optimization algorithm are E50ref, m and (Calvello 2002). For the simulation discussed herein the values of the other model parameters are either kept constant at their initial value (parameters c, , and Rf), or are assumed to be related to one of the other parameters (Eoedref = 0.7 E50ref, Eurref = 3.0 E50ref and k0 = 1 sin ) The sensitivity of the observations to changes in values of Eref, m and determines the parameters that are relevant to the problem simulated herein. Figures 5.12 and 5.13 show the composite scaled sensitivities of the three parameters for layers 1 to 5. In the first figure the bar chart refers to sensitivities computed using all the observations, and the line charts refer to sensitivities computed from the observations of the different

layers. In the second figure the sensitivities are grouped by construction stages. Both figures show that all three parameters (i.e. E50ref, m and ) are important, from a model perspective, in affecting the outcome of the analysis. From a simulation perspective, results show that the parameters that most influence the simulation are the ones relative to layers 1 and 4. Layer 1 is the softest soil layer, thus its major influence on the displacement results is expected. Layer 4 is the stiff clay layer below the bottom of the excavation, into which the wall is tipped. The high sensitivity values relative to this layer indicate that the strength and the stiffness of the clay below the excavation have significant impact on movements.

Figure 6-9 Composite scaled sensitivities of parameters E50ref, m and for layers 1 to 5. As one would expect Figure 6-9 also shows that the observations relative to a soil layer are mainly influenced by changes in that soil layers parameters. For instance, the

values of the sensitivities from layer 4 and layer 5 observations show a clear peak, respectively, at layer 4 and layer 5 input parameters. Likewise, Figure 6-10 shows that the parameters of the deeper layers become more important at later construction stages (i.e. deeper excavation depths).

Figure 6-10 Composite scaled sensitivities of parameters E50ref, m and for layers 1 to 5 grouped by construction stages Other important parameter statistics resulting from a sensitivity analysis are the correlation coefficients. The sensitivity analysis performed on E50ref, m and for layers 1 to 5 indicated that high correlation values occur between parameters E50ref and m. Table 6-7 shows the correlation coefficients between the three parameters at every layer. Results indicate that the two stiffness parameters (i.e. E50ref and m) cannot be simultaneously and uniquely optimized, even though the results of the analysis are sensitive to both parameters. Parameter E50ref, rather than parameter m, was chosen to represent the stiffness of the H-S model. The reasons behind this choice are: (i) m values are bounded between 0 and 1.0, thus they would require the use of a mapping function (see section 2.4) to avoid possible problems with unreasonable updated values during the regression iterations, and (ii) changes in E50ref values also produce changes in the values of parameters Eoedref(equal to 0.7 times E50ref) and Eurref(equal to 3 times E50ref), thus their calibration can be considered as representative of the calibration all three H-S stiffness parameters.

Table 6-7 Highest values of correlation coefficients

The results of the sensitivity analysis seem to indicate that the total number of relevant parameters is 10 (i.e. E50ref an for layers 1 to 5). However a final reduction of the parameters to optimize was necessary to establish a well-posed problem. Table 6-8 shows the parameters that are optimized by inverse analysis for the simulation described herein.

Table 6-8 Parameters optimized by inverse analysis The parameters to optimize were chosen based on the following considerations: 1. Hill (1998) suggests applying the principle of parsimony. Thus, to begin calibration estimating very few parameters that together represent most of the features of interest and to increase the complexity of the parameterization slowly. 2. Poeter and Hill (1997) warn against spreading the data too thin. If all 10 relevant parameters were to be optimized simultaneously, the ratio between the number of observations available at the first construction stage (24 displacement observation points) and the number of parameters estimated (NP=10) would be too low. 3. The first construction stage, which refers to the excavation of the 60-deep secant pile wall, causes movements in all 5 clay layers. Thus, at least one parameter per layer had

to be considered for optimization. 4. The stiffness parameters (E50ref) were chosen over the failure parameters ( ) because the excavation-induced stress conditions in the soil around the excavation were, for the most part, far from failure. Thus, the stiffness parameters were perceived to be more relevant to the simulated problem. 5. Results of the parametric study conducted on the number of input parameters considered for optimization by inverse analysis (section 3.3.2) suggest that a model can be successfully calibrated even if some relevant parameters are excluded from the optimization, provided that their initial estimate is reasonable.

6.4.3 Results This section presents the results of the inverse analysis performed on the finite element simulation of the Chicago & State excavation project. Visual examination of the horizontal displacement distributions at the inclinometer locations plots provides the simplest way to evaluate the fit between computed and measured field response. Figure 6-11 shows the comparison between the measured field data and the computed horizontal displacements when the initial estimates of the parameters are used (see Table 6-2). The comparison shows that the finite element model computes significantly larger displacements at every construction stage. The maximum computed horizontal displacements are about two times the measured ones and the computed displacement profiles result in significant and unrealistic movements in the lower clay layers (Deerfield and Park Ridge). Results indicate that the stiffness properties for the clay

layers have been underpredicted, as was suggested in 6.4.1.2.

Figure 6-11 Measured vs. computed horizontal displacements: initial estimate of parameters

6.4.3.1 Optimization based on observations from construction stage 1 Table 6-9 shows the initial values of the 5 input parameters considered in the analysis and the values of the parameters that best fit the observations relative to stage 1. Results show that the optimized values of the parameters are, at all layers, higher than their initial estimates.

Table 6-9 Initial and optimized input parameters at stage 1 Note that: (i) parameters E1 and E2 were optimized together (i.e. layer 1 and layer 2 were considered as a single layer) and converged to a value slightly higher that the initial estimate of parameter E2, and (ii) parameter E5 was not optimized by inverse analysis, instead its changes were tied to changes in the value of E4 (at every iteration E5 was set equal to 1.5 times E4). This approach was employed after various unsuccessful attempts at optimizing all five parameters simultaneously and independently. The exclusion of parameters E2 and E5 from the regression was based on the results of the sensitivity analysis presented in section 6.4.2.2, which showed that the parameters relative to layer 2 and layer 5 did not affect the computed results as significantly as the parameters relative to layers 1, 3 and 4. Figure 6-12 shows the comparison between the measured field data and the computed horizontal displacements when parameters are optimized based on stage 1 observations. The improvement on the fit between the computed and measured response is significant. Despite the fact that the optimized set of parameters is calculated using only stage 1 observations, the positive influence on the calculated response is substantial for all construction stages. At the end of the construction (i.e. stage 5) the maximum computed displacement exceeds the measured data by only about 15%. These results are

significant in that a successful recalibration of the model at an early construction stage positively affects subsequent predictions of the soil behavior throughout construction. Table 6-10 shows the value of the following statistics that indicate the model fit to the field observations: S(b)in (initial objective function), S(b)fin (final objective function), s2fin (final error variance), CCfin (final correlation coefficient) and FI (fit improvement). The values of these statistics prove that the calibration of the finite element simulation by inverse analysis based on stage 1 observations was extremely effective. The improvement over the initial predicted response is considerable (FI = 99.6%) and the final computed response fits the observations better than one would expect if the inclinometer measurement errors are taken into account (s2fin < 1.0).

Table 6-10 Model fit statistics at optimization stage 1 Note that the statistics presented in Table 6-10 refer to stage 1 observations only. Thus, they cannot be used to assess how the calibrated simulation improves the fit between measured and computed response for the other construction stages. To quantify the predictive improvement of the calibration shown in Figure 6-12, one needs to consider two new model statistics: the global objective function, OF, and the predictive fit improvement, PFI. OF= [y-y'*]T [y*-yi* ] (6.19)

(6.20) where y* is the vector of all available observations (from stage 1 to stage 5); y'*is the vector of the composed values which correspond to the observations; and is the weight matrix used in regression, PFI, is the predictive fit improvement at the end of stage 1, OFbase is the global objective function relative the initial estimates of the parameters and OF is global objective function relative to the parameters on the optimization up to stage i.

Figure 6-13 Measured vs. computed horizontal displacements: parameters optimized based on observations from stages 1 and 2

The global objective function, OF, is computed according to Eq. 6.1 using all the observations available even if they are not used during the regression analysis of the simulation. The predictive fit movement, PFIi, is computed using the value of OF after the optimization (at stage i) and the value of OF relative to the intial estimates of the parameters. Table 6-11 shows the values of the intial OF, the global objective function at the end of stage 1 and the predictive fit movement at the end of stage 1.

Table 6-11 Model statistics quantifying the predictive improvement achieved by the calibration

6.4.3.2 Model fit for all construction stages Section 6.4.3.2 presented the optimization results relative to the calibration of the simulation after construction stage 1. Parameters were recalibrated at every stage until the final construction stage (i.e. stage 5). At every new construction stage, the inclinometer data relative to that stage were added to the observations already available. Figure 6-12 showed the visual fit comparison between inclinometer data and computed displacements when observations relative to stage 1 were used in the regression analysis. Figures 6-13, 6-14, 6-15 and 6-16 show the comparison between experimental and

computed results for the remaining 4 stages (i.e. stage 2 to stage 5, respectively).

Figure 6-14 Measured vs. computed horizontal displacements: parameters optimized based on observations from stages 1, 2 and 3

Figure 6-15 Measured vs. computed horizontal displacements: parameters optimized based on observations from stages 1, 2, 3 and 4

Figure 6-16 Measured vs. computed horizontal displacements: parameters optimized based on observations from all stages When stage 2 observations are used for the regression analysis (Figure 6-13) results do not change significantly. This indicates that the inclinometer readings relative to this construction stage do not provide significant extra information to improve the model calibration. When inclinometer data from stage 3 are added to the observations (Figure 6-14), the fit between the field and the computed results relative to this stage improves, especially in the upper soil layer. These results also improve the fit of stages 4 and 5. That is why observations from these last two construction stages (Figures 5-18 and 519) do not produce any change in the model. The calibrated model cannot be improved any further after stage 3.

Overall, the comparison between the measured field data and the computed horizontal displacements at the different optimization stages shows that the fit between the two sets of data is always remarkable. The inverse analysis performed after the first construction stage significantly improves the initial prediction of displacements and recalibrates the soil models in a way that allows them to capture the main behavior of the soil layers throughout construction. By the end of construction stage 3, the evolution of the horizontal movements at the inclinometer locations is predicted with great accuracy at all stages.

Figure 6-17 Measured vs. computed horizontal displacements: parameters estimated by trial and error based on all observations

These results demonstrate that using inverse modeling techniques enhances the observational method practice. An engineer with detailed knowledge of finite element procedures and constitutive modeling, instead of performing a regression analysis, could modify the input soil parameters by trial-and-error. However, processing displacement data and using them to calibrate the results of the finite element model is an extremely time-consuming and highly subjective task. Figure 6-17 shows the comparison between the measured field data and the computed horizontal displacements in a simulation for which the optimal parameter values were estimated by trial-anderror. This research task was performed before developing the procedure described in this thesis. The fit between computed and observed data is relatively good, yet not every stage of construction is simulated with the same accuracy reached by the inverse analysis. Computed results match the maximum movements well, but they tend to overpredict the displacements of the stiffer layers.

6.4.3.3 Best-fit parameters Table 6-12 shows the initial and optimized values of the 5 input parameters at the different optimization stages. Results show that, after construction stage 3, the values of the optimized input parameters do not change. This indicates that the observations at stages 4 and 5 match the computed results of the model calibrated at stage 3 and therefore further optimization of the model is neither possible nor necessary.

Table 6-12 Best-fit values of input parameters at various optimization stages

Figure 6-18 Best-fit parameter values and normalized excavation depth at different optimization stages In Figure 6-18 the variation of the input parameters at the various optimization stages is shown above a bar chart representing the progress of the excavation. The excavation depth is normalized with respect to the excavation width and plotted for the 5 construction stages. Results show that the maximum changes in parameter values occur at stage 1, when the observations relative to the installation of the secant-pile wall are

used. Note that the excavation was 74 ft wide, the secant pile wall was 60 ft deep and the maximum excavation depth was 39 ft below ground surface. From the results presented in Table 6-12 and Figure 6-18 the following can be inferred: 1. The initial estimates of the stiffness parameters are significantly lower than the optimized values of the parameters. This is due to the fact that the initial values are based on triaxial compression test results, whereas most of the soil around the excavation experiences different stress paths (mainly undrained reduction of stresses). 2. The parameters that vary the most are the ones corresponding to the stiffer clay layers. The stiffer the clay, the more the laboratory test results are affected by sample disturbance and by the underestimation of stiffness values due to global measures of strains in a triaxial apparatus. 3. The simulated excavation is a 3D problem modeled in plane strain. When the excavated depth is small, most of the wall can be adequately modeled as plane strain and hence little changes in parameters are noted between stage 1 and 2. As the excavation deepens (stage 3), the ratio between excavation depth and excavation width increases and higher parameter values compensate for the lack of constraints in the outof-plane direction. 4. The highest changes in parameter values occur at stage 1. The fit between computed and field displacement after stage 1 is extremely satisfactory. Thus, the calibrated parameters only need to change slightly at later stages. By the end of stage 3 the model essentially is calibrated.

6.4.3.4 Model statistics for all construction stages The inverse analysis procedure produces very important by-products that allow one to quantify the effectiveness of the optimization procedure and assess the reliability of the predictions. Table 6-13 shows the values of a series of model fit statistics for the 5 optimization stages. Most of the statistics presented have been already defined elsewhere in this work, yet for convenience their expressions will be presented again in this section. Results show that the calibration improved the fit between observations and computed results up to stage 3, after which there was no improvement.

Table 6-13 Model statistics of inverse analysis at various optimization stages

(6.21)

(6.22)

(6.23)

(6.24)

(6.25)

(6.26)

(6.27)

(6.28)

(6.29) where b is the vector of the estimated parameters, y i is the vector of the observations used at stage i, y'(b)base_i is the vector of the results at stage i computed using the initial (base) estimates of the parameters, is the weight matrix, y'(b)in_i is the vector of the results at stage i computed using the estimates of the parameters at the beginning of that stage, y'(b)fin_i is the vector of the results at stage i computed using the optimized estimates of the parameters at that stage, NDi is the number of observations used at stage i, NP is the number of estimated parameters, y all is the vector of all observations, y'(b)all_i is the vector of all results computed using the optimized estimates of the

parameters at stage i, and OFbase is the global objective function computed using the initial (base) estimates of the parameters.

Figure 6-19 Error variance values at various optimization stages Figure 6-19 shows the base and the final values of the error variance at the various stages. The graph allows one to compare the overall magnitude of weighted residuals relative to the initial estimates of the parameters with the fit resulting from the calibrated models. Results show that error variance values decrease by more than an order of magnitude at every stage. This indicates a significant improvement in the fit between observed and computed displacements. The final error variance values are always very close to 1.0. This indicates that the fit is consistent with the weighting of the observations, and thus with the measurement errors associated with the inclinometer

data.

Figure 6-20 Objective function and relative fit improvement values

Figure 6-21 Global objective function and predictive fit improvement values Figure 6-20 shows the values of relative fit improvement, RFI, at the various stages and the values of the initial and final objective functions from which it is computed. The RFI indicates by what percentage the optimized results improved compared to the predictions at the beginning of a given stage. Results show that stage 1 observations improved the predictions by 2 orders of magnitude (i.e. RFI equal 99%) and that, by the end of stage 3, the recalibration of the model is complete.

Figure 6-22 Initial and optimized E50/Su ratios of the different soil layers computed at different optimization stages: (a) initial, (b) stage 1, (c) stage 2 and (d) stage 3 Figure 6-21 shows the values of the global objective function, OFi, and the values of the predicted fit improvement, PFIi, at the various stages. These statistics can be used to quantify the predictive capabilities of the analysis at the end of a given optimization stage. As one would expect, the global objective function values decrease as more observations become available (i.e. from stage 1 to stage 5). However, the PFI values indicate than most of the improvement happens at stage 1 and nothing is gained by including the observations of stages 4 and 5. These results are extremely promising because they indicate that early stage observations are able to recalibrate the finite element simulation in a way that is beneficial to the predictions at later stages.

6.4.3.5 Comments on the calibrated model In section 6.4.3.5 a procedure was presented to compare the initial stiffness-to-strength ratios, E50/Su, to typical values of undrained E/Su ratios. Figure 6-22 shows the values of the initial and optimized E50/Su ratios of the 5 soil layers at different optimization stages. In Table 6-14 the values of the final equivalent undrained ratios, (Eu/Su)equivalent, are presented. Note that: (i) the stiffness E50 was computed from the values of parameter E50ref (see Table 6-12) according to Eq. 6.13 and considering the stress conditions in the middle of the layers on the two sides of the excavation at the inclinometer locations, and (ii) the equivalent undrained ratios, (Eu/Su)equivalent, were computed from the optimized values of E50/Su following the procedure described in section 6.4.1.2 and assuming G50=Gin/3 and =0.2.

Table 6-14 Stiffness-to-strength ratios for the best-fit values of the input parameters Table 6-14 and Figure 6-22 showed that the optimized values of the stiffness-to-strength ratios increase with depth and are significantly higher than the initial values of the ratios. Note that the final values of (Eu/Su)equivalent for layers 1, 2 and 3 (i.e. Bloldgett and Deerfield layers) are close to values reported by Finno and Chung (1992) for normally

consolidated compressible Chicago clays (i.e. 400-600).

Figure 6-23 E50/Suvs. Su/'1 for the final calibrated model In Figure 6-23 the E50/Su ratios are plotted as a function of Su/ '1. The values in the figure refer to the optimized values of the parameters (see Table 6-14). Results show that the value of E50/Su increases linearly with Su/ '1. The following expression fits the data extremely well:

(6.30) Eq. (6.30) can be conveniently used to estimate, for a given clay layer, the value of E50 from estimates of Su and
' 1,

which are generally more easily available.

In Figure 6-24 the optimized E50/Su ratios are plotted as a function of ( '1- '3)/2Su, an approximate measure of the relative shear stress of the layers. Results show that the value of E52/Su decreases as ( '1- '3/25u increases, until the latter reaches the value of 1.0. Note that the slightly higher than 1.0 ( '1- '3/2Su) values reported in the graph are possible because the undrained shear strength, Su , is not an input parameter of the H-S model, thus its estimate does not influence the "exact" relative shear stress (whose maximum value is 1.0).

Figure 6-24 E50/Suvs. ('1-'3/Su) for the final calibrated model A linear function can be used to interpolate the data:

(6.31) If we assume that, initially, the vertical and horizontal directions are principal directions the ( 1- 3)/Su ratio can be written as:

(6.32) where is the vertical effective stress and ko the coefficient of earth pressure at rest.

Equations 6.31 and 6.32 can be combined to produce an equation to use for estimating the stiffness modulus E50 from values of Su, k0 and :

(6.33)

6.5 SUMMARY This chapter presented a numerical procedure that uses construction monitoring data to recalibrate, by inverse analysis, the input parameters of a finite element model of a

supported excavation. The inverse analysis methodology was developed and tested using data from a 39 ft deep excavation through soft clays in Chicago. The inverse analysis algorithm UCODE was used to optimize the PLAXIS plane-strain finite element simulation of the excavation. Five clay layers, modeled using the HardeningSoil model, were calibrated using inclinometer data that measured lateral movements of the soil behind the supporting walls. One stiffness parameter per layer was optimized in the regression analysis. The other model parameters were either kept constant at their initial value or related to the updated optimized parameters. For the first three construction stages, the set of parameters converged to a new set of optimized values based on the observations available up to that stage. The improvement on the fit between the computed and measured response was always significant. When monitoring data relative to the first construction stage (i.e. installation of the wall) were used, the recalibrated model proved able to adequately predict the behavior of the soil for all construction stages. These results are significant in that a successful recalibration of the model at an early construction stage positively affected the predictions of the soil behavior throughout construction.

CHAPTER 6 SUMMARY AND CONCLUSIONS During the Chicago and State Subway Renovation Project, the existing subway station and tunnels were exposed using the cut and cover excavation technique. The excavation extended down into the soft Chicago clay to a depth of 12.2 m. Space limitations presented significant challenges to the project in that the Frances Xavier Warde School is founded on shallow foundations located within 1.3 m of the face of the excavation. Options for providing lateral support of the excavation support system were further restricted by the presence of the subway station and twin subway tubes. To minimize the excavation-related damage to the adjacent school, a stiff excavation support system consisting of a secant pile wall, cross-lot struts, and tiebacks was chosen as the support system. The excavation support system performance data and the building response data obtained from this project were used to evaluate the conditions that led to ground movement-induced damage to the Warde School. The objectives of this work were to (1) relate excavation and construction activities to the onset of cracking in a structure adjacent to deep excavation in soft Chicago clay, and from this relation, suggest limiting criteria to preclude excavation-related damage to similar buildings, and (2) develop inverse analyses techniques to allow objective updating of finite element predictions of deformations associated with braced excavations. Based on the results of the monitoring effort and subsequent evaluation of the field performance of the excavation support system, the following conclusions are drawn:

1. The secant pile wall with its combined bracing system provided adequate support for the adjacent, shallow foundation-supported Warde School. As planned in the design, minor cracking occurred to nonload-bearing portions of the school. The maximum settlement observed during the project was 40 mm. Of the maximum settlement, 10 mm occurred during wall installation, 18 mm developed as the soil was excavated and 12 mm occurred during tunnel demolition and station renovation as a result of creep and reduction of wall stiffness. Similar percentages of movement components were observed at other monitored sections. Only about half the movements developed during excavation, indicating when the major design criterion is deformation control, movements associated with wall installation and long-term effects can become a significant portion of the total movements for similar stiff support systems. 2. Settlements extended beyond the secant pile wall a distance approximately equal to the depth of the secant pile wall. The effect of excavation was to cause larger settlements within the affected zone, but not to expand its width. This behavior suggests, for secant pile walls where significant movements occur during its installation, settlements should be normalized by wall depth, rather than be the depth of excavation. 3. Significant excavation-related movements did not occur until the excavation was lowered through the stiff clay crust layer at 5.5 m below grade. 4. In cases where little or no volume change is expected during excavation and when the deformations are small such that localized strains do not develop, excavation-related settlement behind a support wall can be estimated reasonably well from inclinometer

data. This is significant because comprehensive survey data are typically not available during construction. 5. When excavating through saturated clays, distortions obtained from inclinometer data can be used to estimate distortions in an adjacent structure when settlement and tiltmeter data are not available. Hence, inclinometer data can be used to monitor and update estimates of the likelihood and the severity of potential cracking to adjacent structures. Based on the results of the observed responses of the Warde school to the adjacent excavation activities, the following conclusions are drawn: 1. The damage to the school consisted mainly of hairline cracks 300 mm to 500 mm long in non-load bearing walls. Only a few cracks had widths greater than 6 mm. This damage is classified as slight according to the Burland et al. (1977) classification. 2. Cracks first were observed at a distortion of approximately 1/920 in a region where the building deformed in a sagging mode. Larger distortions occurred in sagging region than in the hogging zone. Consequently damage to the school, a 3-story reinforced concrete frame with load bearing exterior walls and a stiff floor system, would have been negligible if distortions in the sagging zone were limited to 1/1000. 3. Cracks in the west end of the north foundation wall and north exterior masonry wall were first observed at a distortion of approximately 1/1500 where the school deformed in a hogging mode. While these distortions were smaller than those in the sagging region, they occurred after the distortions reached 1/500 in the sagging zone.

4. Most of the cracking occurred along the west side of the school in the area where distortions were largest, and when the maximum distortions increased from 1/920 to 1/400. This response corresponds to the excavation being advanced from the top of the soft clay layer (at a depth of approximately 5.5 m) to the final depth (at a depth of approximately 12.2 m). Further increases to 1/320 during station renovation and backfill caused little new cracking. Based on the results of the finite element studies presented herein, the following conclusions are drawn: 1. The inverse modeling procedure in UCODE, when coupled with the commercial finite element code, PLAXIS, is able to provide optimized parameters which when used in finite element simulations of the Chicago-State excavation resulted in quantitatively better computations of deformations. 2. When monitoring data obtained during the installation of the secant pile wall were used as the basis of optimization, the recalibrated model was able to adequately predict the magnitude and distribution of lateral wall movements for all subsequent construction stages. These results are significant in that a successful recalibration of the model at an early construction stage positively affected the movement predictions throughout construction.

REFERENCES 1. Amrhein, J. E. (1998). Reinforced Masonry Engineering Handbook: Clay and Concrete Masonry, 5th edition updated, Masonry Institute of America, Los Angeles, CA, pp. 469. 2. Bjerrum, L. (1963). Allowable Settlements of Structures, Proceeding of European Conference on Soil Mechanics and Foundation Engineering, Weisbaden, Vol. 2, pp. 135-137. 3. Boone, S. J. (1996). Ground-Movement-Related Building Damage, Journal of Geotechnical Engineering, ASCE, Vol. 122, No. 11, pp. 886-896. 4. Boone, S. J., Westland, J., and Nusink, R., (1999). Comparative Evaluation of Building Responses to and Adjacent Braced Excavation, Canadian Geotechnical Journal, Canadian Geotechnical Society, Vol. 36, pp. 210-223. 5. Boscardin, M. D. and Cording, E. J. (1989). Building Response to ExcavationInduced Settlement, Journal of Geotechnical Engineering, ASCE, Vol. 115, No. 1, pp. 1-21. 6. Boscardin, M. D., Cording, E. J., and ORourke, T. D. (1979). Case Studies of Building Behavior in Response to Adjacent Excavation, University of Illinois Report for the U.S. Department of Transportation, Report No. UMTA-IL-06-0043-78-2. 7. Bozozuk, M. (1962). Soil Shrinkage Damages Shallow Foundations at Ottawa Canada, The Engineering Journal, Canada, pp. 33-37.

8.Brinkgreve R.B.J. and Vermeer P.A. (1998). Finite Element Code for Soil and Rock Analysis. PLAXIS 7.0 manual. Balkema. 9. Burland, J. B., and Wroth, C. P., (1974). Settlement of Buildings and Associated Damage, Proceeding of a Conference on Settlement of Structures, Cambridge, pp. 611654. 10. Burland, J. B., Broms, B. B., and de Mello, V. F., (1977). Behavior of Foundations and Structures, 9th International Conference on Soil Mechanics and Foundation Engineering, Tokyo, State-of-the-Art Report, Vol. 2, pp. 495-546. 11. Bushell, T. D., McCarthy, D. W., and McCluskey, E., (1999). Performance of Multiple Retention Systems During Cut and Cover Tunnel Construction, GeoEngineering for Underground Facilities, Proceeding of the Third National Conference, ASCE, Geotechnical Special Publication No. 90, pp. 888-899. 12. Calvello,M. (2002). Inverse analysis of supported excavations through Chicago glacial clays. PhD Thesis, Northwestern University, Evanston,IL. 13. Caspe M.S. (1966) Surface settlement adjacent to braced open cuts. Journal of the Soil Mechanics and Foundations Division, ASCE, Vol. 92, p. 51-59. 14. Canadian Foundation Engineering Manual, (1992), 3rd edition, Canadian Geotechincal Society, BiTech Publishers Ltd., Vancouver, BC, Canada, 512 p.

15.. Chung, C. K. and Finno, R. J. (1992). Influence of Depositional Processes on the Geotechnical Parameters of Chicago Glacial Clays, Engineering Geology, Vol. 32, 1992, pp. 225-242. 16. Clough, G. W. and Reed, M. W. (1984). Measured Behavior of Braced Wall in Very Soft Clay, Journal of Geotechnical Engineering, ASCE, Vol. 110, No. 1, pp. 1-19. 17. Clough, G. W., and O`Rourke, T. D. (1990). Construction Induced Movements of In-Situ Walls, Design and Performance of Earth Retaining Structures, Proceedings of a Specialty Conference at Cornell University, ASCE, New York, pp. 439-470. 18. Clough, G. W., Smith, E. M., and Sweeney, B. P. (1989). Movement Control of Excavation Support Systems by Iterative Design, Current Principles and Practices, Foundation Engineering Congress, Vol. 2, ASCE, pp. 869-884. 19. Cunningham, J. A., and Fernandez, J. I. (1972). Performance of Two Slurry Wall Systems in Chicago, Proceeding of the ASCE Specialty Conference on Performance of Earth and Earth-Supported Structures, Lafayette, Indiana, Vol. 1, Part 2, pp. 1425-1449. 20. DAppolonia, D. J. (1971). Effects of Foundation Construction on Nearby Structures, Proceedings of the 4th Pan-American Conference on Soil Mechanics and Foundation Engineering, ASCE, New York, Vol. 1, pp. 189-236. 21. Dulacska, E. (1992). Soil Settlement Effects on Buildings, Developments in Geotechnical Engineering, No. 69, Elsevier, New York, pp. 447.

22.. ETABS (1999). ETABS Linear and Nonlinear Static and Dynamic Analysis and Design of Building Systems, Users Manual, Volumes 1 and 2, 1st ed, Computer and Structures, Inc., Berkeley, CA, 895 pp. 23. Finno, R. J. (1992). Deep Cuts and Ground Movements in Chicago Clay, Excavation and Support for the Urban Infrastructure, Geotechnical Special Publication No. 33, ASCE, New York, pp. 119-143. 24. Finno R.J. and Chung C.K. (1992). Stress-strain-strength responses of compressible Chicago glacial clays. Journal of Geotechnical Engineering, ASCE, Vol. 118, No. 10, p. 1607-1625. 25. Finno, R. J. and Harahap, I. S. (1991). Finite Element Analysis of the HDR-4 Excavation, Journal of Geotechnical Engineering, ASCE, Vol. 117, No. 10, pp. 15901609. 26. Finno, R. J., Atmatzidis, D. K., and Perkins, S. B. (1989). Observed Performance of a Deep Excavation in Clay, Journal of Geotechnical Engineering, ASCE, Vol. 115, No. 8, pp. 1045-1064. 27. Finno R.J., Bryson L.S. and Calvello M. (2002). Performance of a stiff support system in soft clay. Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 128, No. 8, p. 660-671. 28. Goldberg, D. T., Jaworski, W. E., and Gordon, M. D. (1976). Lateral Support Systems and Underpinning, Vols. I, II, and III, Federal Highway Administration, Report No. FHWA-RD-75-127, -128, and -129.

29. Hill M.C. (1998). Methods and guidelines for effective model calibration. U.S. Geological Survey Water-Resources investigations report 98-4005, 90 pp. 30. Hsieh, P-G. and Ou, C-Y. (1998). Shape of Ground Surface Settlement Profiles Caused by Excavation, Canadian Geotechnical Journal, Vol. 35, No. 6, pp. 1004-1017. 31. Kawamura, K.M. (1999). Hardening-Soil Parameters for Compressible Chicago Glacial Clays. Masters Thesis, Northwestern University, Evanston, Illinois, USA. 32. Koutsoftas, D. C., Frobenius, P., Wu, C. L., Meyersohn, D. and Kulesza, R. (2000). Deformations During Cut-and-Cover Construction of MUNI Metro Turnback Project, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 126, No. 4, pp. 344-359. 33. Lamb, T. W. (1970). Braced Excavations, Proceeding of the ASCE Specialty Conference on Lateral Stresses in the Ground and Design of Earth Retaining Structures, Ithaca, New York, pp. 149-218. 34. Lambe T.W. and Whitman R.V. (1969). Soil mechanics. John Wiley and Sons. New York. 35. Lee, F. H., Yong, K. Y., Quan, K. C. N., and Chee, K. T. (1998). Effect of Corners in Strutted Excavation: Field Monitoring and Case Histories, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 124, No. 4, pp. 339-349.

36. Leonards, G. A. (1975). Discussion of Differential Settlement of Buildings, by Grant, R., Christian, J. T., and Vanmarcke, E. H., Journal of Geotechnical Engineering Division, ASCE, Vol. 101, No. GT7, pp. 700-702. 37. MacGregor, J. G. (1997). Reinforced Concrete: Mechanics and Design, 3rd ed., Prentice Hall, Upper Saddle River, NJ, pp. 939. 38. Mair, R. J., Taylor, R. N., and Burland, J. B. (1996). Prediction of Ground Movements and Assessment of Risk of Building Damage Due to Bored Tunneling, Proceeding of the International Symposium on Geotechnical Aspects of Underground Construction in Soft Ground, edited by R. J. Mair and R. N. Taylor, London, Balkema, pp 713-718. 39. Mana, A. I. and Clough, G. W. (1981). Prediction of Movements for Braced Cut in Clay, Journal of Geotechnical Engineering, ASCE, Vol. 107, No. 8, pp. 759-777. 40. Mainstone, R. J. (1971). On the Stiffness and Strengths of Infilled Frames, Proceedings of the Institution of Civil Engineers, The Institution of Civil Engineers, London, England, Supplemental Vol. IV, Paper 7360s, pp. 57-90. 41. Mainstone, R. J., (1974). Discussion on Session No. 5, Proceedings of the Conference on Settlement of Structures, Cambridge, Pentech Press, London, p. 771. 42. Mainstone, R. J., and Weeks, G. A., (1970). The Influence of a Bounding Frame on the Racking Stiffness and Strengths of Brick Walls, Proceedings of the 2nd International Brick and Masonry Conference, Building Research Establishment, Watford, England, pp. 165-171.

43. Meyerhof, G. G. (1982). Limit States Design in Geotechnical Engineering, Structural Safety, No. 1, pp. 67-71. 44. Meyerhof, G. G. (1956). Discussion of Settlement Analysis of Six Structures in Chicago and London, by Skempton, A. W., Peck, R. B., and MacDonald, D. H., Proceedings of the Institution of Civil Engineers, London, England, Vol. 5, No. 1, p. 170. 45. Meyerhof, G. G. (1953). Some Recent Foundation Research and its Application to Design, Structural Engineer, Vol. 31, pp. 151-167. 46. Milligan, G. W. E., (1985). Soil Deformations Behind Retaining Walls, Ground Movements and Structures, Pentech Press, London, p. 702-722. 47. Morgenstern N. (1995). Managing risk in geotechnical engineering. Proc. 10th Pan American Conference on Soil Mechanics and Foundation Engineering, Vol. 4. 48. NAVFAC DM-7.2 (1982) Design Manual: Foundations and Earth Structures, U.S. Department of the Navy, Washington, DC 49. ORourke, T. D. (1981). Ground Movements Caused by Braced Excavation, Journal of Geotechnical Engineering, ASCE, Vol. 107, No. 9, pp. 1159-1178. 50. ORourke, T. D., Cording, E. J., and Boscardin, M. D. (1976). The Ground Movements Related to Braced Excavations and Their Influence on Adjacent Structures, University of Illinois Report for the U.S. Department of Transportation, Report No. DOT-TST-76T-22.

51. Otto, George H. (1963). Engineering Geology of the Chicago Area, Foundation Engineering in the Chicago Area: 3-1 3-24. 52. Ou, C. Y. and Shiau, B. Y., (1998). Analysis of the Corner Effect on Excavation Behavior, Canadian Geotechnical Journal, Canadian Geotechnical Society, Vol. 35, pp. 532-540. 53. Ou, C. Y., Chiou, D. C., and Wu, T. S. (1996). Three-Dimensional Finite Element Analysis of Deep Excavations, Journal of Geotechnical Engineering, ASCE, Vol. 122, No. 5, pp. 337-345. 54. Peck, R. B. (1969). Deep Excavations and Tunneling in Soft Ground, Proceedings 7th International Conference on Soil Mechanics and Foundation Engineering, State-ofthe-Art Volume, pp. 225-290. 55. Peck, R. B. (1969b). Advantages and Limitations of the Observational Method in Applied Soil Mechanics, Geotechnique, Vol. 18, No. 2, pp. 171-187. 56. Peck, R.B. and Reed, W.C. (1954). Engineering Properties of Chicago Subsoils, Bulletin No. 423, Engineering Experiment Station, University of Illinois, Urbana, Illinois. 57. Poeter E.P. and Hill M.C. (1998). Documentation of UCODE, a computer code for universal inverse modeling. U.S. Geological Survey Water-Resources investigations report 98-4080, 116 pp.

58. Poeter EP and Hill M.C. (1997). Inverse Methods: A Necessary Next Step in Groundwater Modeling, Ground Water, v. 35, no. 2, p. 250-260. 59. Poh, T. Y. and Wong, I. H., (1998). Effects of Construction of Diaphragm Wall Panels on Adjacent Ground: Field Trial, Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 124, No. 8, pp. 745-756. 60. Polshin, D. E., and Tokar, R. A. (1957). Maximum Allowable Non-Uniform Settlement of Structures, Proceedings of the 4th International Conference on Soil Mechanics and Foundation Engineering, London, England, Vol. 1, pp. 402-405. 61. Roboski, J. F. (2001). Soil Parameters for Constitutive Models of Compressible Chicago Glacial Clays. Masters Thesis, Northwestern University, Evanston, Illinois, USA. 62. SAP2000 (2000). SAP2000 Three Dimensional Static and Dynamic Finite Element Analysis and Design of Structures, Analysis Reference, version 7.4, Computer and Structures, Inc., Berkeley, CA, 421 pp. 63. Schanz T., Vermeer P.A. and Bonnier P.G. (1999). The Hardening Soil model formulation and verification. Proceedings Plaxis Symposium "Beyond 2000 in Computational Geotechnics", Amsterdam, Balkema, p. 281-296. 64. Skempton, A. W., and MacDonald, D. H. (1956), The Allowable Settlements of Buildings, Proceedings of the Institution of Civil Engineers, London, England, Part 3, Vol. 6: pp. 727-768.

65. Swatek, E. P., Asrow, S. P., and Seitz, A. M. (1972), Performance of Bracing for Deep Chicago Excavation, Proceeding of the ASCE Specialty Conference on Performance of Earth and Earth-Supported Structures, Lafayette, Indiana, Vol. 1, pp. 1303-1322. 66. Terzaghi, K. (1943), Theoretical Soil Mechanics, John Wiley and Sons, Inc., New York, NY. 67. Terzaghi, K. and Peck, R. (1967), Soil Mechanics in Engineering Practice, 2nd ed., John Wiley and Sons, Inc., New York, NY, pp. 729. 68. Tschebotarioff, G. P. (1973), Foundations, Retaining and Earth Structures, 2nd ed., McGraw-Hill Book Company, New York, NY, 642 pp. 69. Viggiani G. and Atkinson J. (1995). Stiffness of fine grained soils at very small strains. Geotechnique, Vol. 45, p. 249-265.

APPENDIX A: CONSTRUCTION ACTIVITY, TIME HISTORY AND STRUCTURAL PLAN

Figure A-1- Chicago Avenue and State Street Subway Renovation Project Structural Staging Plan

Table A-1- Construction Activities Time History Appendix A: Construction Activities Time History and Structural Plan
Date 6/21/1999 6/10/1999 6/11/1999 6/11/1999 6/14/1999 6/14/1999 6/16/1999 6/17/1999 6/29/1999 8/11/1999 10/8/1999 10/22/1999 8/4/1999 8/12/1999 8/14/1999 8/25/1999 8/27/1999 8/30/1999 9/8/1999 9/16/1999 6/22/1999 6/24/1999 6/21/1999 6/23/1999 6/25/1999 6/28/1999 6/21/1999 6/22/1999 6/23/1999 6/25/1999 6/26/1999 6/28/1999 6/29/1999 -11 -10 -10 -7 -7 -5 -4 8 51 109 123 44 52 54 65 67 70 79 87 1 3 0 2 4 7 0 1 2 4 5 7 8 TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity Start of Construction Inclinometer and Piezometer Installation Inclinometer 1 (74 ft deep) Piezometer 1A and 1B (cells @ 33 ft and 43 ft BGS) Inclinometer 2A (only 54 ft deep because of squeezing ) + 3 samples Inclinometer 4A (74 ft deep) + 1 Sample Piezometer 4A and 4B (cells @ 35ft and 45 ft BGS) Inclinometer 3 (74 ft deep) Inclinometer 2B (74 ft deep) Inclinometer 5 (74 ft deep) Inclinometer 4 Damage Inclinometer 4B (80 ft BGS) Inclinometer 4B Destroyed Inclinometer 4C (80ft bgs) Piezometer 4 Damage Piezometer 4 Broken New Piezometer 4 (4C @ 35ft and 4D @ 45ft BGS) Inclinometer 1 Damage Inclinometer 1 broken 5 ft BGS Inclinometer 5 Damage Inclinometer 5 broken Inclinometer 5B (80 ft BGS) Inclinometer 5B cut down Inclinometer 5B read with extension (down to 74') Inclinometer 5B blown out (back down to 80') Southeast Wall Installation PILES E2 PILES E1, E3 East Wall Installation - South PILES E4 PILES E6 PILES E5 PILES E7 East Wall Installation - INCL 2B PILES E8 PILES E10, E14, E18 PILES E6, E12, E16 PILES E5, E9, E13 PILES E17 PILES E7, E11 PILES E15 East Wall Installation - Central

Date 6/21/1999 6/22/1999 6/23/1999 6/24/1999 6/25/1999 6/28/1999 6/29/1999 6/30/1999 6/23/1999 6/24/1999 6/25/1999 6/28/1999 6/29/1999 6/30/1999 7/2/1999 6/24/1999 6/25/1999 7/2/1999 7/14/1999 7/17/1999 7/19/1999 7/21/1999 8/30/1999 8/31/1999 9/1/1999 9/2/1999 9/3/1999 9/4/1999 9/7/1999 9/8/1999 6/21/1999 7/6/1999 7/7/1999 6/30/1999 7/6/1999 7/7/1999 7/8/1999 7/14/1999 7/15/1999 7/17/1999

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 1 2 3 4 7 8 9 2 3 4 7 8 9 11 3 4 11 23 26 28 30 70 71 72 73 74 75 78 79 0 15 16 9 15 16 17 23 24 26 Start of Construction PILES E22(beg.) PILES E22(end), E26, E30 PILES E20, E24, E28, E32 PILES E27 PILES E19 PILES E21, E25, E31 PILES E23, E29 East Wall Installation - INCL 1 PILES E34, E37 PILES E42 PILES E36, E40 PILES E33, E35 PILES E39 PILES E38, E43 PILES E41 East Wall Installation - North 1 (grouped based on dates) PILES E45 PILES E46 PILES E44 East Wall Installation - North 2 (grouped based on dates) PILES E47 PILES E48, E51 PILES E49 PILES E50 Install North Wall - Phase II PILES E67 PILES E55, E59, E63 PILES E57, E61 PILES E53, E65 PILES E60, E64, E68 PILES E52, E56, E66 PILES E62 PILES E54-E58 West Wall Installation - W4 PILES W4 Southwest Wall Installation PILES W2 PILES W1, W3 West Wall Installation - South PILES W7 PILES W6, W8 PILES W5(beg.), W11 PILES W5(end) PILES W12 (beg.) PILES W9, W12 (end) PILES W10

Date 6/21/1999 6/30/1999 7/2/1999 7/15/1999 7/17/1999 7/21/1999 7/23/1999 7/26/1999 7/27/1999 7/28/1999 7/6/1999 7/7/1999 7/8/1999 7/6/1999 7/7/1999 7/8/1999 7/9/1999 7/17/1999 7/19/1999 7/21/1999 7/22/1999 7/23/1999 6/30/1999 7/2/1999 7/15/1999 7/19/1999 7/9/1999 7/14/1999 10/13/1999 10/19/1999 10/20/1999 10/21/1999 10/22/1999 8/16/1999 8/18/1999 8/19/1999 8/20/1999

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity Start of Construction 9 11 24 26 30 32 35 36 37 15 16 17 15 16 17 18 26 28 30 31 32 9 11 24 28 18 23 114 120 121 122 123 56 58 59 60 West Wall Installation - Drift Tunnel PILES W13(beg.)*, W15(beg.)*, W19(beg.)* PILES W17(beg.)* PILES W14 PILES W15(end) PILES W13(end), W17(end) PILES W19(end) PILES W16 PILES W20 PILES W18 West Wall Installation - Central PILES W23, W25, W27, W31 PILES W21, W24, W26, W29, W32 PILES W22, W28, W30 West Wall Installation - INCL 4 PILES W33, W39, W43, W47 PILES W35, W37, W41, W45 PILES W34, W40, W44, W46 PILES W36, W38, W42 West Wall Installation - North PILES W48 PILES W49, W51 PILES W53 PILES W52 PILES W50 Drift Tunnel Drift Tunnel found (W13 thru W19) Sink Hole in Parking Lot Beging Concrete Backfill End Concrete Backfill South Soldier Pile Wall Installation SOLDIER PILE S1, S2, S3, S6, S7, S8, S9, S10, S11 SOLDIER PILE S4, S5 Install Sheet Pile Installed WF PZ27 sheet pile in northwest corner of the excavation North Soldier Pile Wall Installation SOLDIER PILE 2-14, 2-15, 2-17, 2-18 SOLDIER PILE 2-13, 2-12, 2-11, 2-10 SOLDIER PILE 2-9, 2-8, 2-7, 2-6, 2-5 SOLDIER PILE 2-19, 2-20, 2-21, 2-4 Install Struts - Phase I STRUT #1 STRUT #2 STRUT #3 (with strain gauges) STRUT #4 (with strain gauges)

Date 6/21/1999 9/3/1999 9/8/1999 2/14/2000 2/23/2000 2/25/2000 3/5/2000 10/27/1999 11/1/1999 1/13/2000 2/11/2000 2/4/2000 7/16/1999 7/20/1999 7/20/1999 7/24/1999 7/23/1999 2/26/2000 3/5/2000 7/29/1999 7/30/1999 7/29/1999 7/30/1999 2/26/2000 3/5/2000 10/26/1999 2/26/2000 8/3/1999 8/4/1999 8/5/1999

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 74 79 238 247 249 258 128 133 206 235 228 25 29 29 33 32 250 258 38 39 38 39 250 258 127 250 43 44 45 Start of Construction STRUT #5 (with strain gauges) STRUT #6 Remove Struts - Phase I Cut STRUT #4 and STRUT #5 and relocated adjacent to STRUT #3 and STRUT #6 Removed STRUTS #1 and STRUTS #2 Removed STRUTS #3 and SE corner diagonal bracing Removed STRUTS #4, STRUT #5, and STRUTS #6 Install and Remove Struts - Phase II 1st Level of bracing on North Wall Installed 1st Level Bracing at the west end of Stair #3 Lower Level Phase II Strut, inclined at the SW corner and bolted to Mezz. floor Welder cut lower level and upper level Phase II bracing South Soldier Pile Wall Struts and Walers South Soldier Pile Wall: Removed lower level bracing and waler Waler Installation and Removal - East Wall Began excavating trench (10ft-wide x 5ft-deep) for east secant pile wall waler End excavating trench for east secant pile wall waler Began installing upper-level wale on the east wall and southeast corner bracing End installing southeast corner bracing End installing upper-level waler on the east wall Removed east upper level waler from <M1> to <P7> Removed remaining east upper level waler Waler Installation and Removal - West Wall Began excavating trench (9ft-wide x 5ft-deep) for west secant pile wall waler End excavating trench for west secant pile wall waler Excavate to 5ft bgs from W6 to W50 for west wall upper waler installation Installed upper-level waler on the west wall Removed west upper level waler from <M1> to <P7> Removed remaining west upper level waler Waler Installation and Removal - North Wall Installed Waler on North Wall at EL +11.0 Removed Phase II upper level waler Install Sanitary Sewer on West Side Excavate 12 ft deep trench on west side of west wall for 24in SS (W5 to W30) Excavate 12 ft deep trench on west side of west wall for 24in SS (W30 to W40) Excavate 12 ft deep trench on west side of west wall for 24in SS (W40 to W50) Install Sanitary Sewer on East Side Began install and excavate 12 ft deep trench on east side of east wall for 30in

8/12/1999

52

Date 6/21/1999 8/17/1999

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity Start of Construction SS DIP End install and excavate 12 ft deep trench on east side of east wall for 30in SS DIP Install Sanitary Sewer on North Side Began install and excavate 12 ft deep trench on northwest side of FXW for 30in SS DIP End install and excavate 12 ft deep trench on northwest side of FXW for 30in SS DIP Began install and excavate 12 ft deep trench on north side of FXW for 30in SS DIP End install and excavate 12 ft deep trench on north side of FXW for 30in SS DIP Chipping East Wall East Wall Chipped (to flange) at 12ft bgs to accommodate 30in SS DIP from E42 to E50 East Wall Chipped (to flange) at el -1.0 from E6 to E22 East Wall Chipped (to flange) to Pile E24 East Wall Chipped (to flange) to Pile E34 East Wall Chipped (inside flange) from Pile E22 to E32 Per revised shop dwg detail East Wall Chipped (inside flange) from E32 to E34 East Wall Chipped (inside flange) from E34 to E36 East Wall Chipped (inside flange) from E36 to E37 East Wall Chipped (inside flange) from E36 to E42 East Wall Chipped (inside flange) from E42 to E44 East Wall Chipped (inside flange) from E44 to E46 East Wall Chipped (inside flange) from E46 to E50 East Wall Chipped (inside flange) from E46 to E51 East Wall Chipped from E6 to E16 (2nd level) East Wall Chipped to E40 (2nd level) East Wall Chipped to E47 (2nd level) East Wall Chipped to E51 (2nd level) East Wall Chipped from EL +11.0 to -1.0 from E50 to E32 East Wall Chipped from EL -1.0 to -14.0 from E44 to E34 East Wall Chipped from EL -1.0 to EL -14.0 from E50 to E44 Begin East Wall Chipped from EL -8.0 to EL -12.0 from E34 to E14 End East Wall Chipped from EL -8.0 to EL -12.0 from E34 to E14 Begin East Wall Chipped from EL -8.0 to EL -12.0 from E14 to E7 End East Wall Chipped from EL -8.0 to EL -12.0 from E14 to E7 East Wall Chipped to Flange from EL +11.0 to EL -1.0 from E32 to E26 East Wall Chipped to Flange from EL -1.0 to EL -14.0 from E34 to E7 East Wall Chipped to Flange from EL -14.0 to EL -25.0 from E34 to E17 East Wall Chipped to Flange from EL -14.0 to EL -25.0 from E50 to E42 East Wall Chipped from EL +11.0 to -1.0 from E26 to E17 East Wall Chipped to Flange from EL -14.0 to EL -25.0 from E17 to E12 East Wall Chipped to Flange between UL and LL from E51 to E54 Chipping West Wall

57

8/21/1999 8/27/1999 11/3/1999 11/8/1999

61 67 135 140

8/17/1999 8/20/1999 8/21/1999 8/23/1999 8/25/1999 8/26/1999 8/28/1999 9/1/1999 9/2/1999 9/7/1999 9/9/1999 9/10/1999 9/13/1999 9/18/1999 9/22/1999 9/25/1999 10/2/1999 10/9/1999 10/9/1999 10/16/1999 10/16/1999 10/18/1999 10/18/1999 10/21/1999 11/5/1999 11/5/1999 11/5/1999 11/8/1999 11/17/1999 11/17/1999 12/5/1999

57 60 61 63 65 66 68 72 73 78 80 81 84 89 93 96 103 110 110 117 117 119 119 122 137 137 137 140 149 149 167

Date 6/21/1999 8/20/1999 8/23/1999 8/24/1999 9/7/1999 9/10/1999 9/13/1999 9/18/1999 9/22/1999 9/25/1999 10/2/1999 10/16/1999 10/21/1999 10/29/1999 10/29/1999 11/1/1999 11/1/1999 11/2/1999 11/4/1999 11/5/1999 11/5/1999 12/2/1999 12/3/1999

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 60 63 64 78 81 84 89 93 96 103 117 122 130 130 133 133 134 136 137 137 164 165 Start of Construction West wall Chipped at el -3.0 from W6 to W20 (to flange) West Wall Chipped from W20 to W26 (inside flange) West Wall Chipped from W26 to W30 (inside flange) West Wall Chipped from W30 to W34 West Wall Chipped from W34 to W38 West Wall Chipped from W38 to W42 West Wall Chipped from W42 to W50 West Wall Chipped to W24 (2nd level) West Wall Chipped to W38 (2nd level) West Wall Chipped to W46 (2nd level) West Wall Chipped from EL-8.0 to EL -16.0 from W46 to W14 West Wall Chipped from EL-8.0 to EL -16.0 from W50 to W46 West Wall Chipped from EL+11.0 to EL -1.0 from W50 to W32 West Wall Chipped from EL-1.0 to EL -15.0 from W50 to W32 West Wall Chipped from EL-15.0 to EL -26.0 from W24 to W22 West Wall Chipped from EL+11.0 to EL -1.0 from W32 to W20 West Wall Chipped from EL-15.0 to EL -26.0 from W54 to W34 West Wall Chipped from EL-15.0 to EL -27.0 from W18 to W6 West Wall Chipped to Flange from EL -1.0 to EL -15.0 from W32 to W6 West Wall Chipped to Flange from EL -15.0 to EL -26.0 from W34 to W24 West Wall Minor Chipped to Flange Between UL and LL Tieback Walers from W52 to W26 West Wall Minor Chipped to Flange Between UL and LL Tieback Walers from W26 to W16 Chipping North Wall North Wall Chipped to flange at EL +11.0 from E68 to E55 North Wall Chipped to flange from E68 to E55 Northeast Corner of East Wall Chipped Install East Wall Tieback Wales Installed Walers on Flange face E6 to E16 Installed Walers on Flange face E16 to E22 Installed Walers inside Flange E22 to E26 Installed Walers inside Flange E26 to E28 Installed Walers inside Flange E28 to E30 Installed Walers inside Flange E30 to E36 Installed Walers inside Flange E36 to E42 Installed Walers inside Flange E42 to E44 Installed Walers inside Flange E44 to E46 Installed Walers inside Flange E46 to E48 Installed Walers inside Flange E46 to E50 Installed 2nd Level Walers to E34 Installed 2nd Level Walers to E46 Install West Wall Tieback Wales Installed Walers on Flange face W6 to W20 Installed Walers inside Flange W20 to W24 Installed Walers inside Flange to W28

10/26/1999 1/6/2000 1/6/2000 8/20/1999 8/24/1999 8/27/1999 8/28/1999 8/30/1999 8/31/1999 9/7/1999 9/8/1999 9/9/1999 9/10/1999 9/13/1999 9/22/1999 9/27/1999 8/23/1999 8/26/1999 8/27/1999

127 199 199 60 64 67 68 70 71 78 79 80 81 84 93 98 63 66 67

Date 6/21/1999 8/31/1999 9/7/1999 9/18/1999 9/22/1999 9/27/1999 9/1/1999 9/2/1999 9/3/1999 9/7/1999 9/8/1999 9/10/1999 9/13/1999 9/29/1999 9/2/1999 9/3/1999 9/6/1999 9/8/1999 9/9/1999 9/13/1999 9/14/1999 9/30/1999 9/10/1999 9/13/1999 9/14/1999 9/15/1999 9/16/1999 9/17/1999 10/4/1999 9/22/1999 9/23/1999 9/24/1999 9/25/1999 9/27/1999 9/28/1999 10/6/1999 10/7/1999 9/23/1999 9/24/1999 9/25/1999 9/27/1999 9/30/1999

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 71 78 89 93 98 72 73 74 78 79 81 84 100 73 74 77 79 80 84 85 101 81 84 85 86 87 88 105 93 94 95 96 98 99 107 108 94 95 96 98 101 Start of Construction Installed Walers inside Flange W28 to W30 Installed Walers inside Flange W30 to W34 Installed 2nd Level Walers from W6 to W12 Installed 2nd Level Walers to W12 and W20 to W25 Installed 2nd Level Walers to W36 Install 1st Level Tiebacks--East Wall TIEBACKS E7 (instl) TIEBACKS E9 (instl), E11 (instl), E13 (instl), E15 (instl) TIEBACKS E17 (instl), E19(instl), E21(instl), E23 (instl), E31(instl) TIEBACKS E25 (instl), E27 (instl), E29 (instl), E35 (instl), E37 (instl) TIEBACKS E33 (instl), E39 (instl), E41 (instl) TIEBACKS E43 (instl), E45 (instl) TIEBACKS E47 (instl), E49 (instl) TIEBACKS E51 (instl), E53 (instl) Regrout 1st Level Tiebacks--East Wall TIEBACKS E7 (rgrt) TIEBACKS E9 (rgrt), E11 (rgrt), E13 (rgrt), E15 (rgrt) TIEBACKS E17 (rgrt), E19 (rgrt), E21 (rgrt) E23 (rgrt) TIEBACKS E25 (rgrt), E27 (rgrt), E29 (rgrt), E35 (rgrt) E37 (rgrt) TIEBACKS E31 (rgrt), E33 (rgrt), E39 (rgrt), E41 (rgrt) TIEBACKS E43 (rgrt), E45 (rgrt) TIEBACKS E47 (rgrt), E49 (rgrt) TIEBACKS E51 (rgrt), E53 (rgrt) Tension 1st Level Tiebacks--East Wall TIEBACKS E7 (tnsn), E9 (tnsn), E11 (tnsn), E13 (tnsn) TIEBACKS E15 (tnsn), E17 (tnsn), E19 (tnsn), E21 (tnsn) TIEBACKS E23 (tnsn), E25 (tnsn), E29 (tnsn), E31 (tnsn) TIEBACKS E33 (tnsn) TIEBACKS E37 (tnsn), E39 (tnsn), E41 (tnsn), E43 (tnsn) TIEBACKS E27 (tnsn), E35 (tnsn), E45 (tnsn), E47 (tnsn), E49 (tnsn) TIEBACKS E51 (tnsn), E53 (tnsn) Install 2nd Level Tiebacks--East Wall TIEBACKS E7 (instl), E9 (instl) TIEBACKS E11 (instl) E13 (instl), E15 (instl), E17 (instl), E19 (instl) TIEBACKS E21, (instl), E23 (instl), E25 (instl), E27 (instl) TIEBACKS E29 (instl), E31 (instl), E33(instl) TIEBACKS E35 (instl), E37 (instl), E39 (instl) TIEBACKS E41 (instl), E43 (instl) TIEBACKS E45 (instl), E47 (instl), E49 (instl) TIEBACKS E51 (instl), E53 (instl) Regrout 2nd Level Tiebacks--East Wall TIEBACKS E7 (rgrt), E9 (rgrt) TIEBACKS E11 (rgrt), E13 (rgrt), E15 (rgrt), E17 (rgrt), E19 (rgrt) TIEBACKS E21 (rgrt), E23 (rgrt), E25 (rgrt), E27 (rgrt) TIEBACKS E29 (rgrt), E31 (rgrt), E33 (rgrt) TIEBACKS E35 (rgrt), E37 (rgrt), E39 (rgrt), E41 (rgrt), E43 (rgrt)

Date 6/21/1999 10/9/1999 9/27/1999 9/30/1999 10/1/1999 10/4/1999 10/11/1999 10/12/1999 8/25/1999 8/26/1999 8/28/1999 8/30/1999 8/31/1999 9/1/1999 9/7/1999 9/8/1999 9/14/1999 9/15/1999 9/16/1999 9/17/1999 9/30/1999 8/27/1999 8/31/1999 9/1/1999 9/2/1999 9/3/1999 9/9/1999 9/10/1999 9/17/1999 9/21/1999 9/30/1999 10/1/1999 9/2/1999 9/7/1999 9/8/1999 9/9/1999 9/16/1999 9/20/1999 9/22/1999 9/24/1999

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 110 98 101 102 105 112 113 65 66 68 70 71 72 78 79 85 86 87 88 101 67 71 72 73 74 80 81 88 92 101 102 73 78 79 80 87 91 93 95 Start of Construction TIEBACKS E45 (rgrt), E47 (rgrt), E49 (rgrt), E51 (rgrt), E53 (rgrt) Tension 2nd Level Tiebacks--East Wall TIEBACKS E7 (tnsn), E9 (tnsn), E11 (tnsn), E13 (tnsn), E15 (tnsn), E19 (tnsn), E21 (tnsn) TIEBACKS E21 (tnsn), E23 (tnsn), E25 (tnsn) TIEBACKS E29 (tnsn), E31 (tnsn), E33 (tnsn) TIEBACKS E27 (tnsn), E35 (tnsn), E37 (tnsn), E39 (tnsn), E41 (tnsn), E43 (tnsn) TIEBACKS E51 (tnsn) TIEBACKS E45 (tnsn), E47 (tnsn), E49 (tnsn), E53 (tnsn) Install 1st Level Tiebacks--West Wall TIEBACKS W7 (instl), W25 (instl) TIEBACKS W9 (instl), W11 (instl) TIEBACKS W13(instl), W21(instl) TIEBACKS W15(instl), W23 (instl) TIEBACKS W17 (instl), W27(instl), W29(instl) TIEBACKS W19 (instl), TIEBACKS W31 (instl) TIEBACKS W33 (instl) TIEBACKS W37 (instl), W39 (instl), W41 (instl) TIEBACKS W43 (instl), W45 (instl), W47 (instl) TIEBACKS W35 (instl) TIEBACKS W49 (instl) TIEBACKS W51 (instl), W53 (instl) Regrout 1st Level Tiebacks--West Wall TIEBACKS W9 (rgrt),W11 (rgrt) TIEBACKS W15 (rgrt), W21 (rgrt), W23 (rgrt), W25 (rgrt) TIEBACKS W17 (rgrt), W27 (rgrt), W29 (rgrt) TIEBACKS W7 (rgrt), W19 (rgrt) TIEBACKS W17 (rgrt), W23 (rgrt) TIEBACKS W13 (rgrt), W31 (rgrt) TIEBACKS W33 (rgrt) TIEBACKS W35 (rgrt), W37 (rgrt), W39 (rgrt), W41 (rgrt), W45 (rgrt), W47 (rgrt) TIEBACKS W43 (rgrt) TIEBACKS W49 (rgrt) TIEBACKS W51 (rgrt), W53 (rgrt) Tension 1st Level Tiebacks--West Wall TIEBACKS W9 (tnsn) TIEBACKS W11 (tnsn) TIEBACKS W15 (tnsn), W17 (tnsn), W19 (tnsn) TIEBACKS W7 (tnsn), W21(tnsn), W23 (tnsn) W25 (tnsn) TIEBACKS W27 (tnsn), W29 (tnsn) TIEBACKS W31, (tnsn), W33 (tnsn), W35 (tnsn), W39 (tnsn) TIEBACKS W37 (tnsn) TIEBACKS W41 (tnsn), W43 (tnsn), W45 (tnsn), W47 (tnsn)

Date 6/21/1999 10/5/1999 9/21/1999 9/22/1999 10/1/1999 10/4/1999 10/5/1999 10/9/1999 9/22/1999 9/23/1999 10/4/1999 10/5/1999 10/6/1999 10/11/1999 9/29/1999 10/5/1999 10/8/1999 10/9/1999 10/13/1999 6/23/1999 6/24/1999 6/25/1999 6/28/1999 1/6/2000 12/2/1999 12/3/1999 12/16/1999 1/3/2000 1/4/2000 1/11/2000 1/12/2000 1/29/2000 11/24/1999 11/29/1999 12/27/1999 12/28/1999 1/6/2000

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 106 92 93 102 105 106 110 93 94 105 106 107 112 100 106 109 110 114 2 3 4 7 199 164 165 178 196 197 204 205 222 156 161 189 190 199 Start of Construction TIEBACKS W49 (tnsn), W51 (tnsn) Install 2nd Level Tiebacks--West Wall TIEBACKS W7 (instl), W9 (instl) TIEBACKS W11 (instl) TIEBACKS W21 (instl), W23 (instl), W25 (instl), W27 (instl), W29 (instl) TIEBACKS W31 (instl), W33 (instl), W35 (instl), W37 (instl) TIEBACKS W39 (instl), W41 (instl), W43 (instl), W45 (instl) TIEBACKS W47 (instl), W49 (instl), W51 (instl) Regrout 2nd Level Tiebacks--West Wall TIEBACKS W7 (rgrt), W9 (rgrt) TIEBACKS W11 (rgrt) TIEBACKS W21 (rgrt), W23 (rgrt), W25 (rgrt), W27 (rgrt), W29 (rgrt) TIEBACKS W31 (rgrt), W33 (rgrt), W35 (rgrt), W37 (rgrt) TIEBACKS W39 (rgrt), W41 (rgrt), W43 (rgrt), W45 (rgrt) TIEBACKS W47 (rgrt), W49 (rgrt), W51 (rgrt) Tension 2nd Level Tiebacks--West Wall TIEBACKS W7 (tnsn), W9 (tnsn), W11 (tnsn) TIEBACKS W27 (tnsn) TIEBACKS W21 (tnsn), W23 (tnsn), W25 (tnsn), W29 (tnsn), W31 (tnsn), W33 (tnsn), W35 (tnsn), W37 (tnsn) TIEBACKS W39 (tnsn), W41 (tnsn), W43 (tnsn), W45 (tnsn) TIEBACKS W47 (tnsn), W49 (tnsn), W51 (tnsn) Concrete Samples CONCRETE : 2 SAMPLES OF MIX 1(3*6) CONCRETE : 3 SAMPLES OF MIX 2(3*6) CONCRETE : 2 SAMPLES OF MIX 3(4*8) CONCRETE : 2 SAMPLES OF MIX 3 (6*12) Elevator #2 Construction ELEV #2: Poured shaft walls to EL -4.0 Elevator #3 Construction ELEV #3: Began excavating pit for slab ELEV #3: Finished excavating pit for slab ELEV #3: Poured pit slab ELEV #3: Shaft walls formed and poured ELEV #3: Stripped forms from shaft walls ELEV #3: Concrete walls poured ELEV #3: Stripped forms from concrete walls ELEV #3: Poured shaft walls to EL +6.0 Escalator #1 Construction ESCL #1: Poured walls and pit slab ESCL #1: Stripped forms from walls and pit slab Escalator #3 Construction ESCL #3: Excavate to EL -5.0 ESCL #3: Excavate to EL -7.0 ESCL #3: Completed final excavation

Date 6/21/1999 1/10/2000 1/11/2000 1/11/2000 1/12/2000 1/17/2000 2/4/2000 2/10/2000 2/15/2000 2/22/2000 11/30/1999 12/1/1999 12/10/1999 12/28/1999 1/7/2000 1/17/2000 1/26/2000 2/2/2000 2/14/2000 2/22/2000 11/30/1999 12/10/1999 2/9/2000 2/24/2000 12/9/1999 1/31/2000 2/1/2000 2/11/2000 2/23/2000 12/1/1999 12/2/1999 12/15/1999 12/15/1999 1/31/2000 2/8/2000 2/11/2000 3/2/2000 3/7/2000 2/10/2000

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 203 204 204 205 210 228 234 239 246 162 163 172 190 200 210 219 226 238 246 162 172 233 248 171 224 225 235 247 163 164 177 177 224 232 235 255 260 234 Start of Construction ESCL #3: Poured pit slab ESCL #3: Stripped pit slab forms ESCL #3: Concrete walls poured ESCL #3: Stripped forms from concrete walls ESCL #3: Poured inclined slab-on-grade and landing ESCL #3: Poured 12in south bulk head wall from <H> to <K> ESCL #3: Poured 12in thick south bulk head wall from <K> to <L> ESCL #3: Poured 12in south wall ESCL #3: Poured 15in thick inclined roof slab Escalator #4 Construction ESCL #4: Began excavating pit for slab ESCL #4: Finished excavating pit for slab at EL -29.0 ESCL #4: Poured pit slab Stair #3 Construction STAIR #3: Excavate to EL -5.0 STAIR #3: Completed final excavation STAIR #3: Poured inclined slab-on-grade and landing STAIR #3: Poured north wall STAIR #3: Poured 12in roof slab STAIR #3: Poured next lift of north wall STAIR #3: Poured 15in thick inclined roof slab Stair #6 Construction STAIR #6: Poured 2ft thick slab-on-grade STAIR #6: Stripped gang forms (walls) STAIR #6: Poured up to M1 line to 1st landing STAIR #6: Poured from first landing to top of the stairs Stair #7 Construction STAIR #7: Poured 2ft thick slab-on-grade STAIR #7: Finished pour STAIR #7: Stripped roof slab forms STAIR #7: Poured up to first landing STAIR #7: Poured from first landing to top of the stairs General Slab-On-Grade Construction - Phase I ELEV #2 to ESCL #1: Poured 2ft thick slab-on-grade ELEV #2 to ESCL #1: Stripped forms for slab-on-grade STAIR #7 to ESCL #4: Poured 2ft thick slab-on-grade ESCL #4 to ELEV #3: Poured 2ft thick slab-on-grade Finished slab pour on east side from M1 to P7 (see STAIR #7) Poured NE corner of 10in thick Mezz floor slab <F> to <F.5> and <M3> to <P15> Poured 12in thick Mezz. floor slab <M1> to <P15> and <C> to <F> Poured roof deck slab from <M1> to <M4> Poured 3in protective cover over roof deck slab from <M1> to <M4> General Slab-On-Grade Construction - Phase II Poured 15in thick Phase II roof slab <J> to <H> West Exterior Wall Construction

Date 6/21/1999 12/17/1999 12/20/1999 12/21/1999 12/27/1999 12/28/1999 1/6/2000 1/14/2000 1/3/2000 1/4/2000 1/7/2000 1/8/2000 1/10/2000 1/11/2000 1/14/2000 1/18/2000 1/22/2000 1/24/2000 1/25/2000 1/29/2000 1/20/2000 1/26/2000 1/27/2000 1/28/2000 2/10/2000 2/15/2000 2/17/2000 3/18/2000 3/20/2000 2/2/2000 2/3/2000 2/4/2000 2/4/2000 2/5/2000 2/21/2000 2/22/2000 2/23/2000

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 179 182 183 189 190 199 207 196 197 200 201 203 204 207 211 215 217 218 222 213 219 220 221 234 239 241 271 273 226 227 228 228 229 245 246 247 Start of Construction Began pouring 2ft thick west wall from from P9 to P12 from EL -15.0 to EL -5.0 Finished pouring 2ft thick west wall from from P9 to P12 from EL -15.0 to EL -5.0 Poured 2ft thick west wall from from P12 to P14 from EL -15.0 to EL -5.0 Began pouring 2ft thick west wall from from P10 to M3 to EL +6.0 Finished pouring 2ft thick west wall from from P10 to M3 to EL +6.0 Poured west wall from from P14 to M3 to EL -4.0 (2nd lift) Poured 1.5ft thick west wall from from P11 to P15 East Exterior Wall Construction Poured 2ft thick east wall from from P7 to P12 (1st lift) Poured 2ft thick east wall from from P12 to P14 (1st lift) Poured 2nd lift of east wall from from P9 to P13 to EL -4.0 Stripped forms from 2nd lift of east wall from from P9 to P13 to EL -4.0 Poured 2nd lift (inclined wall) of east wall from from P7 to P9 Stripped forms from 2nd lift (inclined wall) of east wall from from P7 to P9 Poured 3rd lift of east wall from from P9 to P13 to EL +6.0 West Interior Wall Construction Poured interior west wall from P9 to M1 Poured interior west wall from M1 to P14 to EL +6.0 East Interior Wall Construction Poured interior east wall from from M1 to P15 to EL +6.0 Poured 1.5ft thick east inclined wall from from M1 to P8 Poured 3rd lift of east wall from M3 to P15 to EL +6.0 M1-Wall Wall Construction M1-Wall: Poured between <C.6> and <D> to 4ft M1-Wall: Poured between <F> and <G> to 3ft M1-Wall: Poured between <C.6> and <F> M1-Wall: Poured between <E> and <G> ComEd Vault Construction COMED VAULT: Poured 15in thick bottom slab COMED VAULT: Poured east and part of the north and south walls COMED VAULT: Finalized vault pour Demo Secant Pile Wall <P9> to <P14>: Demo top 2ft of east and west secant pile wall <P9> to <M3>: Demo top 2ft of east and west secant pile wall Phase I Backfill M1 to P7: Began backfill M1 to P7: Backfilled to EL -3.5 from <C> to <F> M1 to P7: Backfilled to EL -1.0 M1 to M4: Backfilled to grade for proposed Mezz. Floor and grade beams to EL -5.0 Behind (north) ELEV #2 and ELEV #3: Backfilled to EL -5.0 M1 to P7: Continued backfill to EL +1.0 M1 to P7: Backfilled to EL +2.0 M1 to P7: Backfilled to EL +4.0

Date 6/21/1999 2/24/2000 2/25/2000 3/8/2000 3/9/2000 3/10/2000 12/7/1999 12/9/1999 12/29/1999 12/30/1999 1/3/2000 2/25/2000 2/26/2000

TABLE A-1. CONSTRUCTION ACTIVITIES TIME HISTORY Day Activity 248 249 261 262 263 169 171 191 192 196 249 250 Start of Construction M1 to P7: Backfilled to EL +6.0 M1 to P7: Backfilled to EL +8.5 M1 to M4: Backfilled 12in lift on top of Mezz. Deck to EL +8.0 M4 to P7: Backfilled to EL +9.5 M3 to P7: Backfilled to EL +11.0 Phase II Excavation and Backfill North side adjacent to STAIR #3: Began excavating pit for slab and installing lagging North side adjacent to STAIR #3: Finished excavating pit for slab North side: Phase II excavation (see field notes for data) North side: Phase II excavation (see field notes for data) North side: Phase II excavation completed North side: Phase II excavation backfilled to EL +8.5 North side: Phase II excavation backfilled to EL +11.0

APPENDIX B: FIELD INTRUMENTATION DATA Table B-1- Survey Data


TABLE B-1. SURVEY DATA CONSTRUCTION DAYS 9 15 29 49 61 70 74 86 103 111 117 131 137 145 151 162 C1 DEF (MM) 0.305 1.219 2.438 3.048 3.048 3.048 3.048 4.572 5.791 8.230 8.230 9.449 9.754 10.973 10.973 11.887 C2 DEF (MM) 0.000 0.610 2.438 2.438 2.438 2.438 2.438 3.353 3.353 3.353 3.353 4.267 4.572 4.877 5.486 5.791 C3 DEF (MM) 1.829 2.438 4.572 4.572 4.572 4.572 4.572 5.182 6.096 7.925 7.925 9.449 10.058 10.058 11.278 11.582 C4 DEF (MM) 0.610 0.914 2.743 2.743 2.743 2.743 2.743 3.048 3.048 3.658 3.658 4.267 4.572 4.267 5.486 5.791 C5 DEF (MM) 2.438 3.048 4.572 4.572 4.572 4.572 4.572 4.877 5.486 6.706 6.706 7.620 8.230 8.230 8.230 9.449 C6 DEF (MM) 0.000 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.610 0.610 0.914 1.219

TABLE B-1. SURVEY DATA CONSTRUCTION DAYS 169 175 183 189 196 201 208 215 222 228 235 257 312 355 C1 DEF (MM) 12.497 13.106 14.021 13.716 16.764 15.850 18.288 16.154 17.069 17.069 17.069 17.069 17.069 17.069 C2 DEF (MM) 6.096 6.401 6.706 6.401 8.839 7.620 9.754 8.534 9.144 9.449 9.449 9.449 9.449 9.449 C3 DEF (MM) 11.887 12.192 13.411 13.411 14.935 13.411 15.545 15.545 15.545 15.850 15.850 15.850 15.850 15.850 NOTE: (1) C = SURVEY POINTS ON INTERIOR COLUMNS (2) W = SURVEY POINTS ON EXTERIOR WALLS (3) R = SURVEY POINTS ON PARAPET OF ROOF C4 DEF (MM) 6.096 6.096 6.706 7.010 8.230 6.706 7.925 7.925 8.534 8.534 8.534 8.534 8.534 8.534 C5 DEF (MM) 10.058 10.363 10.363 10.668 11.582 10.668 12.192 12.192 12.802 12.802 12.802 12.802 12.802 12.802 C6 DEF (MM) 1.219 1.219 1.219 1.524 1.829 1.524 1.524 1.524 1.829 1.829 1.829 1.829 1.829 1.829

TABLE B-1. SURVEY DATA CONSTRUCTION DAYS 9 15 29 49 61 70 74 86 103 111 117 131 137 145 151 162 169 175 183 189 196 201 208 215 222 228 235 257 312 355 C7 DEF (MM) 2.438 2.438 3.658 3.962 3.962 3.962 3.962 3.962 3.962 5.486 5.486 6.096 6.096 6.401 7.010 7.925 7.925 7.925 8.839 9.144 9.449 7.925 10.363 9.754 10.058 10.058 10.058 10.058 10.058 10.058 C8 DEF (MM) 0.000 0.305 2.134 1.829 1.829 1.829 1.829 1.829 1.829 2.134 2.134 2.134 2.134 2.438 3.048 3.658 3.353 3.048 3.658 3.353 4.877 3.353 5.182 3.658 4.267 4.572 4.572 4.572 4.572 4.572 0.305 0.305 0.305 0.305 -0.914 -0.914 -1.524 -0.305 -0.305 0.305 0.305 0.305 0.305 0.305 0.305 -0.305 -0.305 -0.305 0.305 -0.305 W5 DEF (MM) 0.305 W6 DEF (MM) 0.914 0.914 1.524 1.524 1.524 1.524 1.524 1.524 1.524 2.743 2.743 3.353 3.048 3.962 3.962 3.962 3.962 4.267 3.962 3.962 3.962 4.267 5.486 5.486 5.791 5.791 5.791 5.791 5.791 5.791 W7 DEF (MM) 4.877 5.791 6.401 6.401 6.401 6.401 5.182 5.182 8.839 8.839 11.278 11.278 11.582 11.582 11.582 11.582 12.192 12.497 12.802 13.411 12.192 12.192 11.887 14.021 14.021 14.021 14.021 14.021 14.326 14.326 W8 DEF (MM) 7.925 8.230 9.144 10.363 10.363 10.363 10.363 10.363 18.593 18.593 21.641 21.641 23.774 23.774 25.298 25.298 25.908 27.737 28.042 28.956 28.042 28.346 28.346 29.261 29.870 29.870 29.870 29.870 29.870 29.870

NOTE: (1) C = SURVEY POINTS ON INTERIOR COLUMNS (2) W = SURVEY POINTS ON EXTERIOR WALLS (3) R = SURVEY POINTS ON PARAPET OF ROOF

TABLE B-1. SURVEY DATA CONSTRUCTION DAYS 9 15 29 49 61 70 74 86 103 111 117 131 137 145 151 162 169 175 183 189 196 201 208 215 222 228 235 257 312 355 W9 DEF (MM) 7.925 9.144 9.754 10.973 10.973 10.973 10.973 10.973 21.031 21.031 25.298 25.298 27.737 27.432 29.261 29.566 30.480 32.614 32.918 33.833 34.747 35.052 34.747 34.747 35.662 35.662 36.271 36.576 36.576 36.576 W10 DEF (MM) 9.144 9.449 10.363 11.278 11.278 11.278 11.278 15.850 24.079 26.518 27.432 29.566 29.566 30.480 33.528 35.052 35.357 36.576 37.490 37.490 39.929 39.624 41.148 39.319 39.319 40.234 40.538 41.148 40.538 40.538 W11 DEF (MM) 6.401 7.925 9.144 9.754 9.754 9.754 9.754 14.326 20.117 24.079 24.689 27.127 27.127 28.042 30.480 32.004 32.614 33.528 35.357 34.442 36.881 37.490 39.014 36.576 36.576 37.795 37.795 38.100 37.795 37.795 W12 DEF (MM) 4.572 5.182 5.182 7.010 7.010 7.010 7.010 12.192 17.069 19.202 19.812 22.250 22.555 23.470 25.298 27.432 27.432 29.261 31.394 30.175 32.614 34.138 36.881 33.223 33.528 34.442 34.442 34.747 35.052 35.052 W13 DEF (MM) 2.743 3.353 3.353 5.486 5.486 5.486 5.486 10.668 13.716 15.240 15.240 17.374 17.069 17.983 19.507 21.031 21.031 22.555 24.689 23.470 24.689 26.822 29.261 27.737 27.737 27.737 27.737 27.737 27.737 27.737 W14 DEF (MM) 2.743 3.353 3.353 3.962 3.962 3.962 3.962 7.010 9.754 9.144 9.144 11.278 10.363 11.887 12.802 13.716 13.411 14.935 17.983 15.240 16.459 19.202 20.726 19.202 19.202 19.202 18.898 18.898 18.288 18.288

NOTE: (1) C = SURVEY POINTS ON INTERIOR COLUMNS (2) W = SURVEY POINTS ON EXTERIOR WALLS (3) R = SURVEY POINTS ON PARAPET OF ROOF

TABLE B-1. SURVEY DATA CONSTRUCTION DAYS 9 15 29 49 61 70 74 86 103 111 117 131 137 145 151 162 169 175 183 189 196 201 208 215 222 228 235 257 312 355 W15 DEF (MM) 2.438 2.743 2.743 2.743 2.743 2.743 2.743 3.048 5.486 4.267 4.572 5.486 5.486 5.486 6.706 7.010 6.706 6.096 8.839 7.010 7.620 10.058 11.278 9.144 9.144 8.230 8.534 8.230 7.620 7.620 W17 DEF (MM) 1.829 1.829 1.829 1.829 1.829 1.829 1.829 1.829 3.658 2.743 2.743 3.048 2.743 2.743 3.658 3.962 3.962 3.353 6.096 3.048 4.267 6.401 7.925 5.791 5.791 3.962 3.658 3.658 3.353 3.353 W18 DEF (MM) 0.000 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 W19 DEF (MM) 0.000 0.305 0.914 0.914 0.914 0.914 0.914 0.914 0.914 1.219 1.219 1.219 1.219 1.524 1.829 1.829 1.829 1.829 2.438 3.658 3.658 3.353 4.267 4.267 4.267 3.962 3.658 3.658 3.962 3.962 0.610 1.829 1.829 1.219 1.829 1.829 1.829 1.829 1.829 1.829 3.048 3.658 3.962 4.877 4.267 3.658 3.048 2.743 3.048 3.048 2.743 2.743 W22 DEF (MM) 0.000 R1 DEF (MM) 1.219 1.219 1.219 1.219 1.219 1.219 1.219 1.219 2.743 2.134 2.134 2.134 2.134 2.438 3.658 3.658 3.658 3.353 5.791 5.791 5.791 5.486 6.706 6.706 6.706 5.182 5.182 5.182 5.182 5.182

NOTE: (1) C = SURVEY POINTS ON INTERIOR COLUMNS (2) W = SURVEY POINTS ON EXTERIOR WALLS (3) R = SURVEY POINTS ON PARAPET OF ROOF

TABLE B-1. SURVEY DATA CONSTRUCTION DAYS 9 15 29 49 61 70 74 86 103 111 117 131 137 145 151 162 169 175 183 189 196 201 208 215 222 228 235 257 312 355 R2 DEF (MM) 4.877 5.486 5.486 6.401 6.401 6.401 6.401 6.401 11.582 13.106 13.106 13.106 13.106 14.326 16.459 17.983 18.288 19.812 21.946 21.946 21.946 21.946 24.384 24.384 24.384 24.994 24.994 24.994 24.994 24.994 R3 DEF (MM) 6.096 7.315 8.230 8.230 8.230 8.230 8.230 8.230 13.716 16.154 16.154 16.154 16.154 17.374 19.507 20.726 20.726 21.946 24.079 24.079 24.079 24.384 25.908 25.908 25.908 26.518 26.518 26.518 26.518 26.518 R4 DEF (MM) 5.791 6.706 7.620 7.620 7.620 7.620 7.620 7.620 10.668 11.887 11.887 11.887 11.887 11.887 12.497 12.497 12.802 13.716 13.716 13.716 13.716 15.240 15.850 15.850 15.850 15.850 15.850 15.850 15.850 15.850 R5 DEF (MM) 2.743 2.743 4.267 4.267 4.267 4.267 4.267 4.267 7.620 8.839 8.839 8.839 8.839 8.839 9.144 9.144 9.449 9.754 9.754 9.754 9.754 10.058 11.278 11.278 11.278 11.278 11.278 11.278 11.278 11.278 0.914 0.914 0.914 0.914 0.914 0.914 0.914 0.914 0.610 0.610 0.610 0.610 0.610 0.610 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 0.305 -0.610 -0.610 -0.610 R6 DEF (MM) 0.000 R7 DEF (MM) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

NOTE: (1) C = SURVEY POINTS ON INTERIOR COLUMNS (2) W = SURVEY POINTS ON EXTERIOR WALLS (3) R = SURVEY POINTS ON PARAPET OF ROOF

Table B-2- Pore Water Pressure


TABLE B-2. PORE WATER PRESSURE PORE PRESSURE (M HEAD) DAY -6 -6 -4 -4 -3 0 1 1 2 3 3 3 4 5 6 7 8 10 10 11 15 16 17 18 22 24 28 30 35 36 37 38 39 42 44 49 51 56 PZ 1A 6.647 6.751 6.647 6.599 6.655 6.657 6.640 6.690 6.797 7.026 6.756 5.654 6.144 5.984 3.922 3.637 3.505 2.606 3.434 3.457 4.409 4.600 4.455 4.585 4.755 4.912 5.044 5.113 5.174 5.184 5.217 5.192 5.370 5.357 5.385 5.448 5.497 5.385 PZ 1B 9.235 9.101 8.984 8.941 8.854 8.796 8.809 8.903 8.956 9.220 9.393 7.732 8.192 7.846 5.679 5.047 4.813 3.741 4.115 4.234 5.298 5.522 5.352 5.509 5.911 6.045 6.391 6.723 6.678 6.759 6.845 6.843 6.843 6.944 7.015 7.074 7.203 PORE PRESSURE (M HEAD) DAY -6 -4 -4 -3 0 1 1 3 4 8 10 10 11 15 16 17 18 24 28 30 36 37 39 49 56 60 PZ 4A 7.417 7.493 7.455 7.478 7.671 7.620 7.516 7.407 7.404 7.554 7.496 7.564 7.455 7.490 6.490 4.874 4.204 5.075 5.603 5.908 6.543 6.685 6.822 7.115 7.447 7.419 PZ 4B 10.599 10.317 10.150 10.066 10.221 10.175 10.086 9.931 9.916 9.967 10.356 9.957 9.845 9.792 9.274 7.104 5.593 7.455 7.871 8.240 8.946 9.098 9.281 PORE PRESSURE (M HEAD) DAY 56 59 60 63 71 73 75 78 81 85 87 88 91 92 94 96 99 103 110 113 115 117 121 124 131 145 149 156 164 203 221 267 354 PZ 4C 7.358 7.496 7.201 7.041 7.130 7.234 7.206 6.825 7.272 5.941 7.264 8.031 7.635 7.658 7.511 6.165 6.373 6.066 7.173 11.100 8.407 8.184 7.165 6.556 5.779 4.943 4.978 4.750 4.750 4.496 4.521 4.750 5.283 PZ 4D 10.338 10.086 10.061 9.934 10.084 10.150 10.185 9.886 10.216 9.149 9.924 10.343 9.997 9.926 9.688 8.362 8.763 8.542 9.601 13.437 10.490 10.239 10.094 9.423 8.763 7.783 7.849 7.493 7.493 7.087 7.188 7.620 7.874

TABLE B-2. PORE WATER PRESSURE PORE PRESSURE (M HEAD) DAY 59 60 63 65 66 70 71 73 75 78 PZ 1A 5.502 5.298 5.121 5.108 5.014 5.006 5.032 4.813 4.856 4.928 PZ 1B 7.287 7.130 7.153 7.102 7.173 7.168 7.148 7.082 7.021 6.977 PORE PRESSURE (M HEAD) DAY PZ 4A PZ 4B PORE PRESSURE (M HEAD) DAY PZ 4C PZ 4D

Table B-3- Strut Loads


TABLE B-3. STRUT LOADS STRUT 3 LOAD 58 59 60 61 63 64 65 66 67 68 70 71 72 73 74 78 79 80 81 84 85 86 87 88 91 92 93 94 95 99 100 102 105 106 107 108 109 112 202.348 149.724 190.111 201.790 288.287 300.628 307.773 306.779 313.729 459.878 311.689 365.496 454.721 403.961 448.700 453.071 422.518 515.556 406.046 468.147 420.290 423.258 399.278 368.074 407.858 380.175 581.268 583.866 553.858 534.652 649.546 598.351 655.048 663.147 670.402 653.469 638.809 733.049 33.194 24.561 31.186 33.102 47.291 49.316 50.488 50.325 51.465 75.439 51.130 59.957 74.593 66.266 73.606 74.323 69.311 84.573 66.609 76.796 68.945 69.432 65.498 60.380 66.906 62.365 95.352 95.779 90.856 87.705 106.553 98.155 107.455 108.784 109.974 107.196 104.792 120.251 59 60 61 63 64 65 66 67 68 70 71 72 73 74 78 79 80 81 84 85 86 87 88 91 92 93 94 95 99 100 102 105 106 107 108 109 112 113 STRUT 4 LOAD 0.000 127.447 204.428 80.337 189.571 246.923 223.115 241.912 445.440 239.542 411.302 265.724 245.656 296.199 321.444 447.023 436.776 503.517 525.680 452.969 478.142 657.744 412.000 563.998 490.323 375.821 384.771 380.200 579.978 687.107 665.942 657.536 680.473 748.005 766.559 756.954 748.128 713.163 0.000 20.907 33.535 13.179 31.098 40.506 36.600 39.684 73.071 39.295 67.471 43.590 40.298 48.589 52.730 73.331 71.650 82.598 86.234 74.306 78.435 107.898 67.585 92.519 80.434 61.650 63.119 62.369 95.141 112.714 109.242 107.864 111.626 122.704 125.748 124.172 122.724 116.989 72 72 73 74 78 79 80 81 84 85 86 87 88 91 92 93 94 95 99 100 102 105 106 107 108 109 112 113 114 115 116 119 120 121 122 123 124 126 618.933 630.133 627.599 615.248 727.399 738.666 760.168 751.068 768.088 777.582 101.531 103.368 102.952 100.926 119.324 121.172 124.699 123.207 125.999 127.556 STRUT 5 LOAD 0.000 226.995 81.434 112.081 254.859 426.958 638.558 682.292 550.073 685.495 794.467 781.946 831.078 732.119 776.747 760.062 842.580 785.112 640.226 643.209 0.000 37.237 13.359 18.386 41.808 70.039 104.750 111.924 90.235 112.450 130.326 128.272 136.332 120.098 127.419 124.682 138.218 128.791 105.024 105.513 DAY LOAD (KN) LOAD PER M (KN/M) DAY LOAD (KN) LOAD PER M (KN/M) DAY LOAD (KN) LOAD PER M (KN/M)

TABLE B-3. STRUT LOADS STRUT 3 LOAD 113 114 115 116 119 120 121 122 123 126 127 128 129 130 133 134 135 136 137 140 141 142 143 144 147 148 149 150 151 154 155 156 158 161 162 163 164 165 168 169 650.566 649.462 534.158 489.120 495.959 514.536 427.870 426.350 426.259 431.903 410.540 413.924 403.616 396.433 428.246 425.233 429.234 468.075 478.039 478.338 741.480 745.604 106.720 106.539 87.624 80.236 81.358 84.405 70.189 69.939 69.924 70.850 67.346 67.901 66.210 65.032 70.250 69.756 70.412 76.784 78.418 78.467 121.634 122.310 663.660 654.392 646.389 687.452 650.566 108.868 107.348 106.035 112.771 106.720 114 115 116 119 120 123 128 129 130 133 134 135 136 137 140 141 142 143 144 184 185 189 190 191 192 196 197 198 199 200 203 204 205 206 207 210 211 212 213 214 673.661 665.903 671.740 672.311 666.332 658.978 626.311 627.644 626.740 644.592 645.226 647.741 627.070 621.889 640.542 637.644 645.418 651.081 641.720 633.340 633.889 110.509 109.236 110.193 110.287 109.306 108.100 102.741 102.960 102.812 105.740 105.844 106.257 102.866 102.016 105.076 104.600 105.876 106.805 105.269 103.894 103.984 STRUT 4 LOAD 672.492 665.219 758.972 753.521 766.078 669.655 756.644 853.940 917.180 933.729 110.317 109.124 124.503 123.609 125.669 109.852 124.121 140.082 150.456 153.171 127 128 129 130 133 134 135 136 137 140 141 142 143 144 147 148 149 150 151 154 155 156 158 161 162 163 164 165 168 169 170 171 172 175 176 177 178 179 182 183 STRUT 5 LOAD 666.927 651.945 658.558 646.094 636.433 640.060 638.720 642.106 638.164 642.608 620.435 591.921 585.138 600.745 636.679 642.227 647.965 660.153 665.258 673.650 673.983 678.299 725.242 719.666 732.640 772.703 771.700 771.718 698.582 691.468 692.433 688.449 670.008 653.797 689.279 689.451 707.352 706.079 775.322 774.119 109.404 106.946 108.031 105.986 104.402 104.997 104.777 105.332 104.686 105.415 101.777 97.100 95.987 98.547 104.442 105.352 106.294 108.293 109.130 110.507 110.561 111.270 118.970 118.055 120.184 126.756 126.591 126.594 114.597 113.430 113.588 112.934 109.909 107.250 113.071 113.099 116.035 115.827 127.185 126.988 DAY LOAD (KN) LOAD PER M (KN/M) DAY LOAD (KN) LOAD PER M (KN/M) DAY LOAD (KN) LOAD PER M (KN/M)

TABLE B-3. STRUT LOADS STRUT 3 LOAD 170 171 172 175 176 177 178 179 182 183 184 185 189 190 191 192 196 197 198 199 200 203 204 205 206 207 210 211 212 213 214 217 218 220 221 227 235 241 248 254 761.232 758.049 746.215 675.436 552.949 554.280 587.815 583.996 610.549 606.762 605.476 625.234 663.803 670.812 586.854 518.634 497.655 488.626 494.784 531.995 535.340 503.026 508.521 519.602 516.225 519.628 473.233 453.312 449.999 447.524 438.308 588.017 620.103 626.930 587.809 687.511 598.721 623.195 672.111 815.006 124.874 124.352 122.411 110.800 90.707 90.925 96.426 95.800 100.156 99.534 99.323 102.565 108.892 110.041 96.269 85.078 81.636 80.155 81.165 87.269 87.818 82.517 83.419 85.237 84.682 85.241 77.630 74.362 73.819 73.413 71.901 96.459 101.723 102.843 96.425 112.781 98.215 102.230 110.254 133.695 217 218 221 227 235 STRUT 4 LOAD 908.195 976.358 952.174 665.581 635.171 148.982 160.164 165.706 156.196 109.183 104.195 184 185 189 190 191 192 196 197 198 199 200 203 204 205 206 207 210 211 212 213 214 217 218 220 221 227 235 STRUT 5 LOAD 791.389 824.095 816.987 785.943 754.045 721.194 691.761 689.700 654.134 663.081 650.811 661.453 669.557 667.644 671.042 658.374 692.509 696.618 694.850 687.958 690.433 672.986 686.265 696.191 674.658 772.296 739.450 129.821 135.186 134.020 128.928 123.695 118.306 113.478 113.140 107.305 108.773 106.760 108.506 109.835 109.522 110.079 108.001 113.600 114.275 113.984 112.854 113.260 110.398 112.576 114.204 110.672 126.689 121.301 DAY LOAD (KN) LOAD PER M (KN/M) DAY LOAD (KN) LOAD PER M (KN/M) DAY LOAD (KN) LOAD PER M (KN/M)

220 1010.143

Table B-4A- Tieback Anchor Loads at Inclinometer 2 Location

TABLE B-4A. TIEBACK ANCHOR LOADS AT INCLINOMETER 2 LOCATION 1ST LEVEL TIEBACK (E27) DAY 88 91 92 93 94 95 98 99 100 102 105 106 107 108 109 112 113 114 115 116 119 120 121 122 123 126 127 128 129 130 133 134 135 136 137 140 141 361.117 360.422 350.848 347.066 339.439 338.268 344.498 346.414 343.029 341.414 340.236 339.364 338.054 337.494 336.822 336.789 336.627 336.580 337.471 339.814 340.845 342.298 343.024 344.149 344.172 344.008 343.891 344.219 342.673 341.150 236.954 236.497 230.215 227.733 222.729 221.960 226.048 227.306 225.084 224.025 223.252 222.680 221.820 221.453 221.012 220.990 220.884 220.853 221.437 222.975 223.651 224.605 225.081 225.819 225.835 225.727 225.650 225.865 224.851 223.851 LOAD (KN) 341.297 344.308 347.387 348.050 348.387 348.476 LOAD PER M (KN/M) 223.948 225.924 227.944 228.379 228.601 228.659 DAY 105 106 107 108 109 112 113 114 115 116 119 120 121 122 123 126 127 128 129 130 133 134 135 136 137 140 141 142 143 144 147 148 149 150 151 154 155 2ND LEVEL TIEBACK (E27) LOAD (KN) 430.032 428.961 439.068 438.131 439.221 438.714 438.026 436.937 435.014 433.294 434.191 434.785 434.717 434.785 435.585 436.019 436.932 437.805 439.194 440.427 441.135 441.135 441.523 441.501 442.204 443.442 445.978 442.506 439.719 436.681 436.635 436.635 435.722 436.612 438.577 445.361 446.731 LOAD PER M (KN/M) 282.173 281.470 288.102 287.488 288.203 287.870 287.419 286.704 285.442 284.313 284.902 285.292 285.247 285.292 285.817 286.101 286.701 287.273 288.185 288.994 289.459 289.459 289.713 289.698 290.160 290.972 292.636 290.358 288.529 286.536 286.506 286.506 285.906 286.491 287.780 292.231 293.131

TABLE B-4A. TIEBACK ANCHOR LOADS AT INCLINOMETER 2 LOCATION 1ST LEVEL TIEBACK (E27) DAY 142 143 144 147 148 149 150 151 154 155 156 158 161 162 163 164 165 168 169 170 171 172 175 176 177 178 179 182 183 184 185 189 190 191 192 196 197 198 199 200 LOAD (KN) 336.440 332.386 329.528 327.981 326.552 325.474 327.021 328.426 334.378 335.432 336.323 336.791 336.510 336.744 337.330 337.213 337.705 328.122 327.724 327.091 326.130 325.474 326.271 326.786 324.654 319.663 318.539 311.275 314.462 312.892 311.626 315.680 321.749 324.326 325.896 327.396 328.169 326.224 324.373 323.295 LOAD PER M (KN/M) 220.761 218.101 216.225 215.211 214.273 213.566 214.580 215.503 219.408 220.100 220.684 220.992 220.807 220.961 221.345 221.268 221.591 215.303 215.042 214.626 213.996 213.566 214.088 214.427 213.028 209.753 209.015 204.249 206.340 205.309 204.479 207.139 211.121 212.812 213.842 214.826 215.334 214.058 212.843 212.136 DAY 156 158 161 162 163 164 165 168 169 170 171 172 175 176 177 178 179 182 183 184 185 189 190 191 192 196 197 198 199 200 203 204 205 206 207 210 211 212 213 214 2ND LEVEL TIEBACK (E27) LOAD (KN) 447.713 448.353 449.404 450.021 450.729 451.460 451.916 446.160 445.635 444.973 442.848 442.277 445.018 449.358 447.325 443.077 442.506 438.851 437.800 436.704 435.744 446.663 451.208 453.858 453.173 452.350 456.165 467.791 465.621 467.357 468.659 466.238 464.662 463.223 461.738 458.198 457.695 456.690 455.822 455.069 LOAD PER M (KN/M) 293.775 294.195 294.884 295.289 295.754 296.233 296.533 292.756 292.411 291.977 290.583 290.208 292.007 294.854 293.520 290.733 290.358 287.960 287.270 286.551 285.921 293.086 296.068 297.807 297.357 296.818 299.321 306.950 305.526 306.665 307.519 305.930 304.896 303.952 302.978 300.655 300.325 299.665 299.096 298.601

TABLE B-4A. TIEBACK ANCHOR LOADS AT INCLINOMETER 2 LOCATION 1ST LEVEL TIEBACK (E27) DAY 203 204 205 206 207 210 211 212 213 214 217 218 220 221 227 235 241 248 254 259 LOAD (KN) 322.522 327.255 331.027 327.770 326.693 331.918 334.378 334.565 334.753 335.034 328.755 327.677 325.966 324.232 326.786 327.864 328.380 328.122 328.872 332.808 LOAD PER M (KN/M) 211.628 214.734 217.209 215.072 214.365 217.794 219.408 219.531 219.654 219.839 215.718 215.011 213.889 212.751 214.427 215.134 215.472 215.303 215.795 218.378 DAY 217 218 220 221 227 235 241 248 254 259 2ND LEVEL TIEBACK (E27) LOAD (KN) 452.647 451.574 450.774 449.815 450.158 450.934 451.642 452.487 452.465 452.990 LOAD PER M (KN/M) 297.013 296.308 295.784 295.154 295.379 295.888 296.353 296.908 296.893 297.237

Table B-4B- Tieback Anchor Loads at Inclinometer 1 Location


TABLE B-4B. TIEBACK ANCHOR LOADS AT INCLINOMETER 1 LOCATION 1ST LEVEL TIEBACK (E35) DAY 88 91 92 93 94 95 98 99 100 102 105 106 107 108 109 112 113 114 115 116 119 120 121 122 123 126 127 128 129 130 133 134 135 136 137 140 141 142 143 333.231 336.891 330.013 326.295 319.259 318.197 317.552 317.529 314.502 312.396 311.499 310.477 309.216 308.588 308.305 308.305 308.376 308.210 311.870 312.318 314.160 315.010 315.529 316.544 316.521 316.474 316.426 316.521 315.600 314.136 310.382 305.920 218.656 221.057 216.544 214.104 209.488 208.791 208.368 208.352 206.366 204.984 204.396 203.725 202.898 202.486 202.300 202.300 202.346 202.238 204.639 204.933 206.142 206.699 207.040 207.706 207.691 207.660 207.629 207.691 207.087 206.126 203.663 200.735 LOAD (KN) 319.557 314.124 319.377 320.419 320.461 320.102 LOAD PER M (KN/M) 209.683 206.118 209.565 210.248 210.276 210.041 DAY 105 106 107 108 109 112 113 114 115 116 119 120 121 122 123 126 127 128 129 130 133 134 135 136 137 140 141 142 143 144 147 148 149 150 151 154 155 156 158 426.837 424.310 423.612 422.078 420.985 422.403 423.542 422.589 423.635 423.798 424.589 425.426 427.470 429.193 431.053 432.402 431.728 431.867 432.448 432.230 432.634 432.727 429.286 425.077 421.985 427.449 423.194 422.426 423.659 424.449 434.285 435.471 436.518 438.192 280.077 278.418 277.961 276.954 276.237 277.167 277.915 277.289 277.976 278.083 278.602 279.151 280.492 281.623 282.843 283.728 283.286 283.377 283.759 283.615 283.881 283.942 281.684 278.922 276.893 280.478 277.686 277.183 277.991 278.510 284.964 285.742 286.429 287.527 2ND LEVEL TIEBACK (E35) LOAD (KN) 401.166 LOAD PER M (KN/M) 263.232

TABLE B-4B. TIEBACK ANCHOR LOADS AT INCLINOMETER 1 LOCATION 1ST LEVEL TIEBACK (E35) DAY 144 147 148 149 150 150 151 154 155 156 158 161 162 163 164 165 168 169 170 171 172 175 176 177 178 179 182 183 184 185 189 190 191 192 196 197 198 199 200 203 LOAD (KN) 302.355 300.561 299.758 299.050 300.160 300.160 300.774 307.172 308.635 309.957 309.698 310.028 310.382 310.500 310.760 310.996 301.836 300.868 300.372 298.649 298.129 299.050 299.758 296.028 293.124 292.416 288.096 284.578 286.632 284.932 290.409 294.588 297.232 298.696 300.585 301.364 303.158 300.585 302.662 304.008 LOAD PER M (KN/M) 198.396 197.218 196.692 196.227 196.955 196.955 197.358 201.556 202.517 203.384 203.214 203.431 203.663 203.740 203.911 204.066 198.055 197.420 197.095 195.964 195.623 196.227 196.692 194.244 192.339 191.874 189.039 186.731 188.079 186.963 190.557 193.299 195.034 195.995 197.234 197.745 198.923 197.234 198.597 199.480 DAY 161 162 163 164 165 168 169 170 171 172 175 176 177 178 179 182 183 184 185 189 190 191 192 196 197 198 199 200 203 204 205 206 207 210 211 212 213 214 217 218 2ND LEVEL TIEBACK (E35) LOAD (KN) 439.192 440.238 440.680 441.331 441.633 434.588 433.867 433.006 431.123 430.170 432.634 435.564 434.611 429.077 427.914 424.589 423.566 422.333 421.566 425.147 430.960 436.494 437.797 438.913 443.029 446.540 449.377 459.957 462.584 451.400 449.446 447.191 445.400 442.401 441.889 441.401 440.889 439.857 437.215 436.378 LOAD PER M (KN/M) 288.183 288.870 289.160 289.587 289.786 285.162 284.689 284.125 282.889 282.263 283.881 285.803 285.178 281.546 280.783 278.602 277.930 277.121 276.618 278.968 282.782 286.414 287.268 288.000 290.701 293.005 294.866 301.809 303.533 296.194 294.912 293.432 292.257 290.289 289.953 289.633 289.297 288.620 286.887 286.337

TABLE B-4B. TIEBACK ANCHOR LOADS AT INCLINOMETER 1 LOCATION 1ST LEVEL TIEBACK (E35) DAY 204 205 206 207 210 211 212 213 214 217 218 220 221 227 235 241 248 254 259 LOAD (KN) 305.566 306.723 303.158 301.812 306.133 307.290 307.691 308.848 309.532 303.866 301.907 299.145 297.917 298.908 301.623 304.008 305.259 306.345 313.404 LOAD PER M (KN/M) 200.503 201.262 198.923 198.040 200.874 201.634 201.897 202.656 203.105 199.387 198.101 196.289 195.483 196.134 197.916 199.480 200.301 201.014 205.646 DAY 220 221 227 235 241 248 254 259 2ND LEVEL TIEBACK (E35) LOAD (KN) 435.402 434.355 434.960 436.099 436.471 437.355 438.122 439.215 LOAD PER M (KN/M) 285.696 285.010 285.407 286.154 286.398 286.978 287.482 288.199

Table B-5A- Inclinometer 1 Data: A-Axis Table B-5B- Inclinometer 2 Data: A-Axis Table B-5C- Inclinometer 4 Data: A-Axis Table B-5D- Inclinometer 5 Data: A-Axis Table B-5E- Inclinometer 5 Data: B-Axis

You might also like