You are on page 1of 10

Three-Dimensional Responses Observed in an Internally

Braced Excavation in Soft Clay


J. Tanner Blackburn, M.ASCE1; and Richard J. Finno, M.ASCE2
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Several three-dimensional effects were observed in the performance monitoring data collected during excavation for the Ford
Engineering Design Center in Evanston, Illinois. The elevations of the soil around the excavation varied and the excavator removed the
soil in a nonuniform excavation process, both of which contributed to the observed three-dimensional 共3D兲 effects. This paper describes
the excavation support system and subsurface conditions at the site, summarizes the construction procedures, and presents the lateral soil
movements measured by inclinometers, ground-surface movements measured by an automated total station, tilt of components of an
adjacent structure, and forces in internal braces. These responses are compared with expected responses from current design methods. The
3D nature of the excavation resulted in smaller movements on the side of the excavation, where the retained soil was lowest, an
unexpected pattern of axial strut loads and very slight damage to an exterior wall that paralleled one of the excavation walls.
DOI: 10.1061/共ASCE兲1090-0241共2007兲133:11共1364兲
CE Database subject headings: Three-dimensional analysis; Bracing; Excavation; Soft soils; Clays.

Introduction These case studies, parametric studies, and empirical relations


are valid for cases wherein the ground surface elevation is uni-
A braced excavation develops well-known three-dimensional form around all four walls. However, 3D effects for internally
共3D兲 effects that cause the induced ground deformations to be braced cuts can arise from different levels of ground surface re-
smaller near the corner of an excavation wall than near its center tained around the excavation, the presence of adjacent basements
共e.g., Bono et al. 1992; Wong and Patron 1993; Ou et al. 1993; and nonuniform excavation procedures wherein the soil is not
1996; 2000a; 2000b; Chew et al. 1997; Lee et al. 1998; Finno sequentially removed in a uniform manner. This paper summa-
et al. 2002; and Finno and Roboski 2005兲. Finno and Roboski
rizes the observed performance of the excavation for the Ford
共2005兲 proposed an empirical relation for the distribution of
Engineering Design Center 共FEDC兲 to illustrate these effects. The
ground movements parallel to an excavation wall wherein the
geometry of an excavation is related to the distribution of project was located on the Northwestern University campus in
␦x / ␦Center, where ␦x = lateral movement at any distance x along a Evanston, Illinois and consisted of a 44 m ⫻ 37 m, internally
wall and ␦Center⫽movement at the center of a wall. Furthermore, braced excavation with uneven initial elevations on adjacent
the ground deformations that occur near the center of a wall can sides. A basement of a nearby adjacent building supported on
be smaller than those associated with plane strain conditions, spread and strip footings impacted the stress conditions along that
even when the movements develop perpendicular to the wall 共Ou side of the excavation. Furthermore, space, equipment, and con-
et al. 1993; 1996; Chew et al. 1997; Lee et al. 1998; and Lin et al. tractual limitations resulted in a complex excavation sequence at
2003兲. Parametric studies by Finno et al. 共2007兲 have shown that, the FEDC site. Corners were often excavated prior to upper sup-
for excavations through clays, the plane-strain ratio 共PSR兲, de- port installation and access ramps were frequently placed on in-
fined as the maximum lateral movement behind a wall found from stalled internal bracing. The unbalanced ground levels both inside
the results of a 3D simulation normalized by that from a plane- and outside the excavation resulted in different patterns of lateral
strain simulation, depends on geometry expressed as length of ground movements.
wall at which the movement is reported divided by the excavated This paper describes the excavation support system and
depth, L / He, length to width ratio, L / B, wall system stiffness, and subsurface conditions at the site, summarizes the construction
factor of safety against basal heave.
procedures and presents the lateral soil movements measured by
1
inclinometers, ground-surface movements measured by an auto-
Geotechnical Engineer, Hayward Baker, Inc., Odenton, MD 21401. mated total station, tilt of components of an adjacent structure,
E-mail: jtblackburn@haywardbaker.com
2 and forces in internal braces. These responses are compared with
Professor, Dept. of Civil and Environmental Engineering,
Northwestern Univ., Evanston, IL 60208. E-mail: finno@ expected responses from current design methods. While the de-
northwesetern.edu formations that occurred at the site were within expected values,
Note. Discussion open until April 1, 2008. Separate discussions must the forces in the internal braces were slightly larger than those
be submitted for individual papers. To extend the closing date by one expected based on a Terzaghi and Peck 共1967兲 apparent earth
month, a written request must be filed with the ASCE Managing Editor. pressure diagram for soft clays. The 3D nature of the excavation
The manuscript for this paper was submitted for review and possible
resulted in smaller movements on the side of the excavation
publication on June 12, 2006; approved on March 16, 2007. This paper is
part of the Journal of Geotechnical and Geoenvironmental Engineer- where the retained soil was lowest, an unexpected pattern of axial
ing, Vol. 133, No. 11, November 1, 2007. ©ASCE, ISSN 1090-0241/ strut loads and very slight damage to an exterior wall that paral-
2007/11-1364–1373/$25.00. leled one of the excavation walls.

1364 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Subsurface conditions

had little immediate effect on the ground response because the


Fig. 1. Excavation plan and instrument locations
inclinometer data showed no responses to these preload activities.
A four-story concrete and masonry building, herein referred to
as the tech building, is founded on shallow footings 共+1.5-m
Site Description
ECD兲 and located within 5 m of the north wall of the excavation.
The adjacent soil elevation was lower on the north side 共+3.7-m
Excavation Geometry, Stratigraphy, ECD兲 along an alley between the tech building and the excava-
and Support System tion; the soil elevation on the east, west, and south sides of the
Fig. 1 shows a plan view of the site including initial site eleva- excavation was approximately +5.5-m, +5.5-m, and +5.0-m
tions, dimensions of the excavation, and support-system geom- ECD, respectively. Site constraints dictated that construction
etry. The FEDC excavation reached a final elevation of −3.8-m staging was restricted to the south side of the excavation. The
Evanston City Datum 共ECD兲, corresponding to a depth ranging difference in elevations between adjacent sides of the excavation
from 7.5 to 9.1 m, and was supported by sheet-pile walls and two influenced the reaction of internal bracing support systems, as
levels of internal bracing. The structural elements that comprised discussed later. Prior to excavation, belled caissons were installed
the temporary support system are given in Table 1. Each level of as a deep foundation for the five-story FEDC building. The cais-
internal bracing consisted of a pair of cross-lot pipe struts, sup- sons were drilled from the existing grade and backfilled prior to
ported vertically and horizontally by a steel frame at their mid- excavation. No ground movements were recorded in the incli-
points, and 12 diagonal braces. The T and B labels listed in Table nometers outside the retention system as a result of the caisson
1 correspond to the top and bottom levels of bracing and the installation.
numbers correspond to the strut locations shown in Fig. 1. The Fig. 2 shows the excavation support-system geometry and site
loads from the retained soil were transferred to the struts via stratigraphy. The site stratigraphy consists of 3 to 5 m of lake
wide-flange walers. To ensure a tight support system, vertical deposited sand and fill, overlying 0.9 m of desiccated clay crust.
plates were welded to the sheeting and the walers so that the Three, relatively compressible glacially deposited clay strata—
sheeting was in contact with the walers before the struts were Blodgett, Deerfield, and Park Ridge—underlie the sand and file
installed. The struts were set in place, subjected to an axial load and generally increase in strength and stiffness with depth. The
via a pair of hydraulic jacks at opposite ends of the struts, and upper part of the Blodgett stratum is desiccated. Below the com-
welded to the walers. After welding, the jacks were removed. pressible clays is a hard, gravelly clay glacial till, locally referred
Within a short time, the preload forces in the struts relaxed com- to as “hardpan,” at approximate elevation= −16.8-m ECD. Soil
pletely, presumably as the welds cooled. This behavior apparently strength and stiffness properties for the Blodgett, Deerfield, and
Park Ridge strata were determined from results of standard pen-
etration tests, cone-penetration tests, vane-shear testing, and stan-
Table 1. Internal Support System Member Elements
dard soil borings. Table 2 lists approximate engineering properties
determined through field-testing correlations. The variation in val-
Support type Label Structural elements ues listed in Table 2 are due to scatter in field-exploration data,
Strut T-1 to T-5 0.61 m 共24⬙兲 o.d., 1.27 cm 共0.5⬙兲 thickness soil-strength data compiled in Fig. 2 provides a better overview of
T-6 0.46 m 共18⬙兲 o.d., 1.27 cm 共0.5⬙兲 thickness the soil-strength distribution. Based on these data, no consistent
T-7 W14⫻ 145 variations in stratigraphy or properties were noticeable across the
B-3, 4, 6 0.61 m 共24⬙兲 o.d., 1.27 cm 共0.5⬙兲 thickness site. The water table for the FEDC site consisted of water perched
B-5 0.66 m 共26⬙兲 o.d., 1.27 cm 共0.5⬙兲 thickness
on the clay crust at an approximate elevation of 0-m ECD. No
groundwater drawdown efforts were undertaken. The small
B-7 W14⫻ 193
amount of water that entered the site was removed using a tem-
Waler Top W24⫻ 141
porary pump within the excavation. The bottom of the floor slab
Bottom W36⫻ 230
in the adjacent building was at elevation 1.5-m ECD, further re-

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007 / 1365

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Table 2. Engineering Properties of Soil Stratigraphy, Obtained through Various Field Tests and Correlation Methods
Elevation Engineering
Layer 关m ECD兴 properties Field test and correlation method
Sandy fill 5.2–4.2 ␾⬘ = 44− 48° CPT testing 共Robertson and Campanella 1989兲 SPT testing
Medium silty sand 4.2–2 ␾⬘ = 42− 44° CPT testing 共Robertson and Campanella 1989兲 SPT testing
Silty fine to medium Sand 2–0 ␾⬘ = 30− 38° CPT testing 共Robertson and Campanella 1989兲 SPT testing
Blodgett stratum, soft clay −0.9– −4.9 Su = 30− 120 kPa CPT testing 共Robertson and Campanella 1989兲
Su = 30− 60 kPa Vane shear testing
Su = 29− 63 kPa Water content correlation 共Chung and Finno 1992兲
CPT Testing 共Robertson and Campanella 1989兲
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Deerfield stratum, medium clay −4.9– −13.1 Su = 6 − 127 kPa


Su = 29− 53 kPa Vane shear testing
Su = 62− 106 kPa Water content correlation 共Chung and Finno 1992兲
Park Ridge stratum, stiff silty clay −13.1– −16.8 Su = 60− 251 kPa CPT testing 共Robertson and Campanella 1989兲
Su = 78− 150 kPa Water content correlation 共Chung and Finno 1992兲
Hardpan −16.8– 20.7 Su ⬎ 100 kPa CPT Testing 共Robertson and Campanella 1989兲

ducing the in situ ground stresses next to the north wall. See excavation sequence and placement of access ramps caused load-
Blackburn 共2005兲 for a complete description of the conditions at ing on the support-system members that was different than that
the site. envisioned during design.
Although the excavation of the FEDC site required a complex
and nonuniform excavation sequence, the excavation procedure is
Instrumentation
divided into approximate stages, listed in Table 3. The boundaries
The instrumentation locations at the FEDC site are shown in Fig.
1. Four slope inclinometers were installed prior to sheet-pile wall
installation; two inclinometers 共I-1 and I-2兲 were located in the
alley between the excavation and the tech building and inclinom-
eters I-3 and I-5 were installed on the east and west sides of the
site, respectively. Six permanent surveying prisms were installed
on the east side of the site: Three embedded in the surface soil
共P-6,7,8兲, two placed on the sheeting 共P-3,4兲, and one anchored to
the adjacent concrete steam vault 共P-5兲. Displacements of these
points were monitored remotely on a continuous basis with an
automated, radio-linked total survey station mounted on the roof
of the tech building. Two automated, remote access tiltmeter pairs
were installed on structural components of the tech building to
continuously monitor building response during the excavation.
Thirty-four vibrating wire strain gauges were installed on selected
cross-lot and northwest diagonal bracing members, with 29 sur-
viving the entire excavation process. A complete description of
the instrumentation is given by Blackburn 共2005兲.

Construction Sequence
A complex excavation procedure was followed by the excavator
due to space, equipment, and contractual limitations. Soil was
removed by a combination of large and small back-hoes and a
small front loader, which could pass beneath the support struts.
The use of back-hoes, rather than an exterior crane, required an
access ramp for the majority of the excavation duration. To facili-
tate relatively efficient removal of soil under these conditions,
corners were often excavated prior to upper support installation.
Fig. 3 presents photographs that illustrate the nonuniform exca-
vation sequence and placement of access ramps on top of in-
stalled support members.
The actual excavation sequence consisted of the excavation of
the corners and the placement of the soil in the center of the site,
prior to removal. This excavation sequence differed from the “as-
designed” excavation sequence of uniformly excavating the site
to an elevation just below the first layer of supports, followed by
the installation of all supports at that elevation and repeating the
process until final excavation grade was reached. The irregular Fig. 3. Excavation photographs

1366 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Table 3. Excavation Stage Information
Dates 共all 2004 Construction
Stage Activity unless noted兲 days
1 Sheet-pile wall installation 1/8–2/3 38–64
2 Excavation to +0.9-m ECD 2 / 2 – 3 / 16 63–106
2a Upper support installation 共+1.5-m ECD兲 2 / 17– 3 / 20 78–110
3 Excavation to −1.5-m ECD 3/2–4/7 92–128
3a Lower support installation 共−1.0-m ECD兲 3 / 10– 4 / 8 100–129
4 Excavation to −3.8-m ECD 4 / 8 – 5 / 12 129–163
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

5 Permanent structure construction 4 / 27– Fall共2005兲 148–NA


Backfill between permanent and temporary walls 7 / 21– 9 / 20 233–294
Lower support removal 8 / 2 – 8 / 12 245–255
6 Upper support removal 8 / 23– 9 / 3 266–278

for these stages span several weeks of construction to reflect the the direction of installing the sheeting at each location. For ex-
fact that the site was excavated in several corners prior to install- ample, the sheeting along the north wall was installed from west
ing internal bracing in a conventional sequence. to east resulting in a slight eastward component of wall move-
ment. After wall installation, the soil moved toward the excava-
tion. The vectors shown in Fig. 5 largely reflect movements
Soil Responses to Excavation toward the area that was excavated first. Given the erratic nature
of the excavation sequence, these movements are neither consis-
Soil deformations during construction were monitored with four tent nor perpendicular to the excavation. This pattern of lateral
slope inclinometers and an automated total survey system. Both movements toward the first excavated area was reported by Finno
cantilever and deep-seated soil deformation are evident in the et al. 共1989兲.
lateral soil profiles. Settlement and lateral movements were ob-
tained from the survey instrumentation.

Lateral Soil Deformations Observed in Inclinometers


Lateral soil displacements were monitored by weekly readings of
four slope inclinometers, the locations of which are shown in Fig.
1. Soil displacement perpendicular to the retaining wall can be
related to sheet-pile wall installation, causing movement away
from the excavation, and site excavation, causing soil movement
toward the excavation. Because the inclinometers were installed
prior to sheeting installation, both types of soil responses were
observed.
Fig. 4 shows the lateral soil movements caused by installing
the sheet-pile walls. In the Blodgett and Deerfield clay strata,
these movements varied from 6 to 10 mm away from the wall
after installation. The equivalent thickness of the ZX85 sheeting
is approximately 14 mm 共defined as cross-sectional area per unit
length兲. When installing sheeting through a saturated clay, one
can expect that about half of this area will be displaced away
from the wall 共Finno et al. 1988兲. The observed wall deformation
of 6 – 10 mm reflects the expected deformation of 7 mm. Zero
lateral soil displacement occurred in the sand fill at distances from
3 to 5 m from the wall, thus restricting the deformations in the
immediately underlying clay. This response is reasonable when
one considers that the sand will density locally from vibrations as
the wall is installed. The inclinometer deformation profiles reflect
this restriction, the lateral deformation decreases to zero in the
upper portions of the Blodgett stratum. Similar behavior was ob-
served in Finno et al. 共1988兲.
As the FEDC excavation progressed, adjacent soil incremen-
tally moved toward the excavation. Fig. 5 shows a vector repre-
sentation of incremental soil displacements in the soft clays at
elevation= −4.9-m ECD during construction. The increments cor-
respond to the approximate stages defined in Table 3. As the
sheet-piles were installed, the soil displaced away from the walls,
with a component of movement parallel to the wall that reflected Fig. 4. Lateral movements caused by installing sheet-pile wall

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007 / 1367

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Lateral soil movements in soft clay, observed at elevation


−4.9-m ECD

Furthermore, casing rotation during inclinometer installation


could also contribute to the deviations from perpendicular. Stages
2 and 3 in Fig. 5 correspond to the observed deformation after
installation of the top and bottom bracing, listed as 2a and 3a in
Table 3, respectively. The prestressing of the struts did not affect
the soil movements toward the excavation as the ground moved
toward the excavation throughout this period. In general, the net
movement into the excavation was reduced by the amount of
lateral movement away from the excavation as a result of install-
ing the sheet-pile wall.
Figs. 6 and 7 show the inclinometer soil-deformation profiles Fig. 6. Lateral soil movements observed at I-1 and I-5, reset after
wall installation
for inclinometer pairs 1 and 5 in the northwest corner of the
excavation and pairs 2 and 3 in the northeast corner, respectively
共Fig. 1兲. These displacements were reset to 0 after the wall was
installed to illustrate soil deformations due to excavation. Both lateral deflections, especially along the north and west walls. The
figures show the same pattern of movements. Lateral movements Clough et al. 共1989兲 method yielded somewhat larger, although
extend below the bottom of the excavation and terminate in either reasonable, values. The former approach could be expected to
the stiff Park Ridge or Hardpan strata. The maximum movements yield better results because it was based on observed performance
occur below the bottom of the final excavated grade, and are of a flexible wall system in Chicago 共Finno and Roboski 2005兲.
equal to about 14 mm at inclinometers 1, 2, and 5, and 24 mm at The larger deep-seated movements on the east side of the site
inclinometer 3. These lateral movements correspond to normal- likely arose from the presence of softer clays in the Blodgett
ized 共by the excavated depth, H兲 lateral movements of 0.19% for stratum on the east side of the site, as noted by higher water
the north, 0.14% for the west, and 0.26% for the east sides of the contents—typically 40% versus 25%—compared to those found
excavation adjacent to the inclinometers. on the north and west sides of the site. In any case, the perfor-
These normalized values can be compared to those computed mance in terms of deep-seated movements was generally as ex-
by empirical methods developed by Clough et al. 共1989兲 and pected for the conditions encountered at the site.
Finno and Roboski 共2005兲. The system stiffness parameter, S, However, the movements in the upper 5 m or so were some-
defined by Clough et al. 共1989兲 is what atypical, particularly when one looks at inclinometer Pairs 1
and 5 and 2 and 3. The north wall inclinometers 共1 and 2兲 expe-
S = 共EI兲/共␥waterh4ave兲 共1兲
rienced small cantilever movements—approximately 5 mm when
where EI = elastic modulus and moment of inertia of the wall, the excavation reached final grade in Stage 4. These small move-
have = average spacing between support members, and ␥water = unit ments are representative of response of a well-constructed support
weight of water. The system stiffness for the FEDC excavation system. However, large cantilever movements of 50 and 35 mm
was approximately 100. The factor of safety against basal heave, were observed in the east 共I-3兲 and west inclinometers 共I-5兲, re-
defined by Terzaghi and Peck 共1967兲 and computed using design spectively. Furthermore, the deflected shapes of these two incli-
shear strengths, ranged from 2.1 to 2.6 for the various excavation nometers indicated that the wall rotated about the second support
depths. An average value of 2.25 was used for this comparison. level, rather than the first level, as is typically observed. These
The design chart presented by Finno and Roboski 共2005兲 was cantilever displacements likely arise from the 1.8-m-higher eleva-
created for flexible support systems, like that at the FEDC. The tion of the ground surface on the east and west sides of the exca-
normalized lateral deformations predicted by the Clough et al. vation, as compared to the north side 共see Fig. 1兲 and the
共1989兲 method and the Finno and Roboski 共2005兲 methods were basement excavation for the building adjacent to the north side of
0.33% and 0.16%, respectively. The value computed using Finno the excavation, both of which make the stresses in the soil re-
and Roboski’s 共2005兲 method agree well with the FEDC observed tained behind the east and west walls larger than those behind the

1368 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Lateral soil movements observed at I-2 and I-3, reset after
wall installation
Fig. 8. Surface settlement and deep-seated lateral displacement on
east side 共observed at elevation −4.9-m, ECD兲
north wall. The cantilever movement exhibited in I-3 and I-5
implies that the diagonal bracing transferred the loads from the
higher east and west sides to the lower north side, and thus mini- erage was employed for tracking the survey point position. The
mized cantilever displacements along the north side, as indicated standard deviation of the 10-point moving average data between
in data from I-1 and I-2. These elevation and stress differences the two stationary points was 0.8 mm 共Finno and Blackburn
contributed to the smaller ground deformations in the upper 5 m 2006兲.
of soil on the side of the excavation where excessive ground Settlement data collected for the soil and steam vault survey
movements could have damaged the adjacent building. points are given in Fig. 8, along with the construction history of
The presence on the east side of the utility “pop-out” con- the southeast corner and deep-seated lateral deformation of the
structed directly adjacent to INC-3 further contributed to the dif- east inclinometer 共I-3兲. The deep-seated lateral deformation given
ference in the responses at the east and north sides of the cut. This in Fig. 8 is the displacement occurring at an elevation⫽−4.9-m
pop-out was supported by a ring of walers connected to the upper- ECD. The steam vault settled approximately 80 mm during the
level excavation waler. This ring support would increase the excavation. Soil points P-6, P-7, and P-8 settled approximately
inward load on the first level of internal excavation bracing, pos- 50, 75, and 45 mm, respectively, during the excavation process.
sibly increasing lateral deformation in the northeast corner. Un- The automated, real-time data availability of the survey system
fortunately, no strain gauges were placed on the diagonal bracing allowed for the detection of rapid settlement periods. Large settle-
in the northeast corner. ment increments occurred as the southeast corner was excavated
from elevations⫽+1-m to − 1.5-m ECD and from −1.5-m to
− 3.8-m ECD. The soil deformations at the ground surface corre-
Observed Ground-Surface Movements
spond to periods of deep-seated movement observed by I-3. The
An automated, remote-access, optical survey system was devel- magnitudes of surface settlements were greater than the horizon-
oped and employed to measure surface soil deformation during tal deformation on the ground surface and in the clay layers,
the FEDC excavation process 共Blackburn 2005兲. The survey sys- suggesting that the granular fill and sands compressed either as a
tem monitored vertical settlement and horizontal displacement of result of shearing caused by cutting the temporary slope, con-
six prisms located adjacent to the southeast corner of the excava- struction vibrations, or a combination of the two. This observation
tion. Two additional prisms were mounted to permanent struc- emphasizes that the Clough et al. 共1989兲 and other semiempirical
tures over 200 feet away from the excavation, which were used as methods, based in part on finite-element analyses made with rela-
redundant stationary benchmarks. To minimize the scatter in the tively simple-soil models, are best suited for estimating lateral
surface-deformation monitoring analysis, a 10-point moving av- deformations. Surface settlements must be evaluated taking into

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007 / 1369

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. Location of exterior cracks in adjacent building

the footing and the column, as one would expect for an isolated
column on a spread footing as compared to a thick wall on a strip
footing. After day 140 when the excavation had reached its final
Fig. 9. Perpendicular tilting of tech building support members and depth, the column tilt of less than 1 / 2000 was approximately
slope of I-1 at respective depths equal to the slope of the footing. At these levels of column tilting
and footing deformations, no structural or cosmetic damage to the
tech building was detected in this direction during the excavation.
consideration possible consolidation effects, shear-induced vol- The only cracking that was observed during construction was
ume changes in granular soils or fill, vibration-induced settle- cosmetic in the external stone and mortar facade in smaller sec-
ments, and small strain nonlinearity of the soil affected by the tions of the external bearing wall, parallel to the north excavation
excavation. wall, as indicated in Fig. 10. The x-axis of Fig. 10 corresponds to
the eastward distance from the northwest corner of the Ford ex-
cavation. The cracks were diagonal shear and vertical tensile
Structural Response to Movements along
cracks corresponding to “very slight” damage, according to Bur-
the North Wall
land and Wroth 共1975兲. No cracking was observed on the interior
If the deformation in the underlying clay layers is assumed to be of the tech building.
undrained, then the vertical settlement distribution at a footing Roboski and Finno 共2006兲 proposed an empirically based
elevation can be approximated by rotating the lateral soil defor- method for determining the settlement distribution along an exca-
mation distribution about a pivot point with depth corresponding vation when given the maximum lateral or vertical soil deforma-
to the elevation of the footing 共Finno et al. 2002兲. These estimates tion. As shown in Fig. 10, the cracking occurred at locations
of ground deformations can be compared with the tilting of the along the excavation where the largest distortions 共differential
tech building as directly observed with tilt meters. settlement per unit length兲 are predicted using the results of the
Basement column and exterior wall tilt perpendicular to the Roboski and Finno method with the maximum observed horizon-
excavation is compared to the soil slope 共estimated from I-1 data兲 tal displacement. The diagonal shear cracking occurred at the in-
in Fig. 9. Observed tilt is converted to footing slope by assuming flection point of the computed settlement distribution, which is
the footing and column remain contiguous and rotate as a unit. where shear strains would be expected. If it is assumed that the
Thus, the column tilt would be equal to the inverse tangent of the tech building did not undergo any rigid rotation, the angular dis-
slope. Also shown is the distortion between the exterior wall and tortion of the section that exhibited cracking would be approxi-
the interior column, computed from rotated I-1 data as the differ- mately 0.09% or 1 / 1100. This is slightly less than the critical
ential settlement between these two points divided by the corre- shear strain 共␥crit兲 for masonry structures 共0.11% or 1 / 900兲, given
sponding length. by Burland and Wroth 共1975兲; however, the critical shear strain
Initially, the clay soil heaves away from the excavation as a for the external wall of the tech building possibly decreased due
result of installing the sheeting around day 50, resulting in heave to age and weathering 共Blackburn 2005兲. The vertical crack in the
of the structure. This movement is more pronounced in the clay, external wall occurred at the transition point between the flat and
so the maximum heave occurs at some distance from the exterior sloped settlement distribution, which also coincided with a
wall at the south end of the building, resulting in more heave at change in footing type 共from strip to square兲 and footing eleva-
the column. The distortion to the south, implying that the column tion. The foundation discontinuity and transition in settlement dis-
heaved more than the wall, reflects this response, as does the tribution could indicate that rotation between the two segments of
tiltmeter responses of both the column and wall. As the excava- the tech building caused the vertical crack in this location 共Black-
tion progresses and the soil moves toward the excavation, both burn 2005兲. In any case, the only observed damage was slight
the column and wall footing settle. The wall tilts toward the north cracking in exterior masonry walls that ran parallel to the exca-
at approximately the same rate as the soil slope from the incli- vation along the north wall, and was located in areas where maxi-
nometer data, suggesting an almost rigid connection between the mum distortions in that direction are to be expected.
thick masonry wall and underlying strip footing. The magnitude
of the tilt observed at his location approached a slope of 0.0015,
or 1 / 667. Tilt of the interior column does not track the slope of Internal-Bracing Responses
the inclinometer data at the corresponding depth between days 62
and 140, but rather more closely mimicked the distortion. This Vibrating wire-strain-gauge pairs were mounted at the quarter
pattern of movement suggests a more flexible connection between points of pipe bracing and at the neutral axis of wide-flange sec-

1370 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Comparison of observed upper- and lower-level support


Fig. 12. Comparison of calculation upper- and lower-level support
loading during excavation
earth loading during excavation

tions to permit determination of axial stresses in these members.


Fig. 1 shows the labels and locations of the strain gauge. The
axial strut load in a pipe was calculated from the average of the port, whereas the cross-lot braces 共1, 3, and 4兲 exhibited larger
four strain-gauge readings when all four gauges were operational. axial loads in the lower level when the excavation reached final
The strain-gauge pairs on the pipes also provided an opportunity grade. The axial loads in the upper supports remained constant
to separately identify contributions to the observed responses after the lower supports were installed, rather than decreasing as
from earth pressure, axial thermal loading, thermal bending, and is commonly observed. These responses likely were caused by the
construction-induced bending. The data do not include bending nonuniform excavation sequence. The axial load increase in the
stresses that arise from the self-weight of the member, because upper levels 共T gauges兲 after the lower level supports 共B gauges兲
the gauges were installed after the supports were placed on chan- are removed illustrates that the backfill cannot be assumed to
nel supports, but before they were structurally connected to the absorb all soil loads during bracing removal. In fact, the maxi-
walers. The strain-gauge sensors contained thermal-sensor com- mum load observed in supports T-3, T-4, and T-7 occurred during
ponents, which provided strut-temperature data for each data this stage of construction. The strain gauges installed on T-2 were
point. Support member temperature data were used to separate the destroyed early in the excavation process; therefore, the data from
thermal and earth components of the support-system response. T-2 are not presented in Fig. 11. Also, strain gauges were not
Empirical thermal correction factors were determined by modify- installed on the southern lower cross-lot bracing, corresponding to
ing a procedure proposed by Boone and Crawford 共2000兲, locations B-1 and B-2. B-1 and B-2 were not installed because the
wherein incremental changes in observed axial strut load are cor- ramp construction would have quickly damaged the gauges in that
related with incremental changes in temperature to establish a location.
linear thermal correction factor. The Boone and Crawford 共2000兲 Fig. 12 presents the calculated lateral earth pressure compo-
method was modified by Blackburn 共2005兲 to account for the nent of the observed axial loads. The upper and lower support
welded connection of the support struts, which require thermal member earth loads approached equal magnitudes as the excava-
corrections for both expansion and contraction. The correction tion reached final depth. Examination of the data on Figs. 11 and
factors for expansion and contraction differ because the restraint 12 suggests that the thermal loading component of the strut force
conditions are different. Thermal expansion is resisted by the was as much as 40% of the total observed load 共Blackburn et al.
waler, wall, and earth pressure, but thermal contraction is resisted 2005兲. The decrease in calculated earth loads observed in T-4 and
by the waler connection to the wall and adjacent internal braces. T-5 were due to the large thermal load correction for these two
Fig. 11 presents the observed and axial loads in the support support members. Both members were installed in subfreezing
members during the excavation. The locations of the support conditions, which served as the datum for thermal correction. As
members are labeled in Fig. 1. Axial loads in the diagonal braces temperatures rose throughout the project, the internal thermal
共Braces 5, 6, and 7兲 generally were similar in both levels of sup- loading exceeded 400 MPa. While the observed forces remained

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007 / 1371

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


caused by axial loads were about equal to those caused by bend-
ing. Temperature induced axial loads and bending stresses were
significant and were responsible for about one-half of each com-
ponent. With the exception of the ramp loading, all other causes
resulted in stress levels of about 80 MPa, well below the yield
stress of 250 MPa. The contractor used in-stock structural ele-
ments as the bracing, and did not attempt to optimize their size
based on a TP loading diagram. This was fortunate given the
unanticipated loading on several of the cross-lot braces by the
excavator’s temporary ramp.
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Conclusions

Based on results of the field observations and analysis presented


herein, the following conclusions can be drawn related to the
performance of the FEDC excavation.
While the deep-seated movements varied between 0.14 and
0.26% of the excavated depth and thus were typical of those
expected for the conditions at the site, the movements in the
upper 5 m reflect the influence of the lower ground-surface eleva-
tion on the north side of the support system and the effects of the
adjacent basement. Cantilever movements of 50 and 35 mm were
observed at the higher sides of the excavation in the east 共I-3兲 and
west inclinometers 共I-5兲, respectively. These walls rotated about
the second support level, rather than the first level, as is typically
observed. These cantilever movements imply that the diagonal
bracing transfers the loads from the higher east and west sides to
the lower north side and thus minimizing cantilever displace-
ments along the north side. This load-transfer mechanism was
advantageous for this project because ground deformation was
minimized at the location of the adjacent building on shallow
Fig. 13. Extreme fiber stress components in internal bracing 共adapted foundations.
from Blackburn et al. 2005兲 Axial loads in the upper level of bracing changed little when
the lower level bracing was installed, in contrast to the typical
response where the load drops significantly when a lower brace
constant during the excavation, the corrected earth loading de- level is installed. These responses likely were caused by the non-
creased due to this thermal increase. uniform excavation sequence. The maximum axial loads observed
Figs. 11 and 12 also compare observed total strut loads and in the support members generally matched loads predicted with
corrected earth loads with loads expected when one employs Terzaghi and Peck 共1967兲 apparent pressure diagrams. Empirical
Terzaghi and Peck’s 共1967兲 apparent earth-pressure method for analysis determined that internal thermal stresses in the support
the FEDC site conditions. The Terzaghi and Peck 共TP兲 approach members composed up to 40% of the total support load. This
generally overestimated the strut loads in the top-level supports, stress component must be taken into consideration when numeri-
but showed a reasonable agreement for the bottom-level supports, cal analyses are employed for designing internal bracing
when compared to the total observed axial loads. A comparison structures.
between the expected TP loads and the corrected earth loads il- According to the classification of Burland and Wroth 共1975兲,
lustrates that the earth loads are significantly less than the TP the support system and excavation sequence at the FEDC resulted
loads and the total observed loads. The corrected earth loading is in negligible damage to the adjacent building supported on shal-
what one would obtain from numerical methods that neglect ther- low foundations. Slight cracking did occur in the exterior wall of
mal loading, whereas the TP method considers axial loads that the tech building, which is parallel to the north excavation wall,
arise from all loading mechanisms during an excavation, because where the stiffening effects of the corner of the excavation re-
the TP loading diagrams are envelopes of maximum pressures sulted in the largest ground-surface distortions.
derived from strut load measurements made under varied environ-
mental conditions in the field. This implies that if numerical
methods are employed to determine expected bracing loads, ad- Acknowledgments
ditional thermal loading must be properly included in the final
design load. Otherwise, support members will be structurally in- Financial support for this work was provided by National Science
adequate to support all loading components. Foundation Grant No. CMS-0219123 and the Infrastructure Tech-
A summary of the components of the maximum extreme fiber nology Institute 共ITI兲 of Northwestern University. The support of
stresses in each strut is given in Fig. 13 共Blackburn et al. 2005兲. Dr. Richard Fragaszy, program director at NSF, and Mr. David
The data show that the most severe loading condition arose from Schulz, ITI’s director, is greatly appreciated. Dr. Paul Mayne of
the unanticipated 共in-design兲 ramp construction. Without that Georgia Institute of Technology conducted the cone-penetration
loading, fairly consistent trends were observed. The stresses tests. Special thanks to Phil Blakeman of Turner Construction, Dr.

1372 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373


Michael Wysockey of Thatcher Engineering, Jay Baehr of North- supported excavations.” Proc., 13th Ann. Great Lakes Geotechnical
western University Facilities Management, and Robert Acker of and Geoenvironmental Engineering Conf., H. H. Titi, ed., ASCE,
STS Consultants for granting site access and providing data. Ad- Reston, Va.
ditional assistance was provided by current and past Northwestern Finno, R. J., Blackburn, J. T., and Roboski, J. F. 共2007兲. “Three-
University graduate and undergraduate students: Jill Roboski, dimensional effects for supported excavations in clay.” J. Geotech.
Terence Holman, Cecelia Rechea, Wan-Jei Cho, and Kate Geoenviron. Eng., 133共1兲, 30–36.
Sylvester. Finno, R. J., Bryson, L. S., and Calvello, M. 共2002兲. “Performance of a
stiff support system in soft clay.” J. Geotech. Geoenviron. Eng.,
128共8兲, 660–671.
Finno, R. J., Nerby, S. M., and Atmatzidis, D. K. 共1988兲. “Ground re-
References
sponse to sheet pile installation in clay.” Proc., 2nd Int. Conf. on Case
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 11/05/21. Copyright ASCE. For personal use only; all rights reserved.

Histories in Geotech. Engineering, St. Louis, Mo., 1297–1301.


Blackburn, J. T. 共2005兲. “Automated remote sensing and three-
Finno, R. J., and Roboski, J. F. 共2005兲. “Three-dimensional responses of
dimensional analysis of internally braced excavations.” Ph.D., Dept.
a tied-back excavation through clay.” J. Geotech. Geoenviron. Eng.,
of Civil Engineering, Northwestern University, Evanston, Ill.
131共3兲, 273–282.
Blackburn, J. T., Sylvester, K., and Finno, R. J. 共2005兲. “Observed brac-
Lee, F. H., Yong, K. Y., Quan, K. C., and Chee, K. T. 共1998兲. “Effect of
ing responses at the Ford Design Center excavation.” Proc., 16th Int.
corners in strutted excavations: Field monitoring and case histories.”
Conf. on Soil Mechanics and Geotechnical Engineering, ISSMGE,
J. Geotech. Engrg., 124共4兲, 339–349.
Osaka, Japan.
Lin, D. G., Chung, T. C., and Phien-wej, N. 共2003兲. “Quantitative evalu-
Bono, N. A., Liu, T. K., and Soydemir, C. 共1992兲. “Performance of an
ation of corner effect on deformation behavior of multistrutted deep
internally braced slurry-diaphragm wall for excavation support.”
excavation in Bangkok subsoil.” Geotech. Eng., 34共1兲, 41–57.
Slurry Walls: Design, Construction, and Quality Control, ASTM Rep.
Ou, C. Y., Chiou, D. C., and Wu, T. S. 共1996兲. “Three-dimensional finite
No. STP 1129, D. B. Paul, R. G. Davidson, and N. J. Cavalli, eds.,
element analysis of deep excavations.” J. Geotech. Engrg., 122共5兲,
ASTM, Philadelphia.
Boone, S. J., and Crawford, A. M. 共2000兲. “Temperature, elastic modulus, 337–345.
and strut load relationships for braced excavations.” J. Geotech. Ou, C. Y., Hsieh, P. G., and Chiou, D. C. 共1993兲. “Characteristics of
Geoenviron. Eng., 126共10兲, 870–881. ground surface settlement during excavation.” Can. Geotech. J., 30,
Burland, J. B., and Wroth, C. P. 共1975兲. “Settlement of buildings and 758–767.
associated damage.” Proc. Conf. on Settlement of Structures, Pentech Ou, C. Y., Liao, J. T., and Cheng, W. L. 共2000b兲. “Building response and
ground movements induced by a deep excavation.” Geotechnique,
Press, Cambridge, 611–654.
50共3兲, 209–220.
Chew, S. H., Yong, K. Y., and Lim, A. Y. K. 共1997兲. “Three-dimensional
Ou, C. Y., Shiau, B. Y., and Wang, I. W. 共2000a兲. “Three-dimensional
finite element analysis of a strutted excavation.” Proc. 9th Int. Conf.
deformation behavior of the Taipei National Enterprise Center
on Computer Methods and Advances in Geomechanics, Yuan J. X.,
共TNEC兲 excavation case history.” Can. Geotech. J., 37, 438–448.
ed., Balkema, Rotterdam, Wuhan, China; Balkema, Rotterdam.
Robertson, P. K., and Campanella, R. G. 共1989兲. Guidelines for geotech-
Chung, C. K., and Finno, R. J. 共1992兲. “Influence of depositional pro-
nical design using the cone penetrometer test and CPT with pore
cesses on the geotechnical parameters of Chicago glacial clays.” Eng.
pressure measurement, Hogentogler & Company, Inc., Columbia, Md.
Geol. (Amsterdam), 32, 225–242.
Clough, G. W., Smith, E. M., and Sweeney, B. P. 共1989兲. “Movement Roboski, J. F., and Finno, R. J. 共2006兲. “Distributions of ground move-
ments parallel to deep excavations.” Can. Geotech. J., 43共1兲, 43–58.
control of excavation support systems by iterative design.” Proc.,
Terzaghi, K., and Peck, R. B. 共1967兲. Soil mechanics in engineering
Current Principles and Practices, Foundation Engineering Congress,
Vol. 2, ASCE, New York, 869–884. practice, 2nd Ed., Wiley, New York.
Finno, R. J., Atmatzidis, D. K., and Perkins, S. B. 共1989兲. “Observed Wong, L. W., and Patron, B. C. 共1993兲. “Settlements induced by deep
performance of a deep excavation in clay.” J. Geotech. Engrg., excavations in Taipei.” Proc., 11th Southeast Asian Geotechnical
115共8兲, 1045–1064. Conf., S. L. Lee, K. Y. Yong, and Y. K. Chow, eds., SEAGC, Sin-
Finno, R. J., and Blackburn, J. T. 共2006兲. “Automated monitoring of gapore, 787–791.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2007 / 1373

J. Geotech. Geoenviron. Eng., 2007, 133(11): 1364-1373

You might also like