Open access
Case Studies
Dec 23, 2017

Deep Excavation of the Gate of the Orient in Suzhou Stiff Clay: Composite Earth-Retaining Systems and Dewatering Plans

Publication: Journal of Geotechnical and Geoenvironmental Engineering
Volume 144, Issue 3

Abstract

From 2005 to 2010, a large excavation (approximately 26,000  m2 in plane) was conducted in stiff clay deposits for the Gate of the Orient building in Suzhou, China. To reduce project cost and shorten construction duration, composite earth-retaining systems were designed for this excavation, i.e., removal of the uppermost 7.85 m of soil was conducted mostly by the sloped open-cut method, leaving one side supported by soil nailing wall (SNW); subsequent excavation to a depth of 21.5 m was retained by multipropped continuous bored pile (CBP) wall; final excavation of two inner pits to 30.0–30.7 m deep was supported by a SNW along with jet-grouting of basal soils. In spite of this, lateral wall displacements in this case were comparable to those of multipropped excavations in Suzhou. Because of a high phreatic level and confined artesian water, composite dewatering plans were adopted. Because the waterproof curtain did not extend deeply into the underlying confined aquifer, for the sake of saving cost, discharging artesian water inside the pit incurred a dramatic drawdown in the artesian level outside the pit, accompanied by significant ground settlements. It turned out that the methods in the literature for predicting excavation-induced ground settlements were not applicable to this case any more. For a large excavation not following zoned-construction procedure, corner-stiffening behavior was significant. In this case, deformations near the pit middle span were up to 2.5 times those near corners. Beyond recognition, excavation in stiff clay incurred noticeable basal rebound as well, which was up to one-fourth that in soft clays. Generally, excavations in stiff clay caused smaller lateral wall displacement than excavations in soft clays; their maximum lateral wall deflections mostly occurred above the excavation surface instead of equally above and below the excavation surface in soft clays.

Introduction

From May 2005 to January 2010, a large basement was excavated for a 301.8-m-high landmark skyscraper, the Gate of the Orient, in Suzhou, China. This megastructure would consist of two 68-story main towers, a 7-story podium, and a 5-story basement. The main towers would comprise reinforced concrete core tubes to be connected with perimeter mega columns via steel outriggers; the podium would be typical frame structures. The two main towers would converge at the height of 238 m to form a 301.8-m-high superstructure in an arch shape. Its basement would have dimensions of approximately 214  mlong×133  mwide×21.5  m to 30.7 m deep (see Fig. 1). The site was bounded by several utility pipelines buried at depths of 0.5–1.0 m below ground surface (BGS) to the east, an artificial river (Xiangmen) to the south, and green lands to the north and west.
Fig. 1. Plan layout of the pit along with ground improvement and adjacent environment
The site was located on Yangtze River Delta, approximately 110 km west of Shanghai. Prior to construction, field exploration tests, including 21 standard penetration tests (SPTs) and 41 cone penetration tests (CPTs), had been carried out across the site to characterize its subsurface condition. As shown in Fig. 2, the ground comprised 2-m-thick fill (Layer I), overlying 22-m-thick firm to stiff clayey and silty soils (Layer II). Beneath was stiff to hard silty clay and clayey silt (Layer III) to a depth of 38 m BGS, overlying medium dense to very dense silty sand (Layer IV) to 44 m BGS. Below Layer IV, it was stiff to very hard clay and silty clay with a thickness of approximately 48 m (Layer V), followed by very dense silty fine sand (Layer VI) to the termination depth of field exploration at 130 m BGS. The phreatic water level varied from 1.0 to 2.2 m BGS, fluctuating with weather changes. There were two aquifers: one was a perched aquifer, located at 6–12 m BGS with an elevated water level at 1.3–1.6 m BGS; the other was a confined aquifer located at 38–44 m BGS with an elevated water level at 4.33–4.63 m BGS. This indicated that if the excavation depth was over 20 m, the weight of the soil strata above the confined aquifer would be unable to hold down the underlying aquifer pressure; i.e., rupture or uplift of basal soils would occur (e.g., Moore and Longworth 1979; Chow and Ou 1999; Tan and Lu 2017a). To derive soil parameters for design, laboratory tests (e.g., consolidation, direct shear, triaxial, and seepage analysis tests) were performed on Shelby tube soil samples, and measured soil parameters are presented in Fig. 2. Cohesion cDS and friction angle φDS were derived from consolidated undrained direct shear (CUDS) tests; consolidated undrained cohesion ccu and friction angle φcu as well as effective cohesion c and friction angle φ were measured by consolidated undrained (CU) triaxial tests; void ratio e and constrained modulus M were derived from one-dimensional compression (consolidation) tests; water content ω was measured by the oven-drying method; liquid limit LL and plastic limit LP were measured by photoelectric liquid-plastic testers; horizontal and vertical permeability coefficients kh and kv were measured by constant head permeability tests in the laboratory; and total soil permeability k was measured by in-situ injectivity tests.
Fig. 2. Soil profile along with measured soil parameters at site

Design and Construction

To save project cost and shorten construction duration, composite earth-retaining systems were designed for excavation of this basement. The topmost 7.85 m soils along the north and west pit sides were excavated by the sloped open-cut method (Fig. 3); as for the south side, the adjacent artificial river (18 m wide and 5.85 m deep) was drained, followed by demolishing the north river embankment, and then soils were cut to 7.85 m BGS (Fig. S1). Because of the high phreatic level, light-well points were adopted for open-cut to lower groundwater below the excavation surface. To prevent slope failure, exposed cut surfaces were covered by steel-welded meshes and sprayed shotcrete. To safeguard Xing-Gang Street with buried utilities, a house and a bridge to the east (Fig. 1), upper soil removal along the east pit side was retained by soil nailing wall (SNW), i.e., mixed-in-place (MIP) piles, also known as soil-mixing wall (SMW), anchored by soil nails (Fig. 3).
Fig. 3. Typical cross-sections of the pit
Regarding excavation of the soils from 7.85 to 21.5 m BGS, it was supported by continuous bored pile (CBP) walls braced by concrete struts at three levels. To further restrain lateral wall deflection, the soils at 14.0–25.5 m BGS on the excavation side against the east CBP wall were reinforced by MIP piles (5.2  mwide×11.55  mhigh). Soil removal below the first level of struts followed a basin-type excavation procedure down to 21.5 m BGS; i.e., excavation was conducted at the pit center, immediately followed by pouring concrete struts normal to CBP walls; once concrete reached 28-day strength, residual earth berms against the perimeter walls were removed, accompanied by casting diagonal struts. Fig. 4 shows the plan layout of struts cast at Levels 1 to 3. To facilitate soil disposal, concrete decking slabs (30 cm thick) were cast on normal struts at Level 1 (the shaded areas in Fig. 4). The small inner pits, were retained by SNWs; meanwhile, soils below excavation bases were jet-grouted (see Fig. 5). Detailed construction schedules refer to Table 1.
Fig. 4. Plan instrumentation layout of the pit along with struts cast at Levels 1 to 3
Fig. 5. Cross-sections of the inner pits: (a) N1-N1 of the north inner pit; (b) S1-S1 of the south inner pit
Table 1. Excavation Sequence at the Site of the Gate of the Orient
Construction stagesActivitiesTime (month/day/year)Days spent
S0Construction of continuous bored pile wall (CBPW), bored piles (load-bearing elements) inside excavation, waterproof curtains along the perimeter of CBPW, jet-grouting of the soils inside the inner pits, pumping out the water inside the river on the south side and demolishing the gravity wall along the north river side, and construction of dewatering wells and relief wells05/18/2005 to 03/23/20081,040
S1Sloped excavation of the pit along the north, west, and south pit sides as well as construction of soil nailing wall and open cut of the pit along the east side to elevation of 7.85  m below ground surface (BGS)03/24/2008 to 06/26/200894
S2Excavation to 9.25  m BGS and casting of capping beams and steel-reinforced concrete struts at 8.25  m BGS06/27/2008 to 07/27/200830
S3Excavation to 14.6  m BGS and casting of steel-reinforced concrete struts at 13.6  m BGS07/28/2008 to 09/11/200845
S4Excavation to 18.9  m BGS and casting of steel-reinforced concrete struts at 17.9  m BGS09/20/2008 to 10/28/200838
S5Excavation to 21.5  m BGS and casting of base slab at 21.5  m BGS10/29/2008 to 12/02/200834
S6Excavation of the inner pits at the north and the south towers to 30  to  30.7  m BGS12/03/2008 to 12/16/200814
S7Casting of base slab at the main towers12/17/2008 to 01/13/200931
S8Demolishing concrete struts and casting underground structures01/14/2009 to 01/12/2010363
To build waterproof curtains along the pit perimeter, MIP piles were constructed along the outer perimeter of the CBP walls and gaps between them were compaction-grouted to achieve satisfactory watertightness; see Figs. 1 and 3. Throughout excavation of soils below 9.25 m BGS, groundwater inside the pit was kept at 0.5–1.0 m below the excavation surface via one hundred and thirty 26–28 m deep pumping wells (Fig. 6). Because the confined aquifer located at 38–44 m BGS could cause uplift or rupture failure of basal soils or floating of basement structures, nine 44-m-deep relief wells were constructed inside the pit to reduce artesian pressure (Fig. 6).
Fig. 6. Plan layout of dewatering wells inside and outside pit

Instrumentation

Compared to excavations in soft clays (e.g., Ou et al. 1993; Koutsoftas et al. 2000; Hashash et al. 2008; Wang et al. 2010; Tan and Li 2011; Tan and Wei 2012; Tan and Wang 2013a, b; Tan et al. 2014, 2015b, 2016, 2017; Finno et al. 2015; Whittle et al. 2015; Xu et al. 2015; Tan and Lu 2017b, 2018), excavations in stiff clays have received much less attention (Tedd et al. 1984; Ng 1998; Long 2001; Liao et al. 2015; Tan et al. 2015a). With respect to a large deep excavation supported by composite earth-retaining systems, none was reported in the literature, and current methods for predicting pit performance might not be applicable to such cases any more. With the progress of urbanization, more and more large-sized deep excavations (e.g., Tan and Wang 2013a, b; Tan et al. 2015a) have been, and are to be, constructed in urban areas for using underground space. To save project cost or shorten construction duration, unusual composite earth-retaining systems are frequently used for such large excavations, which inevitably bring uncertainties to design and construction. Therefore, a well-documented case history such as this one would be a useful addition for upgrading the current state of the art and practice in deep excavation.
To ensure project safety and investigate excavation performance, this pit was extensively instrumented (Figs. 1 and 4). The monitored items included: (1) lateral CBP wall displacements at P1 to P14, (2) vertical and lateral displacements at waterproof curtains (Q1 to Q12) along the east side and slope crests along other sides (Q13 to Q23), (3) vertical CBP wall displacements at W1 to W38, (4) vertical column displacements at LZ1 to LZ43, (5) axial strut forces at 19 locations for each level (Zi-1 to Zi-19; i=1, 2, 3), (6) ground settlements along six survey sections (C1-1–C1-6 to C6-1–C6–5) and lateral ground displacements at T1 to T11, (7) basal heaves at HT1 to HT12 [at each location, heaves were measured at excavation base and 3 and 6 m below, designated as HTi-1 and HTi-3 (i=1–12)], and (8) variations in both phreatic levels at SW1 to SW20 (20 m deep and 2 m behind CBP wall) and artesian levels at CY1 to CY4 (40 m deep and 25 m behind CBP wall). Moreover, settlements of the adjacent building (F1–F4), bridge abutments (X1–X4), and those utility pipelines buried below Xing-Gang Street were closely monitored throughout construction as well.
Lateral displacements of CBP walls were monitored by deflection measurement systems (Fig. S2), which consisted of plastic inclinometer casings, standard slope indicators with accuracy of 0.1 mm, vibrating wire readout boxes for taking readings, and computer software for data analysis and processing. The used slope indicators (probes) had a resolution of 0.02  mm/500  mm, and the tilt of the walls was measured at 1-m intervals along depth with an accuracy of 5  mm/25  m. As for the lateral ground displacements, they were measured by inclinometer casings placed inside predrilled boreholes in the ground behind the pit. The measured deflection data by slope indicators were the relative displacements to the casing tops. To know the actual lateral displacement caused by excavation, the lateral movements at casing tops were also surveyed. This was done by a theodolite with an accuracy of 2 s. Therefore, the actual lateral wall or ground displacements were the sum of the measured data of inclinometer casings and the surveyed data of casing tops.
Regarding vertical displacements of CBP walls, at each selected survey point, a stainless steel piece was embedded at the top of CBP walls (Fig. S2), and their vertical movements were measured by a level instrument with an accuracy of 0.01 mm. The survey of vertical and lateral displacements of waterproof curtains, slope crest, and interior columns was adopted in the same way for CBP walls. Ground settlements behind the pit were measured via surface markers. They consisted of round-head steel rebars with a diameter of 20 mm and length of 200 mm, which were encased inside capped protection wells (Fig. S3). The protection wells were cast-iron pipes with a diameter of 150 mm and length of 700 mm. In order to secure the surface markers, the soils around the protection wells were compacted with cement mortar. With respect to the settlement survey for buried utility pipelines, at each survey point, the ground was excavated to the pipeline. Then, the pipeline was enclosed with a cast-iron ring, which was connected with an iron rod located vertically on the top of the pipeline. The iron rod was protected by a capped PVC pipe and the gap between them was filled with expansive clay. The iron rod was approximately 0.5 m below the ground surface (Fig. S4). The survey points for settlements of the building and bridge abutments consisted of steel pieces, which were inserted into predrilled holes on the exterior walls of the buildings or bridge piers to be monitored. Gaps between the steel pieces and the walls or piers were filled with cement mortar (Fig. S5).
Prior to commencement of the work, the ground, building, bridge, and utility pipelines within the predefined influence zones were surveyed to establish benchmarks (reference points) for monitoring the influence of the excavation. These benchmarks were constructed at approximately 100–300 m distant from pit, which was far enough not to be disturbed by the excavation. They consisted of concrete piers, which were embedded into undisturbed natural subgrade and protected by concrete wells. The vertical distance from the ground surface to the top of the piers was 0.2 m. Configuration of a typical benchmark is illustrated in Fig. S6.
Axial forces in reinforced concrete struts were measured by vibrating wire steel stress meters. Before concreting, four stress meters were welded on four reinforcing bars along the longitudinal direction of the struts, which were located at the middle points of the four sides of the rectangular steel reinforcement cage (Fig. S7). Cables of the stress meters for taking readings were protected by PVC pipes. Then, the instrumented rectangular steel reinforcement cage, along with the stress meters and PVC pipes, were poured inside the struts. The changes of axial force in the concrete struts would cause variation of tensile force in the steel wires of the stress meters and then change their vibration frequency. By recording vibration frequency with a frequency meter, the axial force of the concrete struts can be calculated using the calibrated relationship curves between frequency and axial force.
Groundwater levels were monitored by standpipes at approximately 2 m (SW1 to SW20) and 25 m (CY1 to CY4) behind the pit. Holes with a diameter of 100 mm were drilled in the ground at the selected locations (20.5 m deep at SW1 to SW20 for phreatic levels and 40.5 m deep at CY1 to CY4 for artesian levels). Then, 20–40-m-long and 50-mm diameter PVC pipes with many tiny holes, which were wrapped up with nylon nets to facilitate seepage of water, were placed into the holes at SW1 to SW20 and CY1 to CY4, respectively. The heads of the PVC pipes were 50 mm above the ground surface and capped for protection. Then, gaps between the PVC pipes and the holes were filled with soils. The fills consisted of expansive clay in the upper 4 m for both SW1 to SW20 and CY1 to CY4, whereas clean sands were used in the lower sections for SW1 to SW20 and in-situ soils for CY1 to CY4; see Fig. S8. Via a portable water level indicator, water level changes inside the standpipes were measured.
Basal heaves were measured by 12 clusters of magnetic extensometers installed inside 12 boreholes (designated as HT1 and HT2 in Fig. 4). In each borehole, three magnetic loops were installed at 0, 3, and 6 m below the excavation base (designated as HT1-1–HT1-3 and HT12-1–HT12-3 along the depth). The magnets were placed on the outside of a PVC pipe inside each borehole, and the gap between the borehole and the PVC pipe was filled with sandy soils. By pulling up the PVC pipe slightly, the barbs of the magnetic loops could pierce into the fill sands and clinch to the natural subgrade tightly (Fig. S9). By lowering a magnetic probe into the PVC pipe, the vertical displacements of the magnetic loops could be measured.

Analysis of Field Measurements

Variation in Phreatic and Artesian Levels

Fig. 7(a) depicts variations in phreatic levels at SW1 to SW20, approximately 2 m behind the CBP walls. During excavation of the uppermost 9.25 m of soils in Stages S1 to S2, phreatic levels behind the north wall (SW17 to SW20), west wall (SW12 to SW16), and south wall (SW7 to SW9) had 4 to 11 m drawdown; i.e., groundwater levels were satisfactorily lowered below the open-cut surface. In contrast, there were only small decreases, no more than 2.0 m, in phreatic levels behind the east wall (SW1 to SW6). Because no light well point was installed along the east pit side (Fig. 6), the monitored data indicated that the waterproof curtains (Figs. 1 and 3) successfully cut off potential seepage flow during dewatering for removal of the uppermost 9.25 m of soils.
Fig. 7. Variations in water levels over time: (a) phreatic water; (b) confined artesian water
Fig. 7(b) shows variations in the artesian levels at CY1 to CY4, approximately 25 m behind the CBP walls. In Stages S3 to S4, the monitored artesian levels at CY1 to CY4 had 10–14 m drawdown and then maintained these levels in subsequent Stages S4 to S7. Obviously, pumping of the confined aquifer water inside the pit caused significant drawdown in the artesian levels outside the pit. Fig. 8(a) illustrates the locations of the observation wells behind the pit and the pumping wells inside the pit relative to the CBP wall and waterproof curtain along the east pit side. As outlined in this figure, the waterproof curtain had not extended deeply into the confined aquifer layer at 38–44 m BGS; i.e., the water flow path between the pit and the ground outside had not been cut off. Consequently, pumping of the confined aquifer by relief wells inside the pit led to dramatic drawdown in the artesian level outside the pit. As shown in Fig. 7(a), phreatic levels did not show substantial variations in Stages S3 to S7; i.e., dewatering of the confined aquifer water did not cause apparent leaking or seepage flow in the upper clayey strata (aquitard) featuring low permeability (Fig. 2). Following completion of the base slab, all pumping wells were shut off upon the end of Stage S7, and artesian levels recovered rapidly to the original levels within approximately 40 days.
Fig. 8. Cross-section view of the east pit side and instruments for water levels: (a) conceptual model for dewatering, waterproofing, and monitoring of water level; (b) model for evaluation ground settlement because of dewatering

Lateral Wall and Ground Displacements

Fig. 9(a) presents typical lateral CBP wall displacements at P2, P6, P11, and P13 at different construction stages. Apparently, P6, P11, and P13 developed smaller displacements than P2 on the east pit side; this discrepancy should result from the open cut, which reduced lateral earth pressures behind the CBP walls along the other three pit sides. Fig. 9(b) shows corresponding lateral ground displacements behind the CBP walls, i.e., T2 at 1 m behind P2, T5 at 13.5 m behind P6, T8 at 23.5 m behind P11, and T10 at 23.5 m behind P13. Both the CBP walls and ground developed typical deep-seated displacements in Stages S2 to S5, and their displacement developments synchronized with each other. Benefitting from the SNW and jet-grouting piles, excavation of the inner pits just incurred very limited additional wall and ground displacements in Stage S6. During the subsequent 31 days for pouring the base slab (Stage S7), neither the CBP walls nor the ground underwent apparent time-dependent displacement; this differed from those excavations in Shanghai soft clay (Tan et al. 2015b) also constructed by bottom-up (BU) method. It was evident that stiff clay hardly featured creep behavior. However, both the CBP walls and ground underwent noticeable postexcavation lateral displacements up to 20 mm in 1 year (Stage S8). Fig. 10 plots typical time histories of settlements for the buried communication cable at D1 to D13 and bridge abutments at X1 to X4 behind the pit (Fig. 1). The shallowly buried pipelines and bridge abutments experienced remarkable settlements during excavation (Stages S1 to S7); however, all of them developed very limited settlements in Stage S8. This was clearly indicative of limited consolidation-induced ground settlement postexcavation. Moreover, as shown in Fig. 7, the confined aquifer levels recovered rapidly to their original levels after completion of excavation, which would push up the overlying impermeable clayey strata (aquitard) somewhat. In light of these, it can be reasonably deduced that the significant postexcavation lateral wall and ground displacements should be largely associated with demolishing the rigid concrete struts in Stage S8 for casting permanent underground structure elements (e.g., inner side walls and floor slabs) rather than dissipation of excess pore pressure (consolidation) in the ground. Dismantling of the concrete struts caused an immediate reduction in supporting system stiffness; moreover, it would take poured underground structural elements at least 28 days to get fully cured (gain sufficient strength) for resisting lateral wall displacements. As a result, remarkable postexcavation lateral wall displacements occurred in Stage S8, followed by lateral ground displacement directly behind.
Fig. 9. Typical lateral displacement profiles of (a) CBP wall; (b) ground behind wall
Fig. 10. Typical settlement development of adjacent environment behind pit: (a) buried communication cable line along D1 to D13; (b) existing bridge at X1 to X4
By comparing the lateral wall and ground displacements, it can be seen that the lateral ground displacements of T2 immediately behind the east pit side were comparable to the corresponding lateral wall displacements of P2. Even at a distance of 23.5 m, the lateral ground displacements of T8 and T10 behind the open cuts were only slightly smaller than the lateral wall displacements of P11 and P13. These observations reveal that the excavation-induced lateral ground displacement zone extended far beyond 1.1He behind the pit, where He denotes the final excavation depth at 21.5 m BGS. This is demonstrated by the measured ground settlement to be presented later.

Relationship among δhm, H, and Hm

Fig. 11(a) plots the relationship between the maximum lateral displacement δhm of CBP walls and excavation level H; Fig. 11(b) plots the relationship between H and depth Hm with δhm. Because lateral displacements of CBP walls were not measured during removal of the uppermost 7.85 m soils, H in Fig. 11 was the excavation depth depicted in Table 1 subtracted by 7.85 m. To better understand excavation performance in Suzhou stiff clay, field data from another 21 excavations in Suzhou summarized in Table 2 were also included. Suzhou is a city located on the alluvial deposit—Yangtze River Delta plain. Similar geological and hydrological conditions feature across the entire city; i.e., all the cases in Table 2 are comparable. Except for Case 16 (Tan et al. 2015a) and Case 20 following zoned-excavation procedure, as well as Case 17 following the top-down (TD) excavation procedure, all these excavations adopted a typical bottom-up (BU) procedure. Moreover, empirical envelopes in the literature were included in Fig. 11 for comparison, i.e., Clough and O’Rourke (1990) and Moormann (2004) for stiff clays worldwide, Hashash et al. (2008) for the Central Artery/Tunnel (CA/T) project in medium stiff Boston Blue clay (BBC), and Tan and Wang (2013a, b) for BU and TD excavations in Shanghai soft clay. As shown in Fig. 11(a), the east CBP wall with a SNW had slightly greater δhm than the other CBP walls with open cut in this project, and their δhm ranged between δhm=0.10%H and δhm=0.50%H, which agreed with those Suzhou excavations retained by multipropped CBP walls but was greater than those supported by multipropped diaphragm walls (DWs). Compared with the envelopes in literature, the upper boundary for Suzhou excavations (δhm=0.50%H) matched with that of Clough and O’Rourke (1990), but was only half of that (δhm=1.0%H) of Moormann (2004). Compared with those excavations in Shanghai soft clay with δhm1.01.2%H (Tan and Wang 2013a, b), δhm in Suzhou stiff clay were much smaller. Despite BBC being weaker than Suzhou clay, δhm reported in Hashash et al. (2008) was considerably smaller than those in Suzhou. This should mainly arise from the narrow excavation widths (17.7–32.9 m) of the CA/T, whose braced struts were much shorter than those in this case. From the viewpoint of material mechanics, a short rod (strut) is much stronger than a long rod (strut) against buckling when subjected to axial loading because of its smaller slenderness ratio.
Fig. 11. Relationships among δhm, H, and Hm for excavations in Suzhou stiff clay: (a) H and δhm; (b) H and Hm
Table 2. Detailed Information of Another 21 Excavations in Suzhou Stiff Clay
Case numberL(m)×W(m)He (m)Retaining wallPropping systemExcavation method
1140×1409.05SMW (3Ø[email protected]; Lw=19  m)1 level of circular concrete strutsBU
2110×1109.95CBP wall (Ø[email protected]; Lw=18.9  m)1 level of concrete strutsBU
3135×11511.2CBP wall (Ø[email protected]; Lw=20  m)1 level of circular concrete strutsBU
474×5817.1CBP wall (Ø[email protected]; Lw=3032  m)3 levels of concrete strutsBU
552×5211CBP wall (Ø[email protected]; Lw=21  m)2 levels of concrete strutsBU
613.2CBP wall (Ø[email protected]; Lw=26.7  m)2 levels of concrete strutsBU
779×709.2CBP wall (Ø[email protected]; Lw=18  m)1 level of concrete strutsBU
8180×6012.7CBP wallBU
9150×8010.7CBP wall (Ø[email protected]; Lw=28  m)2 levels of concrete strutsBU
10181.6×19.117DW (t=0.8  m; Lw=32  m)2 levels of strutsBU
11285×17.526DW (t=1.0  m; Lw=46  m)Concrete struts at Levels 1 and 4 and steel pipe struts at Levels 2, 3, and 5BU
1241.4×35.620DW (t=1.0  m; Lw=37  m)Concrete struts at Levels 1 and 3 and steel pipe struts at Levels 2, 4 to 6BU
13315.2×32.416DW (t=0.8  m; Lw=29  m)3 levels of concrete strutsBU
14122.65×19.918.96DW (t=0.8  m; Lw=33  m)Concrete struts at Levels 1 to 3 and steel pipe struts at Levels 4 and 5BU
1515.9DW (t=0.8  m; Lw=31  m)4 levels of strutsBU
16240×15015.6–17.6DW (t=0.81.0  m; Lw=34  m)3 levels of concrete strutsZoned BU
17178182×559519.04DW (t=0.81.0  m; Lw=3538  m)3 levels of concrete floorsTD
18122×18.715.5SPW (Ø[email protected]; Lw=27.2  m)Concrete struts at Level 1 and steel pipe struts at Levels 2 to 4BU
19138.5×41.55.15SMW (Lw=10  m)1 level of cable anchorBU
20296×44514.65–20.55DW (t=1.0  m; Lw=3048  m)3–4 levels of concrete strutsZoned BU
21145×1509–9.6SMW (Lw=10  m)1 level of circular concrete strutsBU

Note: @ = pile center-to-center spacing; Ø = pile diameter; BU = bottom-up method; CBP = continuous bored pile; DW = diaphragm wall; He = final excavation depth; L = pit length; Lw = wall or pile length; SMW = soil-cement mixing wall; SPW = secant pile wall; TD = top-down method; t = wall thickness; W = pit width.

Distinct from excavations in Shanghai in which δhm occurred at depths within 7 m below and above excavation surfaces (Hm=H7m to Hm=H+7m), most δhm in Suzhou occurred above the excavation surface (Hm=H6m to Hm=H for those retained by propped CBP walls and Hm=H12m to Hm=H for those by propped DWs). As summarized in Table 2, those cases in Suzhou had final excavation depths He of 5.15–26 m and retaining wall lengths Lw of 10–46 m. With the exception of Case 11, retaining walls of all the cases were completely embedded in the upper clayey strata and wall toes did not penetrate into the underlying silty sand layer. As for Case 11, it had an excavation depth of 26 m and wall embedment length (wall length below excavation base) Le of 20 m; i.e., approximately 12 m of Le was in the upper clayey strata and 8 m of Le was in the underlying silty sand. Both the clayey strata and silty sand layer in Suzhou were stronger than those of Shanghai. Therefore, the discrepancy for Hm in Suzhou and Shanghai should be associated with the much stiffer clay strata or denser silty sand in Suzhou; i.e., the soils below excavation level in Suzhou could provide larger resistance against inward lateral wall deflection than those in Shanghai. Consequently, δhm in Suzhou had a much greater possibility of occurring above excavation surfaces than δhm in Shanghai.

Ground Settlement

The ground behind the pit developed significant settlements δv during excavation, which were up to 168 mm behind the east pit side and 107 mm behind the north and west sides; see Fig. 12. The bar charts in Fig. 13 summarize ground settlement increments δiv in each construction stage along C1-1 to C1-6, C2-1 to C2-4, and C3-1 to C3-4 behind the pit middle spans, in which δiv was normalized by the corresponding final measurement δfv, and distance d between survey point and pit was normalized by He=21.5m. Throughout excavation, δv increased dramatically over time and did not show a sign of stopping until completion of the base slab in Stage S7. Although settlement rates in Stages S7 to S8 were much smaller than those in Stages S3 to S6, the cumulative δiv developed in 1 year (Stages S7 and S8) were considerable, up to 2140%δfv. Considering that the rapid recovery of artesian water after completion of excavation (Fig. 7) would try to push up its overlying impermeable clayey strata (aquitard) somewhat, the majority of these postexcavation δiv should result from postexcavation inward lateral wall displacements associated with dismantling rigid concrete struts (Fig. 9) rather than consolidation of the clayey strata over time. This deduction can be also validated by the very limited postexcavation settlements of the shallowly buried utility pipelines and the bridge piers behind the east CBP wall (Fig. 10).
Fig. 12. Typical ground settlement development over time at the survey sections: (a) C1-1 to C1-6 behind the east wall; (b) C3-1 to C3-4 behind the north wall
Fig. 13. Ground settlement increments at different excavation levels for three survey sections: (a) C1-1 to C1-6; (b) C2-1 to C2-4; (c) C3-1 to C3-4

Ground Settlement Profile

Fig. 14 presents distribution of δfv along distance d from the pit, in which both δfv and d were normalized by He. Furthermore, data available from Cases 11, 12, 16, and 17 and Moormann (2004), as well as the empirical envelopes in literature (Peck 1969; O’Rourke 1976; Clough and O’Rourke 1990; Hashash et al. 2008), were included. Obviously, δfv behind this multipropped excavation had spandrel-type profiles typical for excavations retained by cantilever wall (e.g., Peck 1969; Clough and O’Rourke 1990) instead of the concave type typical for multipropped excavations (e.g., Ou et al. 1993; Tan and Wang 2013a, b). This inconsistency should derive from excavation of the uppermost 7.85 m soils being achieved by open cut or retained by SNWs. Evidently, the composite earth-retaining systems adopted in this excavation made its ground settlements distinctively different from those of other multipropped excavations. Within 1.5He from the pit, the ground along C1-1 to C1-6 behind the east pit middle span had greater δfv than the corresponding C2-1 to C2-6 and C3-1 to C3-4 behind the west and north pit middle span. In spite of this discrepancy, they had similar ground settlement influence zones, which can be bounded by envelope (5). Along the same side, C1-1 to C1-6 behind the middle span had apparently greater settlements than both C5-1 to C5-5 and C6-1 to C6-5 distance away, which exhibited clear evidence of spatial corner stiffening behavior like those cases in Tan et al. (2014).
Fig. 14. Relationship between d/He and δfv/He for excavations in Suzhou stiff clay
For this excavation, zone I of Peck (1969) for sand and soft to hard clay highly underestimated its δfv/He and ground settlement zone; δfv/He in this case was up to 0.8% within zone II of Peck (1969) for very soft to soft clay. Envelope (2) of O’Rourke (1976) for soft to medium stiff Chicago glacial clay would highly overestimate δfv/He within 1.5He for this case, but underestimate its δfv/He beyond 2He. Regarding envelope (3) of Clough and O’Rourke (1990) for stiff to hard clay, it could make a reliable estimation on the ground settlement zone, but would highly underestimate δfv/He. Envelope (4) of Hashash et al. (2008) for the CA/T project in medium stiff BBC would highly underestimate δfv/He, but overestimate the ground settlement zone. As for the other four Suzhou excavations, envelope (3) of Clough and O’Rourke (1990) could make a good prediction in terms of both δfv/He and the ground settlement zone. With the exception of this case, most δfv/He in Suzhou and Moormann (2004) were within zone I of Peck (1969). Apparently, the measured ground settlements in this case were much greater than those of Cases 11, 12, 16, and 17 in Suzhou.
It has been recognized in the literature (e.g., Galloway et al. 1999; Xu et al. 2016; Wu et al. 2017) that discharging of ground water (phreatic and artesian water) could cause pronounced land subsidence. Considering the significant drawdown in artesian water levels behind this excavation (Fig. 7), it was suspected that the exceptionally larger ground settlements of this excavation compared to the other excavation cases in Suzhou might be associated with dewatering. As illustrated in Table 2, Case 11 had He=26m and Lw=46  m; i.e., its DW had penetrated through the confined aquifer at 38–44 m BGS and extended into the underlying impermeable clay and silty clay stratum (aquitard). Because the DW had relatively good watertightness, pumping of aquifer water inside the pit for releasing artesian pressure in Case 11 would not cause substantial drawdown in the water level outside the pit; i.e., dewatering-induced ground settlement would not be significant for Case 11. As for Cases 12, 16, and 17, their He were no more than 20 m, and dewatering for releasing underlying confined aquifer pressure was not required; i.e., as mentioned previously, for excavations with He20  m in Suzhou, the weight of overburden soils was able to resist upward confined aquifer pressure. Like this case, the DWs of Cases 12, 16, and 17 were toed in the upper clayey strata and did not penetrate into the underlying confined aquifer. In light of the good watertightness of the DW and the low permeability of the upper clayey strata (aquitard), pumping of phreatic water inside the pit would not cause a substantial drawdown in the water level outside the pit for Cases 12, 16, and 17. Different from Cases 12, 16, and 17, pumping of confined aquifer water inside the pit was conducted in this case because of its great excavation depth (21.5–30.7 m). However, its waterproof curtain was toed at 27.3 m BGS and did not penetrate deeply through the confined aquifer stratum [see Figs. 3 and 8(a)]. Consequently, significant drawdown in the water level outside the pit was monitored during dewatering (see Fig. 7). On the basis of these analyses, it was reasonable to infer that the dramatically larger ground settlements of this excavation compared to those of Cases 11, 12, 16, and 17 should, to some extent, result from the significant drawdown in the aquifer level.

Discussion on Ground Settlement Caused by Dewatering

The preceding analyses disclose that the drawdown in water level imposed a significant adverse influence on ground settlement. As plotted in Fig. 7(a), the measured drawdown in phreatic level by the 20-m-deep observation wells at 2 m behind CBP walls was up to 2 m along the east pit side and up to 11 m along the other three sides. The drawdown in the underlying confined aquifer level measured by the 40-m-deep observation wells at 25 m behind the CBP walls was up to 14 m along the pit perimeter; see Fig. 7(b). As schematically illustrated in Fig. 8(b), drawdown of the water (phreatic or artesian) level in the ground behind the pit would incur consolidation-induced settlement St for the upper 38 m silty clay and clayey silt strata (aquitard) and compression Sc of the underlying silty sand layer (confined aquifer). Therefore, dewatering-induced total ground settlement S can be given by
S=St+Sc
(1)
As for the upper aquitard [Fig. 8(b)], dewatering of its phreatic water levels from h1 to h2 caused increment of vertical effective soil stress Δσ1 with a magnitude of Δσ1=u1=γω(h1h2) in the aquitard, in which γω = density of water, h1 = phreatic level before dewatering, and h2 = phreatic level after dewatering. Meanwhile, releasing (dewatering) of artesian water in its underlying confined aquifer led to reduction of u2=(γωh3γωh4) in the upward artesian pressure against the aquitard bottom, which can be treated as a downward pressure Δσ2=u2 applied to the bottom of the upper aquitard after dewatering, where h3 = confined aquifer level before dewatering and h4 = confined aquifer level after dewatering. According to the one-dimensional consolidation theory of Terzaghi (1925), the change in pressures acting on the boundaries of soil strata would result in consolidation (compression). Although the water level in the upper perched aquifer at 6–12 m BGS was not monitored during excavation, it should rarely affect ground settlement. As illustrated in Fig. 8, the waterproof curtains (CBP wall and MIP) extended to 27.3–34.05 m BGS, and this perched aquifer was intermediate between the two impermeable aquitards; therefore, dewatering conducted inside the excavation hardly influenced the water level of the perched aquifer outside the excavation. Hence, it can be postulated that the upper perched aquifer should rarely affect ground settlement behind the excavation. Based upon the preceding analyses, it can be concluded that for the upper 38-m-thick soil strata overlying the confined aquifer at 38–44 m BGS, its St primarily consisted of St1 due to drawdown u1 in its phreatic level and St2 due to drawdown u2 in its underlying confined aquifer level. In accordance with the one-dimensional consolidation theory of Terzaghi (1925), St can be approximately estimated by the following equations:
St=St1+St2=U·(S1+S2)=U·S
(2)
S1=av1+e0·u12·H1=u12M·H1
(3)
S2=av1+e0·u22·H1=u22M·H1
(4)
S=S1+S2=av1+e0·u1+u22·H1=u1+u22M·H1
(5)
u1=γω·(h1h2)=γω·Δh1
(6)
u2=γω·(h3h4)=γω·Δh2
(7)
U=18π2exp(π24Tv)
(8)
Tv=cvtH12
(9)
cv=k·(1+e0)γωav=k·Mγω
(10)
where, St = consolidation-induced ground settlement at time t for the upper aquitard; St1=St associated with drawdown in phreatic level; St2=St associated with drawdown in the underlying confined aquifer level; U = consolidation ratio; S = final consolidation-induced ground settlement; S1=S associated with drawdown in the phreatic level; S2=S associated with drawdown in the underlying confined aquifer level; H1 = thickness of the upper aquitard, equal to 38 m; av = compressibility coefficient of the upper aquitard; e0 = initial void ratio of the upper aquitard; M = constrained modulus of the upper aquitard; and k = permeability coefficient of the upper aquitard.
As shown in Fig. 7, the reductions in the phreatic level and underlying confined aquifer level were up to 2 and 14 m behind the east pit side (Δh1=2  m; Δh2=14  m) and 11 and 14 m behind the other three sides (Δh1=11  m; Δh2=14  m). Thereafter, the phreatic levels showed limited variation throughout Stages S3 to S7. According to Eqs. (2)(10), the roughly estimated St during excavation was approximately 68 and 55 mm for the upper aquitard behind the east and the other three sides, respectively.
Regarding the underlying confined aquifer layer (silty sand), drawdown in the aquifer level would increase its effective stress with a magnitude of γω·(h3h4), which would lead to compression settlement Sc of the silty sand layer. It can be approximately estimated by
Sc=γωM·(h3h4)·H2
(11)
where H2 = thickness of the confined aquifer layer, equal to 6 m; and M = constrained modulus of the silty sand (confined aquifer layer). The estimated Sc could be up to 60 mm. Based on the previous analysis, the total ground settlement (S=St+Sc) related to dewatering could be up to 128 mm behind the east pit side and 115 mm behind the other three sides.
As illustrated in Fig. 8(a), the 44-m-deep relief wells for releasing confined aquifer pressure were completely penetrating wells; i.e., they penetrated through the confined aquifer and reached the underlying aquitard. The influence zone due to pumping of aquifer water can be roughly delineated with the empirical equation proposed in Liu and Wang (2009), i.e.,
R=10·Δh·k
(12)
where R = radius of influence from relief well; Δh = drawdown in aquifer level; and k = permeability of confined aquifer layer measured by in-situ injectivity test. The estimated R could be up to 45 m distant from this pit, more than 2He. This in part explains why the ground settlement of this case measured at a distance of 2He behind the pit was still significantly greater than the other cases in Suzhou; see Fig. 14.
The previous theoretical analyses disclose that the majority of the pronounced ground settlement measured behind this excavation should derive from the significant drawdown in the aquifer level outside the pit, whose influence could reach as far as 2He behind the pit. The significant drawdown in the water level largely arose from the previously mentioned deficiency in the waterproof design. This deficiency led to the water flow path in the confined aquifer layer not being cut off between the pit and the ground outside [Fig. 8(a)] during discharging underlying confined aquifer water inside the pit; consequently, a significant drawdown in water level was monitored behind the pit. If such a project were constructed in a densely populated urban environment, a safer design would have to be adopted to minimize ground subsidence to protect adjacent structures or facilities, e.g., (1) extending the waterproof curtain deeply enough to penetrate through the underlying confined aquifer despite the risk of increasing project cost dramatically, or (2) recharging the underlying confined aquifer with the water discharged by relief wells inside the pit in the proximity of the structures or facilities to be protected (see the illustrations in Fig. 15).
Fig. 15. Countermeasures for discharging underlying aquifer water: (a) extending waterproof curtain deep through underlying confined aquifer layer; (b) recharging confined aquifer outside pit

Basal Heave

To date, only limited basal heave data in Taipei and Shanghai soft clays were available in the literature (Ou et al. 1998; Tan and Wang 2013a, b), whereas no data were reported for excavations in stiff clay. Fig. 16(a) plots the development of basal heaves δbh over time for this case. Like soft clays, basal stiff clay rebounded remarkably with the progress of excavation, which was up to 34 mm in Stage S4. To further characterize basal heave, δbh is plotted against H in Fig. 16(b). In addition, δbh data from Taipei and Shanghai soft clays were also included. The value of δbh ranged between δbh=0.10%H and δbh=0.22%H in this case, which matched that for the unpropped circular excavation in Shanghai soft clay but was only 1/4 to 1/2 of those (δbh=0.20%H and δbh=0.90%H) for multipropped excavations in soft clays.
Fig. 16. Observed basal heave δbh at this site: (a) development of δbh over time; (b) relationship between δbh and H

Vertical Displacements of CBP Walls, Waterproof Curtains, and Interior Columns

Fig. 17 presents development of vertical displacements δvw over time for CBP walls. Except for the survey points near the northeast (W1 to W3) and southwest corners (W18 to W20), both the east and south CBP walls moved downward during excavation. On the contrary, the north and west CBP walls moved upward. These opposite movements should be related to the different retaining systems used. As for the east CBP wall, upward force on the inner wall face and wall toe resulting from basal rebounding could not overcome downdrag force on the outer wall face caused by ground settlement; consequently, the east wall settled instead of upheaving. Regarding the north and west CBP walls, because of the open cut, downdrag force on the outer wall face because of ground settlement behind the pit could not suppress wall upheaving resulting from basal rebounding; hence, they moved upward rather than settling. As for the south CBP wall, its open cut was not as large as those behind the north and west walls; therefore, its δvw had an intermediate development tendency. In spite of this inconsistency, following completion of excavation and subsequent construction of underground structures, all CBP walls started to settle gradually over time.
Fig. 17. Summary of vertical displacement development over time at CBP wall heads: (a) east wall; (b) south wall; (c) west wall; (d) north wall
Fig. 18(a) presents typical vertical displacements of waterproof curtains along the east wall (Q1 to Q12). Like their corresponding lateral displacements (the maximum was 92 mm near the pit middle span and the minimum was 53 mm near corners (see Fig. 19), settlement data of the east waterproof curtain exhibited strong corner stiffening behavior as well; i.e., the maximum settlement (130 mm at Q6) occurred near pit midspan and the minimum (52–54 mm at Q1, Q2, and Q12) near corners.
Fig. 18. Typical vertical displacement development over time for (a) the east waterproof curtain; (b) interior columns
Fig. 19. Typical lateral displacement development over time for (a) Q1 to Q12 at heads of the east waterproof curtains; (b) Q21 to Q23 at the north slope crest
Fig. 18(b) plots development of typical vertical displacements δcu over time for interior columns within the northeast pit zone. Generally, δcu data of columns LZ1 to LZ4 near the east CBP wall were much smaller than those of the others. Contrary to expectations, δcu of LZ27 and LZ28 near the north wall was not smaller than those of LZ29, LZ31, LZ32, and LZ42 near the pit center featuring the largest basal heave. This should be the combined results of the following two factors: (1) columns near the pit center were directly below heavy decking slabs—hence, their upheaving was suppressed to some extent; and (2) settling of the east CBP wall restrained upheaving of adjacent columns.

Contour Map of Vertical Ground and Structure Displacements

Fig. 20 presents contour maps of vertical displacements for interior columns, CBP walls, and the ground behind the pit at the end of different construction stages, in which positive and negative magnitudes represent upward and downward movements, respectively. Excavation of the uppermost 9.25 m soils in Stages S1 and S2 hardly incurred vertical ground and structure displacements. As excavation continued to 14.6 m BGS and deeper in Stages S3 to S5, significant vertical displacements took place, although underlying artesian pressure was reduced dramatically [Fig. 7(b)]. Throughout Stages S3 to S5, the southeast zone had smaller displacements than the other zones; this should in part arise from more concrete struts cast at the southeast zone (Fig. 4), which helped constrain column upheaving. With demolishing of concrete struts and completion of underground structures, vertical column displacements became relatively uniform across the entire pit. As expected, interior columns developed larger upward displacements than perimeter walls.
Fig. 20. Contour maps of observed vertical displacements (unit in mm) for interior columns inside pit, retaining walls along pit perimeter, and the ground behind pit at end of different construction stages

Axial Strut Force

Figs. 21 and 22 summarize axial force developments for concrete struts cast at Levels 1 and 2. Because the measured axial strut forces at Level 3 (Fig. S10) had similar magnitudes and development patterns as those of Level 2, they are not presented herein. Distinct from those at Levels 2 and 3, axial forces of normal struts were significantly larger than axial forces of diagonal struts by 10-fold at Level 1. These extremely large axial forces should arise from operation of heavy-duty trucks (Fig. 4), which caused bending of underlying normal struts (i.e., significant bending stresses were generated). Although upper soils outside both the southwest and northwest pit corners were openly cut, axial forces of diagonal struts Z1-15 and Z1-16 against the northwest corner were much larger than those of Z1-6 and Z1-7 against the southwest corner. If soils above 7.85 m BGS behind the CBP wall were treated as surcharge (Fig. 23), much larger additional stresses were generated behind the north and west walls than the south wall. Therefore, Z1-15 and Z1-16 should sustain greater axial forces than Z1-6 and Z1-7, which was evidenced by smaller lateral displacement of the south wall than the north and west walls. As shown in Figs. 3 and 23, the ground behind the east CBP wall sustained a much larger surcharge than the ground behind the other walls. However, axial forces of diagonal struts Z1-1, Z1-2, Z1-3, and Z1-4 against the east wall were not greater. As mentioned previously, the east SNW (Q1 to Q12) moved significantly toward the pit during excavation, and its lateral displacements were up to 53–92 mm (Fig. 19); thus, anchor forces of the soil nails were mobilized to resist lateral SNW displacement. This resulted in the cohesive soil mass anchored by soil nails being pulled upward somewhat (i.e., the vertical load from the anchored soil mass behind the east CBP wall was alleviated to some extent) and meanwhile, the east CBP wall was pushed inward by the displaced SNW. Hence, struts against the east wall did not sustain larger lateral earth pressures than others, whereas the east CBP wall developed greater lateral deflections.
Fig. 21. Development of axial strut forces at Level 1: (a) diagonal struts; (b) normal struts
Fig. 22. Development of axial strut forces at Level 2: (a) diagonal struts; (b) normal struts
Fig. 23. Schematic illustrations of influence of the soils above 7.85 m BGS on CBP wall: (a) the north and west pit sides; (b) the south pit side; (c) the east pit side; (d) mechanical model for SNW
Different from axial strut forces at Level 1, axial forces of diagonal struts at Levels 2 (Z2-6 and Z2-7) and 3 (Z3-6 and Z3-7) against the southwest corner were not smaller than those of diagonal struts at Levels 2 (Z2-15 and Z2-16) and 3 (Z3-15 and Z3-16) against the northwest corner. This indicates that at 13.6 m BGS (Level 2), additional stresses induced by the surcharge (uppermost 7.85 m soils) were already insignificant.
As shown in Figs. 21 and 22, for struts cast at the same levels, their axial force magnitudes varied significantly. Beyond lateral earth pressures, struts might sustain bending stresses. Vertical differential displacements between adjacent columns could be indicative of strut bending. To verify this, vertical column displacements below normal struts at different excavation levels were plotted in Fig. 24. Apparently, there was no differential displacement between the two columns near either Zi-18 or Zi-19. In contrast, the columns adjacent to Zi-9, Zi-10, and Zi-11 had significant differential uplifts. These observations corresponded well with Zi-9, Zi-10, and Zi-11 having the maximum axial forces, whereas Zi-18 and Zi-19 had the minimum.
Fig. 24. Vertical column displacements below normal struts along the east–west and the north–south directions at different excavation levels

Conclusions

Through investigation on this large, deep excavation supported by composite earth-retaining systems in Suzhou stiff clay, as well as through comparison of its performance with another 21 Suzhou excavations, major conclusions can subsequently be drawn:
1.
In spite of the high phreatic level, underlying confined aquifer, and defects of CBP walls with joints, the adopted composite dewatering plans satisfactorily lowered the phreatic level for soil removal and reduced artesian pressure below the pit base to safe levels. A SNW in combination with jet-grouting of basal soils was proved to be an optimal solution for excavation of the 8.5–9.2 m-deep inner pits, in which no additional CBP wall displacement or rupture of basal soils was incurred;
2.
This case provided a good opportunity to investigate performance of a large, deep excavation supported by composite earth-retaining systems. Generally, the east pit side supported by the CBP wall combined with a SNW developed slightly larger lateral wall displacements than the other three sides retained by CBP walls with sloped open cut; the east CBP wall moved downward during excavation, whereas the other CBP walls with open cut moved upward. Although the ground behind the east CBP wall underwent much greater settlements than the ground behind the other three sides, they had similar settlement influence zones;
3.
Removal of the uppermost 7.85 m soils using sloped open-cut and SNW reduced project cost and shortened construction duration; additionally, lateral displacements δh of CBP walls in this case were comparable to those of other Suzhou excavations. Maximum lateral wall displacements δhm in Suzhou stiff clay were approximately δhm=0.10%H to δhm=0.50%H for multipropped CBP walls and δhm=0.06%H to δhm=0.30%H for multipropped DWs; both were far below the upper boundaries δhm=1.01.2%H of excavations in Shanghai soft clay. The δhm in Suzhou tended to occur above the excavation surface (Hm=H6m to Hm=H for CBP walls; Hm=H12m to Hm=H for DWs) instead of equally above and below the excavation surface in Shanghai soft clay;
4.
To save project cost, the waterproof curtain did not extend deeply into the underlying confined aquifer stratum in this excavation; i.e., water flow path in the aquifer inside and outside the pit was not cut off. As a result, significant drawdown of the water level up to 14 m in the artesian level behind the pit was monitored during discharging aquifer water inside the pit. This incurred conspicuously large ground settlements, which might exceed the ground settlements induced by excavation. Therefore, despite δhm of this case not being larger than those of other Suzhou excavations, its ground settlements were much greater. The maximum ground settlements measured within 48 m (2.2He) behind this excavation were 173 mm (0.8%He), up to 2.7 times those of other Suzhou excavations; its normalized δvf/He magnitudes were even comparable to those in Shanghai soft clay. With exception of this case, δvf/He data of all Suzhou excavations were within the envelope of Clough and O’Rourke (1990) for stiff clay. Because of the open cut and SNW, this case had spandrel-type ground settlement profiles instead of the concave type typical for multipropped excavations. Because dewatering was a major contributor to ground settlements during this excavation, those empirical envelopes proposed for excavations in medium stiff and stiff clays (Clough and O’Rourke 1990; Hashash et al. 2008) dramatically underestimated its δvf/He within 2He behind this pit. In the case where existing structures or facilities to be protected are in close proximity to such an excavation, countermeasures (e.g., penetrating waterproof curtain through underlying confined aquifer or recharging confined aquifer behind pit) should be accounted for in the design to minimize ground subsidence; and
5.
Beyond recognition in the literature, this excavation incurred considerable basal heave of δbh up to 34 mm at H=18.9m BGS; its δbh was approximately δbh=0.10%H to δbh=0.22%H, much smaller than that (δbh=0.20%H to δbh=0.90%H) in soft clays. Strut bending would increase axial strut force considerably and should be treated with caution in both design and construction. Differential column displacement is one of the indicators for strut bending. Despite stiff clay not featuring strong creep behavior, underground structures should be completed in a timely manner to prevent postexcavation displacements associated with demolishing of temporary struts. For a large excavation not following zoned-construction procedure, pit corner stiffening behavior was remarkable; in this case, displacements near the pit middle span were up to 2.5 times those near corners.

Supplemental Data

Figs. S1S10 are available online in the ASCE Library (www.ascelibrary.org).

Supplemental Materials

File (supplemental_data_gt.1943-5606.0001837_tan.pdf)

Acknowledgments

The financial support from the National Key Research and Development Plan (Grant 2016YFC0800204), the National Basic Research Program (973 Program) (2015CB057800), National Natural Science Foundation of China (NSFC 41672269), and Natural Science Foundation of Shanghai (16ZR1411900) are gratefully acknowledged. The insightful comments and suggestions from the three anonymous reviewers, Associate Editor, and Editor-in-Chief Dr. Mohammed A. Gabr are sincerely appreciated.

References

Chow, H. L., and Ou, C. Y. (1999). “Boiling failure and presumption of deep excavation.” J. Perform. Constr. Facil., 114–120.
Clough, G. W., and O’Rourke, T. D. (1990). “Construction induced movements of in-situ walls.” Design and performance of earth retaining structures, ASCE, New York, 439–470.
Finno, R., Arboleda-Monsalve, L., and Sarabia, F. (2015). “Observed performance of the one museum park west excavation.” J. Geotech. Geoenviron. Eng., 04014078.
Galloway, D., Jones, D. R., and Ingebritsen, S. E. (1999). Land subsidence in the United States, Vol. 1182, U.S. Geological Survey Circular, Reston, VA.
Hashash, Y. M. A., Osouli, A., and Marulanda, C. (2008). “Central artery/tunnel project excavation induced ground deformations.” J. Geotech. Geoenviron. Eng., 1399–1406.
Koutsoftas, D., Frobenius, P., Wu, C., Meyersohn, D., and Kulesza, R. (2000). “Deformations during cut-and-cover construction of MUNI metro turnback project.” J. Geotech. Geoenviron. Eng., 344–359.
Liao, S., Wei, S., and Shen, S. (2015). “Structural responses of existing metro stations to adjacent deep excavations in Suzhou, China.” J. Perform. Constr. Facil., 04015089.
Liu, G. B., and Wang, W. D. (2009). Deep excavation manual, China Architecture & Building Press, Beijing, China (in Chinese).
Long, M. (2001). “Database for retaining wall and ground movements due to deep excavations.” J. Geotech. Geoenviron. Eng., 203–224.
Moore, J. F. A., and Longworth, T. I. (1979). “Hydraulic uplift of the base of a deep excavation in Oxford clay.” Géotechnique, 29(1), 35–46.
Moormann, C. (2004). “Analysis of wall and ground movements due to deep excavation in soft soils based on a new worldwide database.” Soils Found., 44(1), 87–98.
Ng, C. W. W. (1998). “Observed performance of multipropped excavation in stiff clay.” J. Geotech. Geoenviron. Eng., 889–905.
O’Rourke, T. D. (1976). The ground movements related to braced excavations and their influence on adjacent buildings, Dept. of Transportation, Washington, DC.
Ou, C. Y., Hsieh, P. G., and Chiou, D. C. (1993). “Characteristics of ground surface settlement during excavation.” Can. Geotech. J., 30(5), 758–767.
Ou, C. Y., Liao, J. T., and Lin, H. D. (1998). “Performance of diaphragm wall constructed using the top-down method.” J. Geotech. Geoenviron. Eng., 798–808.
Peck, R. B. (1969). “Deep excavation and tunneling in soft ground. State-of-the-art-report.” Proc., 7th Int. Conf. of Soil Mechanics and Foundation Engineering, ISSMGE, Mexico City, Mexico, 225–281.
Tan, Y., Huang, R., Kang, Z., and Wei, B. (2016). “Covered semi-top-down excavation of subway station surrounded by closely spaced buildings in downtown Shanghai: Building response.” J. Perform. Constr. Facil., 04016040.
Tan, Y., and Li, M. (2011). “Measured performance of a 26 m deep top-down excavation in downtown Shanghai.” Can. Geotech. J., 48(5), 704–719.
Tan, Y., Li, X., Kang, Z., Liu, J., and Zhu, Y. (2015a). “Zoned excavation of an oversized pit close to an existing metro line in stiff clay: Case study.” J. Perform. Constr. Facil., 04014158.
Tan, Y., and Lu, Y. (2017a). “Forensic diagnosis of a leaking accident during excavation.” J. Perform. Constr. Facil., 04017061.
Tan, Y., and Lu, Y. (2017b). “Why excavation of a small air shaft caused excessively large displacements: Forensic investigation.” J. Perform. Constr. Facil., 04016083.
Tan, Y., and Lu, Y. (2018). “Responses of shallowly buried pipelines to adjacent deep excavations in Shanghai soft ground.” J. Pipeline Syst. Eng. Pract., in press.
Tan, Y., and Wang, D. (2013a). “Characteristics of a large-scale deep foundation pit excavated by the central-island technique in Shanghai soft clay. I: Bottom-up construction of the central cylindrical shaft.” J. Geotech. Geoenviron. Eng., 1875–1893.
Tan, Y., and Wang, D. (2013b). “Characteristics of a large-scale deep foundation pit excavated by the central-island technique in Shanghai soft clay. II: Top-down construction of the peripheral rectangular pit.” J. Geotech. Geoenviron. Eng., 1894–1910.
Tan, Y., and Wei, B. (2012). “Observed behavior of a long and deep excavation constructed by cut-and-cover technique in Shanghai soft clay.” J. Geotech. Geoenviron. Eng., 69–88.
Tan, Y., Wei, B., Diao, Y., and Zhou, X. (2014). “Spatial corner effects of long and narrow multipropped deep excavations in Shanghai soft clay.” J. Perform. Constr. Facil., 04014015.
Tan, Y., Wei, B., Zhou, X., and Diao, Y. (2015b). “Lessons learned from construction of Shanghai metro stations: Importance of quick excavation, prompt propping, timely casting, and segmented construction.” J. Perform. Constr. Facil., 04014096.
Tan, Y., Zhu, H., Peng, F., Karlsrud, K., and Wei, B. (2017). “Characterization of semi-top-down excavation for subway station in Shanghai soft ground.” Tunnelling Underground Space Technol., 68, 244–261.
Tedd, P., Chard, B. M., Charles, J. A., and Symons, I. F. (1984). “Behaviour of a propped embedded retaining wall in stiff clay at Bell Common Tunnel.” Géotechnique, 34(4), 513–532.
Terzaghi, K. V. (1925). Soil mechanics based on physical principles, F. Deuticke, Leipzig, Germany.
Wang, J. H., Xu, Z. H., and Wang, W. D. (2010). “Wall and ground movements due to deep excavations in Shanghai soft soils.” J. Geotech. Geoenviron. Eng., 985–994.
Whittle, A., Corral, G., Jen, L., and Rawnsley, R. (2015). “Prediction and performance of deep excavations for Courthouse Station, Boston.” J. Geotech. Geoenviron. Eng., 04014123.
Wu, H. N., Shen, S. L., and Yang, J. (2017). “Identification of tunnel settlement caused by land subsidence in soft deposit of Shanghai.” J. Perform. Constr. Facil., 04017092.
Xu, C., Chen, Q., Wang, Y., Hu, W., and Fang, T. (2015). “Dynamic deformation control of retaining structures of a deep excavation.” J. Perform. Constr. Facil., 04015071.
Xu, Y. S., Shen, S. L., Ren, D. J., and Wu, H. N. (2016). “Factor analysis of land subsidence in Shanghai: A view based on strategic environmental assessment.” Sustainability, 8(6), 573.

Information & Authors

Information

Published In

Go to Journal of Geotechnical and Geoenvironmental Engineering
Journal of Geotechnical and Geoenvironmental Engineering
Volume 144Issue 3March 2018

History

Received: Feb 1, 2016
Accepted: Aug 28, 2017
Published online: Dec 23, 2017
Published in print: Mar 1, 2018
Discussion open until: May 23, 2018

Authors

Affiliations

Yong Tan, M.ASCE [email protected]
Professor, Dept. of Geotechnical Engineering, College of Civil Engineering, Tongji Univ., 1239 Siping Rd., Shanghai 200092, P.R. China. E-mail: [email protected]
Ye Lu, A.M.ASCE [email protected]
Associate Professor, Dept. of Civil Engineering, Shanghai Univ., 149 Yanchang Rd., Shanghai 200072, P.R. China (corresponding author). E-mail: [email protected]
Dalong Wang [email protected]
Senior Engineer, Shanghai Geotechnical Investigations & Design Institute Company Ltd., 681 Xiao-Mu-Qiao Rd., Shanghai 200032, P.R. China. E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share