Vous êtes sur la page 1sur 16

Wongsaroj, J., Soga, K. & Mair, R. J. (2007). Géotechnique 57, No.

1, 75–90

Modelling of long-term ground response to tunnelling under St James’s


Park, London

J. WO N G S A RO J , K . S O G A y a n d R . J. M A I R y

Following a tunnel excavation in low-permeability soil, it Suite à une excavation de tunnel dans un sol de faible
is commonly observed that the ground surface continues perméabilité, il est fréquent d’observer que la surface du
to settle and ground loading on the tunnel lining changes, terrain continue de se tasser et que la charge du sol sur le
as the pore pressures in the ground approach a new revêtement du tunnel change à mesure que les pressions
equilibrium condition. The monitored ground response interstitielles dans le sol atteignent un nouvel équilibre. Le
following the tunnelling under St James’s Park, London, suivi de la réponse du sol après le creusement du tunnel
shows that the mechanism of subsurface deformation is sous le parc St James, à Londres, montre que le méca-
composed of three different zones: swelling, consolidation nisme de déformation sous la surface comporte trois zones
and rigid body movement. The swelling took place in a différentes de gonflement, consolidation et mouvement en
confined zone above the tunnel crown, extending verti- bloc. Le gonflement survient dans une zone confinée au
cally to approximately 5 m above it. On the sides of the dessus de la voûte du tunnel, s’étendant verticalement
tunnel, the consolidation of the soil occurred in the zone jusqu’à près de 5 m au-dessus. Sur les côtés du tunnel, la
primarily within the tunnel horizon, from the shoulder to consolidation du sol intervient dans la zone comprise
just beneath the invert, and extending laterally to a large essentiellement dans l’élévation du tunnel, de l’épaulement
offset from the tunnel centreline. Above these swelling jusqu’à juste en dessous du radier, et s’étendant latérale-
and consolidation zones the soil moved downward ment sur une large partie à partir de l’axe central du
as a rigid body. In this study, soil–fluid coupled three- tunnel. Au-dessus de ces zones de gonflement et de con-
dimensional finite element analyses were performed to solidation le sol descend vers le bas en bloc rigide. Dans
simulate the mechanism of long-term ground response cette étude, des analyses à éléments finis solide-fluide
monitored at St James’s Park. An advanced critical state couplées en trois dimensions ont été réalisées pour simuler
soil model, which can simulate the behaviour of London le mécanisme de réponse du sol à long terme contrôlé au
Clay in both drained and undrained conditions, was parc St James à Londres Un modèle avancé de comporte-
adopted for the analyses. The analysis results are dis- ment à l’état critique du sol, qui peut simuler le comporte-
cussed and compared with the field monitoring data. It is ment de l’argile de Londres dans des conditions drainées
found that the observed mechanism of long-term sub- et non drainées, a été adopté pour ces essais. Les résultats
surface ground and tunnel lining response at St James’s de ces analyses sont ici discutés et comparés avec les
Park can be simulated accurately only when stiffness données de suivi in situ. On a pu constater que le méca-
anisotropy, the variation of permeability between differ- nisme observé de réponse du revêtement du tunnel et du
ent units within the London Clay and non-uniform drain- sol sous la surface à long terme, au parc de St James, ne
age conditions for the tunnel lining are considered. This peut être simulé de façon exacte que lorsque l’anisotropie
has important implications for future prediction of the de la rigidité, la variation de la perméabilité entre les
long-term behaviour of tunnels in clays. différentes unités dans l’argile de Londres et les condi-
tions de drainage non uniforme du revêtement du tunnel
sont prises en compte. Ces observations ont des consé-
KEYWORDS: ground movements; numerical modelling; per- quences importantes en terme de prédiction du comporte-
meability; pore pressures; settlement; tunnels ment à long terme des tunnels creusés dans les argiles.

INTRODUCTION pore pressures will generally be lower than those prior to tunnel
Tunnelling in low-permeability soil often results in ground construction; settlement will therefore occur as pore pressures
surface settlement that continuously increases over a long reduce to their long-term steady-state values, increasing effec-
period of time. The mechanism for these longer term move- tive stresses and thereby inducing consolidation in the clay.
ments at the ground surface is that the tunnel inevitably The evidence that tunnels in low-permeability soil act as
introduces a new drainage boundary condition. This is because, new drainage boundaries was demonstrated from the field
on the inside face of the tunnel lining, the pressure is usually measurements of pore pressure around tunnels made earlier
atmospheric. The pore pressures immediately after construc- by Ward & Thomas (1965) and Palmer & Belshaw (1980).
tion of a tunnel are not in equilibrium with the modified Because of the large time-scale required for field-based
drainage boundary conditions. If the tunnel is not totally research, long-term data of ground movements are limited to
impermeable, a flow of pore water into the tunnel occurs, and a the ground surface, and long-term comprehensive data on
new steady-state flow condition is eventually reached. The final pore pressure changes around a tunnel lined with expanded
concrete segments are especially scarce. The evidence that
the ground surface continues to settle after tunnel excavation
Manuscript received 5 May 2006; revised manuscript accepted 16 has been illustrated by Peck (1969), O’Reilly et al. (1991),
October 2006.
Discussion on this paper closes on 1 July 2007, for further details
Lake et al. (1992), Bowers et al. (1996), Nyren (1998) and
see p. ii. Harris (2002). As most of the post-construction settlement
 data available are only for the ground surface, the mechan-
Geotechnical Consulting Group, London, UK.
y
Department of Engineering, University of Cambridge, Cambridge, ism of the long-term ground response, and in particular at
UK. subsurface level, is not yet well understood.

75
Delivered by ICEVirtualLibrary.com to:
IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
76 WONGSAROJ, SOGA AND MAIR
This paper discusses the ground movements and pore to follow that same trend, based on field measurements made
pressure changes at St James’s Park greenfield instrumented at Westminster (Burland & Hancock, 1977). However, there
site, which was monitored during construction of twin is no pore pressure measurement made at depths greater
tunnels for the Jubilee Line Extension (JLE) route in Lon- than 27 m below the ground surface to confirm that it is
don (Burland et al., 2001). From the comprehensive mon- under-drained.
itoring data obtained from the site (Nyren, 1998), a
mechanism of long-term subsurface ground movements at
this site is proposed. The monitoring data of subsurface OBSERVED LONG-TERM GROUND RESPONSE
ground deformations and measurements of pore pressure in MECHANISM
the vicinity of the tunnel have provided invaluable informa- The pore pressures around the westbound tunnel, meas-
tion to improve the understanding of long-term ground ured at the end of the excavation and then approximately
response mechanism induced by open-face tunnelling in 240 days later (i.e. start and end of period 3), are shown in
London Clay, particularly for those tunnels supported by Fig. 4(a). Those around the eastbound tunnel, measured at
expanded tunnel linings. the end of the excavation and then approximately one year
Because of the complexity of the problem, which involves later (i.e. start and end of period 5), are shown in Figs 4(b)
soil–tunnel–pore fluid flow interactions, three-dimensional and 4(c). Figs 4(a) and 4(b) show that, after the tunnel
(3D) finite element analysis (FEA) of the first excavation of excavations, the pore pressures above the crown of both
the westbound tunnel and subsequent consolidation prior to tunnels rose initially but subsequently stabilised. Neverthe-
the second excavation of the eastbound tunnel under St less, they were still slightly smaller than their pre-construc-
James’s Park was performed to assess the validity of the tion values. This indicates that the tunnels introduced new
proposed mechanism. An advanced anisotropic soil model, drainage boundaries into the soil, which was more evident
for which parameters were calibrated for both undrained and by the reduction in pore pressures at the springline level of
drained behaviour of London Clay, was used for the analysis. the eastbound tunnel one year after the tunnel excavation
The inclusion of stiffness anisotropy in the soil model was (see Fig. 4(c)).
needed to obtain the pattern of ground movements, the Although the tunnels are acting as drains, it seems that
excess pore pressure development and the tunnel lining the inward flow of the pore fluid towards the tunnel is larger
response similar to those observed in the field. The effects at the side of the tunnel than at the crown. This is confirmed
of two parameters on the computed subsurface long-term by the in-tunnel observations by Nyren (1998) that there are
ground response prior to the eastbound excavation were damp and wet patches on both sides of the tunnel. Some
examined. These parameters were permeability variation be- patches are large, and extend from track bed/key level
tween different London Clay units and lining–soil relative upward around the ring to the springline level (see Fig. 2).
permeability. These damp and wet patches below the tunnel springline
were reported for both the west and eastbound tunnels,
particularly nearby the instrumented section. The observation
ST JAMES’S PARK SITE of these patches below the springline and none above
Detailed descriptions of the St James’s Park instrumented correlates well with the measured reduction of pore pres-
site and the tunnel excavations beneath are given in Nyren sures around the springline (Fig. 4(c)) and the rise to almost
(1998); a brief summary is given in this section. St James’s pre-construction values above the crown (Figs 4(a) and
Park is one of the greenfield instrumented sites along the 4(b)). This suggests that the tunnel lining is likely to be
JLE route and is situated between Westminster and Green more permeable below the springline than above.
Park Stations. The alignment of the tunnels and locations of The measurements of subsurface vertical movements after
the instruments in plan and elevation views, together with the westbound tunnel excavation during the rest period
the soil profile interpreted from boreholes near the instru- (Table 1) are shown in Fig. 5(a). The data obtained from
mented section, are shown in Fig. 1. The twin tunnels are extensometer Bx, which is located directly above the west-
referred to as the westbound and the eastbound, with their bound tunnel centreline, indicates that there is a larger
axes located at approximately 31 m and 20.5 m below the increase in displacement at 27 m below the ground surface
ground surface respectively. The tunnels were excavated than that at 22.5 m. This increase in the relative displace-
using a 4.2 m long open-face tunnel boring machine (TBM) ment between these two points suggests that the soil between
with a backhoe excavator that has a 1.9 m reach when fully these two points is swelling. This swelling corresponds well
extended. They are supported by a 200 mm thick expanded to the rise in pore pressure above the tunnel centreline
precast concrete lining, which was erected near the end of during this period (see Fig. 4(a)).
trailing fingers, with a standard assembly to achieve an Although there is no pore pressure measurement at the
external diameter of 4.85 m (see Fig. 2). Measurement of springline level, the wet patches on both sides of the tunnel
ground response was divided into five distinct periods below this level indicate that there must be a reduction in
(Nyren, 1998), as summarised in Table 1. pore pressure, leading to consolidation of the soil on the
Two aquifers exist in the London Basin: (a) a deep aquifer sides of the tunnel. The amount of consolidation decreases
comprising Thanet Sand, Chalk and Basal Sand, overlain by with increasing offset from the tunnel centreline, as indi-
the Lambeth Group (formerly known as the Woolwich and cated by the data from extensometers Ax and Cx to Hx in
Reading Beds), which in many areas is overlain by London Fig. 5(a). These extensometer measurements also show that
Clay; and (b) a perched water table in Terrace Gravel on top the consolidation occurs only within the tunnel horizon in
of either London Clay or the clay of the Lambeth Group. the region close to the side of the tunnel.
This upper aquifer is recharged from surface precipitation Above this consolidation zone (i.e. at a depth less than
and locally from the River Thames. The observations from 20–25 m below the ground surface in this case), the soil
the piezometers installed in London Clay at different depths was settling as a rigid body. This was also noted by
in the Westminster area indicate that the pore pressures at Dimmock (2003). Based on this interpretation, a mechanism
the lower part of London Clay are slightly below hydrostatic, for the long-term ground movement at the westbound tunnel
as shown in Fig. 3 (Nyren, 1998). Hence there is usually can be split into three zones—a swelling zone, a consolida-
under-drainage at depths greater than 30 m below the ground tion zone and a rigid body movement zone—as schemati-
surface. The data for the St James’s Park site are also likely cally illustrated in Fig. 5(b).

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
LONG-TERM GROUND RESPONSE TO TUNNELLING UNDER ST JAMES’S PARK 77

Legend
Tree canopy
To Green Park
Eastbound outline Shallow surface
running tunnel monitoring point (SMP)
(4·85 m diameter)
Electrolevel inclinometer

Rod extensometer

Pneumatic piezometer

Combined pneumatic piezometer/


A B C D E F G H earth pressure cell

2m
2m 2m 4m 4m
2m I J K

2m
8m 6m
2m
8m
2m
4m 6m 6m 5·5 m 4·5 m 6m
4m 1 2 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
3
ST1
ST1

~2·5 m ~21·5 m

N
Westbound 5m 5m To Westminster
running tunnel
(4·85 m diameter)

(a)

0 A B C D E F G H I J K

Made Ground/alluvium
5
Terrace Gravels
10
FP1
Legend
London Clay
FP2 Rod extensometer
15
FP3 Electrolevel inclinometer
20 SP1/
SP4/
Piezometer
SP3/ SP2/ SC1
SC4 SC3 SC2
BP1
25 Combined pneumatic
BP2 spade cell/piezometers

30 Eastbound
running tunnel 5m 5m
(4·85 m OD)
35
Composite borehole log
from JLE BH109 and instrument 40
installation records
Westbound
Woolwich and running tunnel
Reading Beds Clay (4·85 m OD)

(b)

Fig. 1. (a) Plan and (b) elevation view of instrumented section at St James’s Park (after Nyren, 1998)

The extensometer measurements recorded after the east- the soil above the westbound tunnel appears to be consoli-
bound tunnel excavation are presented in Fig. 6(a). A ground dating (see extensometer Bx in Fig. 6(a)), which is probably
movement mechanism similar to that in the rest period was caused by the flow of pore fluid into the eastbound tunnel.
observed: slight swelling above the tunnel crown, which This leads to a mechanism of swelling, consolidation and
corresponds well with the rise in pore pressures above the rigid body movement zones, as depicted in Fig. 6(b).
tunnel, and consolidation at the sides of the tunnel due to In summary, the mechanism of long-term ground move-
the decrease in pore pressures, as shown in Figs 4(b) and ments at St James’s Park, as interpreted from the field data,
4(c) respectively. After the eastbound tunnel construction, is that the soil swells in a concentrated zone above the

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
78 WONGSAROJ, SOGA AND MAIR

D
33·6°

O
mm
4850
200 mm

A A⬘

T1 T2
Observed zone
of damp and 12·0° Observed zone
wet patches of damp and
Key Key wet patches
segment segment
T2 T1

Section A–A⬘
Leading edge

1000 mm
T1 T2 T1 T2

Key details
800 mm

400 mm

B B⬘

Section B–B⬘
Key type Dimension L
L Approximate scale
A 507 mm
B 487 mm 800 mm

Fig. 2. General arrangement of expanded tunnel lining used under St James’s Park (after
Nyren, 1998)

Table 1. Summary of monitoring period timing at St James’s Park (Nyren, 1998)

Period Time span Duration: days Descriptions

1 1 Mar 95 to 27 Apr 95 28.4 Period prior to any tunnelling works


2 27 Apr 95 to 28 Apr 95 1.67 Construction of westbound tunnel drive
3 28 Apr 95 to 8 Jan 96 256 Rest period before eastbound excavation
4 8 Jan 96 to 10 Jan 96 4.6 Construction of eastbound tunnel drive
5 10 Jan 96 to 21 Mar 97 432 Long-term monitoring period

Note that the period is divided according to the survey number of precision levelling surveys (Nyren, 1998, table F1): period 1, survey nos
1–16; period 2, survey nos 16–29; period 3, survey nos 29–62; period 4, survey nos 62–85; period 5, survey nos 85–101.

tunnel crown and simultaneously consolidates on either sides Table 1) to examine the time-dependent behaviour of a
of the tunnel in the zone close to tunnel axis level. Above single tunnel construction. This was necessary to avoid any
these zones, the soil was settling as a rigid body, which complexity due to interactions of the two tunnels, so that a
resulted in continued settlement at the ground surface. time-dependent ground deformation mechanism can be pro-
posed based on a simple boundary condition. As future
research, the numerical study can be extended to include the
FINITE ELEMENT ANALYSIS effects of tunnel–tunnel interaction on the long-term ground
The construction of the westbound tunnel was modelled, deformation behaviour.
and the subsequent ground response prior to the eastbound
excavation was examined. All analyses presented in this
paper were performed using the soil-fluid coupled 3D finite Soil conditions and geometry
element (FE) method with the commercial finite element The FE model for the westbound tunnel excavation with
package ABAQUSTM . For brevity, this numerical study dimensions shown in Fig. 7 was assumed to have the
utilises the data mainly obtained during Period 1 to 3 (see following soil profile:

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
LONG-TERM GROUND RESPONSE TO TUNNELLING UNDER ST JAMES’S PARK 79
Piezometric pressure: kPa Pore pressure: kPa
0 100 200 300 0 50 100 150 200 250
0 0

⫺5 London Clay/gravel
interface
Top of London Clay ⫺10
⫺10

Depth: m
St James’s Park ⫺15

Westminster field data


⫺20
⫺20 (Higgins et al., 1996)

⫺25
z: m

⫺30
⫺30 WB
Initial pore pressures
⫺35 240 days after excavation
Hydrostatic One year after excavation
Bottom of London Clay profile
⫺40 (a)
(estimated) Under-drained
Lambeth Group profile Pore pressure (kPa)
0 50 100 150 200 250
0
⫺50

⫺5 London Clay/Gravel
Fig. 3. Measured pore pressures at instrumented section in St interface
James’s Park and in Westminster area, London (Nyren, 1998)
⫺10
Depth: m

Initial pore pressures


End of excavation
(a) Made Ground and Alluvium, MG (0–5 m) One year after excavation
(b) Terrace Gravel, TG (5–8 m) ⫺15
(c) London Clay, LC (8–43 m)
(d ) Woolwich and Reading Bed Clay, WRBC (43–50 m).
⫺20
EB No recovery of pore pressure because the
It was assumed that the interface between WRB clay and sand pocket of this piezometer was clipped
by the tunnelling shield during excavation
underlying Basal Sand was located at a depth of 50 m ⫺25
(Addenbrooke, 1996; Higgins et al., 1996). (b)
To reduce analysis time, the soil was modelled with two
types of element: Initial pore pressures
End of excavation
(a) 8-node trilinear displacement and pore pressure ele- 200
One year after excavation
ments
Pore pressure: kPa

(b) 20-node triquadratic displacement, trilinear pore pres- 150


sure, reduced integration elements.
100
The 20-node elements were employed in the zone around
the tunnel because they gave smoother pore pressure re-
50
sponse than the 8-node elements. Tunnel lining was mod-
elled with 8-node, doubly curved thick shell elements with EB
reduced integration. The lining and soil elements shared the 0
15 10 5 0
same nodes at the tunnel boundary; the interface was not Offset from centreline: m
modelled. The 3D model consisted of 9625 elements and
(c)
22 938 nodes. The model boundaries were set to minimise
the effects on ground response (Wongsaroj, 2005). Fig. 4. Variation in pore pressures (after Nyren, 1998): (a)
above westbound tunnel crown; (b) above eastbound tunnel
crown; (c) at eastbound tunnel springline level
Modelling of tunnel construction and long-term
consolidation
The model had element length (size of an element in the
This excavation model resulted in 100% reduction of normal
longitudinal direction) of 2 m in the zone where the excava-
stress to the tunnel boundary within the 6 m unsupported
tion was modelled. When the tip of the backhoe is fully
length (including the tunnel face), which leads to stress
extended, the distance from the tunnel face to the shield tail
redistribution and volume loss. This tunnel excavation model
is approximately 6 m during the excavation. Assuming that
sequence was continued until the tunnel face reached 100 m
there is no contact between most of the perimeter of the
into the model. The computed results were taken from a
shield and the surrounding soil (due to over-excavation), the
plane located 40 m from the start of the excavation, which is
excavation process is modelled as a repeated sequence of
referred to as the instrument plane (see Fig. 7). It was
(a) deactivation of elements every 2 m, resulting in 6 m important to model the tunnel construction process in 3D so
unsupported opening, and that realistic short-term stress and pore pressure fields,
(b) placement of 2 m long tunnel linings next to the together with tunnel lining loads, were obtained as the initial
existing tunnel linings (see Fig. 8). condition for the long-term consolidation analysis.

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
80 WONGSAROJ, SOGA AND MAIR

Vertical displacement: mm 0 50 100 150 200 250 300


(⫹ve heave /⫺ve settlement) Days after westbound tunnel construction

Ax Bx Cx Dx Ex Fx Gx Hx
5 0 ⫺5 ⫺10 ⫺15 5 0 ⫺5 ⫺10 ⫺15 5 0 ⫺5 ⫺10 ⫺15 5 0 ⫺5 ⫺10 ⫺15 5 0 ⫺5 ⫺10 ⫺15 5 0 ⫺5 ⫺10 ⫺15 5 0 ⫺5 ⫺10 ⫺15 5 0 ⫺5 ⫺10 ⫺15

⫺5 ⫺5 ⫺5 ⫺5 ⫺5 ⫺5 ⫺5 ⫺5

⫺10 ⫺10 ⫺10 ⫺10 ⫺10 ⫺10 ⫺10 ⫺10

⫺15 ⫺15 ⫺15 ⫺15 ⫺15 ⫺15 ⫺15 ⫺15


Depth: m

⫺20 ⫺20 ⫺20 ⫺20 ⫺20 ⫺20 ⫺20 ⫺20

⫺25 ⫺25 ⫺25 ⫺25 ⫺25 ⫺25 ⫺25 ⫺25

⫺30 ⫺30 ⫺30 ⫺30 ⫺30 ⫺30 ⫺30 ⫺30

⫺35 ⫺35 ⫺35 ⫺35 ⫺35 ⫺35 ⫺35 ⫺35

⫺40 ⫺40 ⫺40 ⫺40 ⫺40 ⫺40 ⫺40 ⫺40


4m 0m 4m 10 m 16 m 21·5 m 25·5 m 31 m
⫺45 ⫺45 ⫺45 ⫺45 ⫺45 ⫺45 ⫺45 ⫺45

After Nyren (1998)


(a)

Offset from westbound


tunnel centreline is Offset from westbound tunnel centreline: m
indicated in the box ⫺5 0 5 10 15 20 25
0

London Clay/Gravel interface ⫺5


Region where ⫺10
swelling occurs

Depth: m
Rigid body movement ⫺15

Region where ⫺20

consolidation ⫺25
occurs ⫺30
⫺35
⫺40

(b)

Fig. 5. (a) Field measurement of long-term ground movement after westbound tunnel construction; (b) schematic diagram
indicating zones of different response during consolidation

0 100 200 300 400 500


Vertical displacement (mm)
(⫹ve heave/⫺ve settlement) Days after eastbound tunnel construction

Ax Bx Cx Dx Ex Fx Gx Hx
0 ⫺10 ⫺20 ⫺30 0 ⫺10 ⫺20 ⫺30 0 ⫺10 ⫺20 ⫺30 0 ⫺10 ⫺20 ⫺30 0 ⫺10 ⫺20 ⫺30 0 ⫺10 ⫺20 ⫺30 0 ⫺10 ⫺20 ⫺30 0 ⫺10 ⫺20 ⫺30

⫺5 ⫺5 ⫺5 ⫺5 ⫺5 ⫺5 ⫺5 ⫺5

⫺10 ⫺10 ⫺10 ⫺10 ⫺10 ⫺10 ⫺10 ⫺10

⫺15 ⫺15 ⫺15 ⫺15 ⫺15 ⫺15 ⫺15 ⫺15


Depth: m

⫺20 ⫺20 ⫺20 ⫺20 ⫺20 ⫺20 ⫺20 ⫺20

⫺25 ⫺25 ⫺25 ⫺25 ⫺25 ⫺25 ⫺25 ⫺25

⫺30 ⫺30 ⫺30 ⫺30 ⫺30 ⫺30 ⫺30 ⫺30

⫺35 ⫺35 ⫺35 ⫺35 ⫺35 ⫺35 ⫺35 ⫺35

⫺40 ⫺40 ⫺40 ⫺40 ⫺40 ⫺40 ⫺40 ⫺40

⫺45 ⫺45 ⫺45 ⫺45 ⫺45 ⫺45 ⫺45 ⫺45


4m 0m 4m 10 m 16 m 21·5 m 25·5 m 31 m

After Nyren (1998)


(a)

Offset from westbound


tunnel centreline is Offset from westbound tunnel centreline: m
indicated in the box ⫺5 0 5 10 15 20 25
0
London Clay/Gravel interface ⫺5

Region where Consolidation caused ⫺10


Rigid body movement
Depth: m

swelling occurs by drainage into


⫺15
eastbound tunnel
⫺20
Region where
consolidation ⫺25
occurs ⫺30
⫺35
Consolidation caused
⫺40
by drainage into
westbound tunnel

(b)

Fig. 6. (a) Field measurement of long-term ground movement after eastbound tunnel construction; (b) schematic diagram
indicating zones of different response during consolidation

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
LONG-TERM GROUND RESPONSE TO TUNNELLING UNDER ST JAMES’S PARK 81
q
Instrument plane q ⫽ Mp⬘
40 m
MG (p⬘*, q*)
TG p⬘ q
50 m ⫽( R R
, )
LC (p⬘, q)

WRBC p⬘
Subloading
surface
140 m
120 m
Region of 20-node
elements Normal yield surface

Fig. 7. Finite element model employed for analysis Fig. 9. Concept of subloading surface within normal yield
surface and ratio R

Unsupported length
during excavation

Fig. 8. Model of excavation sequence

The tunnel lining thickness of 200 mm, which is the Soil constitutive model
actual thickness of the tunnel lining, was used for all All soil units in this study were modelled using the non-
simulations (neglecting the effect of the joints). The tunnel linear elasto-plastic critical state soil model with the sub-
lining was modelled using a linear elastic model with a loading surface concept (Hashiguchi & Ueno, 1977;
Young’s modulus of 28 GPa and a Poisson’s ratio of 0.15. Hashiguchi & Chen, 1998). The concept assumes that there
The expansion of the tunnel lining was not modelled. The is a small yield surface called the subloading surface, which
placement of the tunnel lining was simply modelled by always passes through the current stress point within the
activating the shell elements at the deformed tunnel heading. conventional yield surface (also termed the normal yield
As the expansion of the lining was not modelled, the lining surface), and the ratio of their size is denoted by R, as
was not perfectly circular with the initial opening diameter shown in Fig. 9. The soil behaviour is elastic if the subload-
of approximately 4.85 m. ing surface is contracting (i.e. R is reducing), and is elasto-
When the tunnel face advanced 100 m into the model, the plastic if the subloading surface is expanding (i.e. R is
long-term consolidation was set to begin by assigning a increasing). The ratio R and its evolution law control the
drainage-only flow surface to the tunnel boundary along the amount of plastic strain generated within the normal yield
entire excavated tunnel length, and consolidation was al- surface. Generation of plastic strains within the normal yield
lowed until the new steady-state pore pressure regime was surface was necessary to model accurately the development
reached. The drainage-only flow surface controls the drai- of excess pore pressures during undrained shearing. The
nage condition at the tunnel boundary by assuming that the constitutive model also incorporates the following features:
pore fluid velocity in the direction outwards from the soil non-linear elasticity, anisotropic stiffness, the Matsuoka &
element, vn (m/s), is proportional to the pore pressure at the Nakai (1985) failure criterion, and a re-invoked small-strain
tunnel lining extrados, uw, when it is positive. The propor- stiffness at stress reversal points (SRP; i.e. unloading and
tional constant is termed the seepage coefficient, KT , and is reloading). Further details of the model are described in
equivalent to kl /tl ªw , where kl and tl are the permeability Appendix 1.
and thickness of the tunnel lining respectively. Additionally, The model was implemented in the ABAQUSTM user
no flow is allowed across the tunnel boundary when uw is interface. The model was adopted for all soil units in the
negative (i.e. water is not supplied from inside the tunnel). analyses. As the Made Ground is far from the tunnel, and
Hence varying the magnitude of KT (m3 /kN.s) is the same as the strains occurring within this material are small during
varying the permeability of the tunnel lining for a given tunnel excavation, its behaviour is more or less elastic. For
tunnel lining thickness, which results in different drainage the Terrace Gravel and Woolwich and Reading Beds clay,
conditions at the tunnel boundary. Damp patches on the the parameters were selected such that the stiffness degrada-
tunnel lining suggest that the pore pressure in the tunnel tion curves are within the bounds adopted by other studies
lining segment may be negative. However, the suction within such as Addenbrooke et al. (1996). The critical state lines
the tunnel lining is difficult to determine. Therefore atmo- for the Made Ground and Terrace Gravel were calculated
spheric pressure was assumed inside the tunnel lining for all from the angles of friction given in Potts & Zdravkovic
analyses presented in this paper. (2001). The parameters for London Clay were determined

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
82 WONGSAROJ, SOGA AND MAIR
Assumed preconsolidation pressure: kPa
from data of undrained and/or drained triaxial tests and one-
0 500 1000 1500 2000 2500 3000 3500 4000
dimensional consolidation tests. In particular, emphasis was 0
on calibrating the anisotropic model parameters so that the Made Ground
same model parameters can simulate data of both undrained 5
Terrace Gravel
and drained triaxial compression tests of vertically and 1080 kPa 3000 kPa
10
horizontally cut London Clay specimens (Yimsiri, 2001;

Depth below ground surface: m


Wongsaroj et al., 2004). The soil properties and model 15
parameters used for the simulations are summarised in Table Eastbound tunnel axis level
2 and Table 4 in Appendix 1. In most cases, the earth 20
pressure coefficient at rest, K0 , for London Clay was 25 London Clay
assumed to be 1.2. The preconsolidation profile adopted is
Westbound tunnel axis level
shown in Fig. 10. To investigate the effects of stiffness 30
anisotropy, simulations using a model with isotropic elastic
35
stiffness were also performed.
40
3350 kPa

Permeability model 45 Woolwich and Reading Bed Clay


To examine the influence on the long-term ground re- 50
sponse of different units within London Clay and of the
tunnel drainage condition, two combinations of permeability Fig. 10 Preconsolidation pressure profile assumed for analysis
profiles and tunnel lining drainage conditions were adopted.
The two permeability profiles assumed are summarised in
Table 3, which are also shown in Fig. 11 along with the in surface. The tunnel lining was assumed to have uniform
situ measurement data from Burland & Hancock (1977). permeability.
Although other permeability profiles were examined by An under-drained initial pore pressure profile was derived
Wongsaroj (2005), these two profiles were the extreme cases: for the analyses that adopted permeability profile 2. Four
profile 1 is a ‘simplified’ permeability model case, whereas distinct divisions (B, A3ii, A3i and A2) within the London
profile 2 is a more realistic ‘refined’ permeability model Clay strata, established by King (1981), were identified using
case. the detailed descriptions of split continuous U100 samples
For permeability profile 1, the London Clay has a constant (Standing & Burland, 2006). The qualitative hierarchies of
isotropic permeability of 109 m/s, and for these analyses permeability suggested by Hight et al. (2003) and Standing
the initial pore pressure (prior to tunnelling) was assumed & Burland (2006) are kA2 . kA3ii(top) . kA3i  kA3ii(base) .
hydrostatic with the water table at 5 m below the ground kB . Different anisotropic permeability values were assigned

Table 2. Assumed properties for all soil units

Strata Bulk density, Critical angle of shearing Coefficient of earth Permeability,


ª: kN/m3 resistance, 9cv : degrees pressure at rest, K0 k: m/s

Made Ground 20.0 25.0 0.6 1 3 107


Terrace Gravel 20.0 35.0 0.4 5 3 104

London Clay 20.0 21.0 1.2 See Table 3
Woolwich and Reading Beds Clay 20.0 27.0 1.2 See Table 3

*Only adopted for the simulations to investigate influence of permeability profiles and tunnel drainage conditions.

Table 3. Permeability profiles assumed for analyses

Soil descriptions by Standing & Burland (2006) Permeability profile 1: m/s Permeability profile 2: m/s

kv kh kv kh

London Clay Division B, 8–20.5 m (95–82.5 mPD) 5 3 1010 10 3 1010

London Clay Division A3ii, 20.5–28 m (82.5–75 mPD) 7 3 1010 45 3 1010


to to
9 3 1011 45 3 1011
1 3 109
London Clay Division A3i, 28–35 m (75–68 mPD) 9 3 1011 18 3 1011
to to
3 3 1010 6 3 1010

London Clay Division A2, 35–43m (68–60mPD) 3 3 1010 6 3 1010

Woolwich and Reading Beds Clay 5 3 1011 1 3 1012 2 3 1012

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
LONG-TERM GROUND RESPONSE TO TUNNELLING UNDER ST JAMES’S PARK 83
Permeability (m/s)
and measurements could be further improved by reducing
1 ⫻ 10⫺12 1 ⫻ 10⫺10 1 ⫻ 10⫺8 1 ⫻ 10⫺6 the value of K0 to 1.2 (Wongsaroj, 2005).
⫺13 ⫺11 ⫺9 ⫺7
1 ⫻ 10 1 ⫻ 10 1 ⫻ 10 1 ⫻ 10 The excess pore pressure contours around the tunnel
8 computed with the isotropic and anisotropic soil models are
Profile 1, kh ⫽ kv
Depth below ground surface: m

13 Profile 2, kh LC Div B shown in Fig. 13. The zone of negative excess pore
18
Profile 2, kv pressures with the isotropic model has contours that spread
Burland & Hancock (1977) in the horizontal direction, whereas those computed with the
23 LC Div A3ii anisotropic soil model give contours that spread more in the
28 vertical direction. The computed effective stress paths at
LC Div A3i
33 the tunnel crown and springline closely followed that of an
38
undrained triaxial compression test on a horizontally and
LC Div A2
vertically cut sample respectively (Yimsiri, 2001; Hight et
43 al., 2003). That is, the soil adjacent to the tunnel crown
48 WRB clay extrados experiences a dilation tendency whereas the soil
adjacent to the tunnel springline extrados experiences a
Fig. 11. Assumed permeability profiles tendency to contraction prior to dilation. This initial differ-
ence is due to the initial elastic stiffness anisotropy (e.g.
for these layers based on field measurement data reported in Graham & Houlsby, 1983) before the plastic dilation devel-
Burland & Hancock (1977) and Hight et al. (2003). The ops at larger strains. The isotropic model gave similar stress
assumed anisotropic permeability ratio (kh /kv ) was 2, except paths above the tunnel crown and near the springline: both
for division A3ii, for which the assumed kh /kv ratio was 5. displayed some contraction tendency followed by dilation,
This higher ratio was assigned because water strikes were defined by the isotropic soil model. Soil elements adjacent
encountered in this division (Standing & Burland, 2006). As to the tunnel springline extrados experienced more shearing
discussed before, the tunnel lining above the springline was and greater negative excess pore pressure than soil elements
assumed to have finite permeability whereas the lining below adjacent to the tunnel crown extrados; this was due to the
the springline was assumed to be fully permeable, which is initial K0 (.1) condition.
considered to be reasonably realistic based on the in-tunnel The computed pore pressure profile above the tunnel
observations. The under-drained initial pore water pressure crown is plotted in Fig. 14. The figure shows that, at a depth
profile derived from permeability profile 2 is shown in of 27 m, the simulation with the anisotropic model gave a
Fig. 12. larger reduction in pore pressure than that computed with
the isotropic soil model. The computed pore pressures above
the tunnel crown remain positive at most depths apart from
the zone very close to the tunnel crown. Based on Fig. 14,
IMPORTANCE OF SOIL STIFFNESS ANISOTROPY FOR which includes field measurements of pore pressure, and the
LONG-TERM GROUND RESPONSE development of excess pore pressures during tunnel excava-
It has been reported widely that stiffness anisotropy has tion (not shown), the trend of change in pore pressures
an important influence on predicted short-term ground move- computed with the anisotropic soil model agrees more
ments caused by tunnel excavation (e.g. Addenbrooke, 1996; closely with the field data than that computed with the
Simpson et al., 1996; Lee & Ng, 2002; Franzius, 2003). isotropic model.
Similar findings were made in this study for both surface When the tunnel lining is acting as a partially permeable
and subsurface ground movements, the use of stiffness boundary, the pattern of excess pore pressures generated
anisotropy providing a better match to the field data on around the tunnel with the anisotropic model led to larger
short-term ground movements (Wongsaroj et al., 2004). The consolidation settlement above the tunnel centreline, causing
influence of stiffness anisotropy on long-term ground re- the tunnel lining to squat in the long term. This is indicated
sponse is presented in this paper, based on results of by the increase in the horizontal diameter, as shown in
analyses assuming K0 ¼ 1.5 for London Clay. Further Fig. 15.
investigation showed that the agreement between predictions The direct comparison between this prediction and field
measurements cannot be made owing to the lack of field
data for tunnel lining deformation during the rest period.
Pore pressure: kPa
0 50 100 150 200 250 300 350
0 Isotropic Anisotropic
⫺10
⫺5
0
0
Depth below ground surface: m

⫺10
⫺20
⫺15 ⫺40

⫺20
Depth: m

⫺40
⫺30
⫺25

⫺30 St James's Park


⫺40
⫺35 Westminster field data
(Higgins et al., 1996) Contours at
⫺40 Computed pore pressure 20 kPa intervals
profile ⫺50
⫺45
⫺30 ⫺20 ⫺10 0 10 20 30
⫺50 Offset from tunnel centreline: m

Fig. 12. Pore pressure profile derived from permeability profile Fig. 13. Excess pore pressure contours around tunnel computed
2 (see Fig. 11) with isotropic and anisotropic soil models in 3D

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
84 WONGSAROJ, SOGA AND MAIR
Pore pressure: kPa Pore pressure: kPa
⫺200 ⫺100 0 100 200 300 0
0 ⫺200 ⫺100 0 100 200
⫺5 Consolidation
⫺10 time

Depth: m
⫺5 0 days
⫺15
100 days
Depth below ground surface: m

⫺20 Steady state


⫺10
⫺25
Pre-construction profile
Hydrostatic ⫺30
⫺15 Distance from springline extrados: m
3D isotropic
0 10 20 30
3D anisotropic 300

Pore pressure: kPa


Field data ⫺20 200
100
⫺25 0
⫺100
⫺30 ⫺200
(a)
⫺35
Pore pressure: kPa
Fig. 14. Computed short-term pore pressure profile above west- 0
bound tunnel crown ⫺200 ⫺100 0 100 200
⫺5 Consolidation
⫺10 time

Depth: m
⫺15 0 days
100 days
However, the trend of squatting with time has been observed ⫺20 Steady state
in other tunnels in London Clay (e.g. Ward & Thomas, ⫺25
1965; Nyren, 1998 (during period 5); Dimmock, 2003). The ⫺30 Pre-construction profile
analysis with the isotropic model gave the opposite trend: Distance from springline extrados: m
0 10 20 30
that is, decreasing horizontal diameter with time, as shown 300

Pore pressure: kPa


in Fig. 15. 200
The computed pore pressure profiles immediately after, 100
and 100 days after, the construction, and at the steady-state
0
long term are plotted in Fig. 16; these are shown above the
tunnel centreline and at the springline level. With the ⫺100
anisotropic model (Fig. 16(a)), there is a rise in pore ⫺200
pressure in the zone up to 3 m from the springline extrados. (b)
From this point the pore pressure had to decrease to reach
the long-term steady state. Immediately above the tunnel Fig. 16. Influence of soil stiffness anisotropy on long-term pore
crown there is a recovery of pore pressure from the tunnel pressure profiles: (a) anisotropic soil model; (b) isotropic soil
crown extrados to approximately 3 m above it. In the soil model
within 3 m from the crown extrados there is a small reduc-
tion in pore pressure over the first 100 days of consolidation,
and then there is a rise in the pore pressure to the steady larger at the tunnel springline than at the crown, as reported
state. Above the tunnel crown the soil swelled, whereas at for tunnels in London Clay (Barratt et al., 1994; Dimmock,
the sides and below the tunnel the soil consolidated. The 2003).
volume of consolidation was greater than that of swelling. With the isotropic model there is a substantial increase in
This led to overall consolidation, which in turn decreased pore pressure during the consolidation in the zone from the
the thickness of the soil on both sides of the tunnel. Hence springline extrados to 8 m away from it, as shown in Fig.
the tunnel lining squatted in the long term. This computed 16(b). The majority of the soil above the tunnel crown
long-term deformation agrees with the field measurements underwent consolidation due to the reduction in pore pres-
and coincides well with the computed hoop load, which is sure, except for the soil from the tunnel crown extrados to
approximately 2 m above it, which experienced a rise in
pore pressure. The rise in pore pressure at the springline and
8 the majority of the reduction in the pore pressure above the
tunnel crown caused the soil to swell at the springline and
Change in horizontal diameter: mm

Isotropic soil model


6 consolidate above the crown respectively. This in turn causes
Anisotropic soil model
the tunnel lining to elongate in the vertical direction.
4
In summary, stiffness anisotropy is important not only to
2 evaluate better short-term ground movements, but also to
predict long-term ground and tunnel lining movements in
0 London Clay.
0 100 200 300 400 500 600 700 800 900 1000
⫺2

⫺4
IMPORTANCE OF PERMEABILITY PROFILE AND
TUNNEL DRAINAGE CONDITIONS FOR LONG-TERM
⫺6 GROUND RESPONSE
Days after excavation
Simplified permeability model case
Fig. 15. Influence of stiffness anisotropy on tunnel lining The drainage condition of the tunnel boundary is often
deformation following tunnel excavation (increase of horizontal assumed to be influenced by the relative magnitude of the
diameter shown positive) tunnel lining permeability and the surrounding soil per-

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
LONG-TERM GROUND RESPONSE TO TUNNELLING UNDER ST JAMES’S PARK 85
200
meability. In order to reasonably simulate the ground move- 180
ments during the long-term consolidation in the field, the 160

Pore pressure: kPa


computed changes in pore pressure should have similar 140
values to those observed in the field. 120
The permeability of the expanded concrete tunnel lining 100
KT ⫽ 1 ⫻ 10⫺14
was evaluated by carrying out simulations of the consolida- 80
KT ⫽ 1 ⫻ 10⫺13
tion around the westbound tunnel with different KT values 60
KT ⫽ 1 ⫻ 10⫺12
(1 3 1014 , 1 3 1013 , 1 3 1012 and 1 3 1011 m3 =kN:s). 40
KT ⫽ 1 ⫻ 10⫺11
For the first 300 days of consolidation (i.e. before the 20
Field data
0
passage of the eastbound tunnel excavation beneath the 0 50 100 150 200 250 300
instrumented site), the computed pore pressures at piezo- Days after excavation
meters BP1 and BP2 (see Fig. 1), which are located above (a)
the westbound tunnel crown extrados, are shown in Fig. 17 250
along with the field measurements. The computed pore
pressures at these points show an increase with time for all 200

Pore pressure: kPa


values of KT . The computed response at BP2, where it is
likely to be most affected by the lining-soil relative per- 150
KT ⫽ 1 ⫻ 10⫺14
meability, was closest to the field measurement when the KT ⫽ 1 ⫻ 10⫺13
100
value of KT was 1 3 1011 m3 /kN.s. Non-zero pore pressures KT ⫽ 1 ⫻ 10⫺12
computed at the tunnel crown extrados presented in Fig. kl KT ⫽ 1 ⫻ 10⫺11
50 KT ⫽
17(c) suggest that the tunnel lining was acting as a partially tlγw Field data
permeable boundary. 0
The magnitude of the mass permeability of the tunnel 0 50 100 150 200 250 300
lining can be estimated from the definition of KT given Days after excavation
earlier. For a KT value of 1 3 1011 m3 /kN.s and a lining (b)
thickness of 0.2 m, the mass permeability of the tunnel 250
Hydrostatic
lining becomes 2 3 1011 m/s. This appears to be two orders value
200
Pore pressure: kPa

of magnitude greater than the value of maximum concrete


permeability of 1 3 1013 m/s assumed for the design of the 150
Jubilee Line Extension (Winterton, 1994). The larger per-
meability inferred from the simulation suggests that the 100
KT ⫽ 1 ⫻ 10⫺14
joints between concrete segments of an unbolted expanded
KT ⫽ 1 ⫻ 10⫺13
lining increase the mass permeability of the tunnel lining 50
KT ⫽ 1 ⫻ 10⫺12
substantially, particularly in the portion above the tunnel KT ⫽ 1 ⫻ 10⫺11
shoulder. This may also be true of the lower part, but pore 0
0 50 100 150 200 250 300
pressure measurement is required for such an evaluation. Days after excavation
Using the results from the analysis with the lining per- (c)
meability value evaluated above, the consolidation compo-
nent of vertical displacements was examined by comparing Fig. 17. Computed pore pressure response with simplified per-
the computed values and the field measurements at various meability model compared with field data: (a) BP1 at 24 m
locations of extensometers (Fig. 1). Fig. 18 shows the below ground surface; (b) BP2 at 27 m below ground surface;
computed subsurface vertical ground movements with depth (c) tunnel crown extrados
at various extensometer locations. Despite the close match-
ing of pore pressures near the tunnel (see Fig. 17(b)), the
computed response of vertical subsurface displacements at
extensometers Bx and Cx are quite different from the field response in comparison with the field data. However, field
measurements shown in Fig. 5(a). They show initial upward measurements indicate that consolidation appears to be more
displacements and subsequent downward movements above confined in the zone at the same level as the tunnel axis,
the tunnel springline. Overall, the soil around the tunnel was whereas the analysis seems to predict consolidation that is
predicted to swell significantly owing to the recovery of the more uniformly throughout the clay region. The confined
pore pressure. Although the tunnel is in a partially drained zone of the consolidation displacements that occurs mainly
condition, it is almost impermeable. Outside this swelling close to the tunnel horizon in the field suggests that there
zone, the soil consolidated. With the assumed permeability exist layers of larger permeability within this zone of the
of the soil being homogeneous, the consolidation is pre- London Clay.
dicted to occur more or less radially, which leads to the The simulation results described in this section indicate
consolidation of the London Clay even at a depth close to that modelling the London Clay as a uniform homogeneous
the clay/gravel interface (at 8 m depth). material with isotropic permeability with the tunnel lining as
This long-term ground movement mechanism is comple- a uniform drainage boundary was not able to simulate the
tely different from the mechanism proposed from the field subsurface ground response, especially close to the tunnel
data interpretation described earlier in the paper (see Fig. 5). boundary. A more realistic soil permeability profile and a
The predicted swelling should occur only in a localised zone more appropriate drainage condition for the tunnel lining are
above the tunnel, and consolidation should occur at the sides required in the analysis in order to predict the same mechan-
of the tunnel at the tunnel axis level. The soil above the ism of ground response as that interpreted from the field
swelling and consolidating zones (see Figs 5(b) and 6(b)) extensometer data. This is described in the next section.
settles in a rigid body manner, causing further settlement at
the ground surface.
In the zone at further distance from the tunnel centreline Refined permeability model case
(i.e. from the extensometer Dx outwards in Figs 18(c) and The in-tunnel observations by Nyren (1998) reported that
18(d)), the analysis gives a fairly accurate consolidation there were damp and wet patches only below the springline

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
86 WONGSAROJ, SOGA AND MAIR
Consolidation vertical displacement: mm Consolidation vertical displacement: mm
⫺15 ⫺10 ⫺5 0 5 10 ⫺15 ⫺10 ⫺5 0 5
0 0

⫺5 ⫺5

⫺10 ⫺10
Consolidation
time
⫺15 ⫺15
0 day

⫺20 ⫺20 6 days


27 days
Depth (m)

Depth (m)
⫺25 ⫺25 86 days
175 days
⫺30 ⫺30 215 days
274 days
⫺35 ⫺35

⫺40 ⫺40

⫺45 ⫺45

⫺50 ⫺50
(a) (b)

Consolidation vertical displacement: mm Consolidation vertical displacement: mm


⫺15 ⫺10 ⫺5 0 5 ⫺15 ⫺10 ⫺5 0 5
0 0

⫺5 ⫺5

⫺10 ⫺10
Consolidation
time
⫺15 ⫺15
0 day

⫺20 ⫺20 6 days


27 days
Depth (m)

Depth (m)

⫺25 ⫺25 86 days


175 days
⫺30 ⫺30 215 days
274 days
⫺35 ⫺35

⫺40 ⫺40

⫺45 ⫺45

⫺50 ⫺50
(c) (d)

Fig. 18. Computed subsurface ground vertical movement with simplified permeability model: (a) Bx along tunnel centreline; (b) Cx
4 m from tunnel centreline; (c) Dx 10 m from tunnel centreline; (d) Ex 16 m from tunnel centreline

level. This suggests that the tunnel is significantly more able boundary at the top half and permeable at the bottom
permeable below the springline level: this may be asso- half. Fig. 19(a) shows the distribution of pore pressure
ciated with the key segments being smaller than the other around the tunnel extrados when KT of the upper part was
types (see Fig. 2), because there were holes at the key 1 3 1012 m3 /kN.s. The partial recovery of pore pressure is
locations of the tunnel. Although these holes were filled computed above the tunnel springline, whereas nominally
with in situ concrete, their permeability would probably be zero pore pressure is computed at the bottom half of the
much higher than that of the precast concrete segments tunnel lining extrados.
manufactured under controlled conditions. For demonstra- Figures 19(b) and 19(c) show the recovery of pore pres-
tion purposes, the lower part of the tunnel below the sures with time above the tunnel crown for a range of KT
springline was modelled as a permeable boundary in this values. At BP1, which is approximately 4.5 m above the
analysis. To ensure the permeable condition at the tunnel tunnel crown extrados, the recovery of pore pressure seems
below the springline, a large value of KT relative to the soil to be less dependent on the relative lining–soil permeability.
permeability of 1 3 105 m3 /kN.s was assigned to this This is because the magnitude of pore pressure depends
region. As a result, the tunnel can still act as an imperme- mainly on the horizontal flow of pore fluid due to the

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
LONG-TERM GROUND RESPONSE TO TUNNELLING UNDER ST JAMES’S PARK 87
150
End of construction this division to recover to a higher magnitude than in the
Pore pressure: kPa

100 263 days simplified model case. This is clearly shown by the larger
Steady state ⫺90°
pore pressure at BP1 computed with the permeability profile
50
2 (Fig. 19(b)) than that with profile 1 (Fig. 17(a)).
0 0° The findings from the results of different permeability
models emphasise that appropriate tunnel drainage condi-
⫺50 tions should be assigned for a long-term analysis. The
90°
⫺100 drainage conditions should take into account the presence of
⫺100 ⫺50 0 50 100 joints, gaps and the relatively high-permeability region in
Position from springline: degrees the tunnel lining, although how to account for joints and
(a) gaps is far from straightforward. Furthermore, the rate of the
consolidation settlement depends principally on the per-
180 meability of the soil at the tunnel axis level, because the
160 majority of the consolidation takes place on either side of
Pore pressure: kPa

140 the tunnel. This is important, particularly for the presented


120
100 case where the London Clay consists of various sub-geologi-
KT ⫽ 1 ⫻ 10⫺11
80
KT ⫽ 1 ⫻ 10⫺12
cal units (King, 1981; Hight et al., 2003).
60
40 k KT ⫽ 1 ⫻ 10⫺13
KT ⫽ t γl KT ⫽ 1 ⫻ 10⫺14
20 l w
0 Field data
0 50 100 150 200 250
Days after excavation
(b)
CONCLUSIONS
This study has investigated the long-term ground move-
180 ments following a tunnel excavated in overconsolidated
160
Pore pressure: kPa

London Clay and lined with an expanded concrete lining.


140
120 Based on the comprehensive set of field monitoring data
100
KT ⫽ 1 ⫻ 10⫺11
obtained from St James’s Park, it is found that the mechan-
80 ism of long-term subsurface movements consists of a combi-
KT ⫽ 1 ⫻ 10⫺12
60
40 KT ⫽ 1 ⫻ 10⫺13 nation of swelling, consolidation, and rigid body movement.
20 KT ⫽ 1 ⫻ 10⫺14 The swelling takes place in a small zone above the tunnel
0 Field data
crown. The consolidation takes place on both sides of the
0 50 100 150 200 250
Days after excavation
tunnel, extending to a large offset from the tunnel centreline.
(c) Above the swelling and consolidation zones, the soil settles
as a rigid body.
Fig. 19. Computed pore pressure response with refined per- Modelling of soil layering within London Clay and appli-
meability model compared with field data: (a) along the tunnel cation of non-homogeneous drainage conditions around the
lining extrados; (b) BP1 at 24 m below ground surface; (c) BP2 lining (i.e. with the tunnel impermeable above the springline
at 27 m below ground surface and permeable below the springline) were necessary to
accurately simulate the long-term ground response observed
at St James’s Park. Modelling of soil layering allows the
anisotropic permeability of the layer. Fig. 19(c) shows that change in pore pressures in the soil to occur at the same rate
the recovery of pore pressures at BP2, which is closer to the as those monitored in the field. As the majority of the
tunnel crown extrados than BP1, depends on the permeabil- consolidation takes place on either side of the tunnel, the
ity of the tunnel lining. To achieve the pore pressure values rate of the consolidation settlement therefore depends princi-
close to that observed in the field, the portion of the lining pally on the permeability of the soil at the tunnel springline
above the springline must be made impermeable. Despite level. The application of non-homogeneous drainage condi-
this, such a drainage condition causes the computed pore tions around the lining allows the pore water to flow into the
pressures at the tunnel crown to have slightly lower values tunnel below the springline, causing the soil on either side
than the field measurement. In order to achieve a better of the tunnel to consolidate. Moreover, it also allows the
agreement with the field data, further studies on the tunnel pore pressure above the tunnel crown to partially recover,
lining and surrounding soil permeability are needed. causing the soil to swell. Hence it appears that appropriate
The computed subsurface vertical ground movements with tunnel drainage conditions must be assigned for successful
depth at various extensometer locations are shown in Fig. long-term analysis. It is recognised, however, that these con-
20. The rate of further settlement agrees well with the field clusions are based on the analyses reported in this paper of
measurement (see Fig. 5), and also the computed transverse a case study with only a small number of pore pressure
settlement profile gives a very similar mechanism to that measurements around the tunnel. More long-term pore pres-
shown by the measurements. Non-homogeneous drainage sure measurement data are needed to confirm them.
towards the lining allowed the pore water to flow into the Soil stiffness anisotropy is another factor affecting the
tunnel below the springline, causing the soil on either side long-term ground response for a high-K0 soil. When the
of the tunnel to consolidate. Moreover, it also allowed pore tunnel lining was assigned to be relatively permeable, the
pressures above the tunnel crown to partially recover, caus- magnitudes of negative pore pressures were larger above and
ing the soil to swell. As the water drained into the lower below the tunnel when the anisotropic soil model was used.
half of the tunnel, the increased permeability in division A3i This resulted in a squatting deformation of the tunnel lining
created a slightly faster consolidation rate than the recovery in the long term, which was in good agreement with the
of pore pressure above the tunnel. This enables the down- deformations observed in the field. When the isotropic soil
ward displacements of the soil above the tunnel crown, as model was used, the opposite trend was predicted. The
shown in Fig. 20(a). The higher anisotropic permeability magnitudes of negative excess pore pressure were greater at
ratio assigned for division A3ii compared with the other the sides of the tunnel, which led to a decrease in horizontal
layers allowed the pore pressure above the tunnel crown in diameter in the long term.

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
88 WONGSAROJ, SOGA AND MAIR
Consolidation vertical displacement: mm Consolidation vertical displacement: mm
⫺15 ⫺10 ⫺5 0 5 ⫺15 ⫺10 ⫺5 0 5
0 0

⫺5 ⫺5

⫺10 ⫺10
Consolidation
⫺15 ⫺15 time
0 day
⫺20 ⫺20 5 days
Depth: m

Depth: m
23 days
⫺25 ⫺25
71 days

⫺30 135 days


⫺30
263 days
⫺35 ⫺35

⫺40 ⫺40

⫺45 ⫺45

⫺50 ⫺50
(a) (b)

Consolidation vertical displacement: mm Consolidation vertical displacement: mm


⫺15 ⫺10 ⫺5 0 5 ⫺15 ⫺10 ⫺5 0 5
0 0

⫺5 ⫺5

⫺10 ⫺10
Consolidation
time
⫺15 ⫺15
0 day

⫺20 ⫺20 5 days


Depth: m

Depth: m

23 days
⫺25 ⫺25 71 days
135 days
⫺30 ⫺30 263 days

⫺35 ⫺35

⫺40 ⫺40

⫺45 ⫺45

⫺50 ⫺50
(c) (d)

Fig. 20. Computed subsurface ground vertical movement with refined permeability model: (a) Bx along tunnel centreline;
(b) Cx 4 m from tunnel centreline; (c) Dx 10 m from tunnel centreline; (d) Ex 16 m from tunnel centreline

APPENDIX 1 where u1 and m are the material constants, and då p is the change in
The normal yield surface adopted in this study is the Modified plastic strains tensor, which is obtained from the flow rule. For
Cam Clay yield surface, which can be described in q–p9space as simplicity, the associated flow rule is assumed in the model and,
therefore, the plastic potential is described by equation (2).
F ¼ q 2  M 2 ð p90  p9Þ p9 ¼ 0 (1) The conditions for the elasto-plastic process are given by
9
where M is the gradient of the critical state line and a function of the R ¼ 0: dR ¼ þ1 > >
=
Lode angle using the Matsuoka & Nakai (1985) failure criterion, and 0 , R , 1: dR . 0
for  p 6¼ 0 (4)
p90 is a constant defining the size of the normal yield surface (i.e. R ¼ 1: dR ¼ 0 > >
;
isotropic preconsolidation pressure). For a given value of R, the R . 1: dR , 0
subloading surface can then be expressed as
The elastic modulus of the model is assumed based on experimental
F R ¼ q 2  M 2 ð Rp90  p9Þ p9 ¼ 0 (2)
data plotted in log10 e–log10 p9 space, where swelling behaviour is
Two different evolution laws for R were proposed by Hashiguchi described by the gradient of the swelling line in log10 e–log10 p9 as
& Chen (1998). The one adopted in this study is expressed as (after Pestana, 1994)
   
1 1 þ øs s p9 1=2
dR ¼ u1  1 kdå p k (3) rr ¼ þDð1   r Þ (5)
Rm Cb pa

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
LONG-TERM GROUND RESPONSE TO TUNNELLING UNDER ST JAMES’S PARK 89
where Cb and ø s are material constants, which control the initial
gradient of the swelling line and the non-linearity of the one-

Ghh /Gvh
dimensional (1D) swelling line respectively; D is the gradient of the

1.5
1
1

1
isotropic swelling line at low mean effective pressure; r controls the
rate at which the isotropic swelling line reaches the gradient D; and
atmospheric pressure pa of 100 kPa is assumed for this study.  and
s are used to re-invoke small-strain stiffness at stress reversal points
(i.e. unloading and reloading). They are the dimensionless distances

0.12
0.2
0.2

0.2
9hh
in space as defined by
 
s ¼ ðç  çrev Þ: ðç  çrev Þ 1=2 (6)

p9= p9rev for p9 , p9rev
0.16
¼ ;<1 (7)
0.2
0.2

0.2
9hv

p9rev = p9 for p9 > p9rev


where ç is the ratio of deviatoric stress tensor to mean effective
stress, and çrev is the value of the stress ratio at the most recent
0.07

stress reversal.
0.2
0.2

0.2
9vh

The stress reversal point is defined as (Pestana, 1994)


 ˜1  
  for p 6¼ 0 . 0 loading
_ ¼ ˜1 p p ¼
ås : ås for p ¼ 0 < 0 unloading (SRP)
2
2

(8)
r

˜1
where  is the accumulated strain relative to the previous stress
reversal state, which can be decomposed into its volumetric and
deviatoric components, ˜1 p (¼ p  prev ) and ˜1 ås (¼ ås  åsrev ).
0.05
0.05
D


From equation (5), the tangent bulk modulus, K9, during swelling
is defined as
 
1 þ e p9
0.2476

K9 ¼ (9)
0.556

e rr
0.37
rc

0.3

By starting from the definition of bulk modulus and the assumption


of the symmetry of the elastic D-matrix, it can be shown that
9vh
0.15
0. 1
0. 1

E9h ¼ E9v (10)


º

9hv
 
29vh 29hh 9vh
E9v ¼ K9 1  49vh þ  (11)
9hv 9hv
15
15
20
50
øs

Under the assumption that the soil is isotropic in the plane of the
deposition, the value of Ghh becomes
E9h
Ghh ¼ (12)
2ð1 þ 9hh Þ
100
400
200
900
Cb

The values of 9vh , 9hv , 9hh and Gvh are evaluated from published
data.
For the present model, it is assumed that the normal consolidation
0.1
0.1
0.2
0.1

line is a straight line with a gradient of rc . The hardening rule


m

therefore becomes
1þe
d p90 ¼ p90 dpv (13)
eðrc  rr Þ
100
100
300
100
u1

where p90 is the mean effective preconsolidation pressure, which


controls the size of the normal yield surface, and dpv is the plastic
volumetric strain increment.
0.65

0.65
0.5
0.7
e

NOTATION
Table 4. Model parameters for all soil units

Cb parameter controlling initial gradient of swelling line or


0.984
1.418
0.814

magnitude of small-strain stiffness


1.07
M

D gradient of 1D swelling line at low mean effective stress


E9h drained Young’s modulus in horizontal direction
Woolwich and Reading Beds Clay

E9v drained Young’s modulus in vertical direction


e void ratio
F normal yield surface function
R
F subloading surface function
Ghh shear modulus in horizontal plane
Gvh shear modulus in vertical plane
K9 bulk modulus during swelling
KT seepage coefficient
Terrace Gravel
Made Ground

London Clay

K0 ratio of initial vertical to horizontal effective stress


k permeability of soil
kh permeability of soil in horizontal direction
Strata

kl permeability of tunnel lining


kv permeability of soil in vertical direction

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26
90 WONGSAROJ, SOGA AND MAIR
M gradient of critical state line in triaxial space granular material: constitutive equations for soils. Proc 9 th Int.
m, u1 parameters controlling evolution law of R Conf. Soil Mech. Found. Engng, Tokyo, Special Session 9,
p9 mean effective stress 73–82.
pa atmospheric pressure; taken as 100 kPa Harris, D. I. (2002). Long-term settlement following tunnelling in
p90 mean effective preconsolidation pressure overconsolidated London Clay. Proc. 3rd Int. Symp. on Geotech-
p9rev mean effective stress at most recent stress reversal point nical Aspects of Underground Construction in Soft Ground,
q deviatoric stress in triaxial space Toulouse, 393–398.
R ratio between size of subloading surface and size of normal Higgins, K. G., Potts, D. M. & Mair, R. J. (1996). Numerical
yield surface modelling of influence of the Westminster Station excavation on
r parameter controlling rate at which 1D swelling line reaches the Big Ben clock tower. In Geotechnical aspects of under-
gradient D ground construction in soft ground (eds R. J. Mair and R. N.
tl thickness of tunnel lining Taylor), pp. 523–530. Rotterdam: Balkema.
uw pore pressure at tunnel lining extrados Hight, D. W., McMillan, F., Powell, J. J. M., Jardine, R. J. &
vn pore fluid velocity outward from soil element Allenou, C. P. (2003). Some characteristics of London Clay.
ª soil unit weight Proceedings of the conference on characterisation and engineer-
 p plastic strain ing properties of natural soils, Singapore, Vol. 2, pp. 851–907.
p volumetric strain King, C. (1981). The stratigraphy of the London Basin and asso-
pv plastic volumetric strain ciated deposits, Tertiary Research Special Paper No. 6. Rotter-
prev volumetric strain at most recent stress reversal point dam: Backhuys.
s deviatoric strain Lake, L. M., Rankin, W. J. & Hawley, J. (1992). Prediction and
srev deviatoric strain at most recent stress reversal point effects of ground movements caused by tunnelling in soft ground
˜1
 accumulated strain relative to the previous stress reversal beneath urban areas, CIRIA Funders Report CP/5. London:
state Construction Research and Information Association.
ç deviatoric stress tensor Lee, G. T. K. & Ng, C. W. W. (2002). Three-dimensional analysis
çrev deviatoric stress tensor at most recent stress reversal point of ground settlement due to tunnelling: role of K0 and stiffness
9hh ratio of strain in horizontal direction to other horizontal anisotropy. Proc. 3rd Int. Symp. on Geotechnical Aspects of
direction due to compression in horizontal direction Underground Construction in Soft Ground, Toulouse, 617–622.
9hv ratio of strain in vertical direction to horizontal direction Matsuoka, H. and Nakai, T. (1985). Relationship among Tresca,
due to compression in horizontal direction Mises, Mohr–Coulomb and Matsuoka–Nakai failure criteria.
9vh ratio of strain in horizontal direction to vertical direction Soils Found. 25, No. 4, 123–128.
due to compression in vertical direction Nyren, R. (1998). Field measurement above twin tunnel in London
, s dimensionless distances used to re-invoke small-strain Clay. PhD thesis, Imperial College of Science Technology and
stiffness at stress reversal points Medicine, London.
rr gradient of swelling line in log10 e–log10 p9 O’Reilly, M. P., Mair, R. J. & Alderman, G. H. (1991). Long-term
9cv critical angle of shearing resistance settlements over tunnels: an eleven-year study at Grimsby. Proc.
_ parameter defining stress reversal point Tunnelling ’91, London, 55–64.
øs parameter controlling non-linearity of one-dimensional Palmer, J. H. L. & Belshaw, D. J. (1980). Deformations and pore
swelling line pressure in the vicinity of a precast, segmented, concrete-lined
tunnel in clay. Can. Geotech. J. 17, No. 2, 174–184.
Peck, R. B. (1969). Deep excavations and tunnelling in soft ground.
REFERENCES Proc. 7th Int. Conf. Soil Mech. Found. Engng, Hamburg, State
Addenbrooke, T. (1996). Numerical analysis of tunnelling in stiff of the Art Volume, 225–290.
clay. PhD thesis, Imperial College of Science Technology and Pestana, J. M. (1994). A unified constitutive model for clays and
Medicine, London. sands. PhD thesis, Massachusetts Institute of Technology.
Barratt, D. A., O’Reilly, M. P. & Temporal, J. (1994). Long term Potts, D. M. & Zdravkovic, L. (2001). Finite element analysis in
measurement of loads on tunnel linings in overconsolidated clay. geotechnical engineering: application. London: Thomas Telford,
Proc. Tunnelling ’94, London, 469–481. 35–72.
Bowers, K. H., Hiller, M. D. & New, B. M. (1996). Ground Simpson, B., Atkinson, A. J. & Jovicic, V. (1996). The influence of
movement over three years at Heathrow Express trial tunnel. anisotropy on calculations of ground settlements above tunnels.
Proceedings of the international symposium on geotechnical In Geotechnical aspects of underground construction in soft
aspects of underground construction in soft ground, London, pp. ground (eds R. J. Mair and R. N. Taylor), pp. 591–594.
557–562. Rotterdam: Balkema.
Burland, J. B. & Hancock, R. J. (1977). Underground car park at Standing, J. R. & Burland, J. B. (2006). Unexpected tunnelling
the House of Commons, London: geotechnical aspects. Struct. volume losses in the Westminster area, London. Géotechnique
Engr 55, No. 2, 87–100. 56, No. 1, 11–26.
Burland, J. B., Standing, J. R. & Jardine, F. M. (eds) (2001). Ward, W. H. & Thomas, H. S. H. (1965). The development of earth
Building response to tunnelling. London: CIRIA/Thomas loading and deformation in tunnel lining in London Clay. Proc.
Telford. 6th Int. Conf. Soil Mech. Found. Engng, Toronto 2, 432–436.
Dimmock, P. S. (2003). Tunnelling induced ground and building Winterton, T. R. (1994). Developments in precast concrete tunnel
movement on the Jubilee Line Extension. PhD thesis, University linings in the United Kingdom. Proc. Tunnelling ’94, London,
of Cambridge. 601–633.
Franzius, J. N. (2003). Behaviour of buildings due to tunnel induced Wongsaroj, J. (2005). Three-dimensional finite element analysis of
subsidence. PhD thesis, Imperial College of Science Technology short and long-term ground response to open-face tunnelling in
and Medicine, London. stiff clay. PhD thesis, University of Cambridge.
Graham, J. & Houlsby, G. T. (1983). Anisotropic elasticity of Wongsaroj, J., Soga, K., Mair, R. J. and Yimsiri, S. (2004). Stiffness
natural clay. Géotechnique 33, No. 2, 165–180. anisotropy of London Clay and its modelling: laboratory and
Hashiguchi, K. & Chen, Z. P. (1998). Elastoplastic constitutive equa- field. Proceedings of the Skempton Conference on advances in
tion of soils with the subloading surface and the rotational hard- geotechnical engineering, London,Vol. 2, pp. 1205–1216.
ening. Int. J. Numer. Anal. Methods Geomech. 22, No. 3, 197–227. Yimsiri, S. (2001). Pre-failure deformation characteristic of soils:
Hashiguchi, K. & Ueno, M. (1977). Elastoplastic constitutive law of anisotropy and soil fabric. PhD thesis, University of Cambridge.

Delivered by ICEVirtualLibrary.com to:


IP: 62.189.175.212
On: Thu, 04 Aug 2011 16:10:26

Vous aimerez peut-être aussi