Vous êtes sur la page 1sur 15

Fargnoli, V. et al. (2015). Geotechnique 65, No. 1, 2337 [http://dx.doi.org/10.1680/geot.14.P.

091]

3D numerical modelling of soilstructure interaction during EPB


tunnelling
V. FA R G N O L I  , C . G . G R AG NA N O , D. B O L D I N I  a n d A . A M O RO S I

The paper illustrates a three-dimensional finite-element analysis aimed at back-predicting the response
of a multi-storey reinforced concrete building underpassed by a metro tunnel. The study refers to the
case of the Milan metro line 5, recently built in coarse-grained materials using an earth pressure
balance machine, for which settlement measurements along ground and building sections were
available. The soil behaviour is modelled using an advanced constitutive model that, when combined
with a proper simulation of the excavation process, proves to realistically reproduce the subsidence
profiles recorded in free-field conditions. The building is found to modify the deformative pattern at
the ground surface in relation to its stiffness and weight, reducing the differential settlements as
compared to those calculated under free-field conditions. Results of the numerical simulation carried
out, including the model of the building schematised in detail, are found to be in good agreement
with the monitoring data. They thus indirectly confirm the satisfactory performance of the adopted
numerical approach, which takes into account a unique model of the soil, the tunnel and the building
that is, the key ingredients of this interaction problem. Further analyses are also carried out
modelling the building, adopting different levels of detail. The results highlight that, for the case
under study, the simplified approach based on the equivalent plate schematisation is inadequate to
capture the real displacement field. The overall behaviour of the system proves to be mainly
influenced by the buried portion of the building, including its foundation elements, which plays an
essential role in the interaction mechanism.
KEYWORDS: finite-element modelling; monitoring; settlement; soil/structure interaction; tunnels

INTRODUCTION
The construction of underground tunnels in urban areas
often requires excavation works to be carried out in close
proximity to residential buildings, cultural heritage monuments and underground services. The ability to predict the
tunnelling-induced settlements and the associated impact on
pre-existing structures represents a key aspect to estimate
potential damages and to design protective measures, when
needed (e.g. Mair, 2008; Amorosi et al., 2012; Puzrin et al.,
2012; Rampello et al., 2012).
Soil deformation and structural response are often assumed to be decoupled, so that the building damage is
typically predicted based on free-field settlement profiles
(Peck, 1969; Burland & Wroth, 1974; Burland et al., 1977;
OReilly & New, 1982; Boscardin & Cording, 1989; Burland, 1995). However, such a simplified approach disregards
the influence of the structure stiffness (Potts & Addenbrooke, 1997; Franzius et al., 2006) and the role of its
weight (Franzius et al., 2004), often leading to rather
conservative solutions in terms of estimated differential
settlements and, consequently, of induced damage intensity.
In the last few years two-dimensional (2D) and threedimensional (3D) numerical approaches have been developed
to overcome such limitations. 2D numerical studies were
proposed by Liu et al. (2000) with reference to surface
masonry structures, focusing on the effect of facade weight,
stiffness and position with respect to the tunnel axis. Alongside the same topic, Amorosi et al. (2014) back-analysed the

interaction between the excavation of a tunnel and an


ancient masonry surface structure, adopting an advanced
elasto-plastic constitutive model for the masonry. The dependency of the building response on various structural types
(i.e. brick-bearing structures, open-frame and brick-infilled
frame structures) and different soil conditions was discussed
in detail by Son & Cording (2011).
More sophisticated 3D simulations, again with reference
to masonry surface structures, were presented in Burd et al.
(2000) and Giardina et al. (2010). Soilstructure interaction
studies were also carried out, focusing on 3D framed buildings, schematised as equivalent plates (Maleki et al., 2011)
or adopting simple schemes consisting of beams, columns
and live loads acting at each floor of the reinforced concrete
structure (Liu et al., 2012).
This paper proposes the results of a numerical study
aimed at investigating the soilstructure interaction during
tunnelling using a fully 3D solution scheme based on the
finite-element method. The study refers to a real case
history, the construction of the new Milan (Italy) metro-line
5, carried out in granular soils by an earth pressure balance
(EPB) machine, which guaranteed surface volume losses
lower than 0 .4%. In the examined portion of the route (Figs
1 and 3), the right tunnel of the line diagonally underpasses
a nine-storey reinforced concrete structure dating back to the
end of the 1950s.
In the first part of the paper the tunnelling-induced settlements, as observed along the segment of the route at six
transversal ground sections and in correspondence with the
reference building, are collected and interpreted. Such analysis allows evaluation of the ability of existing closed-form
empirical solutions to back-calculate the observed free-field
settlement troughs along both transversal and longitudinal
directions, also providing direct information on the evolution
of the surface structures response during tunnelling.

Manuscript received 27 May 2014; revised manuscript accepted 23


October 2014. Published online ahead of print 27 January 2015.
Discussion on this paper closes on 1 June 2015, for further details see
p. ii.
 University of Bologna (Italy), Bologna, Italy.
Technical University of Bari, Italy.

23

FARGNOLI, GRAGNANO, BOLDINI AND AMOROSI

24
97 m

Lotto Station

57 m

50 m

GS_1

GS_2

43 m

GS_3

GS_4

50 m

46 m

GS_5
GS_6

Scale: 50 m

Left tunnel
Right tunnel

6m
25

Tun
n

7
el m 0 m
ode
lled

por

tion

Portello Station

Fig. 1. Milan metro-line 5: plan view of the segment of the route between Lotto and Portello stations

The numerical study, described in the second part of the


paper, is performed by the code Plaxis 3D (Plaxis, 2012). An
advanced constitutive model, called Hardening soil with
small strain stiffness (HSsmall, Benz, 2007), is adopted for
the soil, and a detailed numerical step-by-step procedure is
used to model the main features of the tunnel excavation
process. The surface reinforced concrete building interacting
with the metro-line is modelled in detail, taking into account
not only its main structural components, but also its secondary elements that is, the external infill panels. Additional
numerical analyses were also carried out, adopting simplified
building models to highlight the role of the different structural
components in the overall behaviour of the soilstructure
system. They include the equivalent plate schematisation and
a simplified structural model only considering the buried
portion of the building together with its foundation elements.
The outcomes of the finite-element simulations are first
presented with reference to free-field conditions and then
considering the pre-existence of the surface building. The
ability of the proposed numerical approach to reproduce
realistic results is assessed by comparing the computed
settlements with the available geotechnical and structural
monitoring measurements. Finally, the results obtained
adopting different levels of detail in modelling the building
are illustrated and critically discussed.
CASE HISTORY OF NEW MILAN METRO-LINE 5
The new Milan metro-line 5 (Fargnoli et al., 2013) runs
from north to west of the city with a total length of 12 .6 km
and 19 access stations. The monitored portion considered in
this study extends for a length of about 600 m between
Lotto and Portello stations (Fig. 1).
The twin tunnels of the line have a separation between
the two axes of about 15 m and a mean depth of their axes
z0 15 m. This latter reaches its maximum value of about
23 m at Lotto station.
An EPB machine was selected to minimise ground movements in the highly populated areas affected by the construction activities. The EPB shield, with a total length of about
10 m and a thickness of 30 mm, is characterised by an outer
diameter of 6 .69 m at the face and 6 .67 m at the tail. Under

special circumstances, the maximum excavation diameter at


the face could be increased up to 6 .71 m. Six pressure cells
are located on the EPB face, as shown in Fig. 2.
The advancement is provided by 38 hydraulic jacks
located on the external perimeter of the shield body, acting
on the already installed lining. The tunnel lining, set in place
inside the shield tail to support the tunnel as the machine
advances, consists of concrete precast rings characterised by
a length of 1 .4 m and a thickness of 30 cm. The outer and
inner diameters of the lining ring are equal to 6 .40 m and
5 .80 m, respectively. The gap behind the lining segments is
promptly filled by a two-component grouting consisting of
cement paste and grip accelerator, in order to minimise the
volume loss and the related surface settlements.

137 m

193 m

248 m

180 m

197 m

D 669 m

Fig. 2. Location of the pressure cells on the EPB face

MODELLING OF SOILSTRUCTURE INTERACTION DURING EPB TUNNELLING


During the various stages of the tunnel construction,
extensive geotechnical and structural monitoring was carried
out along the line by an accurate levelling survey, with
recording intervals varying between 12 and 24 h. The data
presented and discussed throughout this paper are vertical
displacements recorded during the excavation of the first
tunnel (i.e. right tunnel) at six transversal ground sections
and in correspondence with a nine-storey reinforced concrete

framed structure diagonally undercrossed by the tunnel (Fig.


3). The 30 m high structure dates back to 1959 (Figs 4(a)
and 4(b)) and is characterised by a total weight of about
41 000 kN. Its plan dimensions and the position of its
middle-point M are reported in Fig. 3. The building interstorey height is equal to 3 .2 m, except for the ground floor
and the basement floor, which have heights of 4 .2 m and
2 .5 m, respectively.

GS_6

Ground benchmarks

Building targets

Scale: 10 m

Left tunnel

Right tunnel
L2

L1

L3

L4

L5
T1

M
building
R1

R2

R3

T2

R4

R5

B 12 m

T3

15 m

70

Tun
n

el m

L 30 m

ode

lled

por

tion

Fig. 3. Detail of the examined portion of the route

(a)

25

(b)

Fig. 4. (a) General view of main left side facade and (b) detail of the garage zone on the right longitudinal side of the building

FARGNOLI, GRAGNANO, BOLDINI AND AMOROSI

26

The main structural components of the building have the


following dimensions: the sections of the beams are
40 cm 3 45 cm at the lower floors, and 70 cm 3 20 cm and
45 cm 3 20 cm within or along the perimeter of the upper
floors, respectively; the column section has dimensions of
40 cm 3 40 cm; the thickness of the floor slabs is 26 cm at
the lower floors and 22 cm at the upper ones; the sections of
the reinforced concrete interior panels are 0 .2 m 3 3 .26 m.
The structure is founded on five strip footings (0 .65 m
high, indicated as I, II, III, IV and V in Fig. 5), 4 m below the
ground surface; more specifically, the building rests on the
foundation beams I, II and III, while the garage zone (Fig.
4(b)), situated at the basement floor level along the right

13 m

13 m

13 m

II

III

13 m

13 m

IV

31 m

19 m

VI

30 m

300 m

VII

longitudinal side of the structure, stands on the other ones (IV


and V). Three raft foundations (0 .7 m high, indicated as VI,
VII and VIII in Fig. 5) are located at the same level under the
elevator shafts and the stairwell, both on the right longitudinal
side of the building. Reinforced concrete retaining walls
(40 cm thick and 3 .5 m high) surround the buried portion of
the structure along its three sides, except for the garage zone.
Several ground benchmarks (from five to nine) were
installed on each instrumented ground section, while building targets were placed along the base of the buildings
longitudinal facades and on its transversal right side (Fig. 3).
GEOTECHNICAL CONDITIONS
The city of Milan is located in the central part of the
Padana plain (northern Italy), resting on a deep glacial and
alluvial Pleistocene formation. The upper part of this deposit
mainly consists of sand and gravel, characterised by a silt
percentage that typically increases with depth. A formation
of conglomerate and sandstone underlies this upper deposit,
about 40 m below the ground surface, while sand and clay
are present at greater depth. The new metro-line 5 is located
within the upper granular non-cohesive unit of the formation, which mainly consists of gravel and sand of fluvioglacial and alluvial origin.
Along the metro-line an extensive geotechnical investigation was carried out at the design stage of the work. In
particular, with reference to the portion considered in this
paper between Lotto and Portello stations, it included

42 m
VIII

52 m

64 m

Structure above
ground level

19 m

31 m

51 m

Granulometric analyses were conducted on 14 disturbed


samples retrieved from the drillings. Two main granulometric
grading curves were determined: the first curve is typical of
a gravelly sand soil, S, the second one of a sandy silt, L. At
the location under investigation the gravelly sand soil
emerges as the main component of the deposit, being found
at depths between 0 and 20 m, and between 25 and 30 m,
while a layer about 5 m thick of sandy silt was identified
between 20 m and 25 m. The total unit volume weights
under saturated conditions for these materials are 20 kN/m3
and 17 .5 kN/m3, respectively.

50 m

Garage below
ground level

Fig. 5. Plan view of building foundations


CD_1
z: m
0

g. l. m a. s. l.

GS_1

two core drillings (CD_1 and CD_2) to a depth of 24


30 m from the ground surface (Fig. 6), instrumented with
open pipe piezometers
14 and 12 SPT tests conducted along boreholes CD_1
and CD_2, respectively
three Lenfranc-type permeability tests carried out imposing a constant piezometric level.

GS_2

GS_3 GS_4 GS_5 GS_6

0
R

R
SG
G, S(L)

CD_2
z: m

5
S, G(L)

S, G(L)

10

10
SG
15
S, G
L(S)

w.t._2013
w.t._2007

20

15

Reference monitoring sections


G, S(L)

Tunnel axis
Tunnel invert

25

G(S)

Tunnel crown

G(S) sandy gravel


G, S(L) gravel with silty sand
L(S) sandy silt
SG sand and gravel
S, G sand with gravel
S, G(L) sand with silty gravel
R made ground

20

Horizontal scale: 10 m

L(S)
24

G, S(L)

30

Fig. 6. Soil conditions along the examined portion of the route, depth of tunnel and position of ground monitoring sections

MODELLING OF SOILSTRUCTURE INTERACTION DURING EPB TUNNELLING


Results of SPT tests conducted in the gravelly sand soil
were elaborated following Skempton (1986), leading to a
relative density in the range 6080% and strength parameters
equal to c9 0 kPa and 9 338.
Standard penetration tests (SPTs) were not considered
appropriate to characterise the sandy silt layer; therefore, its
strength parameters were assumed to be equal to c9 5 kPa
and 9 268, based on pre-existing geotechnical characterisations carried out in the Milan area.
At the time of the geotechnical survey the hydrostatic
water level was found at an almost constant depth of about
18 .5 m below the ground level (Fig. 6). Six years after,
during the construction stage, the corresponding level was
detected 15 m below the ground surface, which is consistent
with the slow but continuous increase in the water table
observed in the area in recent years.
The permeability coefficients, k, were observed to vary at
different depths between 5 .5 3 103 and 1 .1 3 102 m/s.
No geophysical investigations were specifically undertaken
along the metro route; the one nearest to the reference
segment of the line is a down-hole test performed at a
nearby construction site: this test resulted in the small-strain
shear modulus (G0) profile shown in Fig. 7. The figure also
illustrates the related stratigraphic profile, which proves to
be rather similar to the one of the tunnel site.

MONITORING MEASUREMENTS: ANALYSIS AND


DISCUSSION
Tunnelling-induced settlements were recorded between
Lotto and Portello stations at six transversal ground sections
(i.e. GS_1, GS_2, GS_3, GS_4, GS_5 and GS_6) and in
correspondence with the surface structures interacting with
the new metro-line.
0
Experimental

Computed:

Gravelly sand

27

Ground monitoring
Transversal settlement profiles due to the excavation of
the first tunnel are shown in Fig. 8. They refer to fully
developed settlements achieved when the tunnel face was at
a sufficient distance from the monitoring sections.
The maximum settlement induced by the excavation never
exceeds 7 mm. Measurements are sufficiently well fitted by a
Gaussian distribution curve (Peck, 1969) for K values in the
range 0 .40 .45, with the exception of the left-most points of
each section, whose values are under-predicted by the empirical relation (Fig. 8), probably due to the influence of nearby
surface structures located along the examined route (see Fig.
1). The above interpolation allowed back-evaluation of the
corresponding volume loss, VL (%), which varies from 0 .3%
to 0 .38%, with an average value equal to 0 .33%, indicating a
well-performing EPB excavation, and consistent with similar
observations reported in the literature (Leblais & Bochon,
1991; Ata, 1996; Mair, 1996; Mair & Taylor, 1997).
The values of maximum settlement, K parameter and
volume loss for all the considered sections are summarised
in Table 1. The table also reports the different depths of the
tunnel axis, z0, at each location and the recording date of
the analysed settlement measurements.
The evolution of settlement above the tunnel centre-line is
presented in Fig. 9 as a function of the face distance for the
two sections, GS_3 and GS_6, for which measurements of
vertical displacements close to the tunnel face (i.e.  1 m)
were available. The tunnel face settlement, Sv,f, at these
locations is equal to about 1 mm, indicating a satisfactory
face support during the excavation process.
Measurements were interpreted at each location by the
cumulative Gaussian probability curve (Attewell & Woodman, 1982) in order to define the longitudinal settlement
trough, assuming the volume loss and K values reported in
Table 1 and considering the longitudinal inflection point, iy,
equal to the transversal one, ix. As shown in Fig. 9, face
settlements are best fitted by the translated Gaussian cumulative curve (Mair & Taylor, 1997), obtained by equating the
ratio Sv,f /Sv,max to the measured one. This translated profile,
however, is not able to capture the further evolution of
settlements, predicting the achievement of steady-state conditions well before what is observed in situ.

Sandy silt
10

Depth, z: m

S_1
15

20
L
25

S_2
30

200
400
600
Small-strain shear modulus, G0: MPa

800

Fig. 7. Experimental and computed small-strain shear modulus


profiles with depth

Structural monitoring
Vertical displacements of the building were gathered during
tunnelling by monitoring targets located along the base of its
three sides, identified by capital letters L, R and T and a
sequential number (see Fig. 3). The target relative distance and
their distances from the tunnel axis are listed in Tables 2 and 3.
As expected, the vertical displacements mainly increase
with the excavation advancement, as shown in Figs 10(a)
10(c), where the evolution at each monitoring point is
displayed at different dates. Final measurements, gathered
when the distance of the right tunnel face from section
GS_6 was about 50 m, that is about 8D, range from 4 .7 mm
to 6 .5 mm along the longitudinal left facade, from 3 .5 mm
to 6 .6 mm along the longitudinal right facade and from
4 .6 mm to 5 .7 mm along the transversal side.
These plots make it possible to highlight the effect of the
distance from the tunnel axis on settlement: the closer the
tunnel axis (i.e. points L1, L2, L3 along the longitudinal left
side, R4, R5 along the longitudinal right side and T2, T3
along the transversal one), the higher the settlement.
The progressive settlement response of the building
evolves during tunnelling, this being particularly evident
along its longitudinal sides: the structure, in fact, is characterised by a hogging-type mode of deformation when the
tunnel-boring machine (TBM) face is located in correspon-

FARGNOLI, GRAGNANO, BOLDINI AND AMOROSI

28

Settlement, Sv: mm

Settlement, Sv: mm

8
40

20
0
20
Distance from tunnel axis, x: m
(a)

20
0
20
Distance from tunnel axis, x: m
(b)

40

20
0
20
Distance from tunnel axis, x: m
(d)

40

20
0
20
Distance from tunnel axis, x: m
(f)

40

Settlement, Sv: mm

Settlement, Sv: mm

8
40

40

8
40

20
0
20
Distance from tunnel axis, x: m
(c)

8
40

40

Settlement, Sv: mm

Settlement, Sv: mm

8
40

20
0
20
Distance from tunnel axis, x: m
(e)

40

8
40

Measurements:

Transversal troughs:

GS_1

GS_3

GS_5

Empirical

GS_2

GS_4

GS_6

Computed

Fig. 8. Transversal settlement troughs: measurements and best-fitting Gaussian curves (the computed profile for
ground section GS_6 is also shown)

Table 1. Values of maximum settlement, K parameter, volume


loss, axis depth and surveying date in the reference monitoring
sections

structure at the end of the excavation works, owing to the


relatively low settlements induced by EPB tunnelling.

Monitoring
sections

NUMERICAL MODEL
Geometry and finite-element discretisation
Different numerical models were set up to simulate the
tunnel excavation under free-field conditions and in the
presence of the building, the latter being modelled with
different levels of detail. In the interaction analyses, for the
sake of simplicity, the presence of the nearby structures was
neglected.
The mesh employed in the present study is shown in Fig.
11(a): it represents a soil volume 68 m wide, 30 m high and
100 m long. These dimensions were selected to minimise the
influence of the boundary conditions on the computed
results. The numerical model adopted for the free-field
analysis is constituted by 84 125 nodes; this number increases to 120 952 in the interaction analysis with the
complete structural model. Nodes at the bottom of the mesh

GS_1
GS_2
GS_3
GS_4
GS_5
GS_6

Sv,max:
mm
5 .6
5 .2
6 .0
5 .5
6 .5
5 .5

VL: %

0 .45
0 .42
0 .45
0 .42
0 .40
0 .45

0 .38
0 .30
0 .37
0 .30
0 .31
0 .34

z0: m

Date

20 .0 16 December 2012
19 .0 18 December 2012
18 .0 20 December 2012
17 .0
8 January 2013
16 .0 10 January 2013
15 .0 15 January 2013

dence with the middle of the building (measurements recorded on 11 January 2013), while the deformative pattern
evolves in a sagging-type mode after the tunnel passage (i.e.
from 12 January 2013 on).
In general, no evidence of damage was detected on this

MODELLING OF SOILSTRUCTURE INTERACTION DURING EPB TUNNELLING

Settlement, Sv: mm

Table 3. Target distance from the right tunnel axis

Monitoring points

40

20
0
20
40
60
80
Distance from tunnel face, y: m
(a)

100

Distance from the right tunnel axis: m


1 .53
0 .33
2 .84
5 .84
9 .82
13 .35
10 .76
9 .26
4 .51
1 .02
11 .23
6 .89
0 .00

L1
L2
L3
L4
L5
R1
R2
R3
R4
R5
T1
T2
T3

8
60

120

Settlement, Sv: mm

29

2
4
6
8
60

40

20
0
20
40
60
80
Distance from tunnel face, y: m
(b)

Measurements:

100

120

Longitudinal troughs:
Original

GS_3

Translated

GS_6

Computed

Fig. 9. Settlements measured above the tunnel centre-line at the


monitoring sections (a) GS_3 and (b) GS_6 with original and
translated longitudinal profiles (the computed profile for ground
section GS_6 is also shown)
Table 2. Target relative distance
Monitoring points

Relative distance: m
5 .54
6 .80
7 .09
8 .44
7 .18
7 .32
6 .97
7 .33
4 .84
6 .38

L1L2
L2L3
L3L4
L4L5
R1R2
R2R3
R3R4
R4R5
T1T2
T2T3

L2

L3

L4

L5

R1

Settlement, Sv: mm

Settlement, Sv: mm

2
4
6
8

R2

R3

R4

R5

T1

(a)

2
4
6
8

T2

T3

Settlement, Sv: mm

L1

are fixed in both vertical and horizontal directions, while the


vertical boundaries are only fixed in the horizontal directions. The soil domain, as well as the foundation elements,
is discretised by ten-node tetrahedral elements, while twonode elastic anchor elements, three-node line beam elements
and six-node triangular plate elements are used to model the
structure and some components of the tunnel (i.e. shield and
lining).
According to the in-situ stratigraphy, the soil profile is
constituted by two layers of gravelly sand (between 0 and
20 m, and between 25 and 30 m) and a layer of sandy silt
(between 20 and 25 m); the imposed hydrostatic water table
is 15 m below the ground surface.
The tunnel axis is located at a depth of z0 15 m as in
the reference case of study, in correspondence with section
GS_6 (see Fig. 6); the tunnel (D 6 .7 m) underpasses the
reference building, which is included in the model. Consistently with the real tunnelstructure relative position, the
inclination of the building longitudinal sides with respect to
the tunnel axis is set as equal to 25 .148 and, according to
the reference system shown in Fig. 11, the x and y coordinates of the structures middle point, M (see Fig. 3), are
equal to 0 m and 35 m, respectively. The building is assumed
to be directly connected to the soil at the foundation level,
situated 4 m below the ground surface.
The numerical analyses were performed in terms of effective stresses, assuming drained conditions for the soil owing
to the relatively high permeability observed during the
geotechnical investigation.

(b)

2
4
6
8

(c)

Recording date:
7 January 2013

10 January 2013

13 January 2013

8 January 2013

11 January 2013

14 January 2013

9 January 2013

12 January 2013

15 January 2013

Fig. 10. Structural vertical displacements recorded in correspondence with the monitoring targets: (a) L1L5; (b) R1R5;
(c) T1T3. Settlements recorded at targets T1T3 before 12 January 2013 are equal to zero

FARGNOLI, GRAGNANO, BOLDINI AND AMOROSI

30

Tunnel axis projection


Tunnel face
final position
30 m

y
x

10
0m

68 m

(a)

y
z

(b)

Fig. 11. (a) Sketch of the mesh used in the interaction analysis.
(b) A detail of the longitudinal section on the finite-element model
is also shown

Soil constitutive model calibration


The mechanical behaviour of the soil strata is described by
the hardening soil model with small-strain stiffness
(HSsmall), a constitutive model capable of taking into account the high soil stiffness observed at very low strain levels,
its variation with strain level and the early accumulation of
plastic deformations. Details of the main aspects of its
formulation are provided in Appendix 1.
A summary of all model parameters and corresponding
values is provided in Table 4. The total unit volume weight and
the strength parameters (c9 and 9) of the soils were determined as discussed in the previous section on Geotechnical
conditions. For the sake of simplicity, the same total unit
volume weight value is assumed for the soils above and below
the water table. The initial profile of the horizontal effective
stress was calculated using K nc
0 values defined in Table 4.
The variation of the small-strain stiffness with depth is
and m against
obtained by calibrating the parameters Gref
0
the down-hole experimental results, as shown in Fig. 6.
The dependency of the shear stiffness on the strain level
(equation (3)) is obtained by referring to the experimental
curves G/G0 proposed by Vucetic & Dobry (1991) for
granular soils (index of plasticity, IP 0) and for material of
low plasticity (index of plasticity, IP 15 %), respectively,
for the layers of gravelly sand and sandy silt. The reference
value of the Youngs modulus at small strains, E9ref
0 , is related
to Gref
0 by the Poisson ratio for unloading/reloading, ur. This
latter is assumed as equal to 0 .2 and 0 .25 for the gravelly
sand and for the sandy silt, respectively. For both materials
the coefficient of earth pressure at rest is estimated with
reference to a normal consolidated state (K nc
0 ). In the absence
of laboratory experimental data, the reference unloading/
ref
.
reloading stiffness, E9ref
ur , is assumed to be 0 24E90 for the
ref
.
gravelly sand and 0 42E90 for the sandy silt, corresponding
to the stiffness values observed along the decay curves at a
shear strain of 0 .1%. The other stiffness parameters, E9ref
50 and
ref
E9ref
oed , are assumed to be three times lower than E9ur : Finally,

Table 4. Soil constitutive model parameters


Parameters

Name

Values
Gravelly sand soil S_1 Sandy silt soil L Gravelly sand soil S_2

: kN/m3

Total unit volume weight

20

17 .5

20

0
33
0

5
26
0

0
33
0

0 .4
48 000
48 000
144 000
0 .2
250 000
0 .0001

0 .85
54 250
54 250
162 750
0 .25
155 000
0 .0002

0 .4
58 944
58 944
176 832
0 .2
307 000
0 .0001

100
0 .455
0 .9
0
0

100
0 .562
0 .9
0
0

100
0 .455
0 .9
0
0

Failure parameters:
c9: kPa
9: degrees
: degrees

Effective cohesion
Effective friction angle
Dilatancy angle

Stiffness parameters:
m
E9ref
50 : kPa
E9ref
oed : kPa
E9ref
ur : kPa
ur
ref
G0 : kPa
0 .7

Power for the stress-level dependency of stiffness


Reference secant stiffness in standard drained triaxial test
Reference tangent stiffness for primary oedometer loading
Reference unloading/reloading stiffness at engineering strains
Poisson ratio
Reference shear modulus at very small strains
Shear strain at which Gs 0 .7G0

Other parameters:
pref: kPa
K nc
0
Rf
tension
cincrement: kPa/m

Reference stress for stiffness


K0 value for normal consolidation
Failure ratio
Tensile strength
Increase of cohesion with depth

MODELLING OF SOILSTRUCTURE INTERACTION DURING EPB TUNNELLING


standard values are taken for the other parameters of Table 4;
for all soils the overconsolidation ratio is fictitiously assumed
to be large enough to exclude yielding during compressive
stress paths (i.e. excluding the activation of the cap surface
included in the constitutive model).
Numerical schematisation of TBMEPB tunnelling
In all the numerical analyses, the following sequence is
considered

initialisation of the stress field in the soil (lithostatic


conditions)
activation of the structure in a single step (only in the
interaction analyses)
tunnel excavation in several steps (in the first step the
displacement field due to the weight of the building is
reset to zero).

The simplified numerical procedure adopted to model the


tunnel construction is illustrated in Fig. 12. The first portion
of the tunnel cavity is lined by a steel shield, which extends
for a total length of 9 .8 m (i.e. seven slices) and it is
connected to the soil by way of an interface characterised by
the strength parameters of the adjacent soil. At the back of
the shield the permanent reinforced concrete lining is installed. The shield and the lining are modelled by means of
plate structural elements, characterised by isotropic linear
elastic behaviour, whose properties are listed in Table 5.
The excavation is simulated by a step-by-step procedure
consisting in 43 advancements, each having the length of
one concrete lining ring (1 .4 m), from y 9 .8 m to 70 m.
The advancement consists in removing one slice of soil
inside the tunnel and imposing dry conditions. A pressure is
applied at the new tunnel face, corresponding to the estimated total horizontal stress acting at rest h0(z), which
ranges from 106 kPa at the tunnel crown to 185 kPa at the
invert, according to the average face pressure values recorded during the tunnel construction.
In the intermediate zone between the shield tail and the
permanent lining, a region of 1 .4 m of unlined soil is
supported by a uniform pressure of 172 kPa, this latter
corresponding to the mean value of the grouting pressure as
recorded during the excavation.
Lining Grouting

31

In order to control the subsidence volume at the ground


surface, a fictitious contraction is applied along the shield
starting from the second slice. Such a contraction, which
induces greater displacements at the top of the circumference and lower ones at the bottom, is characterised by a
constant increment along each slice, aiming at reproducing
in a simplified way the shield conical geometry. The application of a displacement field at the tunnel section, however,
does not exclude the importance of the adopted constitutive
model, especially for the low values of volume loss that
characterise the case under study (Fargnoli et al., 2014).

Numerical modelling of the building


The building is introduced in the numerical scheme (STR
analysis) by modelling its main structural components as
follows

beams and columns are modelled by beam elements


plate elements with isotropic behaviour are used for the
floor slabs, reinforced concrete interior panels, elevator
shafts, stairwell and retaining walls
foundations are represented by volume elements constituted by non-porous material.

A linear-elastic constitutive law is adopted for these


structural components, whose parameters are selected consistently with the reinforced concrete material properties: unit
volume weight, c 24 kN/m3, Youngs modulus, Ec
25 GPa, Poisson ratio, c 0 .2.
The pressure distribution of the structure at the foundation
level is shown in Fig. 13.
The building is characterised by the presence of infill
panels, uniformly distributed in the external frames, which
are modelled in a simplified way by means of equivalent
cross-bracings (Panagiotakos & Fardis, 1996), characterised
by properties defined in Appendix 2.
Simplified structural models were also considered in the
numerical study: the building was first modelled without
cross-bracings (STRwcb analysis), then an analysis was carried out limiting the modelling of the structure to its buried
portion including the foundation elements, reducing the
upper portion to an equivalent load distribution (STRw
analysis). These loads were evaluated with reference to their
influence area and were applied at the ground floor in

Shield

Face pressure

40 kPa

VI

14 m slice
14 m 14 m

80 kPa

98 m

VII

Fig. 12. Numerical procedure adopted to simulate TBMEPB


tunnelling

120 kPa

VII
160 kPa

Table 5. Shield and lining properties


Parameter
Thickness: m
Unit volume weight: kN/m3
Poisson ratio
Youngs modulus: GPa

Shield

Lining

0 .03
75
0 .25
210

0 .3
25
0 .15
35

IV

200 kPa

III
240 kPa

II
I

Fig. 13. Pressure distribution at foundation level

32

FARGNOLI, GRAGNANO, BOLDINI AND AMOROSI


0

8
I
12
30

20

II VII III

IV

10
0
10
Distance from tunnel axis, x: m
(a)

20

30

II

Settlement, Sv: mm

NUMERICAL RESULTS AND COMPARISON WITH


MONITORING DATA
Results of free-field analysis and interaction analysis with a
detailed structural model
A preliminary free-field numerical analysis (FF) was
performed to calibrate the contraction to be applied at the
tunnel shield to reproduce a volume loss equal to 0 .34%,
corresponding to that observed at the monitored ground
section GS_6 (see Table 1) near to the building. Such
contraction induces a maximum vertical displacement at the
tunnel crown, in correspondence with the shield tail, equal
to about 10 mm, which is compatible with the available gap
of the adopted EPB machine.
To validate the model, the computed transversal and longitudinal surface settlement profiles are compared to the measurements and the corresponding empirical curves, as shown
in Figs 8(f) and 9(b). In general the comparison is satisfactory. The computed transversal profile (Fig. 8(f)) results in
fair agreement with the Gaussian distribution; however, the
accordance with the experimental measurements decreases as
the distance from the tunnel axis increases. The numerical
longitudinal subsidence trough (Fig. 9(b)), which is quite
similar in shape to the translated one, is able to capture the
face and final recorded settlements; nonetheless, it predicts
the attainment of steady-state conditions at a shorter distance
from the face when compared to what was measured.
The overall consistency of the computed profiles with the
empirical solutions and with available measurements indicates that the adopted numerical model is amenable to be
adopted for more complex interaction analyses, as illustrated
in the following.
The interaction analysis, defined as STR, was carried out
imposing the same excavation sequence and amount of
contraction as defined above, but in the presence of the
building. It was performed considering the appropriate selfweight and stiffness of the structural elements, including in
the model the presence of the external infill panels by means
of specific cross-bracings.
The numerical final settlement troughs, as computed along
the transversal and longitudinal directions to the tunnel axis
at the foundation level (i.e. z 4 m), are compared to the
corresponding free-field predictions in Figs 14(a) and 14(b),
which also show the plant position of the foundation elements by dashed lines. In particular, Fig. 14 refers to the
subsidence profiles as computed at the barycentre of the
building (point M in Fig. 3).
As expected, the presence of the building influences the
settlement profiles, which deviate from the free-field ones
along both directions, highlighting the stiffer response observed in correspondence with the discrete foundation elements. The two profiles overlap only outside the building
area. It is worth noting that the maximum vertical displacement and volume loss of the interaction analysis are larger
than the free-field one, due to the effect of the building
weight. This observation is particularly evident at the stairwell (VII in Fig. 14(a)) and at the elevator shaft (VI in the
same figure), that is, in correspondence with the heaviest
components of the building.
Figure 15 shows the measured and calculated settlements

Settlement, Sv: mm

correspondence with the columns head, the stairwell and the


elevator shafts.
In a further analysis the structure was strongly simplified and schematised as an equivalent plate (L 30 m and
B 12 m) placed at the foundation level (STR analysis),
whose input parameters were derived adapting the approach proposed by Franzius et al. (2006), as discussed in
Appendix 3.

FF_VL 032%
STR_VL 036%

12

30

20

III

VI

10
0
10
20
30
Distance from tunnel axis, y: m
(b)

40

50

Fig. 14. Comparison between computed final settlement profiles


of FF and STR analyses evaluated along the middle section of the
building in (a) the transversal and (b) longitudinal directions to
the tunnel axis (the corresponding volume loss values are also
reported in the key)

at the ground surface along the left and right longitudinal


facades of the building and along its transversal right side,
as obtained for the tunnel face located at the middle of the
structure (point M in Fig. 3) and at the end of tunnel
excavation. In both cases, a satisfactory agreement between
the numerical outcomes of analysis STR and the monitoring
data is observed, confirming the capability of the proposed
finite-element simulation to provide realistic results. In contrast, the free-field results lead to less intense settlements
and rather overestimated differential ones.
An attempt to analyse the structural response, in terms of
normal compression forces (N) acting at the base (z 0) of
the columns located along the longitudinal facades of the
building, is presented in Figs 16(a) and 16(b) together with
their computed final settlements. The N values predicted
before tunnelling are approximately constant, their distribution being more regular along the left side of the building
due to the corresponding more regular column distribution.
Once the excavation process has been completed, the distribution of N becomes modified: in general, N decreases for
the columns that experience larger settlements, whereas it
increases for those which settle the least. This expected
pattern, which is more evident for the left facade, should be
ascribed to the force-transfer mechanism, which is enhanced
by the presence of the cross-bracings.

Results of interaction analyses with simplified structural


models
Additional interaction analyses were carried out adopting
simplified schematisation of the building, as described in the
previous subsection on Numerical modelling of the building,
in order to highlight the role of the structural components on
the overall stiffness and, thus, on the computed displacement
field. In particular, the following cases were also considered.

The stiffening role of the cross-bracings was investigated

MODELLING OF SOILSTRUCTURE INTERACTION DURING EPB TUNNELLING

Settlement, Sv: mm

Settlement, Sv: mm

12

33

10
15
20
Longitudinal left side, l: m
(a)

25

12

30

10
15
20
Longitudinal right side, l: m
(b)

25

30

Settlement, Sv: mm

Measurements_middle
Measurements_end
FF (point M)
STR (point M)
FF

STR
12

6
9
Transversal right side, l: m
(c)

12

15

Fig. 15. Comparison of monitored and computed settlements of FF and STR analyses on the longitudinal (a) left and
(b) right sides and (c) on the transversal right side of the building

STR analysis

STR analysis

1400

1400

800
600

Normal force, N: kN

Settlement, Sv: mm

1000
800
600

10
15
20
Building left side, l: m
(a)

25

0
30

200

1400

STRwcb analysis
1400

10
15
20
Building right side, l: m
(b)

25

0
30

STRwcb analysis

1200

1000
800

600
400

Normal force, N: kN

1200

Settlement, Sv: mm

Normal force, N: kN

400

400

200

1000

800
600

Settlement, Sv: mm

Normal force, N: kN

1000

Settlement, Sv: mm

1200

1200

200

400
0

10
15
20
Building left side, l: m
(c)
Column position

0
30

25

Sv

200

N before tunnelling

10
15
20
Building right side, l: m
(d)

25

0
30

N after tunnelling

Fig. 16. STR and STRwcb analyses: normal compression force and settlement values at the base of the columns
on the left (a, c) and right (b, d) longitudinal sides of the building

FARGNOLI, GRAGNANO, BOLDINI AND AMOROSI

34

by the analysis denoted as STRwcb, where the building


was modelled without these components.
Analysis STRw investigates the stiffening contribution of
the buried portion of the building, characterised by the
presence of foundation elements and retaining walls, the
latter being 0 .4 m thick and 3 .5 m high, connecting
the foundation level to the ground floor. In this analysis
the weight of the elevated structure was accounted for by
a corresponding distribution of loads, as described earlier
under Numerical modelling of the building.
The building was finally reduced to a plate of equivalent
stiffness and weight in analysis STR , as described later,
in Appendix 3.

Figure 17 summarises all the monitored and computed


settlement profiles as observed and back-predicted along the
longitudinal and transversal sides of the building.
The comparison between the STRwcb and STR results
proves that the absence of the cross-bracings does not significantly affect the overall displacement pattern (Fig. 17),
while it can play a non-negligible role on the structural
forces. As shown in Figs 16(c) and 16(d), the absence of
these elements reduces the force redistribution process within the structure: at the end of excavation, in fact, the N
distribution is characterised by a different pattern as compared to analysis STR (Figs 16(a) and 16(b)), with relatively
higher values in the inner columns for the left side of the
building (Fig. 16(c)) and less intense actions on the external
columns along the right side (Fig. 16(d)).
The displacement curves obtained by the analysis STRw,
carried out disregarding the above-ground portion of the
building, are very similar to the STR ones, thus indicating

that, in this particular case, the buried portion of the


structure provides the most relevant contribution to the
overall stiffness. In particular, the differential settlements
along the transversal sides of the building, in correspondence with the foundational elements, are practically coincident with those computed by the complete structural
model.
Finally, the results obtained using the equivalent plate
schematisation are highly unsatisfactory and on the unsafe
side, since the building stiffness is found to be largely overestimated. In fact, the displacement field at the foundation
level is characterised by almost rigid rotations along the four
sides of the structure, without indicating any sagging or
hogging deformative modes.

CONCLUSIONS
This paper presents the results of a coupled geotechnical
and structural numerical study aimed at investigating the
response of a multi-storey building affected by tunnellinginduced settlements. This topic is currently relevant as the
ever-increasing demand for urban space leads to underground developments, which represent a possible cause of
structural damage for surface structures. The study, conducted using a three-dimensional finite-element code, refers
to the recent construction of the metro-line 5 in Milan
(Italy).
The soil behaviour is described by a non-linear elastoplastic constitutive model (termed hardening soil with
small-strain stiffness) calibrated with reference to in-situ
tests. The main aspects of the excavation process are reproduced in the 3D numerical simulation of the EPB tunnelling.
0

Settlement, Sv: mm

Settlement, Sv: mm

12

10
15
20
Longitudinal left side, l: m
(a)

25

10
15
20
Longitudinal right side, l: m
(b)

25

30

Settlement, Sv: mm

Settlement, Sv: mm

12

30

12

6
9
Transversal left side, l: m
(c)
Measurements_end

12

15

12

6
9
Transversal right side, l: m
(d)

STR

STRwcb

STRW

STR*

12

15

Fig. 17. Comparison of monitored and computed final settlements on the longitudinal (a) left and (b) right sides, and
(d) on the transversal right side of the building. (c) The computed settlement profiles on the transversal left side are
also compared

MODELLING OF SOILSTRUCTURE INTERACTION DURING EPB TUNNELLING


One novel element of the proposed study is the detail
adopted in modelling an existing reinforced concrete building underpassed by the metro-line. The mechanical and
geometrical properties of its principal structural elements are
described in a realistic way, including the external infill
panels, which are schematised by means of weightless crossbracings with equivalent stiffness.
The problem under investigation is initially discussed
with reference to the analysis of the geotechnical and
structural monitoring data, collected during the tunnelling
activities at six ground locations and along three sides of
the reference building. It was found that the ground and
the structure experienced a maximum settlement at the end
of the excavation process always lower than 7 mm, while
the average volume loss value was equal to 0 .33%, indicating a good performance of the adopted EPB machine.
The ground vertical displacements can be well interpreted
by Gaussian curves both along transversal and longitudinal
directions, in this latter case assuming a ratio between the
settlement at the tunnel face and the final settlement much
lower than the standard 0 .5, due to the effect of the support
face pressure.
The paper presents and discusses a class C numerical
prediction of the settlements induced by tunnelling under
free-field conditions and, subsequently, in the presence of a
surface structure. The satisfactory comparison between the
numerical results and settlement measurements, including
those along the structure, proves the reliability of the
proposed finite-element model to capture the essential
mechanisms governing the problem.
As expected, the numerical results for the interaction
analysis clearly highlight the role of the structure stiffness
and weight on the settlement troughs, as compared to the
free-field ones.
It is also demonstrated that the contribution of the infill
panels appears to be negligible in terms of the overall
displacement pattern, whereas it is shown to play a more
significant role in the redistribution of the structural forces
acting in the vertical columns of the building during the
excavation process.
The model which considers only the buried portion of the
building, including the foundation elements, is found to fit
the monitoring data well: in terms of displacement pattern, it
provides almost equivalent results to those obtained by the
complete analysis, highlighting the negligible stiffening role
of the structure above in this reference case study.
In contrast, for this particular building and foundation
typology, the equivalent plate schematisation involves a large
overestimation of the structure stiffness, resulting in a highly
inaccurate displacement field as compared to that observed
in situ.

35

dependent Youngs modulus, which is a function of the effective


stress and strength parameters according to the following expression


c9 cos 9  39 sin 9 m
E9ur E9ref
(1)
ur
c9 cos 9 pref sin 9
where E9ref
ur is the unloading/reloading Youngs modulus at the
reference pressure pref 100 kPa, c9 is the effective cohesion, 9 is
the angle of shearing resistance,  39 is the minimum principal
effective stress and m is a constant that controls the linear or nonlinear dependency of the stiffness on the above quantities. Similar
expressions are used to define the secant stiffness in standard drained
triaxial test, E950 , and the tangent stiffness for primary oedometer
loading, E9oed :
The HSsmall model considers two additional stiffness parameters
to take into account the soil behaviour at very small strains: the
reference small-strain shear modulus, Gref
0 (i.e. the small-strain shear
modulus G0 at the reference pressure pref) and the shear strain at
which the shear modulus is reduced to about 70% of its initial value,
0 .7.
In the case of the small-strain stiffness, the stress dependency is
accounted for by a power law which resembles the ones discussed
above for the other stiffness parameters


c9 cos 9  39 sin 9 m
G0 Gref
(2)
0
c9 cos 9 pref sin 9
The HSsmall model implements the following modified version
(Santos & Correia, 2001) of the stiffness reduction curve proposed
by Hardin & Drnevich (1972)
Gs
1

G0 1 aj=0.7 j

(3)

where Gs is the secant shear modulus and a is a constant equal to


0 .385.
The derivative of equation (3) with respect to the shear strain
provides the tangent shear modulus, Gt, which is bounded by a lower
limit to scale back the stiffness to the value adopted at strain levels.
The lower cut-off of Gt is introduced at the unloading/reloading
shear stiffness, Gur
Gt . Gur

Eur
2(1 ur )

(4)

The HSsmall model is characterised by two yield surfaces which


evolve isotropically: a shear hardening yield surface, fs, which is a
function of the deviatoric plastic strain and a cap yield surface, fv,
which is introduced to bound the elastic region for compressive
stress paths and depends on the plastic volumetric strain. The elastic
region of the model can be further reduced by means of a tensile
cut-off.
The associate flow rule is adopted for the cap yield surface fv,
whereas a non-associate rule is employed for fs, adopting a
formulation inspired by the well-known stress-dilatancy theory.

APPENDIX 2. DEFINITION OF THE CROSS-BRACINGS


PROPERTIES
The width, bw, of the cross-bracings is defined with reference to
the expression proposed by Mainstone (1971)

ACKNOWLEDGEMENTS
Financial support provided by Astaldi S.p.A. in the person
of Eng. Enrico Campa is gratefully acknowledged.
Special thanks go to Eng. Giuseppe Colombo of Milano
Serravalle Milano Tangenziali S.p.A. (formerly Astaldi
S.p.A.) and to Eng. Davide Fraccaroli and Eng. Alessandro
Caffaro of Astaldi S.p.A. for providing the monitoring data
and for the technical support during the site activity.

APPENDIX 1. SOIL CONSTITUTIVE MODEL


In the following, a summary of the main ingredients of the
hardening soil model with small-strain stiffness is provided, based on
Benz (2007).
The elastic behaviour of the soil at medium strain levels (typically
over 0 .1%) is accounted for by isotropic elasticity using a stress-

bw 0.175(h hw )0 4 d w

(5)

where dw is the diagonal length of the panel, hw is the panel height


and the parameter h is equal to
s
4 E w t w sin(2)
h
(6)
4Ec I c hw
with Ew and Ec being the Youngs moduli of the infill panel and of
the reinforced concrete structural elements surrounding the panel,
respectively; is the angle formed by the diagonal of the infill panel
with respect to the horizontal axis; tw is the panel thickness (equal to
40 cm); and Ic is the moment of inertia of the columns adjacent to
the infill panel. A reduced value of Ew 3 GPa was entered in
equation (6) instead of the effective Youngs modulus of the infill
panels (about equal to 6 GPa) in order to take into account the
diffuse presence of voids (doors or windows) on the building

FARGNOLI, GRAGNANO, BOLDINI AND AMOROSI

36

facades, which contribute to reducing the overall stiffness of the


structures (Melis & Ortiz, 2003).
The cross-bracings are modelled as weightless one-dimensional
node-to-node anchor elements reacting solely to axial stresses and
characterised by an axial stiffness equal to Kw Ew 3 bw 3 tw. An
elasto-plastic constitutive law is selected for such elements to
introduce a limit value of the tensile strength equal to zero.
Furthermore, the maximum value of the compression strength is
evaluated according to the following expression
F lim 1.30cr Aw
(7)
where cr is the shear cracking stress of the panel, assumed equal to
0 .2 MPa, and Aw is its transversal area, evaluated as the product of
the panel length, lw, per its thickness, tw.

APPENDIX 3. DEFINITION OF THE EQUIVALENT


PLATE PROPERTIES
The axial (EcA)building and bending (EcJ)building stiffnesses of the
building were calculated by considering that the structure consisted
only of floor slabs and was oriented with the longitudinal sides
parallel to the tunnel axis (such an hypothesis does not significantly
affect the second moment of area of the slab):
(Ec A)building

n
X

(Ec A)slab

(8)

(Ec J )building

n
X

(Ec J )slab Ec

n
X

(J slab Aslab H 2m )

(9)

where n is the reference level of the building; Aslab and Jslab are the
cross-sectional area and the second moment of area of the slab at
each level, respectively; and Hm is the vertical distance between the
slabs and the structures neutral axes (the latter assumed to be
located in correspondence with the structures centroid).
The building foundation system was neglected in this simplified
approach (Franzius et al., 2006). The computed axial and bending
stiffness for each slab are reported in Table 6, together with the
thickness and Hm values.
The input parameters of the plate element used in the finiteelement analysis were then evaluated as
s
12(Ec J )building
tfe
(10)
(Ec A)building
Efe

(Ec A)building
tfe

(11)

with tfe and Efe being the equivalent thickness and the Youngs
modulus, respectively.
The unit volume weight of the plate element, which is equal to
2 .92 kN/m3, was calculated as the ratio between the total building
weight (excluding the weight of the retaining walls modelled in
STR analysis, equal to about 2500 kN) and the plate volume
(B 3 L 3 tfe).
Table 6. Stiffness properties of the slabs at each level
Level, n

Basement floor
Ground floor
First floor
Second floor
Third floor
Fourth floor
Fifth floor
Sixth floor
Seventh floor
Eighth floor
Ninth floor

Slab
thickness: m

Hm: m

EcAslab: kN

EcJslab: kNm2

0 .26
0 .26
0 .22
0 .22
0 .22
0 .22
0 .22
0 .22
0 .22
0 .22
0 .22

15 .70
13 .20
9 .00
5 .80
2 .60
0 .60
3 .80
7 .00
10 .20
13 .40
16 .60

1 .95 3 108
1 .95 3 108
1 .65 3 108
1 .65 3 108
1 .65 3 108
1 .65 3 108
1 .65 3 108
1 .65 3 108
1 .65 3 108
1 .65 3 108
1 .65 3 108

1 .10 3 106
1 .10 3 106
6 .66 3 105
6 .66 3 105
6 .66 3 105
6 .66 3 105
6 .66 3 105
6 .66 3 105
6 .66 3 105
6 .66 3 105
6 .66 3 105

NOTATION
Aslab
Aw
B
bw
c9

cross-sectional area of slab


transversal area of infill panels
structure width
width of cross-bracings
soil effective cohesion
cincrement increase of soil cohesion with depth
D tunnel diameter
dw diagonal length of infill panels
Ec Youngs modulus of reinforced concrete
Efe Youngs modulus of equivalent plate
E9ref
reference tangent stiffness of soil for primary oedometer
oed
loading
Eur unloading/reloading Youngs modulus of soil
E9ref
reference unloading/reloading stiffness of soil
ur
Ew Youngs modulus of infill panels
E9ref
reference Youngs modulus of soil at small strains
0
E9ref
reference secant stiffness of soil in standard drained
50
triaxial test
Flim maximum compression strength of cross-bracings
fs shear hardening yield surface
fv cap yield surface
Gs secant shear modulus of soil
Gt tangent shear modulus of soil
Gur unloading/reloading shear modulus of soil
G0 shear modulus of soil at small strains
Gref
reference shear modulus of soil at small strains
0
H structure height
Hm vertical distance between the slabs and structures neutral
axis
hw height of infill panels
Ic moment of inertia of the columns
IP index of plasticity
ix transversal inflection point
iy longitudinal inflection point
Jslab second moment of area of slab
K trough width parameter
Kw axial stiffness of cross-bracings
K0 coefficient of earth pressure at rest
K nc
coefficient of earth pressure at rest in a normal
0
consolidated state
k permeability coefficient
L structure length
M structure middle-point
m power of stress-level dependency of stiffness
N normal compression force in building columns
n reference level of building
pref reference pressure
Rf failure ratio
Sv settlement
Sv,max maximum settlement
Sv,f settlement at tunnel face
tfe thickness of equivalent plate
tw thickness of infill panels
VL volume loss
z0 depth of tunnel axis
max maximum angular distortion
shear strain, unit volume weight for soil
c unit weight of volume for reinforced concrete
0 .7 shear strain at which Gs 0 .7G0
max maximum relative deflection
(/L)max maximum deflection ratio
smax maximum differential settlement
angle formed by diagonal of infill panel with respect to
horizontal axis
max maximum slope
h parameter depending on geometrical and material properties
of infill panels and reinforced concrete structural elements
c Poisson coefficient of reinforced concrete
ur unloading/reloading Poisson coefficient
tension tensile strength
 39 minimum principal effective stress
cr shear cracking stress of the infill panels
9 effective friction angle
dilatancy angle
rotation

MODELLING OF SOILSTRUCTURE INTERACTION DURING EPB TUNNELLING


REFERENCES
Amorosi, A., Boldini, D., de Felice, G. & Malena, M. (2012).
Tunnelling-induced deformation in a masonry structure: a numerical approach. In Proceedings of the 7th international symposium on geotechnical aspects of underground construction in
soft ground, Rome (ed. G. Viggiani), pp. 353359. London, UK:
Taylor & Francis.
Amorosi, A., Boldini, D., de Felice, G., Malena, M. & Sebastianelli,
M. (2014). Tunnelling-induced deformation and damage on
historical masonry structures. Geotechnique 64, No. 2, 118130,
http://dx.doi.org/10.1680/geot.13.P.032.
Ata, A. A. (1996). Ground settlements induced by slurry shield
tunnelling in stratified soils. In Proceedings of North American
tunnelling 96 (ed. L. Ozdemir), vol. 1, pp. 4350. Rotterdam,
the Netherlands: Balkema.
Attewell, P. B. & Woodman, J. P. (1982). Predicting the dynamics
of ground settlement and its derivatives caused by tunnelling in
soil. Ground Engng 15, No. 8, 1322.
Benz, T. (2007). Small-strain stiffness of soils and its numerical
consequences. PhD thesis, Universitat Stuttgart, Germany.
Boscardin, M. D. & Cording, E. J. (1989). Building response to
excavation induced settlement. J. Geotech. Engng 115, No. 1, 121.
Burd, H. J., Houlsby, G. T., Augarde, C. E. & Liu, G. (2000).
Modelling tunnelling-induced settlement of masonry buildings.
Proc. Instn Civ. Engrs Geotech. Engng 143, No. 1, 1729.
Burland, J. B. (1995). Assessment of risk of damage to buildings
due to tunnelling excavation. Proceedings of the 1st international conference on earthquake geotechnical engineering, IS
Tokyo 95 (ed. K. Ishihara), pp. 11891201. Tokyo, Japan:
ICSMFE.
Burland, J. B. & Wroth, C. P. (1974). Settlements on buildings and
associated damage. Proceedings of conference on settlement of
structures, pp. 611654. Cambridge and London, UK: Pentech
Press.
Burland, J. B., Broms, B. B. & de Mello, V. F. B. (1977). Behavior
of foundations and structures. Proceedings of the 9th international conference on soil mechanics and foundation engineering,
IS-Tokyo 77, vol. 2, pp. 495546. Tokyo, Japan: ICSMFE.
Fargnoli, V., Boldini, D. & Amorosi, A. (2013). TBM-tunnelling
induced settlements in coarse-grained soils: The case of the
new Milan underground line 5. Tunnelling Underground Space
Technol. 38, No. 336, 347347.
Fargnoli, V., Boldini, D. & Amorosi, A. (2014). Numerical prediction of tunnelling-induced settlements in coarse-grained soils.
Numerical methods in geotechnical engineering, NUMGE 2014
(eds M. A. Hicks, R. B. J. Brinkgreve and A. Rohe), vol. 2, pp.
827832. London, UK: CRC Press, Taylor & Francis Group,
Balkema.
Franzius, J. N., Potts, D. M., Addenbrooke, T. I. & Burland, J. B.
(2004). The influence of building weight on tunnelling-induced
ground and building deformation. Soils Found. 44, No. 1, 2538.
Franzius, J. N., Potts, D. M. & Burland, J. B. (2006). The response
of surface structures to tunnel construction. Proc. Instn Civ.
Engrs Geotech. Engng 159, No. 1, 317.
Giardina, G., Hendriks, M. A. N. & Rots, J. G. (2010). Numerical
analysis of tunnelling effects on masonry buildings: the influence of tunnel location on damage assessment. Advd Mater. Res.
133134, 289294.
Hardin, B. & Drnevich, V. (1972). Shear modulus and damping in
soils: design equations and curves. J. Soil Mech. Found. Div. 98,
No. 7, 667692.
Leblais, Y. & Bochon, A. (1991). Villejust Tunnel: slurry shield
effects on soils and lining behaviour and comments on monitoring equipment. Proceedings of Tunnelling 91, London, pp.
6577. London, UK: Institute of Mining and Metallurgy.
Liu, G., Houlsby, G. T. & Augarde, C. E. (2000). 2-dimensional
analysis of settlement damage to masonry buildings caused by
tunnelling. The Struct. Engr 79, No. 1, 1925.
Liu, J., Qi, T. & Wu, Z. (2012). Analysis of ground movement due

37

to metro station driven with enlarging shield tunnels under


building and its parameter sensitivity analysis. Tunnelling Underground Space Technol. 28, 287296.
Mainstone, R. J. (1971). On the stiffnesses and strengths of infilles
frames. Proc. Instn Civ. Engrs 49, No. 2, 5970.
Mair, R. J. (1996). General report on settlement effects of bored
tunnels. In Proceedings of international symposium on geotechnical aspects of underground construction in soft ground (eds R.
J. Mair and R. N. Taylor), pp. 4353. Rotterdam, the Netherlands: Balkema.
Mair, R. J. (2008). 46th Rankine lecture. Tunnelling and geotechnics: new horizons. Geotechnique 58, No. 9, 695736, http://
dx.doi.org/10.1680/geot.2008.58.9.695.
Mair, R. J. & Taylor, R. N. (1997). Bored tunnelling in the urban
environment. State-of-art-report and theme lecture. Proceedings
of the 14th international conference on soil mechanics and
foundation engineering, Hamburg, vol. 4, pp. 23532385. Rotterdam, the Netherlands: Balkema.
Maleki, M., Sereshteh, M., Mousivand, M. & Bayat, M. (2011).
An equivalent beam model for the analysis of tunnel-building
interaction. Tunnelling Underground Space Technol. 26, 524
533.
Melis, M. J. & Ortiz, R. (2003). Consideration of the stiffness of
buildings in the estimation of subsidence damage by EPB
tunneling in the Madrid Subway. Proceedings of international
conference on response of buildings to excavation-induced
ground movements (ed. F. M. Jardine), vol. 199, pp. 387394.
London, UK: Ciria.
OReilly, M. P. & New, B. M. (1982). Settlements above tunnels in
the United Kindom their magnitudes and prediction. In
Proceedings of tunnelling 82 symposium, London (ed. M. J.
Jones), pp. 173181. London, UK: Institution of Mining and
Metallurgy.
Panagiotakos, T. B. & Fardis, M. N. (1996). Seismic response of
infilled RC frames structures. Proceedings of the 11th world
conference on earthquake engineering, Acapulco, paper no. 225.
Oxford, UK: Pergamon.
Peck, R. B. (1969). Deep excavations and tunnelling in soft ground.
Proceedings of the 7th international conference on soil mechanics and foundation engineering, Mexico City, pp. 225290.
Mexico City, Mexico: Sociedad Mexicana de Mecanica.
Plaxis (2012). Plaxis 3D users manual. Delft, the Netherlands:
Plaxis bv.
Potts, D. M. & Addenbrooke, T. I. (1997). A structures influence
on tunnelling-induced ground movements. Proc. Instn Civ. Engrs
Geotech. Engng 125, No. 2, 109125.
Puzrin, A. M., Burland, J. B. & Standing, J. R. (2012). Simple
approach to predicting ground displacements caused by tunnelling in undrained anisotropic elastic soil. Geotechnique 62, No.
4, 341352, http://dx.doi.org/10.1680/geot.10.P.127.
Rampello, S., Callisto, L., Viggiani, G. & Soccodato, F. (2012).
Evaluating the effects of tunnelling on historical buildings: the
example of a new subway in Rome. Geomech. Tunnelling 5, No.
3, 275299.
Santos, J. A. & Correia, A. G. (2001). Reference threshold
shear strain of soil. Its application to obtain a unique strain
dependent shear modulus curve for soil. Proceedings of the 15th
international conference on soil mechanics and geotechnical
engineering, Istanbul (eds A. Anagnostopoulos, M. Pachakis and
Ch. Tsatsanifos), vol. 1, pp. 267270. London, UK: Balkema.
Skempton, A. V. (1986). Standard penetration test procedures and
the effects in sands of overburden pressure, relative density,
particle size, ageing and overconsolidation. Geotechnique 36,
No. 3, 425557, http://dx.doi.org/10.1680/geot.1986.36.3.425.
Son, M. & Cording, E. J. (2011). Responses of buildings with
different structural types to excavation-induced ground settlements. J. Geotech. Geoenviron. Engng 137, No. 4, 323344.
Vucetic, M. & Dobry, R. (1991). Effect of the soil plasticity on
cyclic response. J. Geotech. Engng, ASCE 117, No. 1, 89107.

Vous aimerez peut-être aussi