You are on page 1of 13

Modoni, G., Croce, P. & Mongiovı̀, L. (2006). Géotechnique 56, No.

00, 1–13

Theoretical modelling of jet grouting


G . M O D O N I * , P. C RO C E * a n d L . M O N G I OV Ì †

FS
Theoretical modelling of the mechanical phenomena in- (French summary)
duced by jet grouting is presented. The analysis is devel-
oped for the single-fluid method. The jet propagation
across the space included between the injection nozzles
and the intact soil is first modelled on the basis of the
theory of submerged flows. Different possible interaction
modes between jet and soil are then assumed for gravels,
sands and clays, according to the results of previous
experimental investigations. In the case of gravels, grout
seepage is considered to be the most relevant mechanism.
For sandy soils, the injected fluid is assumed to penetrate,
for a limited extent, into the soil skeleton, producing a
considerable increment of the pore pressures and a
corresponding reduction of the grain-to-grain contact
forces. The removal of the soil particles is then triggered
by the dragging action of the fluid threads, and the
analysis is developed under drained conditions. For clayey
OO
soils, the jet action is considered as a load imposed on
the jet–soil interface, and the erosion process is modelled
as an evolving sequence of undrained failures. Theor-
etical results obtained for the different soil types are
compared with available experimental data, and the mod-
els are thus calibrated by means of back-analysis.

KEYWORDS: grouting; model tests; design; seepage; erosion

INTRODUCTION there is still a relevant degree of uncertainty at the design


Jet grouting is one of the most popular ground improvement stage, arising from the lack of reliable methods for predict-
techniques, and is currently used all over the world for many ing the diameter of the jet columns. In fact, most jet
different purposes, such as increasing the bearing capacity grouting projects are planned on the basis of some empirical
and reducing settlements of new and existing foundations, rules, which may provide only rough estimates of the
supporting open and underground excavations, and creating column diameter (Croce & Flora, 2000).
water cut-offs for dams. The method is based on high-speed Theoretical modelling of the mechanical phenomena in-
grouting of water–cement mixtures and/or other fluids (air, duced by jet grouting has thus been attempted. Considering
water) into the subsoil. The fluids are injected through the complexity of the mechanical phenomena involved, the
small-diameter nozzles placed on a grout pipe, which is analysis has been restricted to the single-fluid system.
continuously rotated at a constant rate and slowly raised The first step of the analysis is devoted to the jet propaga-
towards the ground surface. The jet propagates radially from tion across the space included between the nozzles and the
PR

the borehole axis and, after some time, the injected mortar intact soil. This space is usually filled by some fluid, of
solidifies underground, eventually producing a cemented soil various possible origin (natural groundwater, perforation
body of quasi-cylindrical shape ( jet column). water, previously injected grout, floating soil grains, etc.). At
Currently adopted jet grouting methods can be classified the beginning of treatment this fluid region is relatively thin,
according to the number of fluids injected into the sub- because the soil boundary coincides with the borehole sur-
soil: water–cement grout (single-fluid system), air + grout face. However, if soil erosion takes place, the soil boundary
(double-fluid system), and water + air + grout (triple-fluid will shift, and the fluid region will become larger. The
system). In the double-fluid system, the grout jet is wrapped evolution of the geometrical and kinematical characteristics
by a coaxial air jet, whereas in the triple-fluid system the of the jet within this zone are analysed on the basis of the
grout jet is preceded by a jet of water surrounded by air. theory of submerged flow (Hinze, 1948).
Technical improvements are continuously introduced, for After reaching the soil face, part of the injected grout
each system, in order to increase the dimensions and the may maintain its original direction (radial flow), either by
mechanical properties of the jet columns (e.g. Shibazaki, seeping through the soil pores or by displacing the soil
2003). However, in spite of such rewarding developments, grains, while the remaining grout may flow towards the
ground surface (vertical flow), passing through the annular
space bounded by the perforation hole and the injection
Manuscript received
stem. The measured percentage of vertical outflow, compris-
Discussion ing grout and some eroded soil, increases with decreasing
size of the soil grains, ranging between 0% and 80%
* University of Cassino, Cassino, Italy. (Kaushinger et al., 1992).
† University of Trento, Mesiano, Italy. For very pervious soils, such as coarse gravel, the vertical

1
Article number = 3580
2 MODONI, CROCE AND MONGIOVÌ
flow becomes irrelevant, and almost all the injected grout is the jet is deeply altered along the path from the nozzle (Fig.
retained by the soil. In fact, in the case of gravels (Fig. 2). Immediately out of the nozzle all the jet threads can be
1(a)), all the injected fluid can easily seep through the soil approximately assumed as having constant speed (v0 ) and
pores, following a nearly radial path, without significantly oriented along the x-direction. As the distance from the
displacing the soil grains (Miki, 1985; Croce & Flora, nozzles increases, because of the viscous nature of the fluids
2000). This phenomenon is modelled as a seepage problem, tangential stresses develop at the jet contour, and part of the

FS
and the relevant mechanical property is soil permeability jet energy is transferred to the surrounding fluid. As a
with respect to water–cement grout. consequence, part of the surrounding fluid mass is mobilised
For finer soils, such as sands and clays, whose interstitial in the x-direction, while the speed of the threads located
pores are much smaller, the resistance to grout seepage near the jet contour decreases. Within a small distance,
increases considerably. Therefore the jet threads tend to turn equal to a few nozzle diameters (starting zone in Fig. 2), the
back and displace the grains, which may be removed from velocity decrease affects the external part of the jet cross-
their original position and possibly dragged towards the section, while the velocity on the jet axis remains equal to
ground surface (Shibazaki, 1991; Covil & Skinner, 1994). its initial value (v xmax ¼ v0 ). After a given distance (diffu-
This is essentially an erosion process. The features of this sion zone in Fig. 2) the velocity of all the threads decreases.
complex phenomenon, however, are largely unknown, and The jet energy can be thus assumed as continuously redu-
some simplifying hypotheses are needed in order to imple- cing with the distance from the nozzle, and as being
ment the analysis. In particular, two different erosion me- concentrated in an inner region. Because of the very high
chanisms are postulated, for sands and clays respectively. speed of the injected fluids, these phenomena are highly
For sands (Fig. 1(b)), the injected fluid is assumed to turbulent, and thus the instantaneous fluid velocities are
penetrate by seepage into the soil pores to some limited regarded as a combination of mean and fluctuation compo-
extent, behind the soil face, thus producing a significant nents.
increase of pore pressures and a corresponding reduction of The nozzle diameters are very small (typically from 2 to
the grain-to-grain contact forces. The removal of the soil 4 mm), and thus the extension of the starting zone is very
particles from their initial position is then triggered by the limited. The analysis of jet propagation is therefore confined
OO
dragging action of the fluid, as has been observed from
laboratory tests on sandy soil (Bergschneider & Walz, 2003;
Stein & Grabe, 2003). This phenomenon is modelled by
continuously comparing the jet action with the soil shear
to the diffusion zone only. In particular, it is assumed that
the injected fluid has the same unit weight and viscosity as
the surrounding fluid. This simplifying assumption is based
on the consideration that both fluids are mainly suspensions
strength. The latter analysis is performed in terms of effec- of cement particles, although the surrounding fluid may
tive stresses, and the drained strength parameters c9 and 9 contain also some soil particles.
are assumed as relevant soil properties. Furthermore, with reference to the mean component of
For clayey soils (Fig. 1(c)), whose interstitial pores are fluid velocity, oriented along the x-direction, the following
very small, the jet threads cannot penetrate by seepage into relations can be written accordingly with the theory of
the soil. The jet action is thus considered as a load imposed submerged jets (Hinze, 1948; further developed in the
on the soil wall and is proportional to the momentum of the Appendix).
jet. This erosion process is then described by a sequence of v xmax ¸ d0
undrained failure mechanisms, and the undrained strength cu ¼ pffiffiffiffiffi (1)
is considered as the relevant soil property. Similar conclu- v0 N x
sions have been obtained by Dabbagh et al. (2002) after vx 1
performing a set of laboratory injection tests on clays. ¼ 2 (2)
v xmax 1:33¸2 (r=x)2

N

SUBMERGED FLOW
In the above equations N represents the turbulent kinematic
Starting from the injection nozzles, the injected fluid
viscosity ratio of injected fluid and water (N ¼ f /w ). It has
moves at very high speed (typically some hundreds of
metres per second) before falling onto the soil. Owing to an
Starting Diffusion
intensive energy transfer between the injected fluid and the
PR

zone zone
soil–fluid mix resting in the borehole, the velocity profile of
vxmax 5 v0 vxmax , v0
r

Seepage (a)

vx

Sand
(b)
erosion d0
v0 vxmax x

Clay
(c)
erosion

Fig. 1 Sketch of grout–soil interaction modes Fig. 2. Pattern of grout velocity in fluid region (submerged flow)
THEORETICAL MODELLING OF JET GROUTING 3
ð req
been experimentally observed by Shibazaki (2003) that the
2v x rdr
attenuation of vmax along the jet axis is affected by the 0 2¸ d 0
nozzle shape. In particular, a very sharp reduction of v xmax v eq ¼ ¼ pffiffiffiffiffi v0 C (4)
req2
N x
is produced by a sudden narrowing of the nozzle cross-
sectional area, while a lower rate of velocity decrease occurs where
when the entry and exit section of the nozzle are gently ð req =x pffiffiffi


FS
connected by a cone. It follows that the coefficient ¸ should 1 c
C¼ 2 2 2
dc ¼ (5)
be determined by performing specific experiments adopting (req =x) 0 [1 þ (c =k9N )] 2
various nozzle shapes. In the present study the coefficient ¸
has been evaluated by comparing equation (1) with the Note that all the relevant parameters (req, aeq , veq ) depend
experimental data published by de Vleeshauwer & Maertens on the turbulent kinematic viscosity ratio N of the injected
(2000), who measured the velocity decay along the axis of fluid. However, N is very difficult to measure, because it
several submerged water jets expelled from a nozzle of depends not only on the properties of the fluid but also on
2.2 mm diameter. These experimental data, reported in Fig. the flow characteristics of the jet. In order to estimate this
3, show the reduction of axial velocity against the ratio x/d0 parameter, some simplifying assumptions are proposed, as
for different injection pressures p0 . It can be seen that the follows.
velocity decay does not significantly depend on the feeding For this purpose it is first recalled that the kinematic
pressure p0 , and that the experimental results can be fitted viscosity of a fluid is defined as the ratio between viscosity
with equation (1) by assigning a value of ¸ equal to 16. and density. The density rg of a water–cement mix (grout)
According to equation (2) the velocity pattern in each is a function of the weight ratio W between water and
cross-section is characterised by a bell-shaped surface, pro- cement (Croce et al., 1990).
viding positive values of the mean velocity v x even for
1þW
infinitive distances r from the jet axis. However, it seems rg ¼ (6)
more realistic to assume that the jet is confined inside a 1=r c þ W =rw
conical region, and that the velocity falls to zero at the
OO
border of such cone. For the purpose of the present study
the cone angle has been estimated by neglecting jet threads
where rc and rw are the density of cement and water
respectively. For a typical water–cement ratio W ¼ 1, rc ¼
with velocity smaller than a fraction  of the maximum 3150 kg/m3 and rw ¼ 1000 kg/m3 , a density of about rg ¼
velocity v xmax: The radius of the cone req at each distance x 1500 kg/m3 is obtained.
from the nozzle can be obtained by fixing v x /v xmax ¼  in Concerning viscosity, it is recalled that cement–water
equation (2). suspensions are usually treated in terms of equivalent New-
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi tonian fluids, by defining an apparent viscosity independent
u 0sffiffiffi 1 of the fluid velocity, although it is known that they behave
u
req u N @ 1 A more like Bingham fluids (Winterkorn & Fang, 1975).
¼ f () ¼ t : 1 (3) Furthermore, because of the cement hydration, a variation of
x 1 33¸ 2 
the apparent viscosity should be expected with time, but this
can be neglected if the period between the preparation and
The cross-sectional area of the jet is thus the injection of the grout is relatively short.
It is finally assumed that the ratio N between the turbulent
0sffiffiffi 1 kinematic viscosities () of grout and water is constant with
aeq N @ 1 the jet velocity and is equal to the ratio between the laminar
¼ f () ¼ :  1A (39) kinematic viscosities ().
x 2 1 33¸ 2 
g g
N¼ ¼ (7)
w w
For the sake of simplicity it is assumed that, for a given The simplification introduced by equation (7) has been made
value of x, the velocity v x of the jet threads is constant in considering the lack of data on turbulent viscosities of
the whole cross-section and equal to a mean equivalent value suspensions, and can be removed by conducting specific
PR

veq , calculated as follows. experimental investigations.

1·0

p0 5 150 bar
0·8
p0 5 200 bar

0·6 p0 5 250 bar


vxmax/v0

p0 5 300 bar
0·4
p0 5 350 bar

0·2 p0 5 400 bar

vxmax/v0 5 16 (d0/x)
0
0 50 100 150
x/d0

Fig. 3. Reduction of mean velocity at axis of jet flowing from nozzle of diameter d0 2.2 mm (from de
Vleeshauwer & Maertens, 2000)
4 MODONI, CROCE AND MONGIOVÌ
The apparent laminar viscosity of a water–cement mix
has been measured by Raffle & Greenwood (1961, reported
by Bell, 1993), and is expressed as a function of the weight q0
ratio W (Fig. 4). In particular, for a water–cement ratio W ¼
1 a laminar viscosity ( W ¼1) ¼ 7 cP ¼ 0.007 Pa s can be
q2
found, whereas the laminar viscosity of water, w , is equal

FS
to about 1 cP ¼ 0.001 Pa s. Therefore, considering the ratio
between the viscosities and the densities of grout and water
a relative kinematic viscosity N ¼ 4.6 can be assumed for a
grout having W ¼ 1.
Once an arbitrary value of  is chosen, the evolution of
the geometrical and mechanical characteristics of the equiva- z
lent submerged jet can be simulated by equations (3), (4)
and (5) as functions of the relative kinematic viscosity N
and of the injection parameters (d0, v0 ). As an example, the Rf
variation of the radius and the velocity of an equivalent jet
R1
of water (N ¼ 1) and of grout with water–cement ratio
equal to 1 (N ¼ 4.6), obtained by eliminating the jet threads
whose velocity is lower than half their maximum value
v xmax ( ¼ 0.5), is plotted against the distance from the
nozzle in Fig. 5. It is worth observing that the plotted curves
should not be considered in the lower range of x/d0 where
velocities v xeq are higher than the injection speed v0:
q1
One of the most important features of the submerged jet
is highlighted in Fig. 5, where a much higher spreading of a1
the equivalent jet and a faster reduction of the equivalent
velocity is shown for the grout compared with water. Both
these results depend on the fact that kinematic viscosity of
the grout is higher than that of the water.
OO Fig. 6. Geometrical features of the seepage model

The phenomenon is modelled as a moving front problem,


by assuming that the injected fluid propagates horizontally
JET–SOIL INTERACTION in radial direction. The following hypotheses are also intro-
Seepage model duced.
This mechanism is the most relevant for very pervious (a) Soil skeleton and fluid are incompressible.
soils (gravels and sandy gravels). A typical feature is that (b) The injected fluid moves by filling all the soil pores.
the injected fluid may be totally or partially adsorbed by the (c) The flow is governed by Darcy’s law.
soil. The latter case occurs when the grout flows partly
through the soil pores and partly towards the ground surface, Concerning the third hypothesis, it is recalled that the
moving along the borehole (Fig. 6). Both conditions have validity of Darcy’s law for coarse-grained soils has been
been considered in the model. investigated by several authors (e.g. Rose, 1945, as reported
by Bear, 1972; Muskat, 1946; Taylor, 1948; Scheidegger,
0·14 1957), who tried to find an upper limiting value of Reynolds
Apparent viscosity: Pa s

0·12 number. These investigations have not been conclusive, how-


0·10 ever, as observed by Lambe & Whitman (1969), because the
0·08 threshold value of Reynolds number may vary over a large
0·06 range (from 0.1 to 75, according to Scheidegger, 1957). In
0·04 particular, according to the results of different authors
0·02 collected by Rose (1945), the validity of Darcy’s law may be
PR

0
accepted if the Reynolds number (defined as VD50 /, where
0·3 0·4 0·5 0·6 0·7 0·8 0·9 1·0 V is the seepage velocity,  is the kinematic laminar
Water/cement ratio by weight viscosity of the fluid, and D50 is the mean diameter of the
soil grains) is lower than 10.
Fig. 4. Apparent viscosity of water–cement mixtures measured For jet grouting, the seepage velocity is very large near
by Raffle and Greenwood (1961, as reported by Bell, 1993) the nozzle but decreases sharply as the grout penetrates
further into the soil. It may be thus inferred that the jet flow
1·0 10 propagates under turbulent conditions in the inner region of
veq req the jet column and under laminar conditions in the outer
0·8 8 one. By considering typical values of treatment parameters,
N 5 4·6
6
including grout viscosity, it can be found that Reynolds
0·6 N51
vx eq/v0

req/d0

number for gravels is larger than 10 for radial distances


0·4 4 ranging between 5 and 30 cm, depending on the average
grain size. As the typical column radii of jet grouting in
0·2 2 gravels are usually larger than these values, the present
analysis has been conducted by assuming laminar conditions
0 0
0 20 40 60 80 100 as predominant and occurring throughout the whole fluid
x/d0 path.
The geometrical features of the jet propagation into the
Fig. 5. Variation of equivalent radius and velocity along jet axis soil are very complex. In practice the nozzle is continuously
(for  0.5, ¸ 16) rotated at a fixed depth, for a time interval ˜t, and is then
THEORETICAL MODELLING OF JET GROUTING 5
lifted with a step length equal to ˜s. Therefore the impact meability of the soil to the injected fluid, which is calculated
surface on the borehole face is continuously moving during as
treatment. With regard to the fluid stream it is logical to k
assume that the seepage velocity through the soil has a kf ¼ (14)
component in the radial direction, but also components in N
the vertical and circumferential directions. where k is the Darcy coefficient of the soil, and N is the

FS
For the sake of simplicity, the solution of such a complex relative kinematic viscosity of the fluid, given by equation
problem has been pursued by assuming that the seepage (7).
propagates only in the radial direction with a fixed vertical Equation (13) can be thus integrated from the borehole
height, starting from the impact surface (Fig. 6), and consid- wall (R ¼ R1 ) to the external seepage front (R ¼ Re ) by
ering the flow in each single direction as independent. The considering equation (8).
continuity condition in the radial direction is thus written as  
V1 R1 Re
VR ¼ V1 R1 (8) h1  he ¼ ln (15)
kf R1
where V is the seepage velocity of the fluid at a point
located at distance R from the borehole axis, and the sub- he is assumed equal to the undisturbed head of the water-
script 1 refers to the borehole face. bearing strata surrounding the column, and is nil if treatment
For each radius the flow has been integrated in a time is performed in dry soils. The total head h1 at the borehole
interval t*, equal to the impact period of the jet on the wall is assumed as a function of the energy of the impacting
corresponding point of the borehole face. jet.
ma1
t ¼ (9) v2eq_1
2R1 vs h1 ¼ g (16)
2g
where m is the number of nozzles, vs (¼ ˜s/˜t) is the
average monitor withdrawal speed, and a1 is the impact area. where veq_1 is the equivalent velocity of the jet, defined by
OO
Note that, with respect to the physical phenomenon, the
simplifying assumption of radial flow induces two opposite
equation (4), calculated at the borehole wall (R ¼ R1 ), and
g is a parameter accounting for the loss of energy of the
effects, which are assumed to compensate for each other. In injected fluid while crossing the annular space between the
fact, in each single point of the soil mass, the theoretical injecting rod and the borehole wall.
seepage velocity computed by equation (8) will be higher After introducing the boundary conditions in equation
than the real velocity, but the integration time calculated (15) the seepage velocity V1 at the borehole wall can be
with equation (9) will be lower than the actual seepage time, derived as
which is determined also by the seepage occurring along
directions other than radial. k f  g (v2eq_1 =2 g)  he
V1_h ¼ (17)
The velocity of the advancing front of the grout, vf , can R1 ln(Re =R1 )
be expressed as
dRe Ve The V1 value to be inserted in equation (11) is then the
va ¼ ¼ (10) minimum between V1_q and V1_h, calculated with equation
dt n (12) and equation (17) respectively. It is possible to observe
where Re is the instantaneous radius of the column, com- that, by equalising the two expressions, a limiting distance
puted from the borehole axis, n is the soil porosity, and Ve is Rlim can be derived.
the seepage velocity at the front. " !#
From equations (8) and (10) it follows that k f a1 v2eq_1
Rlim ¼ R1 exp g  he (18)
1 V1 R1 R1 q0 2g
va ¼ (11)
n Re If Re < Rlim the injected fluid is completely adsorbed
The seepage velocity V1 at the borehole wall can de within the soil, and V1 is dictated by the continuity condition
derived by considering two alternative limiting conditions, (V1 ¼ V1_q ). If Re . Rlim the injected fluid is only partly
PR

one based on the continuity of the injected fluid, and the adsorbed by the soil, and the remaining part flows to the
other on its energy. According to the continuity condition, upper surface through the borehole. In this second case the
the fluid entering the soil cannot exceed the injected amount seepage velocity at the borehole wall is dictated by
unless other fluid mass is poured from the top into the the energy condition (V1 ¼ V1_h ). In the proposed algorithm
annular space of the borehole. As this practice is very a passage from the first condition to the second one is
uncommon in jet grouting, the maximum allowable velocity possible, along with increasing of the column radius.
for V1 is The model herein presented also allows for calculation of
q0 the treatment efficiency, defined as the ratio between the
V1_q ¼ (12) volumes of adsorbed and injected grout.
a1
nvs D2
where q0 is the grout flow rate injected from the nozzle and E¼ <1 (19)
a1 is the cross-sectional area of the jet at the borehole wall, mv0 d 20
which is obtained from equation (39) by equating x to the
borehole radius R1 . It is finally worth observing that the results of the
To derive the second conditions on V1 the energy loss of numerical calculations are in principle affected by the dis-
the fluid permeating the soil is considered. cretional parameter  previously defined in equation (3). It is
  worth noting, however, that increase of  determines an
dh
V ¼ kf  (13) increase of the equivalent jet velocity veq (equation (4)) but
dR a simultaneous reduction of the integration time t* depen-
dent on a1 (equation (9)). As a consequence, the column
where h is the total head of the fluid and kf is the per- diameters calculated by varying  between 0.1 and 0.6 differ
6 MODONI, CROCE AND MONGIOVÌ
by an amount smaller than 10%. In the following a value of crater is generated at the soil–jet interface, which propagates
0.5 is adopted for . forward during injection time. It was also observed that the
The results of several calculations performed with the velocity of the advancing front decreases with the distance
method defined above are summarised in the dimensionless from the nozzle, owing to a reduction of the erosive capacity
plot reported in Fig. 7, which shows the ratio between the of the jet. Finally, there is a limiting distance where the jet
column radius Re and the borehole radius R1 as a function is no longer capable of producing erosion. This distance is

FS
of the two dimensionless variables T and B, defined as approximately equal to the maximum allowable column
follows. radius. Both the velocity of the advancing front and the
q0  maximum penetration distance are related to the soil proper-
T¼ t ties and to the jet characteristics. In general they tend to
na1 R1
! increase with fluid speed and decrease with soil strength.
ka1 v2eq_1 In the theoretical model, the evolution of the column
B¼ g  he (20) diameter is simulated on the basis of a comparison between
NR1 q0 2g the erosive action of the jet and the resistive action of the
soil. The former is quantified by the force per unit cross-
Once the soil properties (k), the treatment parameters (v0 , section produced by the equivalent jet (Fig. 8).
N, d0, R1 , vs , m) and the boundary conditions (he ) are ªf
defined, the column radius can be simply calculated by ł ¼ 1 v2f (21)
g
referring to the plot curves. In fact each curve represents the
progress of the column radius along with time. It is worth where ªf is the unit weight of the injected fluid, vf is the jet
noting that each dotted curve detaches from the continuous velocity at the impact points, and 1 is a dimensionless
one at different time factors T, depending on the values of parameter accounting for energy dissipations (Fig. 8).
the dimensionless parameter B. This result can be explained Concerning the resistance of the soil to the erosion, a
by considering that, at the beginning of the injection process, clear distinction is made between sandy and clayey soils,
until the column radius reaches the limiting value defined by considering that different mechanisms take place depending
OO
equation (18), the injected fluid is completely adsorbed by
the soil and the seepage is governed only by the continuity
of the injected fluid. The column radius is then dependent
on the fluid mass q0 and independent of the soil permeabil-
on the soil particle dimensions. In fact, in the case of sand
the injected fluid is supposed to penetrate to some extent
into the soil pores and to surround each single particle. The
analysis is thus conducted in terms of effective stresses by
ity k. Along with increasing integration time t*, which is a assuming a drained failure for the soil. In the case of clays
function of the lifting rate of monitor vs and of the number the fluid does not penetrate into the soil pores, and the soil
of nozzles m according to equation (18), any further increase mass behaves as a continuous body. The analysis is thus
of the column radius exceeding the limiting value Rlim is performed in terms of total stresses by assuming undrained
also determined by the energy content of the injected fluid. failure of the soil.
Therefore the final value of the radius depends on such In both cases, depending on the soil strength, a limiting
energy content and on the permeability of soil, which are velocity vL is defined that identifies the minimum value of
both considered in the second dimensionless parameter B jet velocity capable of producing erosion. Strictly related to
defined in equation (20). such a limiting velocity vL and to the fluid velocity at the
soil–jet interface is the velocity Vc of the advancing erosion
front. According to the experimental results shown by Dab-
bagh et al. (2002), Vc has been expressed by the following
Erosion model empirical rule.
When the injected fluid impacts on soils with permeability
vf  vL
lower than gravel, grout seepage is largely inhibited and the Vc ¼ Æ (22)
jet threads turn back, dragging the soil aside from its initial vL
position. This erosion mechanism, together with the replace- where Æ is an empirical parameter, dependent on the soil
ment of soil particles by grout, is responsible for the growth type, that has the dimension of velocity.
of the jet columns. Experimental laboratory investigations In the following calculations the fluid velocity vf has been
PR

performed to study the effects of water jets impacting on put equal to veq (see equation (4)), thus allowing for a
different soils (Dabbagh et al., 2002; Bergschneider & Walz, reduction along with the distance from nozzles.
2003; Stein & Grabe, 2003), showed that a bell-shaped By adopting equation (22) the erosion process can be
integrated along with time and treated as a moving front
35 problem.
B5`
 Particular care must be taken, however, in defining an
30
B 5 2·0
appropriate integration time. In fact, whereas in the seepage
25 process it can be easily derived as the time when the jet acts
on a single point, these points being located on a cylindrical
20 B 5 1·0
Re/R1

15 B 5 0·5

10
B 5 0·1 Ø
5

0
vf
<
0 100 200 300 400 500
T

Fig. 7. Dimensionless plot reporting results obtained with


seepage model Fig. 8. Sketch of jet–soil interaction for erosion models
THEORETICAL MODELLING OF JET GROUTING 7
fixed surface corresponding to the borehole face, for erosion with time given by equation (23). The velocity of the
the front where the jet impacts on the soil is continuously advancing front, Vc, is continuously calculated by equation
advancing, and no unique time can be retrieved. A reference (22), where vf is put equal to veq (equation (4)), and vL is
time t* is then defined by subdividing the whole height of calculated by equation (27).
the column into slices of finite thickness ˜s, corresponding
to the length of the lifting steps chosen for treatment.

FS
m˜s 2(req =x) Clayey soils
t ¼ (23) The jet erosion for cohesive soils is assumed to be a
vs 2 global collapse, similar to the failure mechanism produced
by a distributed force acting on a circular shallow founda-
The proposed model allows also for computation of the tion. Under such a hypothesis the resistance per unit area
treatment efficiency E. This is defined as the ratio between provided by the soil is dependent on its undrained cohesion
the volume of grout in the column and the volume of cu , and can thus be expressed by the following relation.
injected grout, and is estimated by assuming a homogeneous
łc ¼ 3 cu (29)
composition of Soilcrete and spoil and a nil volume of air in
the Soilcrete. where 3 is a dimensionless parameter that relates the
1 resistance of the soil to the soil strength cu . When the jet
E¼ (24) equivalent pressure ł (equation (21)) equals the resistive
(1  n) þ (mv0 d 20 =vs D2av ) pressure łc (equation (29)), the jet velocity vf attains the
limiting value vL, which depends on soil strength cu .
where Dav is the mean value of the column along with sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
depth. c g
vL ¼ cu (30)
ªf

Sandy soils where c (¼ 3 /1 ) is a dimensionless parameter that


OO
The resistive action given by a sandy soil against the jet
erosive force is expressed in terms of effective stresses by
should be evaluated on the basis of experimental results.
Such a parameter has been experimentally evaluated for soils
the following relation. with different granular composition by Dabbagh et al.
(2002), which found values ranging between 0.36, for a clay,
łs ¼ 2 [c9 þ  9tg(9)] (25)
to 0.0001 for a sandy clay. Such a large difference could be
where c9 and 9 are respectively the effective cohesion and explained as the effect of different mechanisms responsible
friction angle of the soil, 9 is an effective stress representa- for the erosion.
tive of the field conditions, and 2 is a dimensionless It is worth observing that the function relating the critical
parameter relating the resistive action of the soil to the jet velocity to the soil strength introduced in equation (30) is
to the soil strength. In this particular case 9 has been similar to the one found by authors working on jet penetra-
calculated as the difference between the initial vertical over- tion in rocks and soils (Farmer & Attewell, 1964; Dabbagh
burden stress z and the pore pressure generated at the et al., 2002).
impact point by the injected fluid. If the limiting velocity v L is put equal to the equivalent
ªf 2 velocity of the jet, calculated by equation (4), the maximum
 9 ¼ z  v (26) allowable column radius can be obtained.
2g f
2¸Cd 0 v x0
When the jet equivalent pressure ł (equation (21)) equals Rmax ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (31)
the resistive pressure łs (equation (25)), the jet velocity vf c gNcu
attains the limiting velocity vL, which depends on the soil ªf
strength (c9 and 9) by the following function.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi The erosion process on a single direction can be now
s g c9 þ  z tan(9) simulated by integrating the velocity of the advancing front,
vL ¼ (27)
ªf 1 þ s [tan(9)=2] Vc (equation (22)), for the time given by equation (23). In
PR

equation (22) vf is put equal to veq (equation (4)), and vL is


In equation (27) s (¼ 2 /1 ) groups the unknown geome- calculated with equation (30).
trical and mechanical aspects of the jet and the soil.
Correspondingly, the maximum allowable column radius can
be obtained by equating vL with the equivalent velocity of Design chart
the jet, previously defined in equation (4). The two erosion models introduced above give the same
2¸v x0 Cd 0 function for the column radius R, its maximum values Rmax
Rmax ¼ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (28) defined by equations (28) and (31) for sandy and clayey
s gN c9 þ  z tan(9) soils respectively, and integration time t*, expressed by
ªf 1 þ s [tan(9)=2] equation (23). This function, found by performing several
numerical calculations, is plotted in Fig. 9. This plot can be
It is worth observing that, in equation (28), the injection of used as a design tool, because it makes it possible to
more viscous fluids reduces the maximum allowable column estimate the cross-sectional dimensions of columns depend-
radius, thus showing the positive effect of injecting water, as ing on the soil’s properties (c9 and 9 or cu and ªs ), on the
in the cases of pre-washing and the triple-fluid injection treatment parameters (N, ªf , v0 , d0 , vs , m), and on the
system. parameters introduced to calibrate the model ( and Æ).
In accordance with equation (22), the velocity of the These last may be based on the results of laboratory tests, as
advancing front is positive when the actual distance of shown by Dabbagh et al. (2002), or may be defined on the
the eroded wall from the nozzle does not exceed Rmax basis of observations gained directly from site investigations.
(vf . vL ). In the following section the choice of these parameters
The whole erosion process can be now integrated along has been accomplished in this second manner: that is, by
8 MODONI, CROCE AND MONGIOVÌ
3·0 0.03 m. However, the results of calculation have been shown
2·5
to be independent of the value of such distance as well as of
á * t* 5 ` 2·0 m the parameter , which has thus been fixed constant and set
2·0 1·0 m
equal to 0.5. The two parameters ¸ and k9, previously
introduced (equations (1) and (2)) to define the jet cross-
R: m

1·5 0·5 m á * t*
sectional velocity profile, have been fixed respectively at 16

FS
1·0 0·2 m and 0.003, in accordance with the experimental data shown
0·1 m
in Fig. 3. The empirical parameter g has been finally fixed
0·5 equal to 1, thus considering for simplicity the head respon-
sible for seepage as equal to the kinematic unit energy of
0 the fluid at the impact point.
0 1 2 3
Rmax: m
The minimum and maximum diameters obtained by the
calculations, which depend on the soil permeability, are in
Fig. 9. Design chart for jet grouting in sandy and clayey soils good agreement with the measured values, as they range
from 0.62 to 1.38 m.
A parametric analysis has also been performed in order to
calibrating the model on the results of some field trials investigate the influence of Darcy’s permeability coefficient
performed on different soils. k together with the role of the most relevant treatment
Figure 9 shows that the radius of columns R increases parameters (d0, v0 , vs and m) on the diameter of the jet
with time and tends to reach the maximum attainable value columns and on the efficiency of treatments. The results,
Rmax for a very long time t*. In practice t* can be increased shown in Fig. 11, refer to a soil having initial porosity n ¼
by reducing the monitor withdrawal speed or by increasing 0.3, located above the water level and injected with a grout
the number of nozzles. However, it must also be considered having relative turbulent kinematic viscosity N ¼ 4.6
that higher volumes of injected fluid generally correspond to (water–cement ratio equal to 1).
lower treatment efficiency, the latter defined by equation The different curves in each plot refer to different injec-
(24).
OO tion velocities v0 (from 180 to 380 m/s) and to different
vs /m ratios (from 0.0015 to 0.005). The effect of different
nozzle diameters can be observed by comparing Fig. 11(a)
and Fig. 11(b).
MODEL CALIBRATION On a first examination it is noticeable from all cases that
Gravelly soils the seepage mechanism is relevant only for very coarse-
The effectiveness of the seepage model has been tested by grained soils. In fact, for Darcy’s coefficients typical of
comparing the results of simulations with the results of a sands (lower than 101 cm/s), whatever the nozzle diameter,
case history from Polcevera (Croce et al., 1994), where a injection speed and injection time, the column diameters are
soil consisting of more than 90% gravel by weight was very limited. By contrast, for soils with a predominant
treated by single-fluid jet grouting. Four columns were fraction of gravel, column diameters are in the range 0.4–
discovered after treatment by excavating the surrounding 1.8 m for typical values of nozzle diameter.
soil, and their diameters were measured at different depths. Such values, similar to those typically observed on site
The average diameter of each column ranged between 1.06 (Kutzner, 1996), may be considerably increased (up to 3 m
and 1.20 m, but the variation of all the measured values was and more) by doubling the nozzle diameter. It is also notice-
consistently higher (between 0.6 and 1.50 m) (Fig. 10). The able that larger jet columns may be obtained by increasing
set of adopted treatment parameters is reported in Table 1. the injection velocity (v0 ) and the number of nozzles (m),
The soil porosity n was estimated to be equal to 0.30, and and by reducing the monitor withdrawal speed (vs ).
the permeability coefficient k was evaluated as being vari- The efficiency of treatment, which is relevant for evaluat-
able in the range between 0.5 and 4 cm/s by adopting the ing the convenience of jet grouting, is proportional to soil
relation suggested by Hazen (1911), because direct per- permeability k and to the injection velocity v0 (see equation
meability measurements were not performed. (17)). The plots also show that the efficiency reduces by an
The theoretical analysis has been performed by consider- increase in the number of nozzles and a reduction in the
ing a radius R1 of the borehole equal to 0.05 m and a
PR

monitor withdrawal speed.


distance of the borehole wall from the nozzle equal to

D: m Sandy soils
0 0·2 0·4 0·6 0·8 1·0 1·2 1·4 1·6 1·8 The erosion model for sandy soils has been tested by
0
Column 1 comparing the theoretical results with the data obtained by a
Column 2 previous field investigation of single-fluid jet grouting per-
1
Column 18 formed on a pyroclastic deposit of silty sand located above
Column 19 the water level (Croce & Flora, 1998). The average mech-
2
anical soil properties (ªs ¼ 18 kN/m3 , c9 ¼ 70 kPa, 9 ¼
308) were found by means of laboratory tests performed on
Z: cm

3 undisturbed samples.
Seven columns were treated with the single-fluid method,
4 where the grout (w/c ¼ 1) was injected at a constant
pressure (45 MPa) and constant velocity (220 m/s). The
5 number and diameter of nozzles, the average monitor with-
Dmin-max drawal and rotational speeds were changed from column to
from calculations column according to the sequence reported in Table 2. The
6
monitor was lifted by steps of 5 cm each.
Fig. 10. Observed and calculated diameters for Polcevera case The measured diameters are reported along with depth in
history four different plots (Fig. 12(a) to 12(d)), each of them
THEORETICAL MODELLING OF JET GROUTING 9
Table 1. Parameters from Polcevera case history

w/c ratio Injection Number of Nozzle diameter: Monitor withdrawal ksoil : Dav from D from
velocity, nozzles m speed: m/s cm/s site: m model: m
v0 : m/s

286.6 0.002 0.003 0.5–4 1.06–1.20 0.62–1.38

FS
1:1 2

2·0 4.0
vs/m 5 0·0015 m/s v0 5 380 m/s v0 5 380 m/s
1·8 3.5 vs/m 5 0·0015 m/s
1·6 vs/m 5 0·005 m/s v0 5 280 m/s vs/m 5 0·005 m/s v0 5 280 m/s
3.0
1·4
v0 5 180 m/s 2.5 v0 5 180 m/s
1·2
v0 5 380 m/s

D: m
D: m

1·0 2.0 v0 5 380 m/s


v0 5 280 m/s v0 5 280 m/s
0·8
v0 5 180 m/s 1.5 v0 5 180 m/s
0·6
1.0
0·4
0.5
0·2
0 0
0·01 0·1 1 10 100 0·01 0·1 1 10 100
k: cm/s k: cm/s

100 100
90 90 vs/m 5 0·0015 m/s
80
70
v0 5 380 m/s
v0 5 280 m/s
OO 80
70
vs/m 5 0·005 m/s

v0 5 380 m/s
60 v0 5 180 m/s 60
E: %
E: %

50 vs/m 5 0·0015 m/s 50 v0 5 280 m/s


v0 5 380 m/s
40 40
vs/m 5 0·005 m/s v0 5 380 m/s v0 5 180 m/s
30 30 v0 5 280 m/s
20 v0 5 280 m/s 20
v0 5 180 m/s
10 10
v0 5 180 m/s
0 0
0·01 0·1 1 10 100 0·01 0·1 1 10 100
k: cm/s k: cm/s
(a) (b)

Fig. 11. Column diameters and treatment efficiencies calculated with the proposed model for soil porosity n 0.3, grout
composition w/c 1 (N 4.6, ªf 15 kN/m3 ), and for different parameters: (a) d0 0.002 m; (b) d0 0.004 m

Table 2. Parameters from Vesuvius case history

Column w/c ratio Flow rate: Number of d0 : Rotation Monitor v0 : Measured Calculated
number by weight 103 m3 /s nozzles mm speed: withdrawal m/s average average
rad/s speed: mm/s diameter: m diameter: m

C0 1 1.38 2 2.0 1.57 5.71 220 0.657 0.738


PR

C1 1 2.50 1 3.8 0.79 5.00 220 0.959 0.906


C2 1 2.35 2 2.6 1.57 6.67 220 0.748 0.835
C3 1 2.50 1 3.8 1.18 5.00 220 0.968 0.906
C4 1 2.35 2 2.6 1.05 6.67 220 0.711 0.835
C5 1 2.50 1 3.8 0.63 4.00 220 1.111 0.990
C6 1 2.50 1 3.8 0.94 4.00 220 0.951 0.990

grouping the results obtained with a different set of jet measured average diameters is always lower than 15%, thus
parameters (nozzle diameter, number of nozzles and monitor showing that the model is able to catch the main features of
withdrawal speed). the grout–soil interaction. The comparison performed in the
The experimental results are simulated with the proposed four plots (Fig. 12) also shows that the role of each
model by assuming that the jet acts on a single direction for treatment parameters is considered quite satisfactorily in the
a time calculated by equation (23). The parameters , ¸ and proposed model (the effects of number and diameter of
k9, previously defined to evaluate the equivalent jet velocity nozzles can be seen by comparing Figs 12(a), 12(b) and
(equations (1), (2) and (4)) are fixed equal to 0.5, 16 and 12(c), whereas the effects of monitor withdrawal speed can
0.003 respectively. The two unknown parameters ( ¼ 0.2 be seen by a comparison of Figs 12(c) and 12(d)).
and Æ ¼ 0.3 m/s) introduced in the model have been It is finally worth observing that the shape of columns in
evaluated by a trial and error procedure. According to the frictional soils is typically like a funnel, and that this result
results shown in Table 2, the scatter between observed and has been obtained in the model (see Fig. 12) by introducing
10 MODONI, CROCE AND MONGIOVÌ
D: m D: m
0 0·2 0·4 0·6 0·8 1·0 1·2 0 0·2 0·4 0·6 0·8 1·0 1·2
0 0

2 C0 2 C2 C4

Z: m
Z: m

4 4

FS
Nozzles: 2 3 2 mm 6 Nozzles: 2 3 2·6 mm
6
v s 5 5·71 mm/s v s 5 6·67 mm/s
8 8
(a) (b)

D: m D: m
0 0·2 0·4 0·6 0·8 1·0 1·2 0 0·2 0·4 0·6 0·8 1·0 1·2
0 0

C6 C5 2 C6 C5
2

Z: m
4
Z: m

4
Nozzles: 1 3 3·8 mm Nozzles: 1 3 3·8 mm
6 6
v s 5 5·00 mm/s v s 5 4·00 mm/s

8 8 (d)
(c)

0.3; Æ 0.3 m/s)


OO
Fig. 12. Measured and calculated diameters from the Vesuvius case history (

a dependence of the diameter on the soil strength, increasing 1·0


with depth (equations (25) and (26)). 0·9
In order to evaluate the influence of the treatment para- 0·8
meters (v0 , d0, vs , m, and N) on the radius of columns, the 0·7
results of Fig. 12 have been parametrically extrapolated in 0·6
Fig. 13, where the effect of each single parameter is clearly E 0·5
distinguished. As expected, an increase in v0 and d0 is 0·4 d0 5 0·002 m; v0 5 200 m/s
0·3 d0 5 0·002 m; v0 5 300 m/s
reflected in a straight increase of R, while an increase in
0·2 d0 5 0·003 m; v0 5 200 m/s
vs /m produces an overall reduction of R. The effect of the
0·1 d0 5 0·003 m; v0 5 300 m/s
injected fluid is particularly significant is . In fact, when 0
water (C/W ¼ 0, ª ¼ 9.81 kN/m3 , N ¼ 1) is injected 0 0·001 0·002 0·003 0·004 0·005
instead of a grout (C/W ¼ 1, ª ¼ 15 kN/m3 , N ¼ 4.6) the vs/m: m/s
diameters of the columns increase significantly, thus showing
the benefits of selecting the grout composition properly. In Fig. 14. Efficiency of treatment calculated for different injection
particular, the above results confirm the benefit of the pre- parameters (W 1; n 0.3)
cutting of soil by water, which is sometimes performed in
single-fluid jet grouting before treatment.
The efficiency of treatments, defined by equation (24) and
calculated with the same parameters as adopted in Fig. 13, by v0 and d0, the influence of such parameters on the
for C/W ¼ 1 and n ¼ 0.3, is finally plotted in Fig. 14. As efficiency E is the opposite. In fact, whereas increasing the
PR

expected, it decreases when the injection time t* is in- injection speed v0 has a positive consequence, an increase in
creased: that is, it decreases with the number of nozzles and nozzle dimension d0 produces an overall reduction in E.
increases with the monitor withdrawal speed. Differently This last effect is a straight consequence of the fact that an
from the case of column radius, which is similarly affected increase in nozzle diameter produces a large increase of the
volumetric flow rate, which is not fully involved in the
column growth.
1·0 w/c 51, N 5 4·6, ãf 5 15 kN/m3
0·9 vs/m 5 0·001 m/s
w/c 5`, N 5 1, ã f 5 9·81 kN/m3 Clayey soils
0·8
0·7 vs/m 5 0·002 m/s Theoretical column diameters are reported in Fig. 15
0·6 vs/m 5 0·003 m/s against soil undrained cohesion cu for different treatment
R: m

0·5 vs/m 5 0·004 m/s parameters (d0 ¼ 0.002 m; v0 ¼ 200, 300 and 400 m/s; vs /m
0·4 ¼ 0.002 and 0.005 m/s). These diameters have been calcu-
0·3 lated by assuming the previously adopted factors for sub-
0·2 merged flow ( ¼ 0.5, ¸ ¼ 16 and k9 ¼ 0.003). Model
0·1 calibration for clayey soils has been based on laboratory and
0
0 0·2 0·4 0·6 0·8 1·0 in situ observations retrieved from the literature (Tornaghi,
v0/d0: m2/s2 1989; Dabbagh et al., 2002). In particular, the model
calibration factor c has been fixed equal to 0.36 in
Fig. 13. Parametric study performed from the Vesuvius case accordance with the laboratory test results presented by
history ( 0.3; Æ 0.3m/s) Dabbagh et al. (2002). The second calibration factor Æ has
THEORETICAL MODELLING OF JET GROUTING 11
1·20
v0 5 400 m/s 2 vs/m 5 0·002 m/s

1·00 v0 5 300 m/s 2 vs/m 5 0·002 m/s

v0 5 200 m/s 2 vs/m 5 0·002 m/s

0·80

FS
D: m

0·60

0·40

v0 5 400 m/s 2 vs/m 5 0·005 m/s


0·20
v0 5 300 m/s 2 vs/m 5 0·005 m/s

v0 5 200 m/s 2 vs/m 5 0·005 m/s


0·00
0 10 20 30 40 50 60 70 80 90 100
cu: kPa

Fig. 15. Simulated (W 1; d0 0.002 m; v0 200 m/s; c 0.36; Æ 0.12 m/s) and
observed (shaded area) jet columns diameters in cohesive soils
OO
been fixed equal to 0.12 m/s in order to match the range of
observed column diameters reported in Fig. 15 as a shaded
obtained for sands and clays indicate that the column
diameter decreases with the square root of the soil shear
area (from Tornaghi, 1989). strength. In the case of sands, a reduction of column
Independently of the chosen set of treatment parameters, a diameter with depth from the ground level has also been
strong reduction of column diameter occurs when the soil recognised, as previously observed by field investigations
undrained cohesion cu is increased. Such a result is in (Croce & Flora, 1998). With regard to clayey soils, the
agreement with the practice of considering single-fluid jet parametric calculations show that jet grouting is effective
grouting not to be effective on soils having cu higher than only if treatments are performed providing high flow rates
50 kPa. and low monitor withdrawal speeds.
The theoretical results show, however, that a significant In general, theoretical results obtained for all soil types
improvement of column diameters could be obtained, even show a dependence of column diameters on the fluid velo-
for higher cohesion, by increasing the injection velocity (v0 ) city, number and diameter of nozzles, and monitor lifting
and the number of nozzles (m) and/or by reducing the speed. However, by increasing the volume of injected fluid
monitor withdrawal speed (vs ). the amount of spoil will also be increased, thus reducing the
economical efficiency of treatment.
It can be thus concluded that the presented models are
CONCLUDING REMARKS able to simulate the main effects of single-fluid jet grouting,
A comprehensive theoretical analysis of single-fluid jet and that they can be used to forecast the diameter of jet
grouting has been presented. Jet propagation across the columns and to select the most suitable set of treatment
space included between the nozzles and the intact soil has parameters. Finer calibration of the models may be obtained
been modelled first, on the basis of the theory of submerged by further experimental investigations. Model extension to
flow (Hinze, 1948). It has been found that the velocity the double- and triple-fluid systems may be pursued by
cross-sectional profile depends on the following parameters: properly simulating the effect of the air jet that wraps the
PR

distance from the nozzle, nozzle diameter, flow rate and grout or water jets.
kinematic viscosity of the grout.
With regard to jet–soil interaction, three different models
have been developed, for gravels, sands and clays respec- APPENDIX
tively, according to the results of previous experimental The original formulation (Hinze, 1948) analyses the pat-
observations. The results obtained by each model have been tern of the mean component of jet velocity by introducing
used to draw design charts relating the column diameter the following hypotheses.
with the relevant soil properties and the main treatment
parameters. Empirical correcting factors had to be intro- (a) The velocity profile is similarly shaped in each cross-
duced, however, in order to account for possible unknown section all along the distance from the nozzle, and is
phenomena. These factors have been evaluated by comparing expressed by a function of the type
theoretical results with available experimental observations  
r
(Tornaghi, 1989; Croce et al., 1994; Croce & Flora, 1998; v x ¼ v xmax f (32)
d0
Dabbagh et al., 2002).
Following model calibration, parametric calculations have where v x is the velocity at a single point of coordinates
been performed showing the influence of each treatment x and r (see Fig. 2), and v x_max is the velocity
parameter. In particular, results obtained by the seepage computed on the jet axis at a distance x from the
model show that the jet column diameter increases consider- nozzle.
ably with soil permeability, and that significant dimensions (b) The fluid viscosity is proportional to the fluctuation
can be obtained through this mechanism only in the case of velocity component normal to the x-axis, which is
very pervious materials (i.e. for gravelly soils). Results proportional to the difference (v xmax  v xmin ) in each
12 MODONI, CROCE AND MONGIOVÌ
cross-section (vxmin is always equal to 0), and to the NOTATION
mixing length assumed by Prandtl (1942, as reported by a1 cross-sectional area of the jet at the borehole wall
Hinze, 1948) as proportional to the current value of jet c9 soil cohesion
diameter. cu undrained soil cohesion
(c) The momentum of the jet in each cross-section is d0 nozzle diameter
E treatment efficiency defined as ratio between absorbed
constant, along with the distance x from the nozzle.
and injected fluid

FS
h total head of injected fluid
Under these hypotheses the following relation is derived. k coefficient of permeability of soil with respect to
water
vx 1 kf coefficient of permeability of the with respect to
¼ 2 (33) injected fluid
v xmax (r=x)2
1þ m number of nozzles
k N relative turbulent kinematic viscosity of injected fluid
n soil porosity
where  is the turbulent kinematic viscosity of the injected q0 fluid volumetric rate injected by a single nozzle
r radial distance measured from jet axis
fluid (defined by the relation r x ¼ r(@v x /@r)), and k is a
R radial distance measured from treatment axis
parameter usually evaluated by experiments. R1 borehole radius
By introducing a turbulent relative kinematic viscosity req radius of the ‘equivalent jet’
N ¼ f /w (where f is the kinematic turbulent viscosity of Re radius of seepage front
a generic fluid and w is the kinematic turbulent viscosity of Rlim limiting distance for gushed back flow
water), equation (33) can be written for a generic fluid in Rmax maximum jet column radius attainable by erosion
the following manner. ˜s length of lifting steps
t integration time
vx 1 ˜t period between two consecutive lifting steps
¼ 2 (34)
v xmax V1 seepage velocity at borehole wall
(r=x)2
1þ Vc advancing velocity of erosion into soil
k9N
OO
The factor k9 ¼ kw can be experimentally evaluated for
water jets. In the present case k9, together with the variation
va
veq
veq_1
velocity of advancing front in seepage model
average velocity of jet
average velocity of jet at borehole wall (seepage
model)
of v xmax along with x, is obtained by assuming that the vf average jet velocity at soil–fluid boundary (erosion
momentum  of the jet along the x-axis is constantly equal model)
vL ‘limiting velocity’ (minimum jet velocity producing
to the momentum at the nozzle. soil erosion)
ð r¼1 vs average monitor withdrawal rate
ª 2 ª d2
¼ v x 2rdr ¼ v2x0  0 ¼ 0 (35) vx velocity of the injected fluid at a generic point
r¼0 g g 4 vx max velocity of injected fluid along jet axis
v0 velocity of injected fluid at nozzle
x distance from nozzle
where v x0 is the velocity of the jet immediately out of the
W water–cement ratio by weight
nozzle and d0 is the nozzle diameter Introducing v x , calcu- Æ calibration factor for sands and clays
lated by equation (34), into equation (35) and putting y ¼ ªf unit weight of injected fluid
r/x the following relation can be written.  turbulent kinematic viscosity
¸, k9 calibration factors for jet velocity
d0 1
v xmax ¼ v x0 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi (36) f viscosity of injected fluid
x ð 1  laminar kinematic viscosity
y
8 2 4
dy  ratio between minimum effective and maximum jet
0 [1 þ ( y =k9N )] velocity (submerged flow)
rg density of injected fluid
The integration at the second expression can be calculated rc density of cement
as follows. rw density of water
PR

ð1 ð1 , 9 soil total and effective stresses


y 4 y 9 soil friction angle
d y ¼ (k9N ) dy
2
0 [1 þ ( y =k9N )]
4 2 4
0 (k9N þ y ) ł equivalent jet load (erosion model)
ð łs , łc erosion resistance of sands and clays
(k9N )4 1 d(k9N þ y 2 ) k9N g , s , c calibration factors for gravels, sands and clays
¼ 2 4
¼ (37)
2 0 (k9N þ y ) 6

After equation (37), equation (36) becomes REFERENCES


Bear, J. (1972). Dynamics of fluid in porous media. Elsevier.
v xmax ¸ d0 Bell, A. L. (1993). Engineering properties of soils. London: E &
¼ pffiffiffiffiffi (38)
v x0 N x FN Spon.
Bergschneider, B. & Walz, B. (2003). Jet grouting: range of the
grouting jet. Proc. 13th Eur. Conf. Soil Mech. Found. Engng,
where Prague, 53–56.
Covil, C. S. & Skinner, A. E. (1994). Jet grouting: a review of
1
¸ ¼ rffiffiffiffiffiffiffiffi (39) some of the operating parameters that form the basis of the jet
8 grouting process. In Grouting in the ground, pp. 605–627.
k9 London: Thomas Telford.
6
Croce, P. & Flora, A. (1998). Jet-grouting effects on pyroclastic
soils. Riv. Ital. Geotec., No. 2, 5–14.
The value of k9 is thus mutually related to the value of ¸, Croce, P. & Flora, A. (2000). Analysis of single fluid jet-grouting.
which has been experimentally evaluated (see Fig. 3). Géotechnique 50, No. 6, 739–748.
THEORETICAL MODELLING OF JET GROUTING 13
Croce, P., Chisari, A. & Merletti, T. (1990). Indagini sui trattamenti Kutzner, C. (1996). Grouting of rock and soil. Rotterdam: Balkema.
dei terreni mediante jet-grouting per le fondazioni di alcuni Lambe, T. W. & Whitman, R. V. (1969). Soil mechanics. New York:
viadotti autostradali. Rassegna dei Lavori Pubblici, No. 12 (in John Wiley & Sons.
Italian). Miki, G. (1985). Soil improvement by jet grouting. Proc. 3rd Int.
Croce, P., Gajo, A., Mongiovı̀, L. & Zaninetti, A. (1994). Una Geotech. Seminar on Soil Improvement Methods, Singapore.
verifica sperimentale degli effetti della gettiniezione. Riv. Ital. Muskat, M. (1946). The flow of homogeneous fluids through porous
Geotec. 28, No 2. media. Ann Arbor, MI: J. W. Edwards.

FS
Dabbagh, A. A., Gonzalez, A. S. & Peña, A. S. (2002). Soil erosion Scheidegger, A. E. (1957). The physics of flow through porous
by a continuous water jet. Soils Found. 42, No. 5, 1–13. media. New York: Macmillan.
de Vleeshauwer & Maertens (2000). Jet-grouting: state of the art in Shibazaki, M. (1991). The state of art in jet-grouting. Proceedings
Belgium. Proceedings of the conference ‘Grouting – Soil im- of the symposium on soil and rock improvement in underground
provement – Geosystems including reinforcement’, Helsinki, pp. works, pp. 19–46.
145–156. Shibazaki, M. (2003). State of practice of jet grouting. Proc. 3rd
ENV 12716 (2001). Execution of special geotechnical works jet Int. Conf. on Grouting and Ground Treatment, New Orleans.
grouting, Standard European Code. Stein, J. & Grabe, J. (2003). Jet grouting tests and simulation. Proc.
Farmer, W. & Attewell, P. B. (1965). Rock penetration by high 13th Eur. Conf. Soil Mech. Geotech. Engng, Prague, 899–902.
velocity jet. Int. J. Rock Mech. Mining Sci. 2, 135–153. Taylor, D. W. (1948). Fundamentals of soil mechanics. New York:
Hazen, A. (1911). Discussion on ‘Dams on sand foundations’. John Wiley & Sons.
Trans. ASCE 73, 199. Tornaghi, R. (1989). Trattamento colonnare dei terreni mediante
Hinze, J. O. (1948). Turbulence, 1st edn. New York: McGraw-Hill. gettiniezione (Jet grouting). Associazione Geotecnica Italiana,
Kaushinger, J. L., Perry, E. B. & Hankour R. (1992). Jet grouting: XVII Convegno Nazionale di Geotecnica, Taormina.
state-of-the-practice. In Grouting/soil improvement and geosyn- Winterkorn, H. F. & Fang, H. Y. (1975). Foundation engineering
thetics, ASCE Geotechnical Special Publication No. 30, Vol. 1, handbook, 1st edn. New York: Van Nostrand Reinhold, pp.
pp. 169–181. 340–350.
OO
PR

You might also like