Article: Effects of Capillary Number, Bond Number, and Gas Solubility On Water Saturation of Sand Specimens

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

133

ARTICLE
Effects of capillary number, Bond number, and gas solubility on water
saturation of sand specimens
Bruce L. Kutter
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

Abstract: To better understand how to prepare completely water-saturated specimens or centrifuge models from dry sand, the
mechanisms of the infiltration and filling of pores in sand are studied. Complete saturation has been shown by others to be
especially important in studies involving the triggering of liquefaction. This paper discusses how the degree of saturation
obtained during infiltration increases with the “Bond number”, Bo (ratio of body forces and capillary forces), and the “capillary
number”, Ca (ratio of viscous forces and capillary forces), as well as the solubility of gas bubbles in the pore fluid. Bo is varied by
changing the particle size, fluid density, and centrifugal acceleration. Ca is varied by changing the fluid viscosity and infiltration
rate. The dissolution of gas is encouraged by replacing pore air by CO2 (56 times more soluble in water than N2), by de-airing the
liquid prior to infiltration or by increasing the pore fluid pressure after infiltration. Infiltration experiments performed at 1g and in a
centrifuge are presented. A new technique for measuring the degree of saturation is also presented. Quantitative pressure–saturation
relations are presented for different gasses, illustrating the importance of replacement of air by CO2. Spinning a specimen in a
centrifuge during infiltration is also useful for speeding up the saturation process and for achieving higher degrees of saturation.

Key words: degree of saturation, soil mechanics, liquefaction, unsaturated soil, infiltration.

Résumé : Les mécanismes d'infiltration et de remplissage des pores dans du sable sont étudiés afin de mieux comprendre
la façon de préparer des échantillons complètement saturés en eau ou des modèles pour centrifugeuse à partir de sable sec.
For personal use only.

D'autres travaux ont démontré l'importance de la saturation complète lors d'études impliquant le déclenchement de la
liquéfaction. Cet article discute de comment le degré de saturation obtenu durant l'infiltration augmente avec le « nombre
de Bond », Bo (ratio des forces du corps et des forces capillaires), et le « nombre capillaire », Ca (ratio des forces visqueuses
et des forces capillaires), ainsi qu'avec la solubilité des bulles de gaz dans le fluide interstitiel. Bo est varié en changeant la
granulométrie, la densité du fluide et l'accélération centrifuge. Ca est varié en changeant la viscosité du fluide et le taux
d'infiltration. La dissolution des gaz est encouragée par le remplacement de l'air interstitiel par du CO2 (56 fois plus soluble
dans l'eau que N2), par la dé-aération du liquide avant l'infiltration ou en augmentant la pression interstitielle après
l'infiltration. Des essais d'infiltration réalisés à 1g et sur une centrifuge sont présentés. Une nouvelle technique pour
mesurer le degré de saturation est aussi présentée. Des relations quantitatives de pression–saturation sont aussi présentées
pour différents gaz, ce qui illustre l'importance du remplacement de l'air par le CO2. L'utilisation d'une centrifugeuse pour
tourner un échantillon durant l'infiltration permet d'accélérer le processus de saturation et d'atteindre des plus hauts
degrés de saturation. [Traduit par la Rédaction]

Mots-clés : degré de saturation, mécanique des sols, liquéfaction, sol non saturé, infiltration.

Introduction and literature review urated soil. Okamura et al. (2006) observed that undissolved air
The presence of gas bubbles in sand is important in geotechni- bubbles persisted in a sand deposit 26 years after being intro-
cal earthquake engineering because the degree of saturation af- duced during installation of sand compaction piles. Eseller-Bayat
fects compressibility of the pore fluids, which in turn affects (2009) performed experiments to investigate the persistence of air
important soil properties such as permeability and the cyclic bubbles by seeping water through columns of unsaturated sand in
stress ratio required to trigger liquefaction. Chaney (1978), Yoshimi the laboratory. While Okamura et al. (2006) and Eseller-Bayat
et al. (1989), and Xia and Hu (1991) have shown that the cyclic stress (2009) have demonstrated that air bubbles might be persistent,
required to trigger liquefaction generally increases as the degree they did not consider all possible seepage conditions and (or)
of saturation decreases. other factors that may lead to dissolution, compression or re-
In laboratory and centrifuge liquefaction experiments, it is moval of gas bubbles.
therefore important to test completely saturated specimens to Although the present paper is focused on factors controlling
ensure that the cyclic strength is not overestimated. Techniques the degree of saturation of centrifuge models, the principles dis-
such as infiltration under vacuum and back-pressure saturation cussed in this paper may be of interest to others working with
(Lowe and Johnson 1960; Black and Lee 1973) and flushing with unsaturated soils. The mechanisms of bubble formation and dis-
CO2 prior to saturation (Mulilus et al. 1975) have been used to solution are of interest in air sparging (injection of gas) to facili-
facilitate complete saturation in the laboratory. tate removal of volatile organic fluids from ground water (e.g.,
On the other hand, the introduction of air has also been sug- Marulanda et al. 2000). Although it is often implied that the soil-
gested as a method for reducing the liquefaction potential of sat- water characteristic curve, i.e., the water retention curve, relating

Received 6 July 2011. Accepted 16 October 2012.


B.L. Kutter. Department of Civil and Environmental Engineering, University of California, Davis, CA, 95616, USA.
E-mail for correspondence: blkutter@ucdavis.edu.

Can. Geotech. J. 50: 133–144 (2013) dx.doi.org/10.1139/cgj-2011-0250 Published at www.nrcresearchpress.com/cgj on 13 December 2012.
134 Can. Geotech. J. Vol. 50, 2013

Fig. 1. Suggested setup for saturation of centrifuge models CGM (2011) and Ueno (1998).
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

degree of saturation to water suction is a soil property, many have 5. Allow the de-aired pore fluid to seep into the model container
shown that the curve also depends on gravity, infiltration rate, from the bottom up while the specimen is sitting at 1g.
For personal use only.

pore gas solubility, and time (Fredlund and Rahardjo 1993; Blunt
and Scher 1995; Ratnam et al. 1996; Friedman 1999; Lu and Likos Takahashi et al. (2006), using relationships between saturation
2004; Garnier et al. 2007). However none of these references con- and p-wave velocity presented by Ishihara et al. (1998) and
siders the combined effects of all of these factors with specific Kokusho (2000), have demonstrated that measured p-wave veloc-
focus on the residual air content of sand after water infiltration. ities in specimens saturated by this procedure indicate that the
saturation was successful. During the infiltration from the bot-
Background and literature tom up, caution must be exercised to avoid disturbing the speci-
The quality of saturation may be characterized or measured by men due to excessive pressure gradients and erosion. Stringer and
various methods Madabhushi (2009) describe equipment for slowly saturating sand
by carefully controlling the rate of infiltration to ensure that the
• Measuring weights and volumes of solids, water, and gas. hydraulic gradients are not excessive.
• Measuring the Skempton B-value (e.g., Black and Lee 1973). Naesgaard et al. (2007) attempted to emulate the procedure
• Measuring the p-wave velocity (e.g., Ueno 1998 or Kokusho described by Ueno (1998) in a triaxial specimen by introducing
2000). water while the specimen is subjected to vacuum. Naesgaard et al.
• A new method presented in this paper whereby the change in (2007) measured the p-wave velocity and the B-value to character-
volume of the pore fluids is measured as the pore pressure is ize the degree of saturation. They measured high p-wave velocities
varied. and low B-values (approximately 0.5) that were inconsistent with
Ueno (1998) describes the procedure for saturation of centrifuge the relationships relating B-value and p-wave velocity presented
models of dry sand developed at the University of California (UC) by Ishihara et al. (1998). Naesgaard et al. (2007) showed that p-wave
Davis (CGM 2011) and other centrifuge centers (Fig. 1). The proce- velocity is a good indicator of saturation if the bubbles are small
dure has many similarities to the approach developed for triaxial and sufficiently distributed amongst all the pores, but it is a poor
specimens by Mulilus et al. (1975). Briefly, the procedure is as indicator of saturation if the air bubbles are large and scattered
follows: because the p-waves may travel between the bubbles, largely un-
affected by the large bubbles or air pockets. Thus, not only did
1. Create a supply of de-aired water. This may be accomplished by Naesgaard et al. (2007) raise doubts about the effectiveness of the
spraying the water into an evacuated tank or by boiling the procedure described by Ueno (1998), but they also raised doubts
water and preventing access to the atmosphere. about the accuracy of one of the procedures often used by centri-
2. Remove most of the air from the model container by applying fuge modelers to verify satisfactory saturation.
a vacuum to the soil to be saturated. An absolute pressure of It must be pointed out that the procedure described by Naesgaard
10 kPa is recommended. As 10 kPa corresponds to a vacuum of et al. (2007) does not accurately follow that described by Ueno
about 90% of atmospheric pressure, 90% of the air mass is (1998) in two important ways. First, they only used 50 kPa vacuum
removed. when flushing the specimen with CO2 (potentially replacing 50%
3. Flood the model container with CO2 gas at atmospheric pres- of the air by CO2), while Ueno (1998) recommend flushing twice
sure. This displaces air by CO2, which is about 56 times more with 90 kPa vacuum (potentially replacing 99% of the air). Second,
soluble in water than is N2. Naesgaard et al. (2007) do not mention the process for de-airing
4. Re-apply the vacuum (10 kPa absolute pressure). Assuming the water used to infiltrate the specimen.
uniform mixing of the gasses in the specimen, 90% of the Okamura and Inoue (2012) performed centrifuge tests to illus-
remaining air (99% of the original) is removed after the second trate that liquefaction behavior is sensitive to the degree of satu-
phase of re-evacuation and re-flooding with CO2. ration and that degree of saturation obtained during infiltration

Published by NRC Research Press


Kutter 135

depends on centrifugal acceleration and the vacuum pressure ap- Fig. 2. Infiltration at small and large Ca. (a) At low Ca, the small
plied during infiltration. crevices attract water first. (b) If viscous forces dominate, water
preferentially flows faster through the larger pores.
Goals
The main purposes of this paper are to (i) provide a better un-
derstanding of the physical and chemical processes that affect the
degree of saturation of laboratory specimens of sandy soil and
especially large-scale centrifuge models; (ii) provide tools for
quantitative assessment of the importance of de-gassing, back
pressure, gas solubility, g-level, infiltration rate, and time on the
degree of saturation that can be expected for centrifuge models
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

and laboratory sand specimens; (iii) present experiments to test


our hypothesis that the degree of saturation depends on “Bond
number” (Bo), “capillary number” (Ca), and gas solubility; (iv) show
that the procedure described by Ueno (1998) for saturation of
centrifuge models can reliably achieve saturation; and (v) show are relatively large compared to the surface tension forces.
that the saturation process is enhanced by spinning the specimen Ratnam et al. (1996) and Ratnam (1996), for example, observed
in a centrifuge during infiltration. that above a critical Bo value, capillary trapping of a nonwet-
ting fluid (e.g., air) is less than about 3%. Thus the degree of
The physics of saturation by infiltration saturation obtained during infiltration is expected to increase
as Bo increases.
The infiltration and saturation process is affected by many fac- In geotechnical engineering water is the most common liquid,
tors, such as surface tension of the liquid(s), size and size distri- which has a surface tension of about 0.072 N/m and a viscosity of
bution of the pores, buoyant force on the bubbles, density of fluid, approximately 1 cP (1 cP = 0.001 kg/m·s) at standard temperature
gravitational acceleration, rate of infiltration, hydraulic gradient and pressure. In dynamic centrifuge model tests, the viscosity of
and permeability, viscosity of the pore fluid, gas and liquid pres- the fluid is often altered to improve similarity (Kutter 1995; Muir
sures, and solubility of the gasses in the liquids. Wood 2004).
One of the important dimensionless groups that affects the The “Weber number” (We) relates inertial forces to capillary
mechanism of pore filling during infiltration is the Ca (e.g., Blunt forces and controls mechanisms such as entrapment of air due to
and Scher 1995; Ratnam et al. 1996; de Gennes et al. 2004). The Ca
For personal use only.

entrainment, but this mechanism is not considered to be impor-


is proportional to the pressure drop associated with viscous flow tant for the present work. The “Reynolds number” (Re) relates
in a pore tube to the pressure drop across a meniscus in a pore tube viscous forces and inertial forces and determines the transition
between laminar and turbulent flow.
[1] Ca ⫽ ␩v/Ts During preparation of an unconfined specimen for a centrifuge
or triaxial test, piping or boiling due to large fluid pressure gradi-
where ␩ is the viscosity, v is the fluid velocity, and Ts is the surface ents should be avoided. If the fluid pressure (air or water) ap-
tension in the fluid. Blunt and Scher (1995) suggest that for a proaches the total stress, then the effective stress will approach
typical soil, the viscous pressure drop in a single pore throat con- zero and the soil skeleton may boil and erupt due to piping. For
necting two larger pores is approximately equal to the pressure level ground, boiling and piping instability may be avoided if the
drop across a meniscus in that the pore throat for Ca ⬃ 103 or 104. “pore pressure ratio” (Ru) is less than 1.
At larger values of Ca, the viscous drag forces are larger, hence
preferential flow is in the large pores as depicted in Fig. 2b. At [3] Ru ⫽ ue / ␴vo
small values of Ca, surface tension effects dominate the flow, and
wetting tends to advance more rapidly in the crevices than in the
where ue is the excess pore water pressure and ␴vo is the initial
large pores (Fig. 2a). For high Ca values, infiltration could theoret-
effective stress. For sloping ground, instability may occur at
ically be faster through regions of high permeability (larger void
smaller pore pressure ratios. The pore pressure ratio, Ru, may be
spaces) and this might lead to trapping air in regions of low per- considered to be an additional dimensionless group controlling
meability. The critical capillary number depends on the size dis- the saturation infiltration process.
tribution and connectedness of the pores. If water is introduced
very slowly, the capillary tension will tend to be the driving force Dissolution of gas bubbles
for the first phase of infiltration. Crevices will fill first and then
Besides displacement of one fluid by another during infiltra-
larger pores will tend to fill in a sequence from smallest to largest.
tion, the degree of saturation may be affected by dissolving one
If a large pore filled with gas is surrounded by smaller pores filled
fluid in the other or by changing the pressure. LeBihan and
with water, the isolated air in the large pore may be trapped.
Leroueil (2002) also discuss the pressure–dissolution–saturation
Another important dimensionless group that determines the
relations using Henry's Law and Boyle's Law, but in the context of
mechanism of pore filling is the Bo
developing a model for transport of dissolved gasses by advection
and diffusion in the context of calculating the evolution of satu-
[2] Bo ⫽ (␳f ⫺ ␳g) gr2 /Ts ration in the core of an earth dam.
The volume of gas in the gas phase may be estimated by the
where ␳ represents density (subscripts f and g denote the fluid and ideal gas law
gas, respectively), g represents gravity, and r the representative
pore radius. This number can be thought to represent the ratio of [4] pV ⫽ nRT
body forces and surface tension forces for a trapped air bubble or
for a small volume of water held at the particle contacts by capil- where p is the absolute pressure, V is the volume of the gas, n is the
lary tension. If the Bo value is large (i.e., large pores, high gravity number of moles of gas, R is the gas constant, and T is the absolute
field or small surface tension), the body forces that push bubbles temperature. The number of moles in the gas phase, n, may
up through the voids and pull down the water trapped in crevices change as the gas molecules dissolve in or come out of solution in

Published by NRC Research Press


136 Can. Geotech. J. Vol. 50, 2013

the liquid. The equilibrium solubility for a gas in water is de- p2 1 ⫺ S1 ⫹ HccS2
[7b] ⫽
scribed by Henry's Law. p1 1 ⫺ S2 ⫹ HccS2

[5a] caq ⫽ Hcp p


If the concentration of dissolved gas in the influent water is equal to
[5b] caq ⫽ Hcccgas (p2/RT)Hcc (i.e., if it is equal to the final dissolved concentration)

p2 1 ⫺ S1 ⫹ HccS1
where Hcp represents the Henry's constant that relates concentra- [7c] ⫽
p1 1 ⫺ S2 ⫹ HccS1
tion in solution to pressure in the gas phase, Hcc represents the
Henry's constant that relates concentration in the solution to con-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

centration in the gas phase, p represents the partial pressure of a Table 2 presents results showing the absolute pressure ratio
molecular species in the gas phase, caq represents the equilibrium (p2/p1) required to dissolve all of the gas in the system for the three
concentration of the same molecular species dissolved in the wa- alternative assumptions embodied by eqs. [7a]–[7c]. Two different
ter, and cgas represents the molar concentration of the species in Henry's constants (for N2 and CO2) are considered and the initial
the gas phase. Henry's constants along with diffusion coefficients saturation is assumed to be 90%. For small Henry's constants,
for the most plentiful atmospheric gasses are given in Table 1. there is little difference between eqs. [7a] and [7b]. But for CO2,
Henry's Law states that if there is a gas–liquid interface, the which has a much larger Henry's constant, the difference is sig-
equilibrium concentration of a gas species dissolved in the liquid nificant. For CO2, an absolute pressure increase between 2% and
is proportional to the concentration of that gas species present in 13% is sufficient to saturate the specimen, depending on the as-
the gas phase. The interface may be at the surface of a gas bubble sumed quantity of dissolved gas in the influent water. For N2 gas,
or at the free surface above a pool of water. Equilibrium is quickly a pressure increase between 756% and 841% is required. The ben-
obtained at the interface, but significant time may be required for efit of using CO2 gas is evident from these calculations.
concentrations to diffuse throughout the liquid phase.
Nonequilibrium solution
From Table 1, it may be seen that at equilibrium, 0.832 L of CO2
Other useful relationships between saturation and pressure can
gas could be dissolved in 1 L of water if it is the only gas species
be obtained from other idealized assumptions. Suppose the pore
present in the system. Similarly, only 0.0149 L of N2 gas would be
gas pressure is initially p1 and completely de-aired water is rapidly
dissolved in 1 L of water if N2 was the only gas species. It is appar- infiltrated into the pores so that no gas dissolves in the water
ent that O2 and Ar gas are about twice as soluble in water as N2.
For personal use only.

during infiltration and the degree of saturation is S1. According to


But CO2 is 56 times more soluble in water than is N2. For this this idealized scenario, immediately after infiltration, all the gas
reason, Ueno (1998) recommends replacement of air with CO2 molecules are in the gas phase, and none in the liquid phase. Then
before attempting to saturate centrifuge specimens. Much earlier, the gas dissolves gradually in the water so that the saturation is 1
Mulilus et al. (1975) described the replacement of air by CO2 to and all the gas molecules are in the liquid phase. The pressure
enhance the saturation of triaxial test specimens. required to keep the bubbles dissolved is given by
For a soil sample with a total volume of voids, Vv, the volume of
water in the voids is SVv (where S is the degree of saturation) and p2 1 ⫺ Hcc
the volume of gas is (1 − S)Vv. The degree of saturation is the ratio [8] ⫽
p1 S1
of the volume of water to the volume of voids: S = Vw/Vv. Assuming
that the concentration of dissolved gas in the liquid is given by
Henry's Law, that there is only one gas species present, and that As seen in Table 2, if the initial degree of saturation is 90%, the
the number of moles of gas molecules in the system is the sum pressure required to dissolve the 10% CO2 gas is only 12% of the
of the moles in the gas phase, ngas, and the moles of gas dissolved pressure in the pores during infiltration. The pressure required to
in the liquid phase, naq dissolve N2 gas is 667% of the pressure during infiltration; if infil-
tration is performed at p1 = patm/6.67, then the N2 could eventually
p ( dissolve when the vacuum is released and the pressure increases
[6] n ⫽ ngas ⫹ naq ⫽ [ 1 ⫺ S)Vv ⫹ HccSVv] to p2 = patm (atmospheric pressure).
RT
Again using eq. [8], if p2 = p1 (pressure is not changed following
Pressure– dissolution–saturation relations infiltration), and if completely de-aired water is introduced to the
Equation [6] is useful for calculating the change in the equilib- specimen, then the system would eventually become saturated if
rium degree of saturation due to a change in pressure. Let the S1 > (1 − Hcc). For pure N2 gas bubbles to dissolve, the degree of
initial absolute pressure and degree of saturation be denoted by p1 saturation must be greater than Scrit,N2 = (1 − 1.49 × 10−2) = 98.5%.
and S1, respectively, and final values by p2 and S2. If the pressure is For pure CO2 gas bubbles to eventually dissolve, the initial degree
increased slowly while maintaining a constant effective stress on of saturation must be greater than Scrit,CO2 = (1 − 0.832) = 16.8%.
the particles, water would flow into the pores, compressing and
dissolving the bubbles as the pore pressure is increased. Some Nonequilibrium solution kinetics
mass of dissolved gas may also enter the system with the influx of Although eventual saturation may be quantified using eqs. [7]
water. If we assume, however, that the concentration of dissolved and [8], the time required to dissolve gas bubbles must be consid-
gas in the influent water is zero, we obtain ered. The rate of dissolution of bubbles in free water has been
studied extensively (e.g., Wise and Houghton 1966; Weinberg
p2 1 ⫺ S1 ⫹ HccS1 1981; Holocher et al. 2003). Kinetic bubble dissolution models typ-
[7a] ⫽ ically assume that the equilibrium concentration according to
p1 1 ⫺ S2 ⫹ HccS2
Henry's Law occurs at the gas bubble–liquid interface and that the
dissolved gas then diffuses into the fluid according to the diffu-
Lowe and Johnson (1960) derived eq. [7a], but presented it in a sion equation. The diffusion process is enhanced if the mixing is
different form. If the concentration of dissolved gas in the influ- aided by dispersion due to flow of water in tortuous pores. Diffu-
ent water is equal to (p1/RT)Hcc (i.e., if it is equal to the initial sion is slower in stagnant water and even slower in stagnant
dissolved concentration) pores. As seen in Table 1, the diffusion coefficient of N2 and CO2 in

Published by NRC Research Press


Kutter 137

Table 1. Air composition and component solubility in water at T = 25 °C. (1 atm = 101.325 kPa.)
Air Proportion of Hcp = caq/p Diffusion coefficient
constituent atmosphere (%)a (mol/L·atm)b Hcc = caq/cgasb in water (m2/s)a
N2 78 6.1 × 10−4 1.49 × 10−2 1.64 × 10−9
O2 21 1.3 × 10−3 3.18 × 10−2 1.80 × 10−9
Ar 0.93 1.4 × 10−3 3.43 × 10−2 —
CO2 0.031c 3.4 × 10−2 0.832 1.77 × 10−9
aTuve and Bolz (1973).
bSander (1999).
cMore recent measurements show that atmospheric CO has increased to about 0.04%.
2
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

Table 2. Example calculation results for pressure–saturation relationships.


Eq. [7a] Eq. [7b] Eq. [7c] Eq. [8] Eq. [7a] Eq. [7b] Eq. [7c] Eq. [8]
S1 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9
S2 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0
Hcc 0.015 0.015 0.015 0.015 0.832 0.832 0.832 0.832
p2/p1 7.56 7.66 8.407 6.67 1.02 1.12 1.13 0.120

water are quite similar, but because the solubility of CO2 is 56 be achieved by applying a vacuum during infiltration and by
times greater, the gradients will be about 56 times greater and flushing de-aired water through the specimen.
therefore CO2 bubbles should theoretically disappear 56 times Black and Lee (1973) assumed the pore gas consisted of air, not
faster than N2 bubbles. CO2. As can be seen from Table 1, at a given pressure CO2 is 56
The difference between the bubble pressure and the liquid pres- times more soluble than N2; therefore, if identical RBL ratios were
sure, given by the Young–Laplace equation (see, for example, Lu used for CO2, times for saturation should be reduced by a factor of
and Likos 2004) is, for a spherical bubble 56. Furthermore, as is apparent in Table 2, pressures required to
provide a given RBL ratio are much smaller, so rates of dissolution
2Ts may be even greater. This highlights the importance of replace-
For personal use only.

[9] pgas ⫺ pliq ⫽ ment of air by CO2, prior to infiltration and saturation.
r
LeBihan and Leroueil (2002) present an interesting theoretical
model for transport of dissolved gasses by advection and diffusion
where pgas is the pressure inside the gas bubble, pliq is the pressure
that could eventually be useful for predicting the evolution of
in the liquid, Ts is the surface tension, and r is the bubble radius.
saturation in centrifuge models due to flushing of pore water
Weinberg (1981) derived a set of nonlinear equations for bubble
through the specimen, but they did not consider the effects of size
dissolution that accounted for the effect of surface tension on the
and spatial distribution of air bubbles on the dissolution rate.
bubble pressure and the consequent effect of the bubble pressure
on the solubility according to Henry's Law. Assuming a surface
Infiltration experiments
tension for the gas–water interface of 0.075 N/m, the excess pres-
sure in a bubble of radius 0.2 mm is 1.5 kPa, which is small com- Experimental techniques
pared to atmospheric pressure and will result in a small increase A series of experiments was performed to explore the effects of
in dissolution rates. rate of infiltration, viscosity of infiltrating pore fluid, g-level, and
Holocher et al. (2003) describe a kinetic bubble dissolution particle size on degree of saturation that may be achieve by infil-
(KBD) model for seeping groundwater. Their model assumes the tration of water into dry sand. The test set up included a 100 mm
ideal gas law, where the dissolved gas in the liquid obeys Henry's diameter clear acrylic plastic cylinder for the sand specimen, a
Law at the bubble interface, and uses the diffusion equation to water supply reservoir connected by a needle valve, and solenoid
compute the transport of dissolved gas away from the bubble valve as shown in Fig. 3. The syringe was attached after infiltration
through a boundary layer. Beyond the boundary layer, the fluid to measure saturation using a procedure described later. Several
dissolved gas is assumed to be uniformly mixed, but the thickness experiments were conducted in the laboratory at 1g, but experi-
of the boundary layer was shown to depend on the groundwater ments were also performed at 8g and 64g in the mini-drum cen-
flow regime. Holocher et al. (2003) explain that the dispersion trifuge at the Schofield Center at Cambridge University. The sand
associated with water flow in tortuous pores results in a smaller was dry-pluviated into the cylinder and carefully shaped using
effective boundary layer thickness. For stagnant water the bound- vacuum and a template to fit tightly against the conically ma-
ary layer is thicker and therefore dissolution rate is much slower chined sealed acrylic cap. The purpose of the conical top cap was
than for flowing water. to minimize trapping of air bubbles in the corners of the cylinder.
Black and Lee (1973) performed an experimental investigation Achieved densities were very consistent as shown in Table 3. The
to quantify the time required for saturation of trapped gas bub- water supply cylinder was filled with water. As noted in Table 3,
bles. In their experiments, they attempted to create a uniformly sometimes the water was de-aired and sometimes it was initially
unsaturated sample of a coarse Ottawa sand. Then they measured at equilibrium with atmospheric pressure. In some tests, the wa-
the initial degree of saturation and calculated the theoretical back ter was mixed with hydroxypropyl methylcellulose (Stewart et al.
pressure, p2, required to saturate the specimen using an equation 1998) to produce viscous pore fluid. For most tests the gas in the
similar to eq. [7a]. They defined a ratio RBL = (p − p1)/(p2 − p1) to specimen prior to infiltration was air at atmospheric pressure, but
quantify the back pressure during saturation compared with the in two of the experiments, the air was replaced by CO2 gas at
minimum back pressure required to saturate the specimens. For atmospheric pressure.
RBL values between about 2 and 4 and initial degrees of saturation Three different sands were used for the experiments. Fraction A
greater than 88%, they found the time required for saturation sand (D10 = 0.9 to 1.2 mm, where D10 is the grain diameter that
was between 1 week and 1 month. Black and Lee (1973) also corresponds to 10% finer by weight) and Fraction E sand (D10 =
suggested that further speed-up of the saturation process could 0.11 mm) were obtained from David Ball Specialist Sands, Cam-

Published by NRC Research Press


138 Can. Geotech. J. Vol. 50, 2013

Fig. 3. Apparatus for infiltration experiments. Hr, height of water in the rigid tube. (All dimensions in centimetres.)
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19
For personal use only.

Table 3. Test matrix and test results for infiltration experiments.


Saturation
Dry density Pore fluid De-airing CO2 Mass and Syringe Time for
Test ID (Mg/m3) viscosity (N·s/m2) vacuum (kPa) flooding? volume method method g/ge out-flow (min) 103Bo 108Ca
A1 — 0.001 0 No 0.985 0.983 1 3 34 930
A2 1.64 0.1 −85 No — 0.982 1 132 34 2150
A2a 1.64 0.1 −85 No 1.002 0.999 — — — —
E1 1.582 0.001 0 No 0.859 0.859 1 11 0.41 250
E2 1.590 0.001 0 No 0.810 0.837 1 105 0.41 26
E3 1.587 0.001 0 No 0.946 0.953 64 7 26 400
E3b 1.587 0.001 0 No 0.955 0.956 — — — —
E4 1.591 0.001 −85 Yes 0.986 1.000 64 12 26 230
E5 1.591 0.05 −85 Yes 0.988 1.000 64 105 26 1300
E6 1.590 0.001 0 No 0.781 0.823 8 250 3.3 11
H1 1.594 0.1 0 No 0.970 0.965 1 787 2.4 350
H2 1.621 0.001 0 No 0.962 0.984 8 20 20 140
H3 1.623 0.001 0 No 0.958 0.979 1 5.5 2.4 510
H4 1.623 0.001 0 No 0.922 0.940 1 420 2.4 6.6
Note: First initial of test ID identifies the sand used: A, Fraction A; E, Fraction E; H, Hostun sand. The surface tension, Ts, is taken to be 0.072 N/m. ge, acceleration
due to earth's gravity, 9.81 m/s2.
aAfter letting sample sit.

bAfter letting sample sit 6 days at 1g.

bridge, UK. Hostun sand (D10 = 0.27 mm), was obtained from Si- conventional sieves. An estimated gradation curve sketched
belco Hostun, Paris, France. Particle-size distributions for these through three meaningful data points for the Fraction A sand is
sands are shown in Fig. 4. Several curves are shown for Fraction E believed to provide an accurate-enough characterization.
sand indicating the variability of several samples. Hostun sand The needle valve in Fig. 3 was adjusted to restrict the flow rate
data was much less variable. It should be noted that the particle- into the specimen and then the solenoid valve was opened and
size distribution for the two finer sands was determined using a water began to flow into the bottom of the cylinder. As water
“single particle optical sizing” (SPOS) system described by White flowed into the specimen, the water level in the supply dropped
(2003), who stated that relative to conventional sieving, the SPOS and the water level in the soil cylinder increased; no attempt was
system oversized the particles by 20% to 40%. The Fraction A sand made to maintain a constant gradient or flow rate. The infiltration
was too coarse for the SPOS, so it was sized using the available was terminated after water began to fill the rigid clear plastic tube

Published by NRC Research Press


Kutter 139

Fig. 4. Grain-size distribution for the sands used in infiltration experiments.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

above the sand. It should be noted that for the finer sands (Hostun rection to the volume change was made to account for the com-
and Fraction E), and especially in 1g tests, gradients due to capil- pliance of the soil cylinder (⌬Vcor). The gas pressure, p2, after
lary tension had a major effect on the infiltration. In fact, for the pulling (⌬Vsyringe) was calculated assuming isothermal expansion,
finest sand at 1g, the head loss in the capillary fringe due to cap- p1V1 = p2V2, where p1 is atmospheric pressure, V1 = Vtop1, and V2 =
illary tension exceeded the head loss due to the elevation differ- Vtop1 − ⌬Vrt + ⌬Vsyringe is obtained from eq. [12]. Then the degree of
ence. For all of these experiments, the gas pressure was held at saturation is obtained from eq. [11]. The volume of gas in the
atmospheric pressure during injection. (Note that Ueno (1998) rec- specimen (Va1) under the initial absolute pressure (p1) and degree
ommended that the pore gas pressure should ideally be under of saturation (S1), was computed based on the known change in
vacuum during infiltration.) pressure and volume change of the specimen as follows:
For personal use only.

Techniques for assessing saturation p2(⌬Vrt ⫺ ⌬Vcor)


Two techniques were used to measure the degree of saturation.
Va1 p1 ⫺ p2
One of these (the conventional mass and volume method) was [11] S1 ⫽ 1 ⫺ ⫽1⫺
based on careful determination of the volume of the specimen Vv Ms
Vt ⫺
(VT), the mass of the solids (Ms), the total mass of the specimen (Mt) ␳s
and water (Mw = Mt − Ms), the density of the solids (␳s), and the
density of the water (␳w = 0.998 Mg/m3) using eq. [10]. It may be
seen that in one case (E5) the computed degree of saturation was where p2 is given by the gas law assuming isothermal conditions
1.002, proving that the error in the mass and volume method was
at least 0.2% for that specimen. It should be noted that small Vtop1
[12] p2 ⫽ p1
errors in the densities of water, solids, and total volume each Vtop1 ⫺ ⌬Vrt ⫹ ⌬Vsyringe
contribute to this error. The measured density of solids was
2.670 Mg/m3 for the Fraction A sand and 2.650 Mg/m3 for the
The syringe method, as implemented, assumes the following:
fraction E sand. For results presented in Table 3, the density of
solids for the Hostun sand was assumed to be 2.650 Mg/m3. 1. The ideal gas law and isothermal expansion of the gas (pV =
constant) for the air above the water in the rigid tube.
Vw Mw / ␳w 2. The ideal gas law and isothermal expansion of gas bubbles in
[10] S⫽ ⫽
Vv Vt ⫺ (Mw / ␳w) ⫺ (Ms / ␳s) the soil specimen.
3. The initial air pressure in the air bubbles is equal to p1, and
furthermore, that p1 = atmospheric pressure = 101.3 kPa. It is
A new technique called the “syringe method” was developed to understood that, due to surface tension around the bubble,
more quickly and reliably measure small concentrations of air the pressure inside the air bubbles is ua = uw + 2 Ts/r ⬃ 3 kPa
bubbles in a rigid container. This method has many similarities to (where uw is the pressure in the water surrounding the air
the new method presented by Okamura and Inoue (2012). This test bubble) for a 0.1 mm diameter bubble, but the bubble radii
was performed by closing the valve at the bottom of the soil were not measured and this was considered to be negligible.
cylinder and attaching a syringe to the top of the rigid tube with a 4. The volume of gas going into or out of solution while the
flexible tube running through a rubber stopper. The volume of gas syringe test is conducted is negligible (the reasonability of this
between the water level and the syringe is denoted by Vtop1. Then, assumption was confirmed by checking that the computed
using the syringe, a known volume of gas (typically 5 cm3) was saturation was not sensitive to the magnitude or sign of
pulled out of the top of the rigid tube (⌬Vsyringe). This reduces the ⌬Vsyringe).
air pressure by an amount that can be calculated by the ideal gas
law. The change in pressure causes the air bubbles in the speci- Table 4 compares accuracy of the new “syringe method” to the
men to expand and this increases the height of water in the rigid accuracy of the conventional method by investigating the theoret-
tube (Hr in Fig. 3). The change in height of the water in the rigid ical sensitivity of the computed volume of air to potential errors
tube along with the known area of the rigid tube was used to in the measured input parameters for an example specimen with
determine the change in volume of the water within the rigid S = 0.99 (1% air voids). For the conventional method, a 1% error in
tube (⌬Vrt). As the specimen was contained in an approximately density or mass of water leads to an error factor (computed air
rigid container, the change in water height is primarily due to the volume / actual air volume) of 1.98 (98% error). If the density of
expansion of gas bubbles in the sand. Nevertheless, a small cor- solids, volume of specimen or mass was in error by 1%, the error

Published by NRC Research Press


140 Can. Geotech. J. Vol. 50, 2013

Table 4. Results of sensitivity study to compare errors of the conventional and syringe methods.
Conventional method Proposed syringe method
Va/Vv actual Source of error Va/Vv calculated Error factor Va/Vv actual Source of error Va/Vv calculated Error factor
0.01 1% error in ␳w 0.0198 1.98 0.01 1% error in p1 0.0100 1.002
0.01 1% error in ␳s 0.0295 2.45 0.01 3% error in Vtop1 0.0104 1.035
0.01 1% error in Vt 0.0341 3.41 0.01 −3% error in DVsyringe 0.0104 1.045
0.01 1% error in Ms 0.0342 3.42 0.01 10% error in ⌬Vrt 0.0115 1.149
0.01 1% error in Mw 0.0199 1.99 0.01 −50% error in DVcor 0.0110 1.102
Note: The magnitude of the sources of error were qualitatively selected to be consistent with the accuracy of the equipment and techniques used for this research.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

factor would be about 2.4 to 3.4. For the syringe method, the as the pore gas, 99.9% saturation was obtained. For cases E4 and
largest error factor (computed air volume / actual air volume) is E5, prior to infiltration, the air in the sand pores was replaced
1.15 (15% error) corresponding to a 10% error in the measured by flushing with CO2 several times. 100% saturation was
change in volume, Vrt. A 3% anticipated error in input parameters achieved for the finest sand when infiltration was by de-aired
Vtop1 and dVsyringe would lead to a similar 3.5% to 4.5% error in the water into a system with pores initially saturated with CO2 pore
measured volume of air in the specimen. The uncertainty in the gas.
correction factor for compliance of the container would be ex-
pected to produce an error factor of about 1.10 (10% error) in air Snapshots taken during infiltration are shown in Fig. 5. When
volume. This sensitivity study shows that the syringe method is viewing these photographs, keep in mind that pore sizes are in-
likely to be more accurate than the conventional method for ac- fluenced by the rigid wall — the mean pore size would be larger
curate determination of small volumes of gas trapped in the spec- on the boundary and therefore, capillary rise on the boundary is
imen. Furthermore, the syringe method is nondestructive and less than it is in the interior. As such, images of the boundary are
much less tedious than the conventional method of measuring not necessarily representative of infiltration patterns in the inte-
masses and densities to determine total, solid, and liquid volumes rior. In specimen H4, for example, it appears that there are many
and then subtracting to determine the volume of air. The conven- bubbles on the boundary, yet the measured average degree of
tional method requires meticulous attention to careful measure- saturation of this specimen was about 94%. For the fine sand at 1g
ments and still yields greater error. The conventional method (Fraction E), the capillary rise is greater than the specimen height.
requires disassembly and re-assembly of the specimen to repeat At 8g, the apparent capillary fringe for Fraction E sand (the eleva-
For personal use only.

the measurements required to calculate air volume. The syringe tion difference between the top-most wet sand and the elevation
method is nondestructive and can be conducted repeatedly in of the wetted container) is about 40 mm; at 64g, the apparent
case there is some reason to suspect an error. The results of the capillary fringe is about 5 mm. As expected, the height of the
syringe method presented in Table 3 are, in fact, the result of apparent capillary fringe is approximately inversely proportional
several repetitions of the syringe method procedure on each to the g-level (Garnier et al. 2007).
specimen.
Analysis and discussion
Infiltration results Figure 6 plots the degree of saturation obtained (from the sy-
Some of the important information in Table 3 is highlighted ringe method) after infiltration for all the tests. Data (shown in
below. Table 3) obtained after letting the sample sit for an extended
• The pore fluid was water except in specimens A2, E5, and H1 for period of time are not plotted.
which the viscosity of the water was increased by addition of Ratnam (1996) and Ratnam et al. (1996) studied displacement of
hydroxypropyl methyl cellulose as described by Stewart et al. oil from soil pores by water infiltration by plotting the water
(1998). saturation as a function of Ca and Bo. These results, also shown in
• The time for outflow (the time between initial wetting at the Fig. 6, are in remarkable agreement with the results of the present
base and water egress into the rigid tube at the top) varied study in Fig. 6a. For these tests, Ca was varied over only a small
between 3 and 787 min. range, whereas it was varied by several orders of magnitude in the
• The capillary number, Ca = ␩v/Ts, varied by a factor of 300. present study. It should be mentioned that Ratnam (1996) used the
The largest Ca value was obtained in specimen A2 with a mean particle size for his computation of Bo. For Fig. 6, Bo values
viscous fluid injected rapidly in the coarse sand and the were computed using r = D10/2 from eq. [2].
smallest Ca value was obtained in specimen H4 with low Recently, Okamura and Inoue (2012) reported saturation ob-
viscosity fluid injected slowly. Note that Ca is not explicitly tained by infiltration of water into air on the centrifuge for Toy-
dependent on particle size. It is, however, much easier to oura sand (D10 = 0.11 mm) and Keisha sand (D10 = 0.068 mm).
obtain high infiltration velocities for coarse sands because Okamura and Inoue (2012) did not specify how or if the water they
they are more permeable. introduced was de-aired, and did not consider the effect of disso-
• The Bond number, Bo = (␳f − ␳g)gr2/Ts, increases with g and lution of gasses in their analysis or description of the test proce-
decreases with particle size. The surface tension, Ts, has not dure. They did some infiltration tests at atmospheric pressure and
been observed to be very sensitive to addition of methyl cellu- some at vacuum. Dissolution effects are expected to be small for
lose so for the present work, Ts is assumed constant at displacement of air by water under atmospheric pressure; there-
0.072 N/m. Similar Bo values were obtained for the coarse sand fore, only their tests at atmospheric pressure have been added to
at 1ge, the medium sand at 8ge, and the fine sand at 64ge. Fig. 6.
• In most cases, the infiltrating water was at equilibrium with the For all cases in Fig. 6, if Bo > 0.01, the achieved saturation is
atmosphere prior to infiltration and hence gas dissolution did between 95% and 98% when gas dissolution and pressure changes
not play a significant role. But for three cases (A2, E4, and E5), do not play a significant role. The Bond number, Bo, can be
the water was de-aired prior to infiltration. The quality of the thought of as being proportional to particle size divided by the
de-airing is indicated in Table 3 as 85 kPa (the vacuum pressure height of capillary rise. Trapping of gas bubbles is more likely at
at which bubbles were observed to transition between growing small Bo when capillary rise is much greater than the particle size,
and shrinking in the de-aired water chamber). For A2, with air consistent with Fig. 6.

Published by NRC Research Press


Kutter 141

Fig. 5. Observations of saturation front during infiltration experiments.

Note: more fingering


in E2 than E1 and
evidence of horizontal
stratification in E2 and
E1. The bottom of the
apparent capillary
fringe is not visible.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

(a) E1 (1 g, 11 min) (b) E2 (1 g, 105 min)


Two images from test
E3. At 64 g, the
apparent capillary
fringe is about 0.5 cm
thick. (d) is an
enlarged detail of the
image shown in (c).

(c) E3 (64 g, 12 min) (d) E3 (64 g, 12 min)


Two images from test
E6. At 8 g, the
apparent capillary
fringe (distance from
For personal use only.

arrow to arrow is
about 4 cm) The
thickness of the black
stripes is 5 mm.
(e) E6 (8 g, 250 min) (f) E6 (8 g, 250 min)
For Hostun sand
infiltrated at 1 g,
pockets of trapped air
collected at the
boundary of the
specimen, especially
obvious for very slow
infiltration. The
apparent capillary
fringe is about 4 cm.

(g) H3 (1 g, 5.5 min) (h) H4 (1 g, 420 min)


Capillary rise and air
trapping are not
apparent with Fraction
A with D10 = 1 mm. (j)
shows a close up view
of the interface in (i).

(i) A2 (1 g, 134 min) (j) A2 (1 g, 134 min)

Published by NRC Research Press


142 Can. Geotech. J. Vol. 50, 2013

Fig. 6. Saturation obtained after infiltration at varying (a) Bond number, Bo, and (b) capillary number, Ca. O & U, Okamura and Inoue (2012);
Ratnam, Ratnam (1996) and Ratnam et al. (1996).
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19
For personal use only.

For Bo < 0.01, better saturation is achieved at high infiltration air that forms a vapor or gas bubble barrier. Such barriers may
rates (high Ca values). The three points in Fig. 6b for Fraction E form when a material with larger capillary suction overlies a ma-
sand (solid triangles) with saturation between 0.82 and 0.86 had terial with low capillary suction, for example,
Bo < 0.01, showing an increasing trend with Ca. The three points
for Hostun sand (solid squares) at Bo = 2.4 × 10−3 show a similar 1. In layered soil profiles — Gas barriers may develop where a fine
increase with Ca. The capillary number, Ca, compares viscous and soil overlies a coarse soil; air may be trapped under a fine layer
capillary forces. For low Ca and low Bo values, water is pulled by that is saturated by capillary tension. Heterogeneities on the
capillary tension into tiny pores high above the water table, rais- order of a couple millimetres in thickness were exposed dur-
ing the potential for gas to be trapped in the large voids below the ing infiltration of the finer sand (see Figs. 5a and 5b).
wetting front. At high Ca, when viscous forces are large, the flow 2. Where the supply tube hits the sand — For the test set up shown in
in the large pores is increased, less trapping of air occurs, and Fig. 3, it was discovered that if the water supply flow rate was
higher degrees of saturation are obtained. less than the rate where capillary tension could pull the water
Fraction E sand and Keisha sand consistently produced smaller out of the supply tube (as illustrated in Fig. 7), the sand would
degrees of water saturation than Hostun and Toyoura sands even pull water up from the edges of the tube and release air back
though they were tested over overlapping ranges of Bo and Ca, into the supply tube, creating an air bubble in the inlet below
suggesting that there are other factors affecting the process. It is the screen. This air bubble has no place to go once the pores
suspected that this difference may be related to the presence of a above are blocked by water, and the bubble can block the
small percentage of fines in Fraction E sand and 13.8% fines for supply tube and radically slow down the infiltration process.
Keisha sand. It was noticed that during pluviation of Fraction E
sand through air, dust production was much more obvious than Benefits of saturating in a centrifuge
for the other sands. Due to air resistance, the finer particles spread The saturation process can be sped up and difficulties with
into a wider pluviation stream than coarse particles. This pro- vapor barriers can be minimized by performing the infiltration
duced noticeably heterogeneous layering (variable darkness of while the specimen is spinning in a centrifuge. The centrifugal
wet sand) in Figs. 5a and 5b, with a layer spacing on the order of acceleration helps by increasing Bo and effective stresses so that
about 5 mm. This introduces another characteristic length (be-
larger pore pressure gradients can be applied without causing a
sides particle size and specimen size) to the problem, which is not
boiling condition as the pore pressure ratio, Ru, approaches 1. The
accounted for in Fig. 6. Trapping of air under the dust-rich layers
efficacy of infiltration during application of increased gravity was
might explain the lower saturation of Fraction E sand when com-
demonstrated experimentally for the very fine sand spun at 64g.
pared with Hostun sand.
Okamura and Inoue (2012) also clearly demonstrated the benefits
For the two cases where the air was replaced by CO2 and de-
of saturation on a centrifuge. Additional work is also required to
aired water was used to displace the CO2, 100% saturation was
determine how to ensure that pore pressure transducers embed-
achieved. The rate of dissolution can be enhanced by flushing or
ded in sand specimens can be reliably saturated during this alter-
flowing water through the specimen, as this speeds up the diffu-
nate procedure.
sion of dissolved gas away from bubbles in the soil and reduces
dissolution times (Holocher et al. 2003).
Conclusions
Capillary barrier problems Achieving complete water saturation is known to be important
During the course of this work it became apparent that the for centrifuge models, especially with respect to centrifuge mod-
infiltration process can by severely hampered by the trapping of eling of liquefaction problems. A new method of measuring the

Published by NRC Research Press


Kutter 143

Fig. 7. Illustration showing how an air bubble may form at the de-gassed prior to infiltration. Although saturation of pore pres-
water inlet. sure transducers is still a concern, the saturation process could be
sped up and difficulties with vapor barriers could be minimized
by performing the infiltration while the specimen is spinning in a
centrifuge.

Acknowledgments
The author would like to thank his hosts and colleagues Gopal
Madabhushi and Malcolm Bolton, at the Schofield Center of Cam-
bridge University, for enabling his sabbatical visit from June to
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

December 2010. Students and staff at the Schofield Center (espe-


cially Richard Adams, John Chandler, Ulas Cilingir, Stuart Haigh,
Charles Heron, Jennifer Haskell, Chris McGinnie, Brenden McMa-
hon, Kristian Pether, Mark Stringer, Paul Shepley, Philip Wall-
bridge, Tony Wang, and Michael Williamson) all provided
important assistance at various stages of this work. The sabbatical
visit was facilitated by an Overseas Fellowship at Churchill Col-
lege, Cambridge, and a Distinguished Visiting Fellowship award
degree of saturation (the syringe method) of specimens in a rigid from the Royal Academy of Engineering in London. Dan Wilson at
container is presented. The method is better able to accurately the UC Davis Center for Geotechnical Modeling authored CGM
measure small air contents; for S > 99%, the method has a resolu- (2011) that was later adapted by Ueno (1998) and helped develop
tion of about 0.1%. the concepts presented in this paper.
This paper sheds light on the effects of Bo, Ca, and gas disso-
lution on the “soil-water characteristic curve” (the relation be- References
tween water pressure and degree of saturation). For Bo > 10−2, Black, D.K., and Lee, K.L. 1973. Saturating laboratory samples by back pressure.
the height of capillary rise was small enough that the volume of Journal of Soil Mechanics and Foundations Division, ASCE, 99(SM1): 75–93.
bubbles trapped during infiltration was less than about 5% of Blunt, M.J., and Scher, H. 1995. Pore-level modeling of wetting. Physical Review
the pore volume. Bo can be increased by increasing particle size E, 52(6): 6387–6403. doi:10.1103/PhysRevE.52.6387.
or by increasing the g-level by spinning in a centrifuge. Espe- CGM. 2011. Saturation procedures. Wiki article posted by the UC Davis Center for
Geotechnical Modeling. [Accessed July 2011 at http://wiki.cgm.ucdavis.edu/
For personal use only.

cially for Bo < 10−2, the degree of saturation depends on Ca. At wiki/pages/W4D0A1u4/Saturation_Procedure.html.]
low Bo values, and small infiltration flow rates, capillary action Chaney, R., 1978. Saturation effects on the cyclic strength of sands. Earthquake
sucks water far above the water table, increasing the potential Engineering and Soil Dynamics, 1: 342–358.
for bubble entrapment. de Gennes, P.-G., Brochard-Wyart, F., and Quere, D. 2004. Capillarity and wetting
In water, the solubility of carbon dioxide gas is 56 times greater phenomena: drops, bubbles, pearls, waves. Springer Science+Business Media
Inc.
than N2 gas; therefore to obtain high degrees of saturation, it is Eseller-Bayat, E.E. 2009. Seismic response and prevention of liquefaction failure
very advantageous to replace a large fraction of the N2 gas by CO2 of sands partially saturated through introduction of gas bubbles. Ph.D. dis-
prior to infiltration. As indicated in Table 1, de-gassed water can sertation, Dept. of Civil and Environmental Engineering, Northeastern Uni-
dissolve about 83% if its volume of CO2, but only 1.5% of its volume versity, Boston, Mass.
Fredlund, D.G., and Rahardjo, H. 1993. Soil mechanics for unsaturated soils.
of N2. Wiley, New York.
Lowering the fluid pressure prior to infiltration (by application Friedman, S.P. 1999. Dynamic contact angle explanation of flow rate-dependent
of a vacuum) decreases the density of the gas and reduces the mass saturation-pressure relationships during transient liquid flow in unsaturated
of air trapped during infiltration. Increasing the fluid pressure porous media. Journal of Adhesion Science and Technology, 13(12): 1495–
1518. doi:10.1163/156856199X00613.
after bubbles are trapped (either by releasing the vacuum or ap-
Garnier, J., Gaudin, C., Springman, S.M., Culligan, P.J., Goodings, D., Konig, D.,
plying back pressure) decreases the volume of the bubbles and Kutter, B., et al. 2007. Catalogue of scaling laws and similitude questions in
increases the solubility, leading to improved saturation. Equa- geotechnical centrifuge modeling. International Journal of Physical Model-
tions are provided for quantitative calculation of relationships ling in Geotechnics, 7(3): 1–24.
between degree of saturation and pressure for different gasses. Holocher, J., Peeters, F., Aeschbach-Hertig, W., Kinzelbach, W., and Kipfer, R.
2003. Kinetic model of gas bubble dissolution in groundwater and its impli-
The time required to dissolve trapped bubbles depends on the cations for the dissolved gas composition. Environmental Science & Technol-
size of the bubbles, the solubility of the bubbles, and their total ogy, 37(7): 1337–1343. doi:10.1021/es025712z. PMID:18930245.
volume. For air bubbles, Black and Lee (1973) found that the time Ishihara, K., Huang, Y., and Tsuchiya, H. 1998. Liquefaction resistance of nearly
required to dissolve bubbles could be days or weeks. As CO2 is 56 saturated sand as correlated with longitudinal wave velocity. In
Poromechanics: a tribute to Maurice A. Biot. Balkema, Rotterdam, the Neth-
times more soluble, the time required for CO2 to dissolve would
erlands. pp. 583–586.
be expected to be much faster. Kokusho, T. 2000. Correlation of pore-pressure B-value with P-wave velocity and
Vapor barriers can develop at the interface between soils with Poisson's ratio for imperfectly saturated sand or gravel. Soils and Founda-
different particle sizes or where a water supply tube touches a soil tions, 40(4): 95–102. doi:10.3208/sandf.40.4_95.
with significant capillary action; these vapor barriers can practi- Kutter, B. L. 1995. Recent advances in centrifuge modeling of seismic shaking.
State-of-the-Art Paper In Proceedings, Third International Conference on Re-
cally block the saturation process. Some of these mechanisms are cent Advances in Geotechnical Earthquake Engineering and Soil Dynamics,
described, but more detailed study is warranted. St. Louis, Mo. Vol. 2, pp. 927–942.
Naesgaard et al. (2007) questioned the quality of saturation pro- LeBihan, J.-P., and Leroueil, S. 2002. A model for gas and water flow through the
duced using the procedure described by CGM (2011) and Ueno core of earth dams. Canadian Geotechnical Journal, 39(1): 90–102. doi:10.1139/
t01-081.
(1998). By consideration of the effects of Ca, Bo, and the high
Lowe, J., and Johnson, T.C. 1960. Use of back pressure to increase degree of
solubility of CO2, it is concluded that if carefully followed, the saturation of triaxial test specimens. Presented at the ASCE Research Confer-
procedure described by Ueno (1998) is able to reliably produce ence on Shear Strength of Cohesive Soils, Boulder, Colo., June. pp. 819-836.
100% saturation of uniform specimens for laboratory or centrifuge Lu, N., and Likos, W.J. 2004. Unsaturated soil mechanics. Wiley, Hoboken, New
model tests. Jersey.
Marulanda, C., Culligan, P.J., and Germaine, J.T. 2000. Centrifuge modeling of air
As an alternative to the proposed procedure, complete satura- sparging — a study of air flow through saturated porous media. Journal of
tion can be theoretically obtained without application of a vac- Hazardous Materials, 72(2-3): 179–215.
uum during infiltration if air is replaced by CO2 and the water is Muir Wood, D.M. 2004. Geotechnical modelling. Spon Press, New York.

Published by NRC Research Press


144 Can. Geotech. J. Vol. 50, 2013

Mulilus, J.P., Chan, C.K., and Seed, H.B. 1975. The effects of method of sample methylcellulose as a viscous pore fluid in centrifuge models. Geotechnical
preparation on the cyclic stress-strain behavior of sands. University of Cali- Testing Journal, 21(4): 365–369. doi:10.1520/GTJ11376J.
fornia, Berkeley, Calif. EERC Report No. 75-18. Stringer, M.E., and Madabhushi, S.P.G. 2009. Novel computer-controlled satura-
Naesgaard, E., Byrne, P.M., and Wijewickreme, D. 2007. Is P-wave velocity an tion of dynamic centrifuge models using high viscosity fluids. Geotechnical
indicator of saturation in sand with viscous pore fluid? International Journal Testing Journal, 32(6). doi:10.1520/GTJ102435.
of Geomechanics, ASCE, 7(6): 437–443. doi:10.1061/(ASCE)1532-3641(2007)7: Takahashi, H., Kitazume, M., Ishibasi, S., and Yamawaki, S. 2006. Evaluating the
6(437). saturation of model ground by P-wave velocity and modeling of models for a
Okamura, M., and Inoue, T. 2012. Preparation of fully saturated models for liquefaction study. International Journal of Physical Modelling in Geotech-
liquefaction study. International Journal of Physical Modeling in Geotech- nics, 6(1): 13–25.
nics, 12(1): 39–46. doi:10.1680/ijpmg.2012.12.1.39. Tuve, G.L., and Bolz, R.E. 1973. CRC handbook of tables for applied engineering
Okamura, M., Ishihara, M., and Tamura, K. 2006. Degree of saturation and liq- science. 2nd ed. CRC Press, Boca Raton, Fla. ISBN 0-8493-0252-8.
uefaction resistances of sand improved with sand compaction pile. Journal of Ueno, K. 1998. Methods of preparation of sand samples. In Proceedings of the
Geotechnical and Geoenvironmental Engineering, ASCE, 132(2): 258–264. International Conference Centrifuge 98, Tokyo. Edited by Kimura, Kusakabe,
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by UNIVERSITY OF PITTSBURGH on 09/13/19

doi:10.1061/(ASCE)1090-0241(2006)132:2(258). and Takemura. Balkema, Rotterdam. Vol. 2, pp. 1047–1056.


Ratnam, S. 1996. Geotechnical centrifuge modeling of the behavior of Light Weinberg, M.C. 1981. Surface tension effects in gas bubble dissolution and
Nonaqueous Phase Liquids (LNAPLs) in sand samples under hydraulic flush- growth, Chemical Engineering Science, 36(1): 137–141.
ing. S.M. thesis, Massachusetts Institute of Technology, Cambridge, Mass. White, D.J. 2003. PSD measurement using the single particle optical sizing
Ratnam, S., Culligan-Hensley, P.J., and Germaine, J.T. 1996. Modeling the behav- (SPOS) method. Géotechnique, 53(3): 317–326. doi:10.1680/geot.2003.53.3.317.
ior of LNAPLs under hydraulic flushing, Non-Aqueous Phase Liquids (NAPLs) Wise, D.L., and Houghton, G. 1966. The diffusion coefficients of ten slightly
in subsurface environment: assessment and remediation. In Proceedings of soluble gases in water at 10–60°Cs. Chemical Engineering Science, 21(11):
the Specialty Conference in conjunction with the ASCE National Convention, 999–1010. doi:10.1016/0009-2509(66)85096-0.
Washington, D.C., 10-14 November 1996. ASCE. pp. 595-606. 978-0-7844- Xia, H., and Hu, T. 1991. Effects of saturation and back pressure on sand lique-
0203-0 or 0-7844-0203-5. faction. Journal of Geotechnical Engineering, 117(9): 1347–1362. doi:10.1061/
Sander, R. 1999. Compilation of Henry's law constants for inorganic and organic (ASCE)0733-9410(1991)117:9(1347).
species of potential importance in environmental chemistry. Version 3. Yoshimi, Y., Yanaka, K., and Tokimatsu, K. 1989. Liquefaction resistance of a
Available at http://www.henrys-law.org [accessed 26 January 2011. partially saturated sand. Soils and Foundations, 29(3): 157–162. doi:10.3208/
Stewart, D.P., Chen, Y.R., and Kutter, B.L. 1998. Experience with the use of sandf1972.29.3_157.
For personal use only.

Published by NRC Research Press

You might also like