Buchgraber 2012

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Transp Porous Med (2012) 95:647–668

DOI 10.1007/s11242-012-0067-0

A Study of Microscale Gas Trapping Using Etched


Silicon Micromodels

Markus Buchgraber · Anthony R. Kovscek ·


Louis M. Castanier

Received: 9 May 2012 / Accepted: 10 August 2012 / Published online: 28 August 2012
© Springer Science+Business Media B.V. 2012

Abstract Immobilization and trapping of carbon dioxide (CO2 ) enhances the security of
geological storage. Trapping mechanisms have been characterized in four groups: structural,
residual, dissolution, and mineralization. While structural trapping acts immediately when
injection starts and is well investigated, the contribution of residual and dissolution trapping
increases over storage time and these contributions need to be better understood for bet-
ter predictions. This paper focuses on an experimental pore-scale investigation of residual
and capillary trapping. CO2 –water imbibition experiments were conducted in micromodels
whose homogenous pore space is geometrically and topologically similar to Berea sand-
stone. Microvisual data, photographs and video footage, describes the trapping mechanism
and, especially, the disconnection and shrinkage of the CO2 phase. Results show that depend-
ing on the flow rate of the imbibing water different trapping mechanisms are observed. Lower
flow rates, comparable to the trailing edge of a CO2 plume, lead to more snap-off events and
greater trapped residual saturation, whereas rates comparable to the near wellbore area dur-
ing enhanced sequestration showed displacement of gas bubbles and greater dissolution that
ultimately leads to very low or zero gas saturations. Furthermore, complete dissolution events
showed that homogenous as well as heterogeneous dissolution occurs. Whereas the latter is
subdivided into microbubble formation and dissolution on crevices or pore roughness, the
former occurs without the influence of pore walls. Based on the observations we suggest that
the type of rock and its roughness as well as the fines present at the CO2 brine interface are
important factors determining the dissolution mechanism.

M. Buchgraber · A. R. Kovscek (B) · L. M. Castanier


Department of Energy Resources Engineering, Stanford University,
367 Panama St., Stanford, CA 94305, USA
e-mail: kovscek@stanford.edu
M. Buchgraber
e-mail: markus.buchgraber@stanford.edu
L. M. Castanier
e-mail: louisc@stanford.edu

123
648 M. Buchgraber et al.

Keywords Dissolution · Micromodel · Carbon sequestration · Imbibition ·


Residual gas saturation

List of symbols
NCa Capillary number
q Flow rate cm3 /s
U Velocity (cm/s)
σ Interfacial tension (mN/m)
μ Viscosity (Pa s)

1 Introduction

Oil and gas reservoirs, coal seams, and gas shale notwithstanding, saline aquifers represent
an important storage medium for carbon dioxide (CO2 ). Sequestration of CO2 in deep saline
aquifers does not provide any added value, in comparison to enhanced oil or gas recovery,
but it has other advantages. First, the estimated storage capacity of saline aquifers is very
large. The worldwide storage capacity estimation of saline aquifers is between 1,000 and
10,000 GtCO2 and therefore two times greater than all oil and gas reservoirs and 70 times
larger than all unmineable coal seams (IPCC 2007a). Second, most existing large CO2 point
sources are within easy access of a saline formation. Within a saline aquifer, CO2 is seques-
tered as a free phase, as physically trapped bubbles of CO2 , as CO2 dissolved in brine, and
in mineral form. The trapping mechanism that prevents the CO2 phase from escaping the
injection site is divided into (1) structural trapping, (2) residual gas trapping, (3) solubility
trapping, and (4) mineral trapping.
In this paper, we observe directly residual gas and solubility trapping mechanisms from a
pore-level perspective thereby addressing aspects of mechanisms (2, 3). Our experiments use
etched silicon micromodels that contain a 2D sandstone-like pore network. This paper starts
with an overview of residual and dissolution trapping to put our work into context. Then, the
fabrication and characteristics of the micromodels are described followed by the experimen-
tal apparatus and procedures. Results in the form of photographic sequences, image analysis
of the sequences, and a discussion conclude the paper.

1.1 Capillary Trapping

The trapping phenomena for immiscible fluids is a well-researched topic and has been inten-
sively investigated by authors such as Agarwal (1967), Roof (1970), Al Mansoori et al.
(2008), Chierici et al. (1963), Crowell et al. (1966), Firoozabadi et al. (1987), Jerauld (1997),
and Kantzas et al. (2001). Others have studied trapping of CO2 [e.g., Iglauer et al. 2011;
Pentland et al. 2011; Bennion and Bachu 2005, 2006; Bachu 2003]. During CO2 injection,
the trapping process as a result of water imbibition is combined with a dissolution–diffusion
process that leads to a more complex situation. The literature presents either the dissolution of
single bubbles or the trapping and displacement for immiscible cases (Land 1968a; Pentland
et al. 2008; Spiteri et al. 2008; Buchgraber et al. 2011). There is no or little work published
about the combination of both phenomena at the pore scale.
The work that is most related to the phenomenon of dissolution and phase disappearance
during geological CO2 sequestration is the solution gas drive mechanism during oil recovery
(Arora and Kovscek 2003; Bauget and Lenormand 2002; Bora et al. 2000; George et al. 2005;

123
A Study of Microscale Gas Trapping 649

Firoozabadi 2001; Li and Yortsos 1995; Bisperink and Prins 1994). Dissolution trapping is
controlled by phase equilibration. The local equality of fugacities in the aqueous and the gas
phase determines the amount of CO2 to be transferred. A complete dissolution of a single gas
bubble only happens after the continuous gas phase has been disconnected and immobilized
by capillary trapping. This process on the other hand is governed by flow rate and pore body
to throat size ratio in porous media. Although solution gas drive occurs in a reverse order
compared to the sequestration process, similar principles apply. The process is broken into
(1) disconnection of the continuous gas phase by snap off, displacement and dissolution, (2)
collapse of separated and trapped gas bubbles, and (3) complete dissolution of the bubble or
stabilization of a micro bubble on a pore wall.
The disconnection of continuous phases for different mobility ratios and capillary num-
bers has been studied intensively by Lenormand (1990), Morrow and Songkran (1981), and
Wardlaw (1980) however, the mobility ratio and capillary number regime applicable to the
CO2 –water fluid pair is relatively unexplored (Cinar et al. 2009). Laboratory experiments in
heterogeneous rock samples and the influence of heterogeneity on capillary trapping and rel-
ative permeability functions were performed by Perrin et al. (2009) and Krevor et al. (2011).
The effect of local capillary trapping was investigated by Pini et al. (2012). Not only is it
difficult to determine representative relative permeability and capillary pressure curves, but
also there is a lack of pore-scale understanding of the imbibition and dissolution processes,
especially for supercritical CO2 conditions.
Residual trapping of crude oil is a well understood problem in water flooding operations
and results in roughly half of the original oil in place being left in the reservoir. Recovery is
increased by increasing the flow rate and therefore creating viscous forces that overcome the
capillary forces that trap snapped-off oil. Oil mobilization has been studied experimentally
by various authors (Abrams 1975; Larson 1981; Morrow and Songkran 1981). They conclude
uniformly that there is a strong relation between residual nonwetting-phase saturation and
capillary number, NCa ,


NCa = , (1)
σ

where U is velocity, μ is the wetting-phase viscosity, and σ is the interfacial tension. Note
that Eq. (1) is a microscopic definition as appropriate to our pore-scale study. Hilfer and Øren
(1996) discuss scale dependence and the connection between pore and field-scale trapping.
To understand residual trapping, a more detailed understanding of capillary forces and vis-
cous forces at microscale during imbibition is required. Dodd and Kiel (1959) and Wardlaw
(1980) were amongst the first to study the effect of capillary pressure on multiphase flow in
porous media. Lenormand and Zarcone (1984) investigated immiscible two-phase flow under
various capillary numbers and divided pore-scale processes into the IN-type imbibitions, pis-
ton-type motion, and snap-off. IN-type imbibition mechanisms depend upon the nonwetting
fluid occupancy of pores. An I1 imbibition mechanism occurs when the nonwetting phase is
retreating and fills a single pore. An I2 imbibition mechanism occurs when the nonwetting
phase fills two adjacent pores while retreating. Both I1 and I2 mechanisms lead to relatively
rapid pore filling by the wetting phase. Depending upon the configuration of the nonwetting
phase within the pore network these events may lead to disconnection of the nonwetting
phase leading to trapping.
Roof (1970) and Lenormand et al. (1983), were among the first who explained the rela-
tionship between flow in pore corners and the snap-off mechanism. As capillary pressures
decreases the pore-corner water rivulet swells. For appropriately constricted pores with

123
650 M. Buchgraber et al.

sufficient liquid available locally, snap-off occurs that finally leads to residual trapped satu-
rations (Gauglitz et al. 1987; Ransohoff et al. 1987; Kovscek and Radke 1996, 2003).
At reservoir scale, residual trapping occurs when CO2 , driven by buoyancy forces or sub-
surface currents, migrates through the reservoir and is partially replaced by formation water,
water influx to the aquifer, or injected water. The initially continuous CO2 phase is discon-
nected and the local CO2 saturation decreases. Once the phase is discontinuous, it is trapped
by capillary forces. A CO2 bubble that occupies a large pore space cannot pass through a
small pore neck because the capillary entry pressure needed to invade the pore throat is large.
Accordingly, this trapping mechanism is referred to as residual gas or capillary trapping
(Nghiem et al. 2009; Ziqiu et al. 2009). One of the most widely used trapping models is the
Land trapping model (Land 1968a,b). It was developed to predict trapped gas saturation as
a function of the initial gas saturation based on published experimental data from water-wet
sandstone cores. The amount of trapping depends strongly on fluid and rock-type parameters.
Pentland et al. (2008) and Pentland (2010) performed trapping experiments in sandstones
with different porosities. The trapped saturations matched acceptably with the trapping mod-
els from Land (1968a,b), Jerauld (1997), and Spiteri et al. (2008); however, the residual-phase
saturations where greater than observed in the earlier experiments. A pore network simulator
showed good agreement by using intrinsic contact angles for CO2 and water in the range of
35 − 65◦ perhaps indicating that the system is not strongly wetting. The observation that the
brine CO2 system is not strongly water wet is in line with observations by other researchers
(e.g., Chalbaud et al. 2010; Yang et al. 2005). Similarly, Iglauer et al. (2009) used sandpacks
and consolidated rock samples for a trapping experiment. They concluded that the trapping
capacity reaches a maximum of 7 % when the rock porosity is around 20 %. This suggests
an optimal porosity for CO2 storage.

1.2 Dissolution

CO2 is usually injected as a supercritical fluid that forms a separate phase and is not miscible
with the brine present in the reservoir, but over time it dissolves in the aqueous phase. The dis-
solution into the aqueous phase and eventual sequestration as carbonate minerals are highly
desirable processes as they increase permanence and security of storage. Dissolution of CO2
establishes phase equilibrium locally between the overlying CO2 plume and the aqueous
phase beneath. In the case of an immobile aqueous phase, the CO2 dissolution rate is limited
by the rate of molecular diffusion that removes dissolved CO2 from the interface between
CO2 -rich and aqueous phases. This process is very slow as illustrated by Hesse (2008) and
Pruess and Zhang (2008).
At an interface where both the supercritical CO2 and the brine are at rest, the dissolution
of CO2 is slow. In this case, the dissolution rate of the CO2 across the interface is controlled
by molecular diffusion of dissolved CO2 in the brine, and the dissolution rate is expected to
decay as the concentration of dissolved CO2 increases. Much more rapid dissolution of CO2
occurs across an interface where the supercritical CO2 continues to contact fresh brine, due
to advection in either phase. The buoyancy-driven migration of the supercritical CO2 as well
as the convective motion in the brine leads to rapid dissolution (Garcia 2003).
Gas bubbles do not always undergo a complete dissolution into the liquid phase. There are
two mechanisms for stabilization of bubbles: microbubbles in bulk liquids and microbubbles
entrapped in cavities or wall roughness (Lawrence and Crum 1982). For microbubbles in
bulk liquids to become stable there are three described models. The rigid skin model by Fox
and Herzfeld (1954) proposed that organic compounds build a rigid skin about the surface of
the free gas bubble and prevent the bubble from dissolving. Secondly, the ionic skin model

123
A Study of Microscale Gas Trapping 651

Table 1 Experimental parameters

Exp.# Temp. Pressure Flowrate Velocity Nca CO2 phase


[◦ C] [psi] [ml/min] [ft/day] []

#1 22.2 11 psi 0.1 821.6 3.78E-05 Gas


#2 0.01 82.2 3.78E-06 Gas
#3 0.001 8.2 3.78E-07 Gas
#4 0.0001 0.8 3.78E-08 Gas
#5 480 psi 0.1 821.6 3.78E-05 Gas
#6 0.01 82.2 3.78E-06 Gas
#7 0.001 8.2 3.78E-07 Gas
#9 930 psi 0.1 821.6 3.78E-05 Liquid
#10 0.01 82.2 3.78E-06 Liquid
#11 0.001 8.2 3.78E-07 Liquid
#12 44.4 1,150 0.1 821.6 5.44E-05 Supercritical
#13 0.01 82.2 5.44E-06 Supercritical
#14 0.001 8.2 5.44E-07 Supercritical

(Alty 1926) states that the interface of gas bubbles possesses some excess electric charge
and thus creates surface forces that prevent the gas bubble from dissolving. And finally the
film surface active substance model is a modified rigid skin model from Sirotyuk (1970). In
this model, natural surfactants (such as dilute crude-oil components in water) stabilize the
bubble.
Pore-level observations in real time for water–CO2 imbibition experiments are required
to describe and validate the mechanisms stated above. The following experimental section
explains the method used in this paper to analyze trapping and dissolution events.

2 Experimental

Micromodel imbibition experiments with CO2 and house-distilled water were performed to
observe the trapping mechanism (dissolution–diffusion and residual trapping) and to deter-
mine the amount of trapped CO2 . Moreover, the experiments provide more information
about the shrinkage behavior of single bubbles (heterogeneous or homogeneous shrinkage)
and their resulting consequences. In addition, snap shots and videos record the interaction of
displacement and dissolution for the water–CO2 system under various conditions in porous
media. The unique advantages of using micromodels over a microtomograph in this set of
experiments are, first, the continuous recording of the imbibition and dissolution process
at the pore scale, whereas the microtomograph only gives saturation snapshots at different
times. Secondly, the number of pores that are examined in the micromodel is moderately
large. To get a better understanding of process mechanisms, a wide range of pressures at two
different temperatures and various imbibition capillary numbers have been tested. Experi-
mental conditions vary from ambient temperature and pressure where the CO2 is in the gas
phase to conditions where the carbon dioxide is in a liquid phase and finally to the super-
critical state. Table 1 gives an overview of the different conditions and the corresponding
capillary numbers and CO2 phase states.

123
652 M. Buchgraber et al.

Fig. 1 a Scanning electron microscope image that serves as the basis for network extraction. b Etched mi-
cromodel with pore throats, pore bodies, and grains. c Etched micromodel wall surface roughness 100 nm

2.1 Micromodel

Two-dimensional micromodels allow direct observation of pore-scale events. They contain


an etched pore network pattern that is directly observable with a microscope. The pattern
is virtually anything that can be etched onto glass or silicon. Micromodels are ideal for
visualization of pore-level events, such as snap off and dissolution of the nonwetting phase,
while honoring geometric and topologic properties of real rocks. They are also useful for
understanding flow processes within the pore network. One limitation that is inherent to
all micromodels is they are 2D. Extrapolation of results to 3D flow must be done carefully
and such quantitative extension is best achieved using models resulting from micromodel
observations while accounting for the change in dimensionality. In addition, micromodel
observations and trends are often helpful to interpret core-scale information. Micromodels
provide the best means to visualize fluid movement at the pore scale while honoring both
geometric and topologic properties of real rocks. In addition, a micromodel can be repro-
duced and provide identical initial conditions for new or repeated experiments. Detailed
literature reviews of micromodels are provided elsewhere (Buckley 1990; Wan et al. 1996;
Rangel-German and Kovscek 2006).
The micromodels employed here represents a 1:1 realization of pore size and pore shapes.
The pore network is drawn from a portion of a representative Berea sandstone thin section,
Fig. 1a. The creation of this particular pore network is described in detail by Upadhyaya
(2001). We note that the extraction of 3D pore features and reproducing them in two-dimen-
sions tends to increase average micromodel porosity relative to the original rock, as docu-
mented shortly, and increases pore connectivity in two dimensions. Further details on the
transfer of pore network features from thin sections to 2D micromodels are given by Buch-
graber et al. (2012).
The pore network pattern is etched into silicon using standard photolithography tech-
niques. During the etching process with deep reactive ion etching (DRIE), surface roughness
on the vertical as well as on the bottom surface of the silicon wafer is produced. The surface
roughness depends on the etching parameters and on the aspect ratio of the structures. The
root-mean-square surface roughness was estimated to be 100 nm (Pham et al. 2007) and is
illustrated in the SEM picture of a vertical etched structure in Fig. 1c.
Following etching, the micromodel is scrupulously cleaned and then bonded to an optically
flat glass plate to create a 2D porous medium. During the bonding of the etched silicon wafer
to the glass cover plate, a thin oxide layer develops on the surface of the pore structure that
gives the micromodel a water-wet characteristic. Recent experiments that were conducted

123
A Study of Microscale Gas Trapping 653

Outlet port Inlet port


Inlet Pore Inlet fracture
Channel
Inlet Pore

Cover plate
Inlet Pore
Channel
Inlet Pore Outlet port
Outlet fracture
Inlet port

(a) (b) (c)


Fig. 2 a Etched micromodel with cover plate, inlet port and fluid distribution channel. b Base image (indicated
by solid lines) is stitched together 50×50 times in x and y direction to fill an area of 5 × 5 cm2 . c Micromodel
flow schematic

with CO2 and a water UV tracer mixture clearly indicate the strongly water-wet behavior of
the micromodel (Buchgraber et al. 2012).
Grains in Fig. 1b appear as islands and pores appear as channels. The average channel
depth is 25 µm and grain sizes range from 50 to 300 µm. Pore-wall roughness is reproduced
to an extent, but heterogeneous mineralogy, that is clay particles, are not. The pore network
features a pattern of pores that is repeated 50 by 50 times across the plane of the micromodel
to fill an area of 5 by 5 cm as shown in Fig. 2. The overall pore network consists of about 600
by 600 pores. The average permeability is 950 mD whereas the porosity is 46 % resulting
in a pore volume of 0.013 ml. This volume corresponds to the size of a raindrop! Further
information on details of the fabrication process are given by Hornbrook et al. (1991) and
Upadhyaya (2001).
The completed micromodel is mounted in an aluminum or stainless steel holder, which
provides fluid delivery and production from the micromodel. The micromodel has inlet and
outlet ports at the corner points that are connected via a “fracture” or channel along the entry
and exit faces. The fluid distribution channel enables the displacement process to progress
in a linear fashion and not like a radial displacement that is observed in a five-spot pattern.
Figure 2 presents the fluid delivery system.

2.2 Experimental Approach

Although sequestration of CO2 mainly takes place under supercritical conditions, our set
of experiments covers all phase states of carbon dioxide in order to a gain a better under-
standing. In addition to the conventional water–CO2 system experiment, UV dyed water
was used to visualize better the displacement process and especially understand the wetting
characteristics of the micromodel. As a comparison, an immiscible glycerin/water mixture-
heptane imbibition experiment was conducted. It has similar viscosity ratios as compared
to the supercritical CO2 –water system. This experiment investigated pure capillary trapping
under various capillary numbers. In the following section, the different experimental setups
and procedures are described and summarized.

2.2.1 Low-Pressure Experiment

For pressures up to 80–100 psig a conventional micromodel holder is used (Rangel-German


and Kovscek 2006). The experimental apparatus used in these experiments consists of two

123
654 M. Buchgraber et al.

Fig. 3 Experimental setup for low-pressure tests

parts. The first part is the micromodel, water pump, a pressurized CO2 vessel and a back-
pressure valve, in which the CO2 and water are present, and the experiment is performed.
The second part is the microscope and a camera that captures micro and macroscale pictures
and videos during the experiment. The setup is illustrated in Fig. 3.
Carbon dioxide is stored in a pressure vessel and the water is stored in the ISCO 100D
syringe pump that operates at very low flow rates. Both are connected to the same inlet port
on the inlet fracture side. The second inlet port is kept closed. The back pressure valve,
Equilibar EB1LF1, for ultra low flow rates is connected to the outlet fracture side to one
outlet port and the second outlet port is kept closed. Microscale pictures and videos are taken
with a Nikon P5100 camera and a Nikon Eclipse ME 600 microscope with a metal halide
lamp.
At the beginning, the micromodel is fully saturated with CO2 in order to clean or remove
any residual components from the fabrication procedure. After the cleaning, the model is
fully saturated (100 %) with water. The water is degassed for 30 min with nitrogen before
using. At all times, the system is closed and the operating pressure is defined by the back
pressure valve and the water pump. After water saturation, a primary drainage process is
initiated and the model is saturated with CO2 . Note that the very small pores are still occu-
pied by the wetting fluid (water). As a final step, we start the water imbibition at a con-
stant flow rate. During this step, we record the pressure and take microscale pictures and
videos to observe the disconnection and shrinkage of CO2 bubbles due to dissolution and
diffusion. All experiments are stopped when residual gas saturation has completely dis-
solved.

2.2.2 High-Pressure Experiment

Owing to pressure limitations in the conventional low-pressure setup, a special holder is


used for supercritical conditions (George et al. 2005) as shown in Figs. 4 and 5. In addition, a
heating tape and temperature controller were attached to the holder for experiments with CO2
in the supercritical state. The experimental procedure was similar to that of the low pressure
experiments except for the elevated pressures of 950 and 1,150 psig. After starting the imbi-
bition process, the system was pressurized by increasing the pressure of the back pressure
regulator to the desired final pressure because of the limited CO2 pressure of 900 psig in
the CO2 gas tank. In addition, during the 1,150 psig experiments the holder was kept at a
constant temperature of 44.4 ◦ C. Due to the water bath in the confining pressure vessel and
the large thermal mass of the holder, the temperature was easily held constant (±1 ◦ C) over
the entire experiment.

123
A Study of Microscale Gas Trapping 655

Fig. 4 High pressure setup with temperature control and confining pressure vessel and confining fluid (blue)

Fig. 5 Left micromodel for high pressure setup with confining pressure chamber and bar window to prevent
cracking; right schematic of confining pressure chamber with polycarbonate window on top (George et al.
2005)

3 Results

3.1 Low Pressure

A set of four low-pressure experiments with imbibition flow rates starting from 0.0001 to
0.1 ml/min was conducted at 11 psig and 22.2 ◦ C. These rates translate to superficial velocities
of 0.8–820 ft/day. All of the experiments ended with significant trapped gas saturations. As
expected, greater NCa lead to lower gas saturations, Fig. 6, for two reasons. The displacement
efficiency is greater at greater velocities and also the dissolution rate is greater because of the
supply of fresh water is greater. As a consequence, the total dissolution of the trapped bubbles
occurred first in the high flow rate experiments and last in the low flow rate experiments.
Gas saturations were determined using an image analysis tool and defining threshold colors
for different phases as shown in Fig. 7 where white is gas and grey is either water or grains.
This analysis involves enhancing certain color values and then transferring the image into a
binary image by defining a threshold value on a color map histogram. Finally, the black and
white pixels are counted and the edges of the pores are subtracted.

123
656 M. Buchgraber et al.

Fig. 6 Trapped saturation for different capillary numbers at 11 psi. a NCa = 5.64 × 10−5 Strap = 4.3 %. b
NCa = 5.64 × 10−6 Strap = 5.6 %. c NCa = 5.64 × 10−7 Strap = 8.2 %. d NCa = 5.64 × 10−8 Strap = 9.9 %

Threshold

Fig. 7 Image analysis with original image on top left corner and modified black and white image on the top
right corner. Bottom Color histogram with threshold value

Previous authors published that different NCa show different behavior in terms of pore-
scale displacement (Cinar et al. 2009; Lenormand and Zarcone 1984). For several structures,
we could identify snap-off, I1 and I2 type and sweep out mechanisms and also record these

123
A Study of Microscale Gas Trapping 657

Fig. 8 Displacement/dissolution combination series at pore level @ 11 psi and NCa = 3.78 × 10−7 from
a 0 s, b 1.5 s, c 2.2 s, d 2.8 s, e 3.2 s, f 6 s, g 8 s, h 9.3 s, i 9.7 s

mechanisms in terms of videos and snapshots. We have not characterized the frequency of
I1 in comparison to I2 imbibition events, but visually I1 events occur more often.
In addition, a new type of displacement pattern that is a combination of dissolution and
displacement was found during our experiments. For the purpose of presentation and discus-
sion, a specific pore structure from the repeating network was chosen. A pore-scale sequence
of dissolution/displacement is shown in Fig. 8. In this case, it was observed how the combina-
tion of dissolution and displacement takes place. Initially the continuous gas phase spans five
pores. As time proceeds, the left end of the bubble is pushed by a piston-like displacement
inside the center bubble. Although the aspect ratio at this pore throat is very high, no snap off
occurs. Instead, the continuous bubble increases its volume by expanding its edges into the
water-filled throat. With proceeding time, the same process is observed at the two bubbles
on the right side. Because this process is under constant pressure, it is concluded that this
displacement pattern is a result of dissolution and viscous forces alone and not compression.
After the gas phase became discontinuous, two periods of dissolution were identified. The
first is observed when the fluid front hits the observation point and displaces or dissolves
most of the CO2 . The second complete dissolution period is when the trapped gas saturation
dissolves completely. During both periods, various dissolution events were observed. First,
homogeneous dissolution of bubbles in the center of pores was detected. In this case, bub-
bles shrank progressively until they disappeared, Fig. 9. Second, some dissolution sequences
resulted in a residual gas nucleus on the bottom of the silicon wafer, Fig. 10. In this case, the

123
658 M. Buchgraber et al.

(a) (b) (c)

(d) (e)
Fig. 9 Homogenous dissolution at NCa = 3.78 × 10−7 —bubble completely dissolves from a 0 s, b 1.5 s,
c 2.6 s, d 3.0 s, e 3.9 s

(a) (b) (c)

(d) (e)
Fig. 10 Heterogeneous dissolution at NCa = 3.78 × 10−7 ; bubble sticks to the roughness of the pore wall
a 0 s, b 1.3 s, c 2.2 s, d 3.0 s, e 3.3 s

123
A Study of Microscale Gas Trapping 659

(a) (b) (c)

(d) (e)
Fig. 11 Heterogeneous dissolution with stabilized bubble at the bottom of the surface at the end at NCa =
3.78 × 10−7 a 0 s, b 2.5 s, c 3.0 s, d 3.3 s, e 3.8 s

dissolving bubble attached to the pore-wall during dissolution. For this event to happen, it is
either required to have particles, cavities, or roughness on the wafer.
Measurements of the wafer bottom and vertical walls have shown that there exists signifi-
cant roughness on the wafer walls to acts as a potential nucleus trap, Fig. 1. It is also likely that
there are particles within or traces of water-borne minerals that attract the gas bubbles. The
silicon-wafer etching process results in significant particle formation as the flow channels
are created. Most particles are removed by cleaning, but traces of solid particles occasionally
remain. Potentially, trace amounts of other solid material are carried into the micromodel
through the flow lines. It is speculated that this scenario, albeit with different materials, might
also happen in saline aquifers where significant clay, grain roughness, and fine material act
as a sink for nuclei. Lastly bubbles that dissolve partially and become meta-stable, or stable,
microbubbles are found floating in the displacement direction, Fig. 11.

3.2 High Pressure

The behavior of CO2 under high pressure either in the liquid state or in the supercritical state
is of great relevance, because all of the sequestration projects, as of date, are at depth greater
than 800 m. Greater depth takes advantage of the larger CO2 density and therefore the larger
mass of CO2 stored per unit volume of liquid or supercritical fluid.
In comparison to low-pressure results, the quality of the captured images and videos
at greater pressures is lower because of the convex bending behavior of the Lexan win-
dow under pressure. Bending leads to a less focused microscopic view. The set of exper-
iments showed important data of the separation of the continuous CO2 phase to the point

123
660 M. Buchgraber et al.

(a) (b) (c) (d)


Fig. 12 Snap off at low capillary dominated flow; at NCa = 3.78 × 10−7 a–d less than 1 s for the entire
sequence

(a) (b) (c) (d)

(e) (f) (g) (h)


Fig. 13 Sweep out at high capillary dominated flow at NCa = 3.78 × 10−5 a–h less than 1 s for entire
sequence

of complete bubble dissolution. The imbibition process for different flow rates coincides
with the behavior recorded and described in the low-pressure section. In Figs. 12 and
13, two different types of displacements are shown as a showcase for low flow rate, low
NCa behavior where the wetting film (i.e., liquid wetting the corners of the pore) swells
in the pore channel and disconnects the continuous gas phase. This process is snap off
(Roof 1970). At high flow rate, 0.1 ml/min, swelling of the wetting film is followed by a
complete sweep of the pore structure and the pore fills completely with water. This pro-
cess is defined as sweep out. During the complete dissolution, heterogeneous as well as
homogeneous dissolution events were observed for all three pressure steps at different flow
rates.
The triggering mode (or action) for gas bubble dissolution after trapping is another impor-
tant factor investigated. The local concentration of CO2 in the aqueous phase appears to
play a significant role as discussed next. In the example in Fig. 14 two equal-sized bub-
bles are shown, however the lower bubble activates first and starts shrinking, whereas the
upper bubble stays constant during the first two images. This effect seems to be related to an
Ostwald ripening effect. If the aqueous phase is saturated with CO2 , the Ostwald effect states
that smaller bubbles shrink and larger bubbles accept CO2 molecules and grow because the
smaller bubbles have greater CO2 pressure that enhances their dissolution. In Fig. 14, the
lower bubble is clearly dissolving and transferring CO2 to the aqueous phase. It is conjectured

123
A Study of Microscale Gas Trapping 661

(a) (b) (c) (d)


Fig. 14 Bubble dissolution activation—Ostwald rippening effect: a 0.0 s, b 2.1 s, c 3.5 s, d 5.0 s

Fig. 15 Dissolution kinetics trend line for different capillary numbers including error bars

that the upper bubble at the same time experiences no (or very small) net change in volume
because the aqueous phase is locally saturated around the bubble. Dissolved CO2 is trans-
ported away by the aqueous phase. Therefore, the volume of the upper bubble remains
constant. Once the lower bubble has dissolved, the upper bubble begins to shrink as shown
in Fig. 14d.
Additional analysis of the dissolution kinetics showed basic trends during these experi-
ments. At the beginning when bubbles are large, or still in cluster form, dissolution follows a
linear trend. As soon as the bubble reaches a threshold size, depending on the capillary num-
ber of the experiments, the linear trend changes slope and bubble dissolution is accelerated
due to smaller surface area and greater dissolution, Fig. 15. At this point, further quantitative
analyses are not possible due to inaccuracies within the measurement and image analysis
technique. Error bars are illustrated on Fig. 15.

3.3 UV dyed water–CO2 displacement

For this experiment the low pressure setup was used and a UV light source instead of the
microscope light source (Fluorescent Ring Light, Cole Parmer) was attached on top of the
micromodel. A UV dye (BlackLightWorld, Cub Run KY) was added to the water at a concen-
tration of 0.5 wt.%. The purpose of this experiment was to visualize the imbibition process
especially the changing saturation conditions in small pores and channels that under normal

123
662 M. Buchgraber et al.

Fig. 16 Imbibition experiment with UV water (green) and CO2 ; small pores getting imbibed first at NCa =
3.78 × 10−7 . Water injection rate= 0.001 ml/min. a 0.0 s, b 45 s, c 210 s, d 500 s

conditions are difficult to observe. As expected, the CO2 occupies the larger pores. During
the imbibition stage, the water reoccupies smaller pores and channels first. As soon as the
front reaches the observation point, dissolution and imbibition occurs simultaneously in the
larger pores, Fig. 16.

3.4 Glycerin/water–Heptane Imbibition Experiment

All the previous experiments had significant interphase mass transfer. The water and the
CO2 phases interact with each other and exchange molecules; however, the exchange from
the water to the CO2 phase is of minor importance in our case. The last experiment is
an immiscible imbibition run at various capillary numbers and low-pressure conditions.
The fluid pair is a 78 wt.% glycerin and water mixture as the displaced fluid and heptane
as the displacing fluid. Imbibition of the glycerin/water mixture follows. The phases are
pre-equilibrated to satisfy any interphase solubility. The fluid pair represents viscosity ratios

123
A Study of Microscale Gas Trapping 663

similar to supercritical CO2 and water. At room temperature, the viscosity ratio of the fluids
was 13.
To obtain better contrast between the fluid pair, and enable better processing of the images
and therefore more accurate saturation calculations, the glycerin water mixture was dyed
green (Brilliant Green, Hartman-Leddon Co.). The experiment was conducted at room tem-
perature and with a 10 psi back pressure valve at the outlet of the micromodel. To observe
the changes in trapped saturation for different flow, 25 pictures were taken on a regular grid
throughout the micromodel matrix area. For both miscible and immiscible experiments, the
25 microsaturation pictures cover approximately 15 % of the entire the 5 × 5 cm2 area.
Pictures were taken at the end of each flooding period when steady-state conditions were
reached. This corresponded to about ten pore volumes of liquid injection. Phase saturations
were calculated using image analysis software. The color images were thresholded and then
converted to black and white. The number of black and white pixels in the image matrix was
then counted. A reference image at fully saturated conditions provides the number of pixels
associated with micromodel “grains” and these pixels are subtracted. Image analysis details
are similar to previous work in similar micromodels (Buchgraber et al. 2011, 2012).
The starting flow rate was chosen to be 0.0001 ml/min that corresponds to a capillary
number of 1.6 ×10−8 . Flow rate was increased in six equal-sized steps to a final flow rate
of 0.05 ml/min or a capillary number of 7.96×10−6 . The maximum flow rate was limited
by the micromodel material properties from the high pressure drop of the viscous displacing
fluid. Results are shown in Figs. 17 and 18 and the figure captions list the saturation of the
trapped phase for each capillary number. The trapped saturation declines by roughly 30 %
due to increasing NCa .
Now compare the trapped saturations from the low-pressure experiment with CO2 dis-
solution and the immiscible experiment. We find that dissolution has significant impact on
trapped saturation and storage capacity. That is, the greater the fraction of dissolved CO2 ,
the lower the pressure build and therefore the greater CO2 injectivity when the maximum
pressure is the limiting factor.

4 Discussion

For CO2 sequestration, it is essential to understand and to quantify trapping mechanisms.


Capacity estimation is one of the determining factors in site selection. Dissolution events
therefore are one of the key factors to injectivity, storage security, and capacity estimation.
This paper showed the differences in dissolution events and their possible consequences.
The theory of bubble nucleation is used in a reverse mode to describe and characterize the
dissolution events in these experiments. The onset of nucleation (Volmer and Weber 1926;
Farkas 1927) is well understood and in porous media is explained by either heterogeneous (Li
or Yortsos (1995)) or homogenous (Firoozabadi and Kashchiev 1996) nucleation theories.
Nucleation in porous media appears to depend strongly on pore roughness (Kamath and
Boyer 1995), solid substrate present in the liquid bulk, and gas embryo size.
The observations made in these dissolution experiments are comparable to heterogeneous
and homogeneous nucleation in reverse mode. The silicon wafer wall roughness acted as
pore roughness for the nucleation of gas bubbles whereas the formation of microbubbles
and complete dissolution in the bulk liquid showed homogenous dissolution behavior. Pore
roughness and rock texture therefore significantly contribute to storage security by trapping
bubbles in crevices within grain roughness and providing sites for dissolution. Hence, het-
erogeneous nucleation events may lead to more secure storage. Microbubble formation and

123
664 M. Buchgraber et al.

Fig. 17 Trapped immiscible saturation at different capillary numbers: a NCa = 2.71 × 10−8 Strap = 84 %.
b NCa = 1.35 × 10−7 Strap = 83 %. c NCa = 2.71 × 10−7 Strap = 80 %. d NCa = 1.35 × 10−6 Strap = 74 %.
e NCa = 2.71 × 10−6 Strap = 65 %. f NCa = 1.35 × 10−5 Strap = 58 %

homogenous dissolution potentially increase risk when considering subsurface flows that
transport the saturated brine to lower pressure region where supersaturation levels are lower
and the bulk brine or microbubbles bubbles are activated and form a moving gas phase again.
Therefore, the selection and screening of rock type and solid concentration present in brine
is of major importance.

5 Summary

This paper described the basics of pore-level displacement and dissolution of carbon dioxide
as relevant to carbon sequestration. The key findings are summarized as follows.

• The displacement/dissolution of a continuous gas/supercritical phase cluster happened


in three consecutive steps: (1) disconnection/displacement of the continuous gas/super-
critical phase by snap off, I1- and I2-type displacement, or sweep out (2) shrinkage of

123
A Study of Microscale Gas Trapping 665

Fig. 18 Trapped saturation versus capillary number including error bars

individual gas/supercritical phase bubbles, and (3) complete dissolution in the form of
heterogeneous or homogenous dissolution.
• Different capillary numbers during imbibition triggered different pore-level displacement
mechanisms such as sweep out and snap off.
• During imbibition, water fills up smaller pore spaces by corner and thin wetting film flow
followed by filling of larger pore spaces.
• Both a heterogeneous and homogenous dissolution process were observed. For the mi-
cromodel, inspection of video and still images indicated that each type of dissolution
occurred in roughly similar quantity.

References

Abrams, A.: The influence of fluid viscosity, interfacial tension and flow velocity on residual oil saturation
left by waterflood. Soc. Pet. Eng. J. Oct. 15(5), 437–447 (1975)
Agarwal, R.G.: Unsteady-state performance of water-drive gas reservoirs, PhD thesis, Texas A&M University,
College Station (1967)
Al Mansoori, S.K., Iglauer, S., Pentland, C.H., Bijeljic, B., Blunt, M.J.: Measurements of Non-wetting phase
trapping applied to carbon dioxide storage. Energy Procedia. 1(1), 3173–3180. doi:10.1016/j.egypro.
2009.02.100 (2008)
Alty, T.: The Origin of the Electrical Charge on Small Particles in Water. Proc. R. Soc. A. 112, 235–251 (1926)
Arora, P., Kovscek, A.R.: A mechanistic modeling and experimental study of solution gas drive. Transp. Porous
Media 51, 237–265 (2003)
Bachu, S.: Screening and ranking of sedimentary basins for sequestration of CO2 in geological media in
response to climate change. Environ. Geol. 44, 277–289 (2003)
Bauget, F., Lenormand, R.: Mechanisms of bubble formation by pressure decline in porous media. In: SPE
77457 presented at the SPE Annual Technical Conference and Exhibition, 29 Sept–2 Oct, San Antonio
(2002)
Bennion, B., Bachu, S.: Relative permeability characteristics for supercritical CO2 displacing water in a variety
of potential sequestration zones. In: Society of Petroleum Engineers SPE 95547, San Diego (2005)

123
666 M. Buchgraber et al.

Bennion, D.B., Bachu, S.: Dependence on temperature, pressure, and salinity of the IFT and relative perme-
ability displacement characteristics of CO2 injected in deep saline aquifers, SPE 102138. In: Society of
Petroleum Engineers SPE 102138, San Antonio (2006)
Bisperink, C.G.J., Prins, A.: Bubble growth in carbonated liquids. Colloids Surf. A 85, 237–253 (1994)
Bora, R., Maini, B.B., Chakma, A.: Flow visualization studies of solution gas drive process in heavy oil
reservoirs with a glass micromodel. Soc. Pet. Eng., Res. Eval. Eng. 3(3), 224–229 (2000)
Buchgraber, M., Clemens, T., Castanier, L.M., Kovscek, A.R.: A microvisual study of the displacement of
viscous oil by polymer solutions. SPE Res. Eval. Eng. 14(3), 269–280. SPE-122400-PA. (2011). doi:10.
2118/122400-PA
Buchgraber, M., Al-Dossary, M., Ross, C.M., Kovscek, A.R.: Creation of a dual-porosity micromodel for
pore-level visualization of multiphase flow. J. Petroleum Sci. Eng. (86–87) 27–38. (2012). doi:10.1016/
j.petrol.2012.03.012
Buckley, J.S.: Multiphase displacements in micromodels. In: Interfacial Phenomena in Oil Recovery, Marcel
Dekker Inc., New York, pp. 157–189 (1990)
Chalbaud, C., Robin, M., Lombard, J.-M., Bertin, H., Egermann, P.: Brine/CO2 , interfacial properties and
effects on CO2 storage in deep saline aquifers. Oil Gas Sci. Technol.: Rev. IFP. 65(4), 541–555 (2010)
Chierici, G.L., Ciucci, G.M., Long, G.: Experimental research on gas saturation behind the water front in gas
reservoirs subjected to water drive. In: Proceedings of the World Petroleum Conference, Frankfurt am
Main, Section II, paper 17, PD6, pp. 483–498 (1963)
Cinar, Y., Riaz, A., Tchelepi, H.A.: Experimental study of CO2 injection into saline formations. Soc. Pet. Eng.
J. 14, 588–594 (2009)
Crowell, D.C., Dean, G.W., Loomis, A.G.: Efficiency of gas displacement from a water-drive reservoir. Bureau
of mines. Report of Investigations, 6735, United States Bureau of the Interior, Washington, DC (1966)
Dodd, C.G., Kiel, O.G.: Evaluation of Monte-Carlo methods in studying fluid-fluid displacement and wetta-
bility in porous rocks. J. Phys. Chem. 63(10), 1646–1652 (1959)
Farkas: The velocity of nucleus formation in supersaturated vapors. Z. Physik. Chem. 125, 236 (1927)
Firoozabadi, A.: Mechanisms of solution gas drive in heavy oil reservoirs. J. Can. Pet. Technol. 40(3),
15–20 (2001)
Firoozabadi, A., Kashchiev, D.: Pressure and volume evolution during gas phase formation in solution gas
drive processes. Soc. Pet. Eng. J. 1(3), 219–227 (1996)
Firoozabadi, A., Olsen, G., van Golf-Racht, T.: Residual gas saturation in water-drive gas reservoirs. in: SPE
16355 proceedings of the SPE California Regional Meeting, Ventura, 8–10 Apr 1987
Fox, F.E., Herzfeld, K.: Gas bubbles with organic skin as cavitation nuclei. J. Acoust. Soc. Am. 26,
984–989 (1954)
Garcia, J.E. (2003) Fluid dynamics of carbon dioxide disposal into saline aquifers. PhD thesis, University of
California, Berkeley
Gauglitz, P.A., St. Laurent, C.M., Radke, C.J.: An experimental investigation of gas-bubble breakup in con-
stricted square capillaries. J. Petr. Tech. 39, 1137–1146 (1987)
George, D.S., Hayat, O., Kovscek, A.R.: A microvisual study of solution gas drive mechanisms in viscous
oils. J. Pet. Sci. Eng. 46(1–2), 101–119 (2005). doi:10.1016/j.petrol.2004.08.003
Hemmingsen, E.E.: Cavitation in gas-supersaturated solutions. J. Appl. Phys. 46, 213–218 (1975)
Hesse, M.: Mathematical modeling and multiscale simulation of CO2 storage in saline aquifers. PhD thesis,
Stanford University. http://pangea.stanford.edu/departments/ere/faculty-research/ms-and-phd-theses-
search (2008)
Hilfer, R., Øren, P.E.: Dimensional analysis of pore scale and field scale immiscible displacement. Trans.
Porous Media 22, 53–72 (1996)
Hornbrook, J.W., Castanier, L.M., Petit, P.A.: Observation of foam/oil interactions in a new, high-resolution
micromodel. Paper SPE 22631 presented at SPE Annual Technical Conference and Exhibition held in
Dallas, 6–9 Oct 1991
Iglauer, S., Wülling, W., Pentland, C.H., Mansoori1, S., Blunt, M.J.: Capillary trapping capacity of rocks and
sandpacks. In: SPE 120960, presented at the EUROPEC/EAGE in Amsterdam, 8–11 June 2009
Iglauer, S., Paluszny, A., Pentland, C.H. et al.: Residual CO2 imaged with X-ray micro-tomography, Geophys.
Res. Lett. (2011) 38, L21403. ISSN: 0094-8276 (2011)
Intergovernmental Panel on Climate Change: Climate Change 2007—The Physical Science Basis: Contribu-
tion of Working Group I to the Fourth Assessment Report of the IPCC. Cambridge University Press,
Cambridge. ISBN 978 0521 88009-1 (2007a)
Jerauld, G.R.: General three-phase relative permeability model for Prudhoe Bay. SPE Reserv. Eng. 12(4), 255–
263 (1997)
Kamath, J., Boyer, R.E.: Critical gas saturation and supersaturation in low permeability rocks. Soc. Pet. Eng.
Form. Eval. 10(4), 247–253 (1995)

123
A Study of Microscale Gas Trapping 667

Kantzas, A., Ding, M., Lee, J.: Residual gas saturation revisited. SPE. Reserv. Eval. Eng. 4(6), 467–476
(2001)
Kovscek, A.R., Radke, C.J.: Gas-bubble snap-off under pressure driven flow in constricted noncircular capil-
laries. Colloids Surf. A 117, 55–76 (1996)
Kovscek, A.R., Radke, C.J.: Pressure-driven capillary snap off of gas bubbles at low wetting-liquid con-
tent. Colloids Surf. A 212(2&3), 99–108 (2003)
Krevor, S., Pini, R., Li, B., Benson, S.: Capillary heterogeneity trapping of CO2 in a sandstone rock at reservoir
conditions. Geophys. Res. Lett. 38, L15401 (2011). doi:10.1029/2011GL048239
Land, C.S.: Calculation of imbibition relative permeability for two- and three-phase flow from rock properties.
Soc. Pet. Eng. J. 8(2), 149–156. Petrol. Trans. AIME, 243 (1968a)
Land, C.S.: The optimum gas saturation for maximum oil recovery from displacement by water. In: SPE 2216
Proceedings of the 43rd Annual Fall Meeting of SPE of AIME, Houston, 29-Sept–Oct-2 1968b
Larson, R.G., Davis, H., Scriven, L.: Displacement of residual non wetting fluid from porous media. Chem.
Eng. Sci. 36, 75–85 (1981)
Lawrence, A., Crum, L.A.: Nucleation and stabilization of microbubbles in liquids. Appl. Sci. Res. 38(1),
101–115 (1982)
Lenormand, R.: Liquids in porous media. J. Phys. Condens. Matter 2, SA79 (1990)
Lenormand, R., Zarcone, C.: Role of roughness and edges during imbibition in square capillaries. In: SPE
Annual Technical Conference and Exhibition, Houston 16–19 Sept 1984
Lenormand, R., Zarcone, C., Sarr, A.: Mechanisms of the displacement of one fluid by another in a network
of capillary ducts. J. Fluid Mech. 135, 337–353 (1983)
Li, X., Yortsos, Y.C.: Theory of multiple bubble growth in porous media by solute diffusion. Chem. Eng.
Sci. 50(8), 1247–1271 (1995)
Morrow, N.R., Songkran, B.: Effects of Viscous and Buoyancy Forces on Non-wetting Phase Trapping in
Porous Media, Surface Phenomena in Enhanced Oil Recovery. Plenum Press, New York (1981)
Nghiem, L., Yang, C.D., Shrivastava, V., Kohse, B., Hassam, M., Card, C.: Risk mitigation through the
optimization of residual gas and solubility trapping for CO2 storage in saline aquifers. Greenh. Gas
Control Technol. 1(1), 3015–3022 (2009)
Pentland, C., Al-Mansoori, S., Iglauer, S.,Bijeljic, B., Blunt, M.J.: Measurement of non-wetting phase trap-
ping in sand packs. In: SPE 115697 presented at the SPE Annual Technical Conference and Exhibition,
Colorado, 21–24 Sept 2008
Pentland, C., Tanino, Y., Iglauer, S., Blunt, M.J.: Capillary trapping in water-wet sandstones: coreflooding
experiments and pore-network modeling. In: SPE 133798 presented at the SPE Annual Technical Con-
ference and Exhibition, Florence, 19–22 Sept 2010
Pentland, C.H., El-Maghraby, R., Iglauer, S. et al.: Measurements of the capillary trapping of super-criti-
cal carbon dioxide in Berea sandstone. Geophys. Res. Lett. 38, ISSN:0094-8276 (2011). doi:10.1029/
2010GL045662
Perrin, J.-C., Krause, M., Kuo, C.-W., Miljkovic, L., Charob, E., Benson, S.M.: Corescale experimental study of
relative permeability properties of CO2 and brine in reservoir rocks. Energy Procedia 1, 3515–3522 (2009)
Pham, H.T.M., de Boer Charles R., Sarro P.M.: Roughness treatment of silicon surface after deep reactive ion
etching, pp. 535–538. In: Proceedings of SAFE Prorisc 2004, Veldhoven (2007)
Pini, R., Krevor, S., Benson, S.: Capillary pressure and heterogeneity for the CO2 /water system in sandstone
rocks at reservoir conditions. Adv. Water Resour. 38, 48–59 (2012)
Pruess, K., Zhang, K.: Numerical modeling studies of the dissolution-diffusion-convection process during
CO2 storage in saline aquifers. Lawrence Berkeley National Laboratory University of California, Berke-
ley (2008)
Rangel-German, E.R., Kovscek, A.R.: A micromodel investigation of two-phase matrix-fracture transfer mech-
anisms. Wat. Res. Res. 42, W03401 (2006)
Ransohoff, T.C., Gauglitz, P.A., Radke, C.J.: Snap-off of gas bubbles in smoothly constricted noncircular
capillaries. AlChE J. 33, 753–765 (1987)
Roof, J.G.: Snap-off of oil droplets in water-wet pores. Soc. Pet. Eng. J. 10, 85–90 (1970)
Sirotyuk, M.: Stabilization of gas bubbles in water. Sov. Phys. Acoust. 16, 237–240 (1970)
Spiteri, E.J., Juanes, R., Blunt, M.J., Orr, F.M.: A new model of trapping and relative hysteresis for all wetta-
bility characteristics. Soc. Pet. Eng. J. 13, 277–288 (2008)
Upadhyaya, A.: Visualization of four-phase flow using micromodels. 2001 MS Report Stanford University.
http://pangea.stanford.edu/departments/ere/faculty-research/ms-and-phd-theses-search (2001)
Volmer, M., Weber, A.: Nucleus formation in supersaturated systems. Z. Phys. Chem. 119, 277 (1926)
Wan, J., Tokunaga, T.K., Tsang, C.-F., Bodvarsson, G.S.: Improved glass micromodel methods for studies of
flow and transport in fractured porous media. Water Resour. Res. 32(7), 1955–1964 (1996)

123
668 M. Buchgraber et al.

Wardlaw, N.: The effects of pore structure on displacement efficiency in reservoir rocks and in glass
micromodels. In: SPE-8843 presented at the 1st symposium on enhanced oil recovery, Tulsa, 20–23
Apr 1980
Yang, D., Tontiwachwuthikul, P., Gu, Y.: Interfacial interactions between reservoir brine and CO2 at high
pressures and elevated temperatures. Energy Fuels 19, 216–223 (2005)
Ziqiu, X., Saeko, M., Keigo, K.: Case study: trapping mechanisms at the pilot-scale CO2 injection site,
Nagaoka, Kyoto University. Energy Proc I 1, 2057–2062 (2009)

123

You might also like