Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Edith Cowan University

Research Online

Research outputs 2014 to 2021

2021

CO2 – brine – sandstone wettability evaluation at reservoir


conditions via nuclear magnetic resonance measurements
Auby Baban
Edith Cowan University

Ahmed Al-Yaseri
Edith Cowan University

Alireza Keshavarz
Edith Cowan University

R. Amin

Stefan Iglauer
Edith Cowan University

Follow this and additional works at: https://ro.ecu.edu.au/ecuworkspost2013

Part of the Civil and Environmental Engineering Commons

10.1016/j.ijggc.2021.103435 Baban, A., Al-Yaseri, A., Keshavarz, A., Amin, R., & Iglauer, S. (2021).
CO2–brine–sandstone wettability evaluation at reservoir conditions via nuclear magnetic resonance
measurements. International Journal of Greenhouse Gas Control, 111, article 103435. https://doi.org/10.1016/
j.ijggc.2021.103435
This Journal Article is posted at Research Online.
https://ro.ecu.edu.au/ecuworkspost2013/10986
International Journal of Greenhouse Gas Control 111 (2021) 103435

Contents lists available at ScienceDirect

International Journal of Greenhouse Gas Control


journal homepage: www.elsevier.com/locate/ijggc

CO2 – brine – sandstone wettability evaluation at reservoir conditions via


Nuclear Magnetic Resonance measurements
A. Baban a, *, A. Al-Yaseri a, *, A. Keshavarz a, R. Amin b, S. Iglauer a
a
Petroleum Engineering, School of Engineering Department, Edith Cowan University, 270 Joondalup Drive, Joondalup WA 6027, Australia
b
Zana Enterprise, 2 ST. Georges Terrace, Perth WA 6000, Australia

A R T I C L E I N F O A B S T R A C T

Index terms: CO2-rock wettability is a key parameter which governs CO2 trapping capacities and containment security in the
5114, Permeability and porosity context of CO2 geo-sequestration schemes. However, significant uncertainties still exist in terms of predicting
3947, Surfaces and interfaces CO2 rock wettability at true reservoir conditions. This study thus reports on wettability measurements via in­
5139,Transport properties
dependent Nuclear Magnetic Resonance (NMR) experiments on sandstone (CO2–brine systems) to quantify
Keywords: Wettability Indices (WI) using the United States Bureau of Mines (USBM) scale. The results show that CO2 (either
Wettability
molecularly dissolved or as a separate supercritical phase) significantly reduced the hydrophilicity of the
Nuclear Magnetic Resonance
Carbone dioxide (CO2)
sandstone from strongly water-wet (WI ≈ 1) to weakly water-wet (WI = 0.26), and associated with that the
geo-sequestration water-wetness of the rock for the two-phase systems. This was caused by additional protonation of surface silanol
Carbon Capture and Storage (CCS) groups on the quartz, induced by carbonic acid. Capillary pressure and relative permeability curves and residual
Sandstone CO2 saturation were also measured; these results were compared with literature data, and general consistency
was found. NMR T2 distribution measurements also demonstrated preferential water displacement in large pores
(r > 1 µm) following scCO2 flooding, while no change was observed for smaller pores (r < 1 µm). These insights
add confidence to the assessments of CO2-rock wettability and therefore reduce project risk. This work thus aids
in the implementation of large-scale CO2 sequestration.

security (e.g. Abramov et al., 2019; Al-Khdheeawi et al., 2018;


1. Introduction Al-Menhali et al., 2015; Broseta et al., 2012; Garing and Benson, 2019;
Iglauer et al., 2015a; Krevor et al., 2012; Rahman et al., 2016, Ali et al.,
Carbon geo-sequestration (CGS) has been recognized as a strategic 2019). Therefore, it is of vital importance to accurately measure rock
technology to minimize the impact of carbon dioxide (CO2) emissions CO2 wettability. However, conventional methods, such as Amott-Harvey
arising from anthropogenic activities, thus mitigating global warming (AH), US Bureau of mines (USBM), contact angle and tomographic im­
(IPCC, 2005). Injecting CO2, at high pressure and temperature into aging measurements have serious limitations. amott–harvey (AH) and
geological storage formations, where it can be permanently immobi­ US Bureau of mines (USBM) are expensive, time-consuming and prone to
lized, has been identified as the most effective way to capture and error (so far they have not succeeded for scCO2-brine-rock systems)
securely store carbon for millions of years into the future (IPCC, 2005; (Iglauer et al., 2015b). Similarly, 2D contact angle measurements
Bachu and Adams, 2003; Orr, 2009). Once injected, the supercritical overlook 3D pore geometry, surface roughness and chemical heteroge­
carbon dioxide (scCO2) is more buoyant than local fluids present in the neity (at least in a classic contact angle experiment) (Al-Yaseri et al.,
pore space (CO2 has a lower density than local brine); thus, it percolates 2016a;; Valori and Nicot, 2019; Espinoza and Santamarina, 2010), while
upwards through the porous rock until it reaches an impermeable medical CT resolution is too low (only averaged saturations can be
caprock layer, where it can be trapped by capillary forces (Iglauer, 2018; measured), and microCT curvature measurements may be biased due to
Le Guen et al., 2007). Furthermore, CO2 can be trapped in the pore their voxelized nature (Iglauer et al., 2019).
network of the storage rock (“residual trapping”), again by capillary NMR is fast, robust and can measure valuable structural insights on
forces, where it dissolves in formation water and sinks deep into the fluid molecules in confined environment. Transverse relaxation time
reservoir (Riaz et al., 2006). For all these storage mechanisms, CO2-rock (T2) can be used to determine petrophysical properties (porosity,
wettability is a key parameter that influences storage capacity and permeability, water saturation, wettability) and measures pore size

* Corresponding author.
E-mail addresses: a.baban@ecu.edu.au (A. Baban), a.yaseri@ecu.edu.au (A. Al-Yaseri).

https://doi.org/10.1016/j.ijggc.2021.103435
Received 22 December 2020; Received in revised form 22 May 2021; Accepted 12 August 2021
Available online 3 September 2021
1750-5836/© 2021 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
A. Baban et al. International Journal of Greenhouse Gas Control 111 (2021) 103435

the core fluid from the confining fluid. This setup was housed in a cor­
Nomenclature list eflood apparatus (see Figure 1) with an integrated NMR (GeoSpec+
2.4MHz) instrument that consisted of four high precision syringe pumps
WI Wettability index (Teledyne Isco pumps, model 260D; pressure accuracy of 0.1 % full
Nc Capillary number scale). The first two pumps provided and maintained a confining pres­
Krnw Relative permeability of non-wetting phase sure of 13 MPa around the sample. Fluorine based oil (Fluorinert™ FC-
Krw Relative permeability of wetting phase 40) was used as the confining fluid (this is a non-hydrogenated fluid that
Sw Water saturation does not interfere with the NMR signal and has superior heat trans­
Pe Capillary entry pressure mission characteristics). The third pump injected fluids into the core,
Pc Capillary pressure while the fourth production pump applied backpressure and received
Sw* Normalised water saturation the produced fluids. A mixing reactor was used to equilibrate CO2 and
λ Pore size distribution index brine, yielding live brine (i.e. fully CO2-saturated brine at high pressure
T1 Longitudinal relaxation time and elevated temperature) as described by El-Maghraby et al. (2012).
T2 Transverse relaxation time The temperature of the pumps was controlled by a Jabala oil bath
r radius of the pores with heating tape wrapped around delivering and confining tubes to
ρ Surface relaxivity control the core temperature throughout the experiment (i.e. to
isothermal conditions of 333 K). Two high-accuracy pressure sensors
(Keller 33X, accuracy of ±1500 Pa) were attached to both ends of the
core to measure absolute pressure and pressure drop across the core
distribution (PSD) (Zhao et al., 2020; Lyu et al., 2018). T2 is very sen­ plug.
sitive to fluid-surface interactions and very responsive to paramagnetic During CO2 sequestration processes, initially a fraction of injected
ions from rock matrix minerals (Mitchell et al., 2013; Connolly et al., CO2 dissolves in the local brine and creates a live brine (brine saturated
2019; Korb, 2009; Washburn, 2017). Here we therefore used an inde­ with CO2) (Iglauer 2011). To simulate reservoir conditions, the core
pendent technique (NMR), which has a molecular resolution, to over­ sample inside the NMR coreflood apparatus was vacuumed for 48 h
come some of the abovementioned limitations, and to add to the before injecting 10 pore volumes (PV) of dead brine at a preset flowrate
fundamental understanding of flow processes of scCO2 through brine value of 1 ml/min, with 8 MPa backpressure. The saturated core was left
saturated porous rock. This work will thus aid in the implementation of for 24 h to ensure full brine saturation was achieved (effective stress was
large scale CO2 geo-sequestration and increase the certainity of constant throughout the experiment at 8 MPa). T1-T2 and T2 measure­
geo-sequestration schemes and carbon capture and storage projects. ments (note that T1 is the longitudinal relaxation time and T2 is the
transverse relaxation time) were performed and then 10 PV of live brine
2. Experimental Methodology were injected into the core at the same conditions as above to displace
all dead brine with live brine. Once the pressure drop stabilized, T1-T2
All experiments were performed on a cylindrical homogenous San and T2 responses for live-brine saturated core were obtained. Prior to
Saba sandstone sample (5 cm in length and 3.5 cm in diameter). The scCO2 injection, the outlet pressure was maintained constant at 8 MPa
sample consisted of 94 wt% α-quartz, 2 wt% Plagioclase and 4 wt% of (using an ISCO pump as a back-pressure regulator). Subsequently,
Alkali Feldspar; measured via a Bruker-AXS D8 Advance Diffractometer. capillary drainage pressures were measured by injecting 20 PV of scCO2
The sample had a porosity of 21 % and a permeability of 35 mD into the core plug at a constant flowrate of 0.1 ml/min, which was
(measured by a CoreLab UltraPoroPerm-910). The core plug was incrementally increased to 0.5, 1, 2, 4, 6 and 8 ml/min (Ramakrishnan
wrapped in polytetrafluoroethylene (PTFE) tape and sealed with double and Cappiello, 1991) with multiple T1-T2 and T2 scans taken for each
PTFE heat-shrink tubes, and a rubber sleeve (ends were open for the flow rate. Note that the capillary number (NC=Qμ/γ) varied between
injection and desaturation of fluids) was applied to adequately separate 6.23 × 10− 7 and 3.74 × 10− 5 consequently capillary forces dominated

Figure 1. Schematic diagram of the experimental apparatus

2
A. Baban et al. International Journal of Greenhouse Gas Control 111 (2021) 103435

(Dullien et al., 1989; Garing and Benson, 2019). (Dunn et al., 2002; Freedman et al., 2003; Connolly et al., 2019), while
Once scCO2 injection started, water was displaced from the core until diffusion is induced by the presence of magnetic field gradient caused by
steady-state was reached, i.e. no more water production was observed at variation in magnetic susceptibility between the fluid and grains
steady-state and the pressure drop across the core reached a constant composing the rock matrix (Brownstein and Tarr, 1979). However, for
value. Further scCO2 flow maintained the pressure gradient within the short echo times (TE) and low magnetic field strengths, as used in this
core and at each point in the core the pressure difference between CO2 study, the internal gradients are relatively weak, their effects are
and water equalled the capillary pressure (ΔP = PCO2 at inlet – Pwater at negligible (Dunn et al., 2002; Connolly et al., 2019; Korb, 2009).
outlet) (Ramakrishnan and Cappiello, 1991; Pini et al. 2013; Andersen Furthermore, the bulk fluid (water) is diamagnetic and is mainly
et al., 2017). Furthermore, the capillary pressure measurements were controlled by spin – spin interactions between water molecules (Dunn
made at the inlet of the core sample by successively increasing the CO2 et al., 2002; Connolly et al., 2019). Thus, the surface transverse relax­
flow rate, while average fluid saturations were measured via NMR T2. To ation is more dominate and T2 can be written as function of the ratio of
account for the capillary end-effect (Gupta and Maloney; Virnovsky the pore’s specific surface area to the total pore volume.
et al., 1998; Andersen et al., 2017), and to accurately determine the Pc ( )
1 S
and Krg, the capillary end-effect analytical correction for saturation and = ρ2 (6)
T2 V pore
relative permeability were performed as detailed in Supplementary 1
(Virnovsky et al., 1998). The experiment was conducted twice, and A simple limitation of using NMR to measure pore size distribution
reproducibility of the experiments and results was good. (PSD) is unavailability of a fixed value of surface relaxivity (Almen­
Finally, 10 PV of live brine were reinjected (forced imbibition) at the ningen et al., 2020). Furthermore, the pore network in a porous medium
same flow rates with capillary numbers between 4.17 × 10− 6 and 6.6 × conceptually consist of pore-throats (cylindrical) and pore-bodies
10− 5 until residual CO2 saturation (SCO2,r) was achieved. (spherical) (Dullien, 1992) and no geologic porous medium will be
In addition, Darcy’s law can be generalized as a simplified rela­ perfectly symmetric or have spherical pores. This is typically dealt with
tionship between flux and pressure gradient with relative permeability a shape factor (c) and c is typically 2 for cylindrical (Mesquita et al.,
described in terms of local saturation. Hence, the relative permeability 2016) and 3 for spherical pores, (Looyestijn, 2008; Djebbar and
of CO2 (Kr,CO2) as a function of water saturation (Sw) was measured, Donaldson, 2015).
using below equation (with CO2 = i) (Dullien, 1992). The measured T2 surface relaxation time is used to determine the
ui L μi distribution of water in the different pores of the rock (Coates et al.,
kri = (1) 1999; Akitt and Mann, 2000; Li et al., 2019).
ΔPi K
Where ui is the volumetric CO2 fow rate, K is the absolute core r = cρ2 T2 (7)
permeability, µi is the CO2 viscosity, L is the core length and ΔPi is the r is the radius of the pores (µm), c is the shape factor and sphereical
pressure drop at the inlet of the core. Water saturations (Sw) were pores assumed (c = 3) (Li et al., 2019), and ρ2 is relaxivity ≈ 18.3 µm.s −
measured via NMR, and corrected for the capillary end effect (see 1
for homogenous sandstone (Lucas-Oliveira et al., 2020; Washburn
Supplementary 1). As outlined above, the viscous pressure drop across et al., 2017; Li et al., 2019).
the sample was taken as the capillary pressure (Pc) at the inlet face of the The T1/T2 ratio is a quantitative measurement of the strength of the
core, Ramakrishnan and Capiello 1991). interaction between the fluid molecules in the pore space and the solid
Equation (2) is a general representation of capillary pressure (Pc) as a surface (Katika et al., 2017; Guan et al., 2002). Thus, fast moving mol­
function of saturation (Dullien, 1992). ecules and isotropic, i.e. nonviscous fluid, have a very short correlation
( )− 1 times (τc) at Larmor frequency and T1 and T2 relaxation times should in
Pc = Pe S∗w λ
(2)
principle be similar or equal (T1/T2 = 1). However, in complex mole­
Pe is the minimum pressure required for the non-wetting phase to cules or for slow motion systems such as anisotropic (very viscous fluid
enter the pores of the sample, i.e. the capillary entry pressure, and λ is a such as asphaltene or solid surfaces) T1 and T2 are affected differently
fitting parameter (known as pore size distribution index) that tunes the and hence T1/T2 differs from unity (Valori et al., 2017; Zielinski et al.,
curvature of the Pc – Sw curves. The value of λ corresponds to the degree 2010; Lessenger et al., 2015). This is primarily due to affinity of the
of homogeneity of the pores size distribution in the sample; uniform wetting fluid to the surface of the minerals which significantly shorten
pore-size reservoir rocks have large values of λ, whereas broad pore size T2 relaxation time (Korb et al., 2009; Katika et al., 2017; Wang et al.,
distribution yield lower values of λ; Laliberte (1969), S∗w is the normal­ 2018).
ized wetting phase saturation (Dullien, 1992). The acquired T1/T2 ratios were selected at the point of highest in­
tensity in the T1 - T2 spectra (which correspond to hydrogen intensity
Sw − Swr
S∗w = . (3) distribution of the pore fluid), are correlated to the USBM scale as an
1 − Swr independent wettability indicator [− 1 strongly oil-wet, 0 intermediate
Where Swr is the residual saturation of the wetting phase. The and +1 strongly water-wet], via Eq. (8) which was deriven from Valorie
Brooks-Corey model is usually used to predict relative permeability of et al. 2013 data (Valori et al., 2013);
the wetting and non-wetting ina multiphase flow systems (Dullien,
T1/T2 = − 2.25 WI + 2.25 (8)
1992). The non-wetting (Krnw) relative permeability was thus predicted
via Eqs. 5, (Corey, 1954). The T1/T2 value is obtained from Fig. 2 and wettability index (WI)
( )2+3λ can readily be determined, (If 0 < WI < 1, the rock is water-wet (WI ≈ 1
Krw = S∗w λ (4) strongly water-wet), conversely if − 1 < WI < 0 the rock oil-wet (WI ≈
( ] − 1 strongly oil-wet) (Looyestijn, 2008; Valori et al., 2017). A practical
[ ( )2+λ limitation of Eq. (8) arises when measuring mixed-wet systems as NMR
Krnw = 1 − S∗w )2 1 − S∗w λ
(5)
uses an inhomogeneous magnetic field which cannot differentiate be­
tween water and oil (resonance frequency); the NMR response is
The transverse relaxation time (T2) of a spin-echo, in a confined
therefore a combination and an overlap of the responses of all 1H con­
environment, such as pore spaces, fluids exhibit three independent
taining fluids (oil and water) in the rock (Looyestijn, 2008; Valori et al.,
relaxation mechanisms; diffusion, bulk, and surface. The bulk fluids
2017).
(water, brine, light oil) have long transverse relaxation times and are
mainly controlled by spin – spin interactions between fluid molecules

3
A. Baban et al. International Journal of Greenhouse Gas Control 111 (2021) 103435

3. Results and Discussion

3.1. NMR wettability indices

Initially, the sandstone sample was strongly water-wet a T1/T2 ratio


equal to 1 was measured, (see Table 1 and Figure 2a). However, the T1/
T2 ratio increased significantly to 1.7 following live brine saturation,
and remained so for the rest of the experiment (i.e. after primary
drainage with CO2 and live brine imbibition and secondary drainage).
This increase in T1/T2 ratio indicates a drastic reduction in sandstone
hydrophilicity due to live brine exposure. This reduction in hydrophi­
licity was caused by additional protonation of surface silanol groups
(Stumm and Morgan, 2012) (which are located on the quartz surface;
Chen et al., 2015), induced by the carbonic acid (note: CO2 chemically
reacts with H2O to carbonic acid which partially dissociates into H+,
HCO−3 , and CO2− 3 , Adamczyk et al. (2009); the pH value of live brine
under the experimental conditions was 3.6; Peng et al. (2013)). This led
to a lower surface potential and lower polarity of the quartz surface (the
surface become more electrically neutral), which again reduced the
surface’s affinity to (polar) H2O, and thus reduced hydrophilicity
(Iglauer, 2017; Abramov et al., 2019). Importantly, the system remained
weakly water-wet after CO2 was introduced as a separate supercritical
phase (for both, initial CO2 saturation (SCO2,i) – drainage, and residual
CO2 saturation (SCO2,r) – imbibition), i.e. WI = 0.26 was measured,
Table 1. This is consistent with most data in the literature as measured
via video contact angle, microCT and micromodel imaging, capillary
pressure and relative permeability measurements; see also further dis­
cussion below (and compare (Akbarabadi and Piri, 2013; Al-Yaseri et al.,
2016a; Andrew et al., 2014; Chalbaud et al., 2009; Chiquet et al., 2007;
Krevor et al., 2012). Further, as Figure 2a shows, the vertical dashed line
intersects with T2 at 20 ms as compared to T2 = 10 ms for all other
conditions (Figure 2b, c and d). The reduction of high intensity areas
(red band) in Figure 2c, 2d is due to decrease of hydrogen contents in the
fluid inside the rock which demonstrates displacement of water in the
large pores by CO2. This is a clear indication that scCO2 displaced water
in the large pores, which is again an indication of a water-wet system
(Blunt, 2017), consistent with all above observations.
Table 1 looks very confusing in its current format below. There
should be 3 columns x 5 rows (refer to the original format in the
manuscript). Separating the numbers from the text by wide spaces will
make the table neat and tidy.

3.2. Capillary pressure and relative permeability curves

The capillary drainage pressure (Pc) for San Saba sandstone ranged
from 10 to 124 kPa, as shown in Figure 3. The difference between the
San Saba Pc − Sw data and data in the literature (also shown for com­
parison) can be attributed to variations in petro-physical characteristics
of the samples and differences in applied thermo-physical conditions
that can change fluid properties (Abdoulghafour et al., 2020; Al-Yaseri
et al., 2016a; Iglauer et al., 2015a; Iglauer et al., 2015b; Saraji et al.,
2013). A total of 20 PV of scCO2 was injected to displace brine

Table 1
T1
NMR ratio and NMR wettability indices measured for San Saba sandstone for
T2
Figure 2. NMR T1− T2 maps used to determine the T1/T2 ratio and the each process. Note that a wettability index and a NMR T1/T2 ratio of 1 indicate a
wettability indices (WI) for San Saba sandstone (measured at 8 MPa and 333 K), completely water-wet system, while a wettability index of 0 indicates an
the solid diagonal line indicates T1 = T2, a) fully saturated with dead brine, b) intermediate-wet system and a wettability index of − 1 indicates a strongly oil-
fully saturated with live brine, c) initial CO2 saturation (after drainage). d) wet (Al-Muthana et al., 2012; Valori et al., 2017).
residual CO2 saturation (after r drainage). d) residual CO2 saturation (after
Process T1 Wettability index
imbibition). Note that the T1/T2 ratios were selected at the point of highest NMR ratio
T2
intensity in the map (red bands) for WI calculation.
Dead brine 1 -
Live brine 1.7 -
Drainage (CO2 injection) 1.7 0.26
Imbibition (brine injection) 1.7 0.26

4
A. Baban et al. International Journal of Greenhouse Gas Control 111 (2021) 103435

Table. 2
CO2 saturations during primary and secondary drainage and secondary imbi­
bition (measured via NMR) as a function of flow rate.
Flowrate (ml/min) 0.1 0.5 1 2 4 6 8

Primary drainage 0.08 0.15 0.20 0.242 0.288 0.31 0.31


(SCO2,i)
Secondary drainage 0.09 0.13 0.20 0.246 0.292 0.34 0.34
(SCO2,i)
Imbibition (SCO2,r) 0.23

T2 measurements were used to determine the pore-size distribution


via Eq. (7) and to observe the variation of the brine distribution in the
rock’s pore network. It is clear that the scCO2 displaced the (live) brine
from large pores indicating water-wet behaviour (represented by the
shaded area in Figure 5); thus, the T2 relaxation time peak shifted to the
Figure 3. Primary and secondary drainage (scCO2) capillary pressure versus
left (from pores with radii larger than 20 µm to pores with radii less than
brine saturation measured for the San Saba sandstone sample. □ Blue open
10 µm). The preferiential CO2 occupancy by large pores after CO2
squares: primary drainage for San Saba, o red circles: secondary drainage for
San Saba, Black symbols: literature data, added for comparison. The dotted blue
saturated (live) brine flooding is consistent with work reported by
line is the Brooks-Corey model fit. (Connolly et al., 2017) and is also consistent with the observations of
more surface wetting characteristic for CO2 reported by contact angle
method (Al-Yaseri et al., 2016; Iglauer et al., 2015; Iglauer et al., 2014)
(drainage) until SCO2,i (initial CO2 saturation) was established, SCO2,i was
and with results of pore scale residual trapping by x-ray micro-computed
31% at 110 KPa (irreducible water saturation Sw,irr was 69%) and SCO2,r
tomography (μCT) (Igaluer et al., 2011; Niu et al., 2014; Iglauer et al.,
(residual CO2 saturation) (after imbibition) was 23%, consistent with
2019). In contrast, water content in the small pores, (pore radii r < 1
literature values for water-wet sandstone (Akbarabadi and Pini, 2013;
µm), remained unchanged. The lower water saturation of the imbibition
Al-Menhali and Krevor, 2016; Iglauer et al., 2011; Krevor et al., 2012;
T2 peak (incremental porosity black curve in Figure 5) is attributed to
Pentland et al., 2011). As Figure 3 shows, the secondary drainage
the presence of scCO2 in large pores (CO2 is not detected by the NMR
capillary pressure curve shifted very slightly to the right, indicating a
signal).
slight increase in CO2-rock wettability.
Steady state drainage relative permeability data measured as a
4. Conclusions
function of water saturation at representative reservoir temperature and
pressure are presented in Figure 4 and Table 2. Brooks – Corey model
CO2-wettability of storage and caprock is a vital parameter as it
curves (shown as black solid lines ___, obtained via Equations (4) and 5,
determines storage capacities and containment security (Chiquet et al.,
were fitted to the measured datapoints. The Krw - Krg intersection point
2007; Iglauer 2017; Broseta et al., 2012; Iglauer et al., 2015). However,
(for primary drainage) at Sw = 0.7 (Figure 4) indicates strongly water-
there still exists significant uncertainty regarding this parameter, which
wet characteristics of the sample, and is comparable to values stated
leads to high project risk. Thus, we measured the wettability of a
in the literature (Abdoulghafour et al., 2020; Krevor et al., 2012; Niu
quartz-rich sandstone via NMR measurements to reduce this uncer­
et al., 2014; Perrin and Benson, 2009; Pini and Benson, 2013). The slight
tainty. Interestingly, the initially strongly hydrophilic sample turned
variation of the wettability estimation (relative permeability curves
significantly less hydrophilic when exposed to CO2 (either in the form of
intersection), when compared to some of the curves in the literature, are
molecularly dissolved or free supercritical CO2) − the T1/T2 ratio
due to the impact of rock structure and composition; samples with low
increased from 1 to 1.7. These changes in hydrophilicity of the
porosity have overall lower CO2 saturations, rocks with lower void
spaces permit less fluid to pass through them, and thus lower
non-wetting relative permeabilites (Knw), whereas formations with
larger porosity have higher average CO2 saturations and thus higher Knw
values (Perrin and Benson, 2009; Krevor et al., 2012).

Figure 5. Water saturation in the pore network of the San Saba sandstone after
drainage (CO2 flooding) and imbibition (secondary live brine flooding). The
Figure 4. Steady-state drainage relative permeability curves for the San Saba entire T2 drainage curve (red) shifted left which depicts that CO2 displaced
sandstone (o red open circles). black lines () are best fit Brooks – Corey (BC) water from the large pores during drainage. This indicates water-wet behav­
curves. Literature data is also plotted for comparison. iour. Note that T2 times were converted into pore radii via Eq. 8.

5
A. Baban et al. International Journal of Greenhouse Gas Control 111 (2021) 103435

sandstone surface we explain by additional protonation of surface sila­ Andersen, P.Ø., Standnes, D.C., Skjæveland, S.M., 2017. Waterflooding oil-saturated core
samples - Analytical solutions for steady-state capillary end effects and correction of
nol groups and consequently reduced affinity of the quartz surface to
residual saturation. SPE J 157, 364–379. https://doi.org/10.1016/j.
polar H2O, which reduces hydrophilicity (Chen et al., 2015). Related to petrol.2017.07.027.
this a weakly water-wet behaviour for all two phase (CO2 and brine) Bachu, S., Adams, J.J, 2003. Sequestration of CO2 in geological media in response to
systems (primary and secondary drainage and secondary imbibition) climate change: capacity of deep saline aquifers to sequester CO2 in solution. Energ.
Convers. and Manage 44 (20), 3151–3175. https://doi.org/10.1016/S0196-8904
was observed, i.e. WI = 0.26. The capillary pressure, relative perme­ (03)00101-8.
ability and sequential NMR T2 distributions (for each process step) were Broseta, D., Tonnet, N., Shah, V., 2012. Are rocks still water-wet in the presence of dense
comparable with results in the literature acquired via independent X-ray CO2 or H2S? Geofluids 12 (4), 280–294. https://doi.org/10.1111/j.1468-
8123.2012.00369.x.
tomography experiments (Abdoulghafour et al., 2020; Iglauer et al., Brownstein, K.R., Tarr, C.E., 1979. Importance of classical diffusion in NMR studies of
2011; Krevor et al., 2012; Niu et al., 2014; Pentland et al., 2011). As here water in biological cells. Physical Review A 19 (6), 2446–2453. https://doi.org/
an independent technique (NMR) was used, the confidence in the pre­ 10.1103/PhysRevA.19.2446.
Chalbaud, C., Robin, M., Lombard, J.M., Martin, F., Egermann, P., Bertin, H., 2009.
vious datasets is signficiantly increased. Interestingly, water-wettability Interfacial tension measurements and wettability evaluation for geological CO2
was further slightly reduced during secondary drainage. Finally, we storage. Adv. Water Res. 32 (1), 98–109. https://doi.org/10.1016/j.
used the NMR T2 relaxation time distributions to demonstrate water advwatres.2008.10.012.
Chen, C., Zhang, N., Li, W., Song, Y., 2015. Water Contact Angle Dependence with
displacement by CO2 in large pores only (r > 1 µm), while water Hydroxyl Functional Groups on Silica Surfaces under CO2 Sequestration Conditions.
remained unchanged in small pores (r < 1 µm). Environ. Sci. Technol 49 (24), 14680–14687. https://doi.org/10.1021/acs.
The overall conclusion is that CO2 significantly reduces sandstone est.5b03646.
Chiquet, P., Broseta, D., Thibeau, S., 2007. Wettability alteration of caprock minerals by
hydrophilicity and water-wettability. This work therefore provides
carbon dioxide. Geofluids 7 (2), 112–122. https://doi.org/10.1111/j.1468-
fundamental insights into CO2-water displacement processes, and will 8123.2007.00168.x.
aid in the large-scale implementation of CO2 geo-sequestration schemes Coates, G.R., Xiao, L., Prammer, M.G., 1999. NMR logging. Ebooks, 253. https://doi.org/
and carbon capture and storage projects. 10.1016/S0950-1401(02)80008-8.
Connolly, P.R.J., Vogt, S.J., Iglauer, S., May, E.F., Johns, M.L., 2017. Capillary trapping
quantification in sandstones using NMR relaxometry. Water Resour. Res. 53 (9),
7917–7932. https://doi.org/10.1002/2017WR020829.
Declaration of Competing Interest Connolly, Paul R.J., Vogt, S.J., Mahmoud, M., Ng, C.N.Y., May, E.F., Johns, M.L., 2019.
Capillary Trapping of CO2 in Sandstone Using Low Field NMR Relaxometry. Water
Resour. Res 55 (12), 10466–10478. https://doi.org/10.1029/2019WR026294.
The authors declare that they have no known competing financial Corey, A.T., 1954. The Interrelations Between Gas and Oil Relative Permeabilities.
interests or personal relationships that could have appeared to influence Producers Monthly 19, 38–41.
Djebbar Tiab Erle C. Donaldson, Petrophysics (Theory and Practice of Measuring
the work reported in this paper.
Reservoir Rock and Fluid Transport Properties. 4th Ed. SBN: 978-0-12-803188-9.
Dullien, F., 1992. Porous Media. Fluid Transport and Pore Structure. Academic Press Inc.
Supplementary materials Dullien, F.A.L., Zarcone, C., Macdonald, I.F., Collins, A., Bochard, R.D.E., 1989. The
effects of surface roughness on the capillary pressure curves and the heights of
capillary rise in glass bead packs. J. Colloid and Interface Sci 127 (2), 362–372.
Supplementary material associated with this article can be found, in https://doi.org/10.1016/0021-9797(89)90042-8.
the online version, at doi:10.1016/j.ijggc.2021.103435. Dunn, K.J., Bergman, D.J., Latorraca, G.A., 2002. Nuclear Magnetic Resonance
petrophysical and logging applications. Seismic exploration V32. Pergamon. ISBN 0-
08-043880-6.
References El-Maghraby, R.M., Pentland, C.H., Iglauer, S., Blunt, M.J., 2012. A fast method to
equilibrate carbon dioxide with brine at high pressure and elevated temperature
including solubility measurements. J. Supercritical Fluids 62, 55–59. https://doi.
Abdoulghafour, H., Sarmadivaleh, M., Hauge, L.P., Fernø, M., Iglauer, S., 2020. Capillary
org/10.1016/j.supflu.2011.11.002.
pressure characteristics of CO2-brine-sandstone systems. Intl J. Greenhouse Gas
Espinoza, D.N., Santamarina, J.C., 2010. Water-CO2-mineral systems: interfacial tension,
Control 94, 102876. https://doi.org/10.1016/j.ijggc.2019.102876. July 2019.
contact angle, and diffusionImplications to CO2 geological storage. Water Resour.
Abramov, A., Keshavarz, A., Iglauer, S., 2019. Wettability of Fully Hydroxylated and
Res. 46 (7), 1–10. https://doi.org/10.1029/2009WR008634.
Alkylated (001) α-Quartz Surface in Carbon Dioxide Atmosphere. J. Physical
Freedman, R., Heaton, N., Flaum, M., Hirasaki, G.J., Flaum, C., Hürlimann, M., 2003.
Chemistry C 123 (14), 9027–9040. https://doi.org/10.1021/acs.jpcc.9b00263.
Wettability, Saturation, and Viscosity From NMR Measurements. SPE Journal 8 (4),
Adamczyk, K., Premont-Schwarz, M., Pines, D., Pines, E., Nibbering, E.T.J., 2009. Real-
317–327. https://doi.org/10.2118/87340-PA.
time observation of carbonic acid formation in aqueous solution”. Science 326,
Garing, C., Benson, S.M., 2019. CO2 Wettability of Sandstones: addressing Conflicting
1690- 1694.
Capillary Behaviors. Geophy. Res. Lett. 46 (2), 776–782. https://doi.org/10.1029/
Akbarabadi, M., Piri, M., 2013. Relative permeability hysteresis and capillary trapping
2018GL081359.
characteristics of supercritical CO2/brine systems: an experimental study at reservoir
Iglauer, S., Pentland, C.H., Busch, A., 2015. CO2 wettability of seal and reservoir rocks
conditions. Adv. Water Res. 52, 190–206. https://doi.org/10.1016/j.
and the implications for carbon geo-sequestration. Water Resour. Res. 51, 729–774.
advwatres.2012.06.014.
https://doi.org/10.1029/eo066i003p00017-03.
Akitt, J.W., Mann, B.E., 2000. An introduction to modern NMR spectroscopy, 4th Ed.
Guan, H., Brougham, D., Sorbie, K.S., Packer, K.J., 2002. Wettability Effects in a
Taylor & Francis Group. ISBN 0 7487 4344 8.
Sandstone Reservoir and Outcrop Cores From NMR Relaxation Time Distributions.
Al-Khdheeawi, E.A., Vialle, S., Barifcani, A., Sarmadivaleh, M., Iglauer, S., 2018. The
Journal of Petroleum Science and Engineering 34 (1–4), 35–54. https://doi.org/
effect of WACO2 ratio on CO2 geo-sequestration efficiency in homogeneous
10.1016/S0920-4105(02)00151-1.
reservoirs. Energy Procedia 154, 100–105. https://doi.org/10.1016/j.
Iglauer, S., Sarmadivaleh, M., Al-Yaseri, A., Lebedev, M., 2014. Permeability evolution in
egypro.2018.11.017.
sandstone due to injection of CO2-saturated brine or supercritical CO2 at reservoir
Al-Menhali, Ali, Niu, Ben, Krevor, Samuel, 2015. Capillarity and wetting of carbon
conditions. Energy Procedia 63, 3051–3059. https://doi.org/10.1016/j.
dioxide and brine during drainage in Berea sandstone at reservoir conditions. Water
egypro.2014.11.328.
Resources Research 51 (10), 7895–7914. https://doi.org/10.1002/2015WR016947.
Iglauer, S., Al-yaseri, A.Z., Rezaee, R., Lebedev, M., 2015b. Storage Capacity and
Almenningen, S., Roy, S., Hussain, A., Seland, J.G., Ersland, G., 2020. Effect of Mineral
Containment Security. Geophy. Res. Lett. 42 (21), 9279–9284. https://doi.org/
Composition on Transverse Relaxation Time Distributions and MR Imaging of Tight
10.1002/2015GL065787.
Rocks from Offshore Ireland. Minerals 10 (3), 232. https://doi.org/10.3390/
Iglauer, S., 2017. CO2-Water-Rock Wettability: variability, Influencing Factors, and
min10030232.
Implications for CO2 Geostorage. Acc. Chem. Res. 50 (5), 1134–1142. https://doi.
Al-Muthana, A.S., Hursan, G.G., Ma, S.M., Valori, A., Nicot, B., Singer, P.M, 2012.
org/10.1021/acs.accounts.6b00602.
Wettability as a function of pore size by NMR. In: International Symposium of the
Iglauer, S., 2018. Optimum storage depths for structural CO2 trapping. Int. J. Greenhouse
Society of Core Analysts. Aberdeen, Scotland, UK, 27-30 August. SCA2012-31.
Gas Control, 77, 82–87. X-ray micro-tomography. Geophy. Res. Lett. 38 (21), 1–6.
Al-Yaseri, A.Z., Roshan, H., Lebedev, M., Barifcani, A., Iglauer, S., 2016. Dependence of
https://doi.org/10.1029/2011GL049680.
quartz wettability on fluid density. Geophys. Res. Lett. 43 (8), 3771–3776. https://
Iglauer, Stefan, Paluszny, Adriana, Pentland, Christopher.H., Blunt, Martin.J., 2011.
doi.org/10.1002/2016GL068278.
Residual CO2 imaged with X-ray micro-tomography Stefan. Geophysical Research
Ali, M., Arif, M., Sahito, M.F., Al-Anssari, S., Keshavarz, A., Barifcani, A., Stalker, L.,
Letters 38 (21), 1–6. https://doi.org/10.1029/2011GL049680.
Sarmadivaleh, M., Iglauer, S., 2019. CO2 -wettability of sandstones exposed to traces
Iglauer, S., Paluszny, A., Rahman, T., Zhang, Y., Wülling, W., Lebedev, M., 2019.
of organic acids: implications for CO2 geo-storage. Int. J. Greenhouse Gas Control 83,
Residual Trapping of CO2 in an Oil-Filled, Oil-Wet Sandstone Core: results of Three-
61–68. https://doi.org/10.1016/j.ijggc.2019.02.002. June 2018.
Phase Pore-Scale Imaging. Geophy. Res. Lett. 46 (20), 11146–11154. https://doi.
Andrew, M., Bijeljic, B., Blunt, M.J., 2014. Pore-scale imaging of trapped supercritical
org/10.1029/2019GL083401.
carbon dioxide in sandstones and carbonates. Int. J. Greenhouse Gas Control 22,
1–14. https://doi.org/10.1016/j.ijggc.2013.12.018.

6
A. Baban et al. International Journal of Greenhouse Gas Control 111 (2021) 103435

IPCC, 2005. No Title. Intergovernmental PanelonClimateChange(IPCC). Cambridge Perrin, J.C., Benson, S., 2009. An experimental study on the influence of sub-core scale
University Press, 2005. IPCC SpecialR Eporton Carbon Dioxide Capture and Storage, heterogeneities on CO2 distribution in reservoir rocks. Transp. Porous Med. 82 (1),
Prepared by Working Group III of the Intergovernmental Panel on Climate Change. 93–109. https://doi.org/10.1007/s11242-009-9426-x.
Krevor, Pini, R., Zuo, L., Benson, S.M, 2012. Relative permeability and trapping of CO2 Pini, R., Benson, S.M., 2013. Simultaneous determination of capillary pressure and
and water in sandstone rocks at reservoir conditions. Water Resour. Res 48 (2), 1–17. relative permeability curves from core-flooding experiments with various fluid pairs.
https://doi.org/10.1029/2011WR010859. Water Resour. Res. 49 (6), 3516–3530. https://doi.org/10.1002/wrcr.20274.
Katika, K., Saidian, M., Prasad, M., Fabricius, I.L., 2017. Low-Field NMR Spectrometry of Rahman, T., Lebedev, M., Barifcani, A., Iglauer, S., 2016. Residual trapping of
Chalk and Argillaceous Sandstones: rock-Fluid Affinity Assessed from T1/T2 Ratio. supercritical CO2 in oil-wet sandstone. J. Colloid and Interface Sci 469, 63–68.
Petrophysics 58 (2), 126–140. https://doi.org/10.1016/j.jcis.2016.02.020.
Korb, J.-.P., Freiman, G., Nicot, B., Ligneul, P, 2009. Dynamical surface affinity of Ramakrishnan, T.S., Cappiello, A., 1991. A new technique to measure static and dynamic
diphasic liquids as a probe of wettability of multimodal porous media. Physical properties of a partially saturated porous medium. Chem. Eng. Sci 46 (4),
Review E 80, 061601. https://doi.org/10.1103/PhysRevE.80.061601. 1157–1163. https://doi.org/10.1016/0009-2509(91)85109-B.
Laliberte, G.E., 1969. Mesure de profils d’humiditÉ et de tensions au cours d’infiltrations Riaz, A., Hesse, M., Tchelepi, H.A., Orr, F.M., 2006. Onset of convection in a
mono et bidimensionnelles. Int. Assoc. Sci. Hydrology. Bulletin 14 (2), 131–149. gravitationally unstable diffusive boundary layer in porous media. J. Fluid Mech.
https://doi.org/10.1080/02626666909493724. 548, 87–111. https://doi.org/10.1017/S0022112005007494.
Lessenger, M., Merkel, D., Medina, R., Ramakrishna, S., Chen, S., Balliet, R., Xie, H., Saraji, S., Goual, L., Piri, M., Plancher, H., 2013. Wettability of supercritical carbon
Bhattad, P., Carnerup, A., Knackstedt, M., 2015. Subsurface Fluid Characterization dioxide/water/quartz systems: simultaneous measurement of contact angle and
Using Downhole and Core NMR T1-T2 Maps Combined with Pore-Scale Imaging interfacial tension at reservoir conditions. Langmuir 29 (23), 6856–6866. https://
Techniques. Petrophysics 56 (04), 313–333. doi.org/10.1021/la3050863.
Li, M., Vogt, S.J., May, E.F., Johns, M.L., 2019. In Situ CH4-CO2 Dispersion Stumm, W., Morgan, J.J., 2012. Aquatic chemistry: chemical equilibria and rates in
Measurements in Rock Cores. Transp Porous Media (129), 75–92. https://doi.org/ natural waters. Choice Reviews Online 33 (11). https://doi.org/10.5860/choice.33-
10.1007/s11242-019-01278-y. 6312, 33-6312-33–6312.
Looyestijn, W.J., 2008. Wettability index determination from NMR logs. Petrophysics. Valori, A., Hurlimann, M., Nicot, B., Bachman, H., 2013. Obtaining wettability from T1
V49, 2, P130 –145. https://doi.org 10.2118/93624-ms. and T2 measurements. Patent Application Publication. W02013180752.
Lyu, Chaohui, Ning, Zhengfu, Wang, Qing, Chen, Mingqiang, 2018. Application of NMR Valori, A., Hursan, G., Ma, S.M., 2017. Laboratory and Downhole Wettability from NMR
T2 to Pore Size Distribution and Movable Fluid Distribution in Tight Sandstones. T1/T2 Ratio. Petrophysics 58 (04), 352–365.
Energy and Fuels 32 (2), 1395–1405. https://doi.org/10.1021/acs. Valori, A., Nicot, B., 2019. A review of 60 years of nmr wettability. Petrophysics 60 (2),
energyfuels.7b03431. 255–263. https://doi.org/10.30632/PJV60N2-2019a3.
Mesquita, P., Souza, A., Carneiro, G., Boyd, A., Ferreira, F., Machado, P., Anand, V., Virnovsky, G., Vatne, K., Skjæveland, S., Lohne, A., 1998. Implementation of multirate
2016. Surface relaxivity estimation and NMR-MICP matching in diffusionaly coupled technique to measure relative permeabilities accounting for capillary effects. In: SPE
rocks. In: Presented at the Interna- tional Symposium of the Society of Core Analysts. Annual Technical Conference and Exhibition. New Orleans, Louisiana, 27–30
Snowmass, Colorado, 21–26 AugustPaper SCA2016-059. September.
Mitchell, J., Chandrasekera, T.C., Holland, D.J., Gladden, L.F., Fordham, E.J., 2013. Wang, J., Xiao, L., Liao1, G., Zhang, Y., Guo1, L., Arns, C.H., Sun, Z., 2018. Theoretical
Magnetic resonance imaging in laboratory petrophysical core analysis. Phys Rep 526 investigation of heterogeneous wettability in porous media using NMR. Scientific
(3), 165–225. https://doi.org/10.1016/j.physrep.2013.01.003. Repors 8, 13450. https://doi.org/10.1038/s41598-018-31803-w.
Niu, B., Al-Menhali, A., Krevor, S., 2014. A study of residual carbon dioxide trapping in Washburn, K.E., Sandor, M., Cheng, Y., 2017. Evaluation of sandstone surface relaxivity
sandstone. Energy Procedia 63 (0), 5522–5529. https://doi.org/10.1016/j. using laser-induced breakdown spectroscopy. J. Magnetic Resonance 275, 80–89.
egypro.2014.11.585. https://doi.org/10.1016/j.jmr.2016.12.004.
Orr, F.M., 2009. Onshore geologic storage of CO2. Science 325 (5948), 1656–1658. Zhao, Jinsheng, Wang, Pengfei, Zhang, Yanming, Ye, Liang, Shi, Yu, 2020. Influence of
https://doi.org/10.1126/science.1175677. CO2 injection on the pore size distribution and petrophysical properties of tight
Peng, C., Crawshaw, J.P., Maitland, G.C., Trusler, Martin, J., P., Vega-Maza, D, 2013. The sandstone cores using nuclear magnetic resonance. Energy Science and Engineering
pH of CO2-saturated water at temperatures between 308 K and 423 K at pressures up 8 (7), 2286–2296. https://doi.org/10.1002/ese3.663.
to 15 MPa. J. Supercritical Fluids 82, 129–137. https://doi.org/10.1016/j. Zielinski, L., Saha, I., Freed, D.E., Hürlimann, M.D., Liu, Y., 2010. Probing Asphaltene
supflu.2013.07.001. Aggregation in Native Crude Oils with Low-Field NMR. Langmuir 26 (7),
Pentland, C.H., El-Maghraby, R., Georgiadis, A., Iglauer, S., Blunt, M.J., 2011. Immiscible 5014–5021. https://doi.org/10.1021/la904309k.
displacements and capillary trapping in CO2 storage. Energy Procedia 4, 4969–4976.
https://doi.org/10.1016/j.egypro.2011.02.467.

You might also like