You are on page 1of 19

Journal of Petroleum Science and Engineering xxx (xxxx) 109318

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

F
OO
An experimental and field case study to evaluate the effects of shut-in on
well performance
Taregh Soleiman Asl, Ali Habibi, Mahmood Reza Yassin, Obinna Daniel Ezulike,
Hassan Dehghanpour ∗

PR
Civil and Environmental Engineering Department, University of Alberta, Edmonton, Canada

ARTICLE INFO ABSTRACT

Keywords: There is still debate on how shutting fractured wells for a period of time can affect the well performance. In this
Well shut in study, we combined two approaches to better understand the effects of shut-in time on well performance. First,

D
Water imbibition we analyzed flowback and post-flowback production data from a horizontal well drilled in the Montney Forma-
Hydraulic fracturing
tion, which was fractured with water containing a microemulsion (ME) additive. After the shut-in time for 7
Tight rock wettability
months, the oil and solution gas rates significantly increased by 750% and 671%, respectively. However, the free
gas rate decreased by 95% in 65 days, before it started to build up again to exceed the values at the start of the
TE
production. Second, we performed a series of imbibition oil-recovery, dynamic liquid-liquid contact angle, and
interfacial tension measurements to investigate how the interplay of (i) capillary suction and (ii) osmotic pres-
sure affects oil production from core plugs during the counter-current imbibition tests.
Combined analyses of field and laboratory results suggest that the increase in oil production rate after the shut-
in period is due to combined effects of (i) free-gas dissolution into the oil (ii) capillary imbibition of fracturing
water containing ME solution into the rock matrix driven by wettability alteration and osmotic pressure, and (iii)
EC

reduction in phase trapping near fracture face due to interfacial tension reduction.

1. Introduction Adding chemical additives such as surfactants to the fracturing wa-


ter is one of the effective improved oil recovery (IOR) techniques cur-
Unconventional tight oil and gas reservoirs with specific petrophysi- rently used for shale and tight-oil reservoirs (Yassin et al., 2018;
RR

cal properties (low permeability, low porosity, and high organic car- Alvarez and Schechter, 2017a, 2017b; Habibi et al., 2016). These chem-
bon) are crucial energy resources in North America. Horizontal drilling ical additives can alter the rock wettability towards water-wet condi-
and multi-stage hydraulic fracturing are the key technologies for hydro- tions and reduce phase trapping by reducing interfacial tension (IFT).
carbon recovery from these resources (Franz et al., 2005). Unconven- These mechanisms can effectively enhance the spontaneous imbibition
tional tight oil reservoirs in Western Canadian Sedimentary Basin of the fracturing fluid into the reservoir matrix, displace hydrocarbon
(WCSB) have received significant attention due to the latest completion to the fracture, and consequently increase the oil RF (Austad and
technologies, and their desired light crude oil, which make them eco- Milter, 1997 ). Several experimental studies confirm that using differ-
CO

nomically feasible for production (National Energy Board, 2011; ent types of surfactants (anionic, cationic, nonionic, and amphoteric)
Clarkson and Pedersen, 2011). For example, the Montney (MT) Forma- can significantly enhance oil recovery by wettability alteration and
tion in WCSB covers an area of over 130,000 km2. Its estimated hydro- counter-current imbibition (Kewen and Abbas, 2000; Sharma et al.,
carbon-in-place is 4274 Tcf of natural gas and 268 billion bbl of oil and 2014; Alvarez and Schechter,. 2017a, 2017b; He and Xu, 2017; Habibi
natural-gas liquids (Rokosh et al., 2012; Reynolds et al., 2015). How- et al., 2020). Wijaya and Sheng (2020) conducted a numerical model-
ever, its oil recovery factor (RF) is very low (<10% of original oil in ing to evaluate the impact of rock wettability, interfacial tension (IFT)
place) due to ultra-low permeability, complex pore structure, and oil- between water and oil, and surfactant concentration of fracturing fluids
UN

wet or mixed-wet behavior. That means a significant part of hydrocar- on improved oil recovery factor from the Middle Bakken shale reser-
bon remains in the reservoir, and producing the residual hydrocarbon is voir. They showed that the oil recovery factor is maximized when the
a significant challenge. wettability changes toward the water-wet with IFT value as high as

∗ Corresponding author.
E-mail address: dehghanpour@ualberta.ca (H. Dehghanpour).

https://doi.org/10.1016/j.petrol.2021.109318
Received 24 February 2021; Received in revised form 28 July 2021; Accepted 30 July 2021
0920-4105/© 2021
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

20 mN/m. However, in other cases adding surfactants to the fracturing after the shut-in period: i) Dissolution of the free gas into the oil during
fluid can significantly decreases IFT values to less than 1 mN/m and the shut-in period (Self-EOR), ii) reduction of water blockage/satura-
Young-Laplace equation cannot solely explain the counter-current im- tion and formation damage near matrix-fracture interface, and iii) en-
bibition oil recovery only by the capillary pressure (Pc). Therefore, hancing imbibition oil recovery due to wettability alteration by the ME
other factors must be considered in the spontaneous-imbibition process. solution and osmotic pressure. To evaluate these hypotheses, we per-
Osmotic pressure is an important phenomenon that can affect the formed i) data-driven analyses on well post-flowback production data
spontaneous imbibition. Osmosis phenomenon is defined as the flow of to evaluate how shutting the well affects hydrocarbon production, ii)

F
fluid molecules (solvent) through a semi-permeable membrane from spontaneous imbibition oil recovery tests to simulate rock-fluid interac-
low-salinity side into the high-salinity side, until reaching the chemical tions during well shut-in and evaluate the effects of osmotic and Pc on

OO
potential equilibrium in the solution (Cath et al., 2006; Li et al., 2016). oil RF, and iii) dynamic contact-angle measurements to evaluate the
In the reservoir rock, clay minerals can act as semi-permeable mem- possibility of the wettability alteration by the ME solution. Fig. 1 shows
branes, allowing water molecules to pass through, and restricting the the experimental workflow followed in this study. Finally, we estimated
inorganic solutes to pass (Kemper and Rollins, 1966; Malusis et al., osmotic and Pc to evaluate the contribution of each driving mechanism
2003; Olsen et al., 1990). Xu and Dehghanpour (2014) compared the on imbibition-oil recovery.
spontaneous-imbibition rates of fracturing fluids with different salini-
ties to investigate the effects of osmotic pressure on water imbibition in 2. Reservoir, well, and completion descriptions
gas shales of the Horn River Basin. They found that fracturing fluids

PR
with low salinity imbibe significantly higher than fluids with high salin- The target reservoir in the Montney Formation is 165 m thick and its
ity. Using low-salinity water as fracturing fluid is an effective way to en- top is about 1900 m below ground level. It contains saturated black oil
hance the counter-current oil recovery in the low-permeability reser- whose bubble-point pressure and density are 20 MPa and 36.5° API, re-
voirs due to the osmosis effect (Vledder et al., 2010; Austad et al., 2010; spectively. Average dead-oil viscosity measured in the laboratory at
Fakcharoenphol et al., 2014; Padin et al., 2016; Teklu et al., 2018). Dif- temperatures between 20 and 40 °C, is 2.2 cP. The reservoir is com-
ferent studies concluded that in addition to Pc, osmotic pressure signifi- prised of upper (Zone 1) and lower (Zone 2) zones. Appendix A shows
cantly contributes to the spontaneous imbibition of the fracturing fluid the reservoir model built for the reservoir simulation, and it shows
into the rock matrix, enhancing oil RF (Li et al., 2016; Zhou et al., Zones 1 and 2 that are 36 m and 128 m thick, respectively. The perme-

D
2016). ability of Zones 1 and 2 falls within 0.001–0.74 mD and 0.00006–0.007
Several laboratory studies investigated the effects of (i) adding sur- mD, respectively. Both zones have an average porosity of 6% (percent
factants on wettability alteration and increasing oil recovery after soak- of bulk volume) and initial oil saturation values fall within 10–20% and
TE
ing periods (Rao et al., 2006; Yarveicy et al., 2018; Yuan et al., 2019; 50–70% ranges (percent of pore volume), respectively. Zone 1 contains
Liu et al., 2019), and (ii) using low-salinity water on oil production immobile oil and mobile water and is more permeable than Zone 2.
(Fakcharoenphol et al., 2014; Pollen and Berg, 2018). However, there However, Zone 2 is the target production zone due to its mobile oil sat-
are limited studies on comparing the laboratory results with actual field uration. A horizontal well with 2035 m lateral length is drilled in Zone
production data. 2 and fractured by injecting 12,376 m3 of total injected volume (TIV) of
In this study, we analyzed hydrocarbon production data of a hori- water. The well was completed in 65 single-cluster stages with 32 m av-
EC

zontal well completed in the MT Formation and fractured by water con- erage spacing between adjacent clusters. The average half-length of
taining an ME additive. This well was shut-in for 7 months after 5 each fracture stage is 25 m.
months of post-flowback production. After the shut-in period, both oil
and gas rates were significantly increased. Four different hypotheses 2.1. Flowback data
are proposed for the observed increase in oil and gas production rates
After hydraulic fracturing, the well was put on flowback for 33 days.
RR

Fig. 2 shows that the well undergoes early single-phase water produc-
tion for about 17 days, before start of oil and gas production. The well
was eventually shut-in for 2 months after flowback. Its load recovery
(the ratio of cumulative water production to TIV) after flowback is
about 48% that is quite high for 33 days of flowback. However, it could
be due to the delayed hydrocarbon breakthrough and the intercommu-
nication between the upper and lower zones, where the mobile water in
CO

Fig. 1. The experimental workflow is composed of imbibition oil recovery, IFT,


and contact-angle measurements.
UN

Fig. 2. Flowback production data: (a) Flow rates and measured bottom-hole pressure. (b) Cumulative produced volume of water, oil, and gas. sm3 stands for standard
cubic meter.

2
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

Zone 2 can invade Zone 1 and reach the production well through the before the shut-in period, the gas in solution starts to release as the
fractures. pressure declines. However, during the shut-in period, the pressure
builds, leading to dissolution of the free gas back into the oil, increasing
2.2. Post-flowback data the oil volume. This can explain the jump in the rate of gas in solution,
from 783 m3/day before the shut-in to 6039 m3/day after the shut-in
After 2 months of post-flowback shut-in, the well was put on pro- (671% increase). This process can be described as Self-EOR during the
duction using a jet pump. Appendix B shows the post-flowback produc- shut-in period, where the increase in solution gas due to pressure build-

F
tion data for oil, gas, and water at surface conditions. Here, we present up during the shut-in period helps to enhance the oil production rate
the production flow rates under downhole pressure conditions. We con- after the shut-in period. Enhancement of hydrocarbon permeability

OO
verted the surface flowrate data to downhole conditions by using the during the shut-in time can be another effective factor responsible for
well-known correlations available in the literature (Beggs and Brill. the increase of the dissolved gas amount. Fig. 3d shows the production-
1973, Al-Marhoun, 1985; Al-Shammasi, 1999; Guo and Ghalambor. rate ratios of the well. It shows that after the shut-in, the water cut
2005). Fig. 3a shows the post-flowback production data of the well for 5 (WCUT) decreases by around 5 time compared with its value before the
months using the jet pump, and then it was shut-in to complete surface shut-in, while post-shut-in oil-water ratio (OWR) is 20 times more than
facilities and pipeline constructions. The well remained shut-in for 7 pre-shut-in OWR, suggesting significant rise in oil rate after the shut-in.
months, and it was reopened for production again. The jet pump was re- However, 38 days after reopening the well, WCUT gradually returns to
placed with gas lift to minimize water cut. The volume of injected gas almost the same values before the shut-in. Similar to WCUT, OWR re-

PR
was subtracted from the volume of total produced gas to estimate the turns to its values before the shut-in. The gas-oil ratio (GOR) after re-
gas volume produced from the reservoir. The well was monitored using opening the well, decreases 5 times compared with its value before the
a bottom-hole pressure gauge. However, the pressure was not recorded shut-in, then it increases to values even higher than those before the
during the shut-in period. shut-in.

2.3. Shut-in effects on production 2.4. Hypotheses of oil-production rate increase

D
Fig. 3a shows that before the shut-in period, water, oil and total gas The changes in production profiles after the shut-in can be due to
rates decline rapidly along with the bottom-hole pressure decline. The the folowing reasons:
7-month shut-in period significantly affects the instantons oil flow rate
after the shut-in, where it jumps from 14 m3/day before the shut-in, to 1) Dissolution of free gas back into the oil during the shut-in period
TE
approximately its initial rate at start of production, 119 m3/day (750% increases oil saturation and compressibility, and also results in
increase). In contrast to the oil rate, Fig. 3b shows that when the well is solution gas drive mechanism during reopening of the well, as
reopened, free gas rate decreases from 278 m3/day right before the discussed in the previous section.
shut-in to 6 m3/day right after the well is reopened. After 20 days, the 2) IFT reduction by the ME solution reduces water-blockage near the
free gas starts to increase again to values even higher than those at the fracture face, and thus, enhances the oil relative permeability. The
start of the production due to the sudden pressure drop. To understand fluid used for fracturing this well contains an ME additive, which
EC

the change in oil and gas profiles after the shut-in period, we calculated could reduce matrix-fracture skin during the shut-in and improve
the amount of gas dissolved in the oil for this well. Fig. 3c shows that hydrocarbon production rate. One may argue that during the
RR
CO
UN

Fig. 3. Post-flowback production data under downhole conditions. (a) Flow rates and bottom-hole pressure, (b) free gas flowrate (qgfree), (c) solution-gas flow rate
(qgsol), and (d) profiles of rate ratios under downhole conditions (OWR, WCUT, and GOR) where e3 stands for 103.

3
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

flowback and post-flowback periods, the load recovery of water is from 2083 to 2099 m. The well-log and core-log data of these two wells
relatively high for this well, and thus, most of the injected are presented in Appendix C.
fracturing fluid is produced back before the shut-in period. This can
be explained by water production from Zone 1 as an additional 3.2. Porosity, permeability, and pore-throat size distribution of the core
water source. This zone with its mobile water and immobile oil plugs
contributes to the water production through the hydraulic fractures
connecting it to Zone 2. The changes in the matrix-fracture skin Table 1 lists the petrophysical properties of the core plugs collected

F
during shut-in period were discussed with details in a previous from both wells. Porosity was measured using helium porosiometer
study (Soleiman Asl et al., 2020). while the permeability was measured using the pulse-decay method.

OO
3) Counter-current imbibition due to Pc, wettability alteration, and Both permeability and porosity were measured by a commercial labo-
osmotic pressure. Spontaneous imbibition of the fracturing fluid ratory. Core plugs from Well 1 are more porous and permeable than
into the reservoir matrix and counter-current production of oil may those from Well 2. This variation in porosity is due to the differences in
lead to an increase in the early production rate of oil after the shut- their pore-throat size distribution and grain sorting. Appendix D shows
in period. The higher salinity of the formation brine compared with the mercury intrusion capillary pressure (MICP) data and calculated
the injected fracturing water can enhance water imbibition into the pore-throat size distribution of these core plugs.
matrix due to the osmosis effect.
3.3. Mineralogy of the core plugs

PR
In the next section, we present the experimental results to evaluate
the counter-current imbibition of aqueous phases with and without the Table 2 lists the mineralogy of the core plugs obtained from the x-
ME solution into the oil-saturated core plugs. We conducted the imbibi- ray diffraction (XRD) analysis. It shows that the core plugs are mainly
tion tests under four different conditions to investigate the effects of the composed of quartz, dolomite, and clay minerals with concentrations in
ME solution and osmosis pressure on oil recovery. In addition, we mea- the range of 29–56 wt%, 6–20 wt%, 3–32 wt%, respectively. The clay
sured dynamic oil contact angle to evaluate wettability alteration dur- content of the core plugs of Well 1 are significantly higher than that of
ing the addition of the ME solution to water. Well 2, and the most abundant clay mineral is illite/mica. Appendix E

D
shows the scanning electron microscopy (SEM) images of samples close
3. Materials to plugs MT 2-1 and MT 2-2. The SEM images show that the pores of
both core plugs are intergranular, and mainly surrounded by quartz,
3.1. Core properties dolomite, and clay minerals, consistent with the XRD results.
TE
To conduct imbibition oil recovery tests, six core plugs were col- 3.4. Fluid properties
lected from two wells drilled in the MT Formation. These two wells
drilled in the same field are close to each other. We used four core plugs In this study, we used crude oil samples produced from Well 1. It is a
of Well 1, which is the target well and we analyzed its production data light oil sample (36.5° API) with viscosity and surface tension of
in the previous subsection. We used another two core plugs from the 4.97 cP and 24 mN/m, respectively, at room temperature and atmos-
EC

other well (Well 2) to confirm the repeatability of the results performed pheric pressure. We used synthetic brine because the produced water is
on the target well. The depth of four core plugs from Well 1 range from a mixture of formation brine and fracturing fluid containing some
2026 to 2073 m while that of the two core plugs from Well 2 ranges chemical additives that affect the experimental results. The total salin-
ity of the synthetic brine is 134,688 ppm, which is equivalent to the
salinity of the reservoir brine. Table 3 lists the composition of the syn-
thetic brine. Density, viscosity, and surface tension of the synthetic
Table 1
RR

Petrophysical properties of the core plugs from Wells 1 and 2. Permeability


brine are 1.089 g/cm3, 1.10 cP, and 68 mN/m, respectively at room
data were measured by pulse-decay method. temperature and pressure. The TDS (total dissolved solids) of the tap
water used in this study is approximately 1000 ppm.
Well Sample Depth Diameter Length Permeability Porosity Bulk
ID (m) (cm) (cm) (μd) (%BV) density,
(g/cm3) 3.5. Microemulsion properties

Well 1 MT 1-1 2026 2.33 1.88 7.32 10.72 2.45


We used a complex nano-fluid (CnF) sample to conduct the experi-
CO

(Target
well) ments. It is the same chemical product that was added to the water used
MT 1-2 2044 2.33 1.67 6.70 7.96 2.56 for hydraulic fracturing of the target well. This environmental-friendly
MT 1-3 2071 2.33 1.99 1.22 11.51 2.44 additive is comprised of nonionic surfactants (with hydrophilic-
MT 1-4 2073 2.33 1.83 3.01 6.30 2.56 hydrophobic balance of 13–15) surrounding citrus terpenes as solvent
Well 2 MT 2-1 2083 3.81 6.00 1.80 5.10 2.57 (10–25 vol%). The CnF product is diluted with water at room tempera-
MT 2-2 2099 3.81 6.00 3.85 6.18 2.53
UN

Table 2
Mineralogy of the core plugs evaluated by XRD analysis.
Well Sample Core Quartz K- Plagioclase Dolomite Pyrite Fluorapatite Calcite Total Illite/Smectite Illite + Mica Kaolinite Chlorite Total
ID depth (m) Feldspar Non- (I/S) Clay
Clay

1 1 2007.00 56 12 11 17 1 0 0 97 0 3 0 0 3
2 2041.00 36 13 13 18 2 0 0 82 0 18 0 0 18
3 2065.00 55 9 9 14 1 0 2 90 0 10 0 0 10
4 2075.00 29 15 16 6 3 0 0 68 0 29 2 0 32
5 2088.00 54 9 13 20 1 0 1 96 0 4 0 0 4
2 MT 2-1 2083.83 59 7 11 19 1 0 0 97 0 3 0 0 3
MT 2-2 2099.66 56 9 10 18 1 1 0 95 4 0 0 1 5

4
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

Table 3 Table 4
Concentration of the ions used in the synthetic brine prepared for the experi- Calculated pore volume of core plugs and the corresponding soaking fluids
ments. used in Scenarios A, B, and C.
Cations Concentration, mg/L Anions Concentration, mg/L Sample Scenraio Pore volume, Soaking fluid(s)
ID cm3
Na+ 49,750 Cl− 80,945
K+ 1220 HCO3- 0 MT 1–1 A&B 0.85 ME solution in Scenario A, tap water in

F
Ca2+ 1120 SO42- 623 Scenario B
Mg2+ 596 – – MT 1–2 A&B 0.56 ME solution in Scenario A, tap water in
Scenario B

OO
MT 1–3 A&B 0.97 Tap water in Scenario A, ME solution in
ture with 1 cc/L concentration, forming surfactant-solvent nanodisper- Scenario B
sions. The range for specific gravity of this product is 0.904–0.964 be- MT 1–4 A&B 0.49 Tap water in Scenario A, ME solution in
fore mixing it with water. The viscosity and surface tension of the ME Scenario B
MT 2–1 C 3.48 Synthetic brine
solution at room conditions are 0.98 cP and 30 mN/m, respectively.
MT 2–2 C 2.56 Tap water
More details about the chemical structure of this ME additive are pre-
sented elsewhere (Soleiman Asl et al., 2018; Eghbalvala et al., 2020).
the ME solution. The produced oil volume was accumulated at the top
of the imbibition cells due to buoyancy, and it was recorded with re-

PR
4. Methodology
spect to time using a graduated tube at the top of the imbibition cells
with 0.02 cm3 accuracy.
4.1. Imbibition oil recovery tests

We hypothesized that the counter-current imbibition due to wetta- 4.2.1. Scenario a repeated
bility alteration, capillary suction, and osmosis phenomena can partly In the next step, we repeated the imbibition tests using these core
explain the enhanced hydrocarbon production after the shut-in period. plugs to check the repeatability of the oil recovery results in Scenario A
To investigate the effect of these factors on the imbibition oil recovery, and ensure the results are independent of petrophysical properties of

D
we considered three test scenarios. In Scenario A, we performed the the selected core plugs. Here is the procedure for preparing the core
tests using four oil-saturated side-wall core plugs of the target well, im- plugs for repeating the imbibition tests:
mersed in the tap water and the ME solution. This is to investigate if
• Clean the core plugs with toluene and methanol using a Soxhlet
adding the ME to tap water, which affects the interfacial tensions (IFT)
TE
and wettability, can enhance the spontaneous imbibition of the aqueous extractor.
• Place the core plugs in the oven at 90 °C for 48 h to ensure there is
phase into the plugs and improve the oil RF. In Scenario B, we repeated
the previous Scenario A to compare the oil RF of the same core plugs, no water and oil left in the core plugs.
• Saturate them with the reservoir oil by following the saturation
but this time in the absence of the initial salt in the pore network of the
core plugs. In Scenario C, we conducted imbibition oil-recovery tests on procedure mentioned earlier.
• Place the two core plugs (previously immersed in the tap water in
EC

the core plugs of Well 2 (close to the target well) using tap water and
synthetic brine to investigate how salinity of water affects the oil recov- Scenario A) in the imbibition cells filled with the ME solution after
ery from oil-saturated core plugs. Before presenting the experimental cleaning and restoring the initial oil-saturation conditions.
procedure of these three scenarios, we describe the procedure for satu- Similarly, the other two core plugs (previously immersed in the ME
rating the as-received core plugs with oil: solution in Scenario A) were placed in the tap water.

• Place core plugs in an oven at 90 °C for 24 h to dry out residual oil 4.3. Scenario B: effects of the osmotic pressure
RR

and brine left in the core plugs.


• Measure the weight of the dry core plugs. To understand how osmotic pressure affects oil recovery from the
• Place dry core plugs in a high-pressure accumulator filled with the core plugs, we used the following procedure to make sure that the pore
reservoir oil and increase pressure applied on the accumulator to walls of the cleaned core plugs are covered with the salt in the reservoir
3550 kPa using a pulse-free pump connected to the bottom of the brine:
accumulator. Soak the core plugs inside the accumulator at
• Clean the core plugs by using the Soxhlet extractor, as mentioned in
CO

3550 kPa for 24 h.


• Measure the weight change of the core plugs after 24 h. previous section (Scenario A repeated).
• Increase the applied pressure to 20,785 kPa with 3550 kPa as the • Place the core plugs in the oven at 90 °C for 48 h to dry out toluene
incremental pressure step and calculate the weight change at each and methanol.
step. • Saturate the plugs with the high-salinity synthetic brine by
• Leave the cores in the accumulator for 24 h at each pressure step. pressurizing them in the accumulator filled with brine at 20,785 kPa
Stop the saturating process at 20,785 kPa. for 2 days.
• Place the brine-saturated core plugs in the oven at 90 °C for 7 days
UN

• Measure the weight of oil-saturated core plugs and calculate the


volume of oil in place. to ensure that water is mostly evaporated and only salts are left in
the pore space.
• Place the dry plugs in the accumulator filled with the oil and follow
4.2. Scenario A: effects of the ME on capillary pressure
the saturation process.
• Immerse a pair of the oil-saturated core plugs in tap water
After saturating the four as-received side-wall core plugs of Well 1
with the reservoir oil, we immersed them in the imbibition cells filled (previously immersed in ME solution in Scenario A) and the other
with (a) tap water (base case) and (b) the ME solution prepared in tap pair in the ME solution (previously immersed in tap water in
water. Tap water is considered as the base case since in the field, the ME Scenario A) and measure the imbibition oil recovery with respect to
was added to fresh water for hydraulic fracturing of the target well. time.
Since the individual side-wall plugs have low pore volumes (Table 4),
we soaked two of the plugs in the tap water and the two others two in

5
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

4.4. Scenario C: effects of salinity on oil recovery IFT measurement was repeated three times and the mean value was re-
ported.
To further investigate the effects of salinity on the oil recovery re-
sults, we conducted another set of imbibition oil recovery tests on core 4.6. Contact-angle measurement
plugs of Well 2, which is close to the target well. The core plugs from
Well 2 have larger bulk and pore volumes compared to those from Well To investigate how the ME additive affects the rock-fluid interac-
1 and their SEM images are available for interpreting experimental re- tions, we performed liquid-liquid contact angle tests using end-pieces of

F
sults. We used two oil-saturated plugs from Well 2 that were saturated the core plugs, tap water, ME solution, synthetic brine, and oil sample.
with oil, by following the same procedure used for saturating the as- Attension Theta (Biolin Scientific) instrument was used to measure con-

OO
received side-wall core plug of Well 1 in Scenario A. One of the plugs tact angles and investigate wettability alteration by the ME additive.
was soaked in the synthetic brine and the other one in tap water. Table This instrument has Navistar lens (1984 × 1264-pixel resolution, max
4 lists the calculated pore volume from measured porosity and the cor- 3009 FPS) and an LED light source. To remove dirt and mitigate the ef-
responding soaking fluids used for each core plug. fects of surface roughness on contact-angle results, the surface of the
end-pieces was polished using 440 and 600 grit sandpapers. We per-
4.5. IFT measurement formed two types of liquid-liquid contact-angle tests. In Type 1 (limit-
ing conditions), we saturated the end-pieces of the core plugs by im-
To understand the effect of IFT changes on the oil recovery results, mersing them in the oil sample for 14 days. Then, we immersed the oil-

PR
and to estimate the value of Pc, we measured the IFT between the aque- saturated end-pieces into a cell filled with the aqueous phase (tap water
ous and oleic phases. We used a spinning drop tensiometer (SDT, Krüss, or ME solution). We injected an oil droplet below the end-piece using a
Germany) to measure the IFT between (i) tap water and the reservoir oil J-shape needle and monitored the change in the droplet shape and con-
sample (reference case), (ii) The ME solution and the reservoir oil sam- tact angle, in the presence of tap water or ME solution. Results of con-
ple, and (iii) the synthetic brine and reservoir oil sample. We used AD- tact angle measurements performed at limiting conditions do not neces-
VANCE™ software to analyze the shape of the oil droplet spinning in sarily represent rock-fluid interactions under real field conditions. Dur-
the capillary tube and to calculate the IFT values using Vonnegut equa- ing fracturing operations, fracturing fluids containing additives such as

D
tion (Vonnegut, 1942): surfactants come to contact with reservoir oil that is already at equilib-
rium conditions with the reservoir brine and rock. To mimic rock-fluid
(1) interactions at these conditions, we conducted Type 2 (dynamic condi-
tions) contact-angle measurements in the following steps:
TE
Here, is interfacial tension (N/m), is density of aqueous phase • The oil-saturated end-piece was immersed in the reservoir brine
(kg/m3), is density of oleic phase (kg/m3), is angular velocity (ra- (60 cc).
dians/s) and R is radius of the spinning oil droplet (m). IFT values were • An oil droplet was injected from the bottom of the end-piece to
recorded every minute until arriving at equilibrium conditions. Each equilibrate on the rock surface in the presence of the brine.
• The ME solution in tap water was gradually injected into the cell
EC

gradually. We added 30 cc of the ME solution to 60 cc of brine.


• The contact angle of the oil droplet was monitored with respect to
time during diffusion of the ME solution through the reservoir brine
towards the oil droplet.

We monitored the change in the contact angle of the oil droplet for
RR

48 h. This test was repeated three times and the average value was re-
ported. This technique mimics reservoir conditions where the reservoir
rock is initially in contact with the reservoir brine, and then the ME so-
lution is introduced to the rock-fluid system after the fracturing opera-
tion.

5. Results and discussions


CO

5.1. Scenario A: effects of the ME solution on capillary pressure

Fig. 4a and b are close-up shots of the surfaces of the oil-saturated


core plugs immersed in the tap water and the ME solution after 6 h. The
oil droplets were expelled from the core plugs immersed in the ME solu-
tion while no oil expulsion was observed from the core plugs immersed
UN

in tap water after 6 h. Fig. 5a shows the oil recovery profiles for the im-
bibition tests performed with tap water and the ME solution. The oil RF
for the core plugs immersed in the ME solution is almost 12% of the
original oil-in-place after 6 h. Then, the oil RF gradually increased to
24% after 25 days and remained constant. The oil RF for the base case
(tap water) slowly increased and reached the plateau of around 2% af-
ter 3 days. Therefore, one may conclude that the oil RF of the ME solu-
tion is 12 times higher than that of tap water.

6
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
PR
D
TE
Fig. 4. Scenario A: (a) core plugs immersed in the ME solution and (b) in tap water (reference case) after 6 h. Scenario B: (c) oil droplets expelled from the core plugs
immersed in the ME solution and (d) tap water after 20 h.

5.1.1. Scenario a - repeated in the repeated Scenario A, there was almost no oil recovery from the
In this step, to test the consistency and repeatability of the results in core plugs. However, when we reintroduced the precipitated salts in the
Scenario A, we repeated the test by following the procedure described pore network of the core plugs in Scenario B, imbibition oil RF was sig-
EC

in the methodology section. Surprisingly, we observed negligible oil nificantly increased. So far, we tested the effects of the osmotic pressure
production (less than 0.5% of the initial oil-in-place) in both imbibition by changing the salt concentration in the pore network only. In the next
cells after 14 days. We repeated the tests again, and the same results scenario, we changed the salinity of soaking fluids without changing
were obtained. The main difference between Scenario A and the re- the initial salinity of the pore fluid. For that, we soaked the core plugs
peated Scenario A was the initial conditions of the core plugs. In Sce- of Well 2 in the high-salinity synthetic brine and tap water (1000 ppm
nario A, we performed the tests on as-received core plugs and did not salinity) in Scenario C.
RR

clean them with solvents. However, in the repeated Scenario A, we


cleaned the plugs to remove the water and the ME solution imbibed into 5.3. Scenario C: effects of salinity on oil recovery
the plugs. This process also removed salts originally precipitated in the
plugs. Since the crude oil used in this study is very light (36.5oAPI), we Fig. 5c shows the oil RF profiles for Scenario C. After 24 h, the oil
do not expect to have precipitation of heavy-oil components on the pore droplets appeared on the surface of the core plug immersed in the tap
walls after heating the core plugs at 90 °C. We hypothesize that the sig- water, while no oil droplets were observed on the surface of the core
nificant reduction in oil recovery is due to the reduction in osmotic plug immersed in the synthetic brine. Only 7% of the original oil-in-
CO

pressure by removing the salts precipitated in the pore network of the place is produced for the core plug immersed in the synthetic brine and
plugs. To test this hypothesis, we conducted experiments of Scenario B the oil RF reached the plateau after 600 h. However, the oil RF for the
and evaluated the role of osmotic pressure. core plug immersed in tap water is 15% of the original oil in place after
1400 h and reached the plateau after 300 h. The higher oil RF for tap
5.2. Scenario B: effects of osmotic pressure water in this scenario compared to the previous scenarios is due to the
difference in the petrophysical properties of the core plugs.
Overall, the imbibition oil recovery tests from the three scenarios
UN

Fig. 4c and d shows close-up shots of oil droplets expelled from the
oil-saturated core plugs immersed in the tap water and the ME solution lead to the following key results:
after 20 h. Fig. 5b shows the oil RF profiles for the imbibition tests per-
formed in Scenario B. The results show that after 300 h, the oil RF from (i) Using the ME solution as the soaking fluid can enhance the oil RF
the core plugs immersed in the tap water reached the equilibrium value compared with tap water without ME.
of 5%, while that from the core plugs immersed in the ME solution (ii) The oil RF is negligible from the core plugs immersed in either tap
reached 14% at this time. After 960 h (40 days), 34% of the original oil water or the ME solution, in the repeated Scenario A after
was recovered from the core plugs soaked in ME solution. This result cleaning the cores, leading to removal of the precipitated salts. In
was similar to the one obtained from Scenario A. By comparing Sce- Scenario B, after saturating the pore space of the core plugs with
nario B and the results of repeated Scenario A, it is concluded that os- brine and oil, we observed 34% and 5% of oil RF when the core
mosis is a key driving mechanism for imbibition oil recovery from the plugs were immersed in ME solution and tap water, respectively.
core plugs. When precipitated salts in the pore network were removed

7
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
PR
D
TE
Fig. 5. Imbibition oil recovery profiles for: (a) Scenario A, where the oil RF from core plugs immersed in the ME solution is 12 times higher than that from the plugs
immersed in tap water. (b) Scenario B, where 34% of the oil was recovered from the plugs immersed in the ME solution while only 5% of oil was recovered by tap
water. (c) Scenario C, where 15% of the total oil in place was recovered from the core plug immersed in the tap water, compared to only 7% of oil RF when the plug
EC

was immersed in the synthetic brine.

The results indicate a strong relationship between the oil RF and force in the spontaneous imbibition process. Therefore, we compared
the salinity gradient between the imbibing and residing water. the wetting behavior of ME solution, tap water, and synthetic brine by
(iii) The core plugs soaked in the ME solution in both Scenarios A and comparing their contact-angle values.
B resulted in oil RF of 24% and 34%, respectively. In other words,
RR

the core plugs of Scenario B which contain higher precipitated 5.5. Contact-angle measurement results
salts compared with the core plugs of Scenario A, have higher oil
RF. This suggests that the osmotic pressure can be an additional 5.5.1. Liquid-liquid contact angle results (Type 1 – limiting conditions)
driving force for delivering the ME to the pore network of the To understand the effects of different soaking fluids on wettability
plugs. More analyses are required to quantify the extent of of the rock, we measured the contact angle of oil droplets in the pres-
osmotic pressure on improving oil displacement from oil- ence of the soaking fluids used in the oil recovery tests. Fig. 6a–c shows
saturated pores. In the next subsections, we calculate and
CO

oil droplets equilibrated on the surface of oil-saturated end-pieces of


compare the values of osmotic and Pc representing the test the core plugs when the aqueous phases are synthetic brine, tap water,
conditions. and ME solution, respectively. The oil contact angles for the synthetic
(iv) The results obtained from Scenario C demonstrate that oil RF by brine and tap water cases (46° ± 1° and 54° ± 5°) are lower than that
the low-salinity tap water is around two times of that by the high- for the ME case (60° ± 4°). However, the contact angle results still
salinity synthetic brine. show an oil-wet behavior ( ) for the ME solution case, and this
does not explain the significant oil recovery by spontaneous imbibition
5.4. IFT measurement results of the ME solution into the core plugs.
UN

The equilibrium IFT values for the oil-tap water, oil-ME solution and
oil-synthetic brine are 8.57 ± 0.44 mN/m, 0.55 ± 0.33 mN/m, and
1.25 ± 0.02, respectively. Although IFT value for the oil-ME solution is
15 times less than that for the oil-tap water, the ME solution results in
12 times more imbibition oil RF. However, the IFT value for oil-ME so-
lution (IFT > 0.1 mN/m) is not within the ultra-low range
(IFT < 0.01 mN/m) (Rosen et al., 2005). Therefore, we expect suffi-
cient capillary suction for imbibition of ME solution into the oil-
saturated core plugs. According to the Young-Laplace equation, IFT and Fig. 6. Type 1 - Limiting conditions: Oil droplets equilibrated on the surface of
oil-saturated end pieces immersed in (a) synthetic brine, (b) tap water, and (c)
wettability are two factors controlling the capillary suction as a driving
ME solution.

8
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

5.5.2. Liquid-liquid contact angle results (Type 2 – dynamic conditions) of rock surface towards water-wet conditions. This can be observed in
Fig. 7a to c shows the change in oil contact angle with respect to Fig. 7c and d, where the contact angle increases to 71° ± 3° and 92° ±
time in Type 2 tests. Before adding the ME solution to brine, the oil 7° after 24 and 48 h, respectively. The results suggest the gradual alter-
contact angle is 76° ± 2°, indicating an oil-wet behavior (Fig. 7a). The ation of the wetting state from oil-wet to water-wet conditions by diffu-
oil contact angle reduces to 67° ± 3°, 30 min after adding the ME solu- sion of the ME solution through the brine toward the oil droplet, which
tion into the cell filled with brine (Fig. 7b), indicating stronger oil-wet is consistent with observations of Yuan et al. (2021).
behavior. The oil contact angle increases to 71° ± 3° and 92° ± 7° after

F
24 and 48 h (Fig. 7c and d), respectively. The results suggest the grad- 6. Capillary pressure and osmotic pressure as driving forces
ual alteration of the wetting state from oil-wet to water-wet conditions

OO
by diffusion of the ME solution through the brine toward the oil The experimental results led to three important outcomes that re-
droplet. quire further discussions: (i) By adding the ME to the soaking fluid, the
At early time of adding ME solution prepared in tap water to the rock samples become slightly water-wet, and the IFT of the oil-ME solu-
cell, the reduced density of the brine surrounding the oil droplets may tion decreases significantly. Although the IFT of the oil-ME solution is
assist the IFT reduction between oil and brine (Equation (1)) and the oil lower than that of oil-tap water and oil-synthetic brine cases, the oil RF
contact angle reduction based on Young equation. However, there are by the ME solution is higher than those by the other fluids. (ii) Oil RF
other mechanisms responsible for altering the wettability of rock to- was negligible in repeated Scenario A in the absence of precipitated
wards the water-wet conditions. The first mechanism is increasing the salts in the pore network. (iii) In Scenario C, higher oil RF was obtained

PR
double layer repulsion between the oil layer and rock surface in a rock- from the core plugs immersed in tap water compared with those im-
oil-brine system by dilution of brine when ME solution diffuses into the mersed in the synthetic brine.
brine surrounding the oil droplets. This promotes the water-wetness of These observations suggest the existence of other driving forces in
the rock surface. According to Yuan et al. (2021), the high salinity of addition to Pc, responsible for the strong imbibition of the ME solution
aqueous phase (brine) can decrease the stability of the aqueous film be- into the oil-saturated core plugs. Here, we hypothesize that the addi-
tween the oil and rock surfaces due to reduction of the total interaction tional driving force is due to the osmosis effect. Osmosis phenomenon is
energy, resulting in a higher affinity of the rock towards the oil. How- defined as the movement of a fluid through a semi-permeable mem-
brane from the low-salinity side of the membrane to the high-salinity

D
ever, the added ME solution diffuses through the brine and dilutes the
aqueous phase over the time. Therefore, the diluted aqueous phase can side of it. The reason behind the fluid movement is the difference in
result in a higher total interaction energy, more stable film, and thus al- chemical potential of the fluid on both sides of the membrane (Fritz,
tering the wettability of the rock towards water-wet conditions. The 1986; Schmid et al., 2014). In the tight oil reservoirs, clay minerals can
TE
other mechanism is interaction of strongly hydrophilic nonionic surfac- act as semi-permeable membranes that allow water to pass through
tants of the ME solution with the rock surface, altering the wettability (Fakcharoenphol et al., 2014). When only water is allowed to move
through the membrane, it is considered as ideal membrane and when in
addition to water, salts can diffuse in the opposite direction of the water
flux, it is considered a non-ideal membrane (Fritz, 1986; Kooi et al.,
2003). Fig. 8 schematically illustrates the osmosis mechanism, where
EC

only the water can flow from the low-salinity side of an ideal semi-
permeable membrane to the high-salinity side. The water flow creates
an external pressure on the higher-salinity side, which is referred to as
the osmotic pressure.
In this part of the study, to understand the effects of Pc and osmotic
pressure on the oil RF from the side-wall core plugs of Well 1, we will
RR

calculate the Pc and osmotic pressure resulted during spontaneous im-


bibition process during Scenarios A and B.

6.1. Calculation of capillary pressure and osmotic pressure

To calculate the Pc, we use the Young-Laplace equation, and assume


that the pore-throat size ranges from 20 to 200 nm in all the core plugs.
CO

This assumption is made based on the pore-throat size distribution data


UN

Fig. 7. Type 2 - Dynamic conditions: Oil droplets equilibrated on the surface of oil-saturated end piece: (a) At the start of the experiment where the aqueous phase is
synthetic brine only. (b) 30 min after adding the ME solution the contact angle is reduced and the rock shows stronger affinity toward oil. (c) After 24 h of the ex-
periment in the presence of synthetic brine and the ME solution, the contact angle slightly increased (d) After 48 h of the experiment, the oil contact angle increased
to 92°. The results of Type 2 contact-angle measurements show that the rock wettability alters from oil-wet to slightly water-wet with time.

9
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
Fig. 8. Schematic illustration of osmosis mechanism through an ideal semi-permeable membrane. The fluid moves from the low-salinity side of the membrane to the
high-salinity one, creating an external pressure on the high-salinity side, referred to as osmotic pressure (Figure adapted from Marbach and Bocquet, 2019).

PR
obtained from the MICP tests (Appendix D). The IFT and contact-angle where, is the ion mole fraction of the solvent and is the solvent ac-
values used in the Pc calculations are obtained from experimental re- tivity coefficient. The activity coefficient for the ideal solutions is ,
sults presented in previous sections. Table 5 lists the calculated Pc re- and for the real solutions is less than 1. To calculate the activity of syn-
sults. The Pc values for the tap water case are in the range of 7–73 psi, thetic brine filling the rock pore space, initially we considered it as an
and for the ME solution case, the maximum calculated Pc is only 0.3 psi. ideal solution, and the synthetic brine is proportional only with the ions
The results indicate that the calculated Pc cannot be the only driving mole fraction. Next, since the synthetic brine is a non-ideal solution in
force for the measured imbibition oil recovery. Next, we calculate the the rock pore space, we corrected the estimated osmotic pressure by as-

D
values of osmotic pressure corresponding to the test conditions includ- suming values of 20%, 50% and 70% for its non-ideality. We analyzed
ing salinity difference. the changes in the osmotic when the imbibing fluid dilutes the synthetic
In the imbibition tests, we used tap water and ME solution prepared brine by the factor of 10. Table 6 lists the calculated values of the os-
by tap water. The only factor changed in Scenarios A, Scenario A re- motic pressure, assuming temperature of 25 C.
TE
peated and Scenario B, was the salinity inside the pore spaces caused by Comparing the range of osmotic pressure (474–1660 psi) with that
the injected synthetic brine. Therefore, we have a system of a low- of Pc calculated by the Young-Laplace equation (0.3–73 psi), one may
salinity side (the imbibing fluid) which imbibes inside the core plug conclude that osmotic pressure plays a key role in imbibition oil recov-
(the core plug is already saturated with oil at the irreducible brine satu- ery from the oil-saturated core plugs. Overall, osmotic and Pc comple-
ration) and a high-salinity side (the reside brine inside the pores of the ment each other for water imbibition into the oil-saturated core plugs.
core plugs), where both sides are separated by a semi-permeable mem-
EC

brane (the clay minerals inside the core plugs). To calculate the osmotic 7. Summary and conclusions
pressure due to the difference in salinity of imbibing and residing wa-
ter, we used the concentration values of the ions in the synthetic brine 7.1. Field data analysis
listed in Table 3. The osmotic pressure equation derived from the Gibbs
energy equation (Kuhn et al., 2009) is given by In this paper, we presented flowback and post-flowback production
data from a horizontal well drilled and fractured in the Montney Forma-
RR

(2) tion. This well was shut-in for 7 months after 5 months of the post-
flowback production. By analyzing the field production data of this
well, we found that after reopening the well there was an increase of
Here, is the osmotic pressure, is the ideal gas constant, T is the
750% and 671% in oil and solution-gas rates, respectively, while there
temperature in K, is the molar volume of the solvent, and is the
was a 98% decrease in the free-gas rate for the same period. Analysis of
activity of the solvent which is obtained from
the production data leads to the following conclusions:
CO

(3)
• Most of the free gas is dissolved back into the oil as a result of the
pressure buildup during the shut-in period.
• The increase in oil production rate after reopening of the well is
Table 5 partly due to the increase in oil saturation and compressibility, and
The values of the capillary pressure calculated using the Young-Laplace equa- solution gas drive mechanism.
tion for oil-tap water and oil-ME solution.* We used contact angle of 54 for
tap water, and 92 for the ME solution case.
UN

Pore throat size *Pc(psi) when tap water *Pc(psi) when ME solution
(nm) used used
Table 6
20 73.03 0.31 The results of calculated osmotic pressure. All the calculations are performed
40 36.51 0.14 at the room temperature.
60 24.35 0.10 Salinity of the residual (psi) (psi) (psi)
80 18.26 0.06 brine inside the pores assuming 20% Assuming 50% assuming 70%
100 14.61 0.05 (ppm) non-ideality non-ideality non-ideality
120 12.18 0.04
140 10.43 0.03 134,688 (Base case) 0.89 474 1185 1660
160 9.13 0.03 13,000 0.96 50 121 170
180 8.12 0.02 1300 0.98 5 12 17
200 7.30 0.02 130 0.99 0.5 1.2 2

10
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

7.2. Laboratory experiments • From all the experimental and calculation results, one may conclude
that Pc and osmotic pressure are complementary driving forces for
To investigate the effects of the ME solution on the well perfor- imbibition oil recovery, in a way that without the osmotic pressure,
mance during the shut-in period, we performed a set of the spontaneous the capillarity is not effective.
imbibition oil recovery tests with three different scenarios, that led to
the following results: Author contributions

F
• In “Scenario A″, using microemulsion solution can lead to 12 times Taregh Soleiman Asl: Conceptualization, Methodology, Experi-
higher oil RF compared to the base case without using it, indicating ments, Data Formal analysis, Writing original draft. Ali Habibi: Writ-

OO
the significant role of the microemulsion solution in enhancing the ing and Experimental Data Interpretation. Mahmood Reza Yassin:
imbibition oil recovery. Writing and Field Data Interpretation. Obinna Daniel Ezulike: Writing
• Results of “Scenario A repeated” and “Scenario B″, indicate a crucial and Field Data Interpretation. Hassan Dehghanpour: Supervision, Con-
role of the osmosis in the counter-current imbibition process. ceptualization, Writing- Reviewing and Editing
• From “Scenario C″, we observed that, the oil RF is two times higher
when the core plugs are soaked in tap water compared with the case Declaration of competing interest
when they are soaked with the synthetic brine, confirming the
significant role of the osmotic pressure. The authors declare that they have no known competing financial

PR
• Using microemulsion solution can decrease the IFT of oil- interests or personal relationships that could have appeared to influ-
microemulsion significantly up to 15 times compared to IFT ence the work reported in this paper.
between oil-tap water or oil-synthetic brine. Moreover,
microemulsion solution can alter the rock wettability from oil-wet Acknowledgment
to more water-wet.
This research was supported by Flotek Chemistry, Natural Sciences
7.3. Capillary and osmotic pressure and Engineering Research Council (NSERC) of Canada, and the Univer-
sity of Alberta, Future Energy Systems Program under Project No. FES

D
At the end of this study, we calculated the Pc and the osmotic pres- T07-P05. The initial results of this study were presented at the SPE Hy-
sure, and we concluded that: draulic Fracturing Technology Conference and Exhibition, The Wood-
lands, Texas, USA, 2019, SPE-194363-MS.
TE
• Osmotic pressure can be 50 times higher than the Pc.

Appendix A. Reservoir Geological map


EC
RR
CO

Fig. A.1. Geological map of the reservoir. The reservoir is consisted of an upper zone (Zone 1) and a lower zone (Zone 2). Although the permeability of Zone 1 is
higher than the other zone, but it contains immobile oil and mobile water only. However, both oil and water are mobile in Zone 2, which makes it as the target
zone for well completion.

Appendix B. Post-flowback production (surface condition)


UN

Here we show the post-flowback production data for Well 1 under surface conditions.

11
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
Fig. B.1. Post-flowback production data under surface condition. (a) Flow rate and Bottom-hole pressure. (b) Cumulative production.

PR
D
TE
EC

Fig. B.2. Profile of rate ratio (oil-water ratio, gas-oil ratio, and water cut).
RR
CO
UN

12
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

◀ Fig. B.3. Slopes before and after the shut-in period from decline-curve analysis: (a) Oil, (b) gas, and (c) water.

F
OO
PR
Fig. B.4. Estimated ultimate recovery from decline curve analysis for oil, gas and water before & after shut-in.

Appendix C. Well logs and core plugs locations

Track 1 shows the gamma ray log of the wells. The log shape is uniform and slightly deflected to the left for all wells. This indicates that the lithol-
ogy of the core plugs is shaly siltstone with low concentration of shale. The resistivity log on the second track deflects to the right side, which indi-

D
cates presence of hydrocarbon. The neutron porosity and density-porosity logs on the third track overlap, with slight deflection for neutron porosity
log to the left, which indicates presence of oil in the intervals of the core plugs. Based on neutron and density-porosity results, well 1 has higher aver-
age porosity than well 2, and this is consistent with laboratory results in Table 1.
TE
EC
RR
CO
UN

13
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
PR
D
TE
EC
RR
CO

Fig. D.1. Gamma ray log, resistivity log, neutron porosity with density-porosity logs, and location of the core plugs used in this study for: (a) Well 1 and (b) Well 2.

Appendix D. MICP and pore-throat size distribution

Figs. D.1 and D.2 show the Mercury injection capillary pressure (MICP) measurement and pore size distribution obtained from MICP data, re-
spectively. Fig. D.1 shows displacement pressures required for mercury to invade the pores filled with air. The trend of capillary pressure versus
UN

mercury saturation is similar for all the core plugs indicating the uniform pore-throat size distribution. The pore size distribution in Fig. D.2
demonstrates that most of the pore throat sizes in all the core plugs, are micro-pores, with median pore throat size (the median pore throat size is
described as the pore throat diameter at 50% of mercury injection) ranges between 22 and 193 nm.

14
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
PR
Fig. D.1. MICP curve for the different core plugs from wells 1, 2, and 3.

D
TE
EC
RR
CO
UN

15
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
PR
D
TE
EC
RR
CO

Fig. D.2. Pore throatsize distribution measured during air drainage process in MICP for core plugs: (a) MTY 1-1. (b) MTY 1–2. (c) MTY 1–3. (d) MTY 1–4. (e) MTY 2-
1. (f) MTY 2-2.

Appendix E. SEM images, permeability, and porosity


UN

Fig. E.1 and E.2 show the SEM images for MT 2-1 and MT 2-2, respectively. These images show pore structure of the scanning area along with
its mineralogy by energy dispersive x-ray spectroscopy (EDS) analysis. Both core plugs are strongly cemented with quartz and dolomite, which min-
imize pore sizes. Most of the pores are intergranular blocked by cements or coated with clay minerals. Illite and mica are the abundant clays in both
core plugs, where illite with its ribbon-like shape can fill some of the pores and reduces the permeability.

16
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
PR
D
TE
Fig. E.1. The SEM images for core plug MT 2-1, shows that the framework grains for this dolomitic siltstone are mainly quartz with minor amounts of albite and
EC

feldspar (Fd in the figures). In total, the texture of this rock is granular to microcrystalline. Some of the pores are found in between pyrites (py) crystals. Illite and
mica are the presented clays in this rock. The dominant pores are intergranular which surrounded with cements such as dolomite. Some of the pores are found in be-
tween illite minerals, where illite coats part of these pores.
RR
CO
UN

17
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

F
OO
PR
D
TE
EC

Fig. E.2. The lithotype of MT 2-2 is dolomitic siltstone with dominant quartz grains. The mixed composition of framework grains with cement, made the texture of
this rock granular and microcrystalline. Illite is abundant in this rock and it is coats the grains and fills the pores. Rare amounts of chlorite and mica plates are ob-
served in this rock sample. Dolomite is the most abundant cement with minor amount of quartz in form of silica cement (red arrows in figure E2.d) with amounts of
barite (ba in the figures). The intergranular pores (yellow arrows in figure E2.b) are the most common pore type which less than 10 μm in size and surrounded with
cements and clay.

Uncited References
RR

;;;;;;.

References Eghbalvala, M., Habibi, A., Dehghanpour, H., 2020. A laboratory protocol for evaluating
microemulsions for enhanced oil recovery while fracturing. In: Paper Presented at the
CO

Society of Petroleum Engineers - SPE Canada Unconventional Resources Conference


Al-Marhoun, M.A., 1985. Pressure-volume-temperature Correlations for Saudi Crude Oils.
2020, URCC 2020.
Univ. of Petroleum and Minerals.
Fakcharoenphol, P., Kurtoglu, B., Kazemi, H., Charoenwongsa, S., Wu, Y.S., 2014, April.
Al-Shammasi, A.A., 1999, January 1. Bubble Point Pressure and Oil Formation Volume
The effect of osmotic pressure on improve oil recovery from fractured shale formations.
Factor Correlations. Society of Petroleum Engineers. https://doi.org/10.2118/53185-
In: SPE Unconventional Resources Conference. Society of Petroleum Engineers.
MS.
Franz, Jr., J.H., Jochen, V., 2005. When Your Gas Reservoir Is Unconventional, SO IS OUR
Alvarez, J.O., Schechter, D.S., 2017a. Wettability Alteration and Spontaneous Imbibition
Solution, Shale Gas White Paper. Schlumberger, Houston, TX.
in Unconventional Liquid Reservoirs by Surfactant Additives. Society of Petroleum
Fritz, S.J., 1986. Ideality of clay membranes in osmotic processes: a review. Clay Clay
Engineers February 1 https://doi.org/10.2118/177057-PA.
Miner. 34 (2), 214–223.
Alvarez, J.O., Schechter, D.S., 2017b. Improving oil recovery in the Wolfcamp
UN

Gant, P.L., Anderson, W.G., 1988, March 1. Core Cleaning for Restoration of Native
unconventional liquid reservoir using surfactants in completion fluids. J. Petrol. Sci.
Wettability. Society of Petroleum Engineers. https://doi.org/10.2118/14875-PA.
Eng. 157, 806–815.
Guo, B., Ghalambor, A., 2005. In: Natural Gas Engineering Handbook, vol. 22.
Austad, T., Milter, J., 1997, January 1. Spontaneous Imbibition of Water into Low
Habibi, A., Dehghanpour, H., Binazadeh, M., Bryan, D., Uswak, G., 2016, October 1.
Permeable Chalk at Different Wettabilities Using Surfactants. Society of Petroleum
Advances in Understanding Wettability of Tight Oil Formations: A Montney Case Study.
Engineers. https://doi.org/10.2118/37236-MS.
Society of Petroleum Engineers. https://doi.org/10.2118/175157-PA.
Austad, T., Rezaeidoust, A., Puntervold, T., 2010, January 1. Chemical Mechanism of Low
Habibi, A., Esparza, Y., Boluk, Y., Dehghanpour, H., 2020. Enhancing imbibition oil
Salinity Water Flooding in Sandstone Reservoirs. Society of Petroleum Engineers.
recovery from tight rocks by mixing nonionic surfactants. Energy Fuel. 34 (10),
https://doi.org/10.2118/129767-MS.
12301–12313 2020 https://doi.org/10.1021/acs.energyfuels.0c02160.
Beggs, D.H.U., Brill, J.P.U., 1973. A study of two-phase flow in inclined pipes. J. Petrol.
He, K., Xu, L., 2017. Unique mixtures of anionic/cationic surfactants: a new approach to
Technol. 25, 607–617.
enhance surfactant performance in liquids-rich shale reservoirs. In: SPE International
Cath, T.Y., Childress, A.E., Elimelech, M., 2006. Forward osmosis: principles, applications,
Conference on Oilfield Chemistry. Society of Petroleum Engineers: Montgomery, Texas,
and recent developments. J. Membr. Sci. 281 (1–2), 70–87.
USA.
Clarkson, C.R., Pedersen, P.K., 2011, January 1. Production Analysis of Western Canadian
Hensen, E.J., Smit, B., 2002. Why clays swell. J. Phys. Chem. B 106 (49), 12664–12667.
Unconventional Light Oil Plays. Society of Petroleum Engineers. https://doi.org/10.
Kemper, W.D., Rollins, J.B., 1966. Osmotic efficiency coefficients across compacted clays
2118/149005-MS.
1. Soil Sci. Soc. Am. J. 30 (5), 529–534.

18
T. Soleiman Asl et al. Journal of Petroleum Science and Engineering xxx (xxxx) 109318

Kewen, L., Abbas, F., 2000, April 1. Experimental Study of Wettability Alteration to Saarenketo, T., 1998. Electrical properties of water in clay and silty soils. J. Appl.
Preferential Gas-Wetting in Porous Media and its Effects. Society of Petroleum Geophys. 40 (1–3), 73–88.
Engineers. https://doi.org/10.2118/62515-PA. Schmid, K.S., Gross, J., Helmig, R., 2014. Chemical osmosis in two-phase flow and
Kooi, H., Garavito, A.M., Bader, S., 2003. Numerical modelling of chemical osmosis and salinity-dependent capillary pressures in rocks with microporosity. Water Resour. Res.
ultrafiltrationacross clay formations. J. Geochem. Explor. 78, 333–336. 50 (2), 763–789.
Kuhn, H., Försterling, H.D., Waldeck, D.H., 2009. Principles of Physical Chemistry. John Sharma, T., Suresh Kumar, G., Sangwai, J.S., 2014. Enhanced oil recovery using oil-in-
Wiley & Sons. water (o/w) emulsion stabilized by nanoparticle, surfactant and polymer in the presence
Lan, Q., Dehghanpour, H., Wood, J., Sanei, H., 2015, August 1. Wettability of the Montney of NaCl. Geosyst. Sng. 17 (3), 195–205.
Tight Gas Formation. Society of Petroleum Engineers. https://doi.org/10.2118/171620- Soleiman Asl, T., Habibi, A., Ezulike, O.D., Eghbalvala, M., Dehghanpour, H., 2018. The

F
PA. role of microemulsion and shut-in on well performance: from field scale to laboratory
Laplace, P.S., 1805. In: Traité de mécanique celeste, vol. 4. Supplements au Livre X. scale. In: Paper Presented at the Society of Petroleum Engineers - SPE Hydraulic
Li, X., Teklu, T.W., Abass, H., Cui, Q., 2016, August. The impact of water salinity/ Fracturing Technology Conference and Exhibition 2019, HFTC 2019.

OO
surfactant on spontaneous imbibition through capillarity and osmosis for Soleiman Asl, T., Moussa, T., Ezulike, O., Dehghanpour, H., 2020. An integrated field and
unconventional IOR. In: Unconventional Resources Technology Conference, San simulation study to understand the role of shut-in on a montney well performance. In:
Antonio, Texas, 1-3 August 2016 (Pp. 3058-3076). Society of Exploration Geophysicists, Paper Presented at the Society of Petroleum Engineers - SPE Canada Unconventional
American Association of Petroleum Geologists, Society of Petroleum Engineers. Resources Conference 2020, URCC 2020.
Liu, J., Sheng, J.J., Huang, W., 2019. Experimental investigation of microscopic Teklu, T.W., Li, X., Zhou, Z., Alharthy, N., Wang, L., Abass, H., 2018. Low-salinity water
mechanisms of surfactant-enhanced spontaneous imbibition in shale cores. Energy Fuel. and surfactants for hydraulic fracturing and EOR of shales. J. Petrol. Sci. Eng. 162,
33 (8), 7188–7199. https://doi.org/10.1021/acs.energyfuels.9b01324. 367–377. https://doi.org/10.1016/j.petrol.2017.12.057.
Malusis, M.A., Shackelford, C.D., Olsen, H.W., 2003. Flow and transport through clay Vledder, P., Gonzalez, I.E., Carrera Fonseca, J.C., Wells, T., Ligthelm, D.J., 2010. January
membrane barriers. Eng. Geol. 70 (3–4), 235–248. Low Salinity Water Flooding: Proof of Wettability Alteration on A Field Wide Scale.
Marbach, S., Bocquet, L., 2019. Osmosis, from molecular insights to large-scale Society of Petroleum Engineers. https://doi.org/10.2118/129564-MS.

PR
applications. Chem. Soc. Rev. 48 (11), 3102–3144. Vonnegut, B., 1942. Rotating bubble method for the determination of surface and
National Energy Board, 2011. Tight Oil Developments in the Western Canada interfacial tensions. Rev. Sci. Instrum. 13 (1), 6–9.
Sedimentary Basin. National Energy Board, Calgary 2011. Canadian Electronic Library/ Wijaya, N., Sheng, J.J., 2020. Surfactant selection criteria with flowback efficiency and oil
desLibris. Absolute Page 1. Downloaded 01-06-2021. recovery considered. J. Petrol. Sci. Eng. 192. https://doi.org/10.1016/j.petrol.2020.
Olsen, H.W., Yearsley, E.N., Nelson, K.R., 1990. Chemico-osmosis versus Diffusion- 107305.
Osmosis. Transportation Research Record, (1288). Xu, M., Dehghanpour, H., 2014. Advances in understanding wettability of gas shales.
Padin, A., Torcuk, M.A., Katsuki, D., Kazemi, H., Tutuncu, A.N., 2016. Experimental and Energy Fuel. 28 (7), 4362–4375.
theoretical study of water-solute transport in organic-rich carbonate mudrocks. In: Paper Yarveicy, H., Habibi, A., Pegov, S., Zolfaghari, A., Dehghanpour, H., 2018, March 13.
Presented at the Proceedings - SPE Annual Technical Conference and Exhibition, , 2016- Enhancing Oil Recovery by Adding Surfactants in Fracturing Water: A Montney Case
January. https://doi.org/10.2118/181585-ms. Study. Society of Petroleum Engineers. https://doi.org/10.2118/189829-M.

D
Pollen, E.N., Berg, C.F., 2018, November 12. Experimental Investigation of Osmosis as a Yassin, M.R., Dehghanpour, H., Begum, M., Dunn, L., 2018. Evaluation of imbibition oil
Mechanism for Low-Salinity EOR. Society of Petroleum Engineers. https://doi.org/10. recovery in the duvernay formation. SPE Reservoir Eval. Eng. 21 (2), 257–272.
2118/192753-MS. Young, T., 1805. III. An essay on the cohesion of fluids. Phil. Trans. Roy. Soc. Lond. 95,
Rao, D.N., Ayirala, S.C., Abe, A.A., Xu, W., 2006, January 1. Impact of Low-Cost Dilute 65–87.
Surfactants on Wettability and Relative Permeability. Society of Petroleum Engineers. Yuan, L., Dehghanpour, H., Ceccanese, A., 2019, April. Imbibition oil recovery from the
TE
https://doi.org/10.2118/99609-MS. montney core plugs: the interplay of wettability, osmotic potential and microemulsion
Reynolds, Bachman, R.C., Buendia, J., Peters, W., 2015. The full montney - a critical effects. In: SPE Western Regional Meeting. Society of Petroleum Engineers.
review of well performance by production analysis of over 2,000 montney multi-stage Yuan, L., Zhang, Y., Dehghanpour, H., 2021. A theoretical explanation for wettability
fractured horizontal gas wells. In: SPE/CSUR Unconventional Resources Conference, alteration by adding nanoparticles in oil-water-tight rock systems. Energy Fuel. 35 (9),
Society of Petroleum Engineers, Calgary, Alberta, Canada. 7787–7798.
Rokosh, C.D., Lyster, S., Anderson, S.D.A., Beaton, A.P., Berhane, H., Brazzoni, T., Zhou, Z., Abass, H., Li, X., Bearinger, D., Frank, W., 2016. Mechanisms of imbibition
Pawlowicz, J.G., 2012. Summary of alberta’s shale-and siltstone-hosted hydrocarbon during hydraulic fracturing in shale formations. J. Petrol. Sci. Eng. 141, 125–132.
EC

resource potential. Energy Rers. Conserv. Board ERCB\AGS Open File Report 6, 327.
Rosen, M.J., Wang, H., Shen, P., Zhu, Y., 2005. Ultralow interfacial tension for enhanced
oil recovery at very low surfactant concentrations. Langmuir 21 (9), 3749–3756.
RR
CO
UN

19

You might also like