Next Article in Journal
An Innovative Approach on Directed Energy Deposition Optimization: A Study of the Process Environment’s Influence on the Quality of Ti-6Al-4V Samples
Previous Article in Journal
Effect of Ultraviolet Irradiation on Osseointegration of Dental Implants: A Comparative Histomorphometric Study in Canine Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Using Low Salinity Waterflooding to Improve Oil Recovery in Naturally Fractured Reservoirs

by
Faisal Awad Aljuboori
1,
Jang Hyun Lee
1,*,
Khaled A. Elraies
1 and
Karl D. Stephen
2
1
Petroleum Engineering Department, Universiti Teknologi PETRONAS, Seri Iskandar, Perak Darul Ridzuan 32610, Malaysia
2
Institute of GeoEnergy Engineering, Heriot-Watt University, Edinburgh EH14 4AS, UK
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(12), 4211; https://doi.org/10.3390/app10124211
Submission received: 25 February 2020 / Revised: 27 March 2020 / Accepted: 27 March 2020 / Published: 19 June 2020
(This article belongs to the Section Energy Science and Technology)

Abstract

:
Low salinity waterflooding is an effective technique to accelerate and boost oil recovery. The impact of this technique has been investigated widely in laboratories for various scales and rock typing, most of which have demonstrated a potential improvement in oil recovery. This improvement has been attributed to several chemical and physical interactions that led to a change in the wettability to become more water-wet, as well as a reduction in the residual oil saturation. Meanwhile, it is rare to find a discussion in the literature about the efficiency of low salinity flooding in naturally fractured reservoirs. Therefore, in this work, we investigate the potential advantages of this method in fractured reservoirs using numerical simulations. A new approach to estimate the weighting factor using a tracer model has been proposed to determine the brine salinity and, hence, its properties in the mixing region. We have also used the relative permeability curves as a proxy for any physical and chemical mechanisms which are not represented explicitly in the model. The simulation outcomes highlighted the advantage of low salinity waterflooding in fractured reservoirs. An increment in oil recovery by 10.7% to 13% of Stock Tank Oil Initially In Place (STOIIP) was obtained using the dual- and single-porosity model, respectively. Therefore, the low salinity waterflooding technique represents a promising low-cost, effective method in fractured reservoirs.

1. Introduction

Carbonate reservoirs in the Middle East hold a significant proportion of remaining conventional oil reserves, most of which are classified as naturally fractured [1,2]. These reservoirs are well known for their low recoveries compared with their counterpart clastic unfractured reservoirs [3]. Most of these reservoirs have already produced for several decades using natural depletion mechanisms or undergone secondary conventional waterflooding projects. However, developing fractured reservoirs requires an affordable technique that is able to sweep the oil and improve recoveries, as well as controlling the early breakthrough of water that is often observed in conventional waterflooding. It is often the case that enhanced oil recovery techniques are not feasible due to the high cost of such projects.
Low salinity waterflooding is a promising low-cost technique to increase oil recovery. Experimental outcomes have proven a significant improvement in oil recovery when low salinity waterflooding schemes were applied [4]. During low salinity waterflooding, multiple recovery mechanisms have been attributed to be responsible for the increment in oil recoveries, such as cation exchange [5,6], an increase in the pH [7,8,9], changes of the charges at the rock surfaces [10], wettability alteration [11,12,13,14], fine migration [14], a reduction in the interfacial tension [15], a micro-dispersion [16], and the osmosis [17]. Therefore, the increment in oil recovery may be due to one or more of the mentioned recovery mechanisms.
Various impacts of low salinity corefloods were observed on both sandstone and carbonate samples, such as wettability alteration, a reduction in residual oil saturation, and interfacial tension (e.g., [18,19,20,21,22]). This impact has been attributed to the chemical interactions between the injected low salinity brine, crude oil, and rock surfaces. McGuire et al. have referred to the in-situ forming of surfactants due to the saponification reaction between the elevated pH of the low salinity brine and the acid or polar components in the crude oil [23]. Acidity increment (i.e., pH) occurred from the reaction between the low salinity water and natural minerals of the reservoir, which generated hydroxyl ions and increased the pH up to 10 or more. As a result, the system wettability was shifted to a more water-wet state. In addition, a reduction in both residual oil saturation and interfacial tension occurred.
Alteration to the system wettability has been reported at ambient temperature [18]. A laboratory experiment was carried out under these conditions to measure the contact angle of a sessile oil droplet, as well as the interfacial tension and the potential at various temperatures, using limestone and dolomite samples. The outcome was an apparent alteration of the wettability towards a more water-wet system. Furthermore, the chemistry of the crude oil, the rock type, and particularly clay minerals are among the factors that control the system’s wettability and residual oil saturation [19]. These factors also influence how wettability and residual oil may change. When present, heavier asphaltene components tend to stain the rock surfaces and make the system strongly oil-wet. Therefore, using the low salinity scheme in an asphaltenic oil system can lead to less oil recovery compared with a low asphaltenic oil system. It has been reported that, in this case, the wettability does not change significantly with the use of low salinity methods [19].
During low salinity water injection, the high salinity formation water tends to form as a waterfront due to its high mobility, followed by the low salinity injected brine [24]. A mixing salinity region (i.e., dilution region) often forms between the formation water and the injected brine. The dilution region is essential for the chemical interaction between the low salinity brine and crude oil/rock to take place. Moreover, there is a range of salinity in which the wettability, fluid mobility, and residual oil saturation can change [25,26]. This is called the effective salinity range [24,27]. It is also observed that there is occasionally a change to the maximum relative permeability of water. These changes resulted in an increment in oil mobility and, hence, improve oil recovery. In the simulation, two sets of relative permeability should be used to represent the regions of high salinity and low salinity brine, respectively. A weighting factor is used to calculate relative permeability in the mixing zone [24].
In this paper, we presented an approach to simulate the impact of low salinity waterflooding on reservoir behaviour. These steps include:
  • A new weighting factor has been proposed based on a 1D tracer model.
  • The results of reported coreflood experiments were utilised to generate the relative permeability sets of the flooded regions by low salinity brine.
  • Various scales were employed to demonstrate the impact of the low salinity waterflooding in fractured reservoirs as well as to highlight the influence of the dual-porosity model on the simulation outcome.
Initially, we collected and evaluated the published results of coreflood experiments to conclude a range of change to the system wettability (in the form of saturation exponents) and the residual oil saturation (e.g., the published results of [20,21,23,25,26]). Then, we proposed a new approach to determine the weighting factor using a tracer model, which can improve the simulation accuracy in the mixing zone. On the other hand, it is challenging to model the chemical interactions that have been proposed to occur during low salinity water flooding at sector or field scales accurately. Therefore, we utilised the relative permeability curves as a proxy for any chemical and physical interactions that have not been represented explicitly in the simulation model. The generated sets of relative permeability enable us to emulate the observed behaviour of the low salinity coreflood in the laboratory using various modelling scales. Besides, the impact of the change to the relative permeability on the fractional flow displacement in naturally fractured reservoirs has been presented to illustrate how the alteration process to the flow curves can delay the water flow; and, hence, reduce the water cut as well as boost oil recovery.

2. Methodology

The purpose of this work is to emulate the impact of the low salinity waterflooding process in naturally fractured reservoirs using numerical simulation. We have approached this objective by using two steps. The first is collecting the coreflood experimental results and estimating the change in the relative permeability curves due to low salinity brine injection using a similar approach to our previous work [2]. The alteration to the relative permeability has been captured through saturation exponents of oil and water phases, as well as by the saturation endpoints. Then, Corey formulae for the oil–water system were employed to calculate new flow curves to represent the region flooded by the low salinity brine.
Meanwhile, in the second step, an appropriate weighting factor formula is proposed to calculate the average salinity of the brine mixture in the mixing region accurately. A precise weighting factor is essential because it controls the salinity of the brine mixture, which, in turn, determines the brine properties such as density, viscosity, compressibility, and formation volume factor. Besides, the weighting factor determines the mobility of the brine mixture. In this part, a normalised brine salinity has been used to obtain a range between zero and one with a proper increment step, which enables capturing of the soaring of the salinity mixture in the mixing zone. Then, the weighting factor and brine properties calculated accordingly using the provided equations, as illustrated in Table 1.
In the later work, both water and brine will be used interchangeably for the injected water or formation water, and they have the same meaning and the term low salinity has been shortened in the figures to LoSal for simplicity. However, the above-explained methodology has demonstrated the following points:

2.1. The Relative Permeability of the Flooded Regions by Low Salinity Brine

The results of the published laboratory experiments have been collected and analysed to calculate a percentage of change in the parameters of the relative permeability. The estimated rate of change in each variable due to the impact of the low salinity brine is summarised in Table 2. The listed percentages of change were used to generate three sets of the relative permeability to address the uncertainty in the laboratory results using Corey Equations (1) and (2). The calculated flow curves of the regions flooded by the low salinity brine exhibited a shift to the right (i.e., towards a more water-wet state), as illustrated in Figure 1.
k r o _ a l t e r e d = ( k r o ) S w c   [ 1 S w S o r _ a 1 S w c S o r _ a ] n o _ a
k r w _ a l t e r e d = ( k r w _ a ) S o r   [ S w S w c 1 S w c S o r _ a ] n w _ a
where:
kro_altered, krw_altered
Relative permeability of oil and water, respectively, for the region flooded by low salinity.
kro at Swc
Oil relative permeability at connate water saturation.
krw_a at Sor
Water relative permeability at residual oil saturation after low salinity waterflooding.
Sw, Swc
Water saturation and connate water saturation.
Sor_a 
Residual oil saturation of the flooded region by low salinity.
no_a, nw_a
Saturation exponents of oil and water.
The units of the above parameters are fraction except for the saturation exponents which are dimensionless.

2.2. Determination of Weighting Factor (WF)

The weighting factor plays a significant role in determining the salinity of the brine mixture, hence, its properties, as illustrated in Table 1. It controls the dilution process or blending between the high salinity and the injected low salinity water in the mixing zone. Al-Ibadi [24,27] has referred to a linear relationship of the weighting factor, as shown in Equation (3):
W F = 1 ( S C m i x S C L o S a l S C m a x   e f f S C L o S a l )
where:
WF
Weighting factor, fraction
SCmix
Brine mixture salinity, ppm.
SCLoSal
Injected brine salinity, ppm.
SCmaxeff
Maximum effective salinity, ppm.
This linear relationship might be valid when the differences between the minimum salinity and the maximum effective salinity ranges are not very high. However, in a significant salinity contrast, an exponential relationship is more appropriate to represent the salinity increment in the mixing zone. This contrast is applicable in the case of formation water in the petroleum reservoirs, where its salinity can be as high as 200,000 ppm or more (e.g., Nahr Umr reservoir in Iraq has a formation water salinity of 200,000 ppm [28]). In contrast, the salinity of the injected fluid is several thousands of ppm.
In this paper, we present an approach to estimate the weighting factor using a one-dimensional model. The model simulated a highly idealised fractional flow displacement process of Buckley–Leverett. The dimensions of the model were (1000 ft × 25 ft × 50 ft) with 5 ft resolution in the x-direction. A relative permeability set and capillary pressure data of a fractured formation were used along with a proper fluid model to initialise the simulation model. The model represents an immiscible oil–water system and comprises an injector and a producer at the grid ends. A steady-state displacement was designed using controlled surface rates for both production and injection.
A conventional water injection scenario was simulated using a tracer concentration of 3000 for the injected water and a tracer of 100,000 for the formation water. Meanwhile, both injected water and formation water have the same high salinity. The results of the tracer model are illustrated through the profile of tracer concentration in Figure 2A and the profile of water saturation in Figure 2B. The mixing region of the tracer concentration during water injection occurred in a limited interval in the mixing zone. Meanwhile, high and low tracer concentrations were dominant at both sides of the mixing interval within the mixing zone region.
A significant width of the tracers’ mixing zone was indicated by the simulation result, as illustrated in Figure 2. The width of the mixing zone depends on the petrophysical properties of the medium, as well as on the flow properties of the existing fluids. Moreover, physical phenomena can also influence the width of the mixing region, such as the dispersion effect that tends to spread the injected waterfront out. Therefore, a normalised distance was used to represent the change in the tracer concentration, as illustrated in Figure 3.
In Figure 3, the tracer concentration curve contained two dominant intervals (Region 1 and Region 3) of approximately 23% of the normalised distance for each region of low and high tracer concentration. The change in the tracer concentration in these regions is minimal. However, a significant change occurs between 23% and 77% in the dimensionless distance, as the tracer concentration exponentially raises in Region 2. Therefore, regardless of the interval width of the mixing zone, the dilution of the solute tracer concentration has the illustrated shape in Figure 3. Likewise, the mixing zone during low salinity brine injection is expected to have the same pattern. Therefore, the obtained mixing curve pattern of the tracers can be utilised to conclude a weighting factor formula.
The estimated weighting factor formula, which was based on the tracer simulation results, has employed the normalised salinity (NS) to obtain a general form, as shown in Equation (4). Meanwhile, the normalised brine salinity can be calculated using Equation (5). A comparison between the linear weighting factor normally used and the current work of the exponential weighting factor has been illustrated in Figure 4. The linear approach uses Equation (3), while the proposed weighting factor formula is given in Equation (4). The outcome of the simulated tracer model was validated using the analytical solution of the fractional flow, as explained in the Appendix A.
W F = a b × e c N S d + g × e h N S d k × e N S n d
The values of the parameters are listed below;
  • a = 0.59; b = 0.4276; c = 0.9807; d = 0.2351
  • g = 1.046; h = 0.08132; k = 0.2126; n = 3.887
N o r m a l i s e d   S a l i n i t y   ( N S ) = S C m i x S C L o S a l S C M a x e f f S C L o S a l

2.3. Relative Permeability and Capillary Pressure of the Mixing Zone

At the current stage of work, two sets of relative permeability were defined to represent the region of high salinity formation water and the region flooded by low salinity injected brine, as illustrated in Region 3 and Region 1 in Figure 3, respectively. Moreover, the proposed weighting factor formula was used to determine the flow curves in the mixing zone, which controls the brine mixture properties and fluid mobility. Therefore, estimating the relative permeability curves in the mixing region, as illustrated by Region 2 in Figure 3, required the two previously discussed relative permeability sets, using the following Equations [29,30,31]:
k r o _ M i x i n g   Z o n e = W F × k r o L o S a l + ( 1 W F ) × k r o H i g h S a l
k r w _ M i x i n g   Z o n e = W F × k r w L o S a l + ( 1 W F ) × k r w H i g h S a l
Similarly, the capillary pressure in the mixing region (Pcow_Mixing zone) can be computed using the capillary pressure curves of the low salinity waterflooding region (PcowLoSal) and the high salinity region (PcowHighSal) of the formation water, as explained in Equation (8)
P c o w _ M i x i n g   Z o n e = W F P c × P c o w L o S a l + ( 1 W F P c ) × P c o w H i g h S a l

3. Setup of Numerical Models

Fine-scale modelling has been used widely to investigate various problems of recovery mechanisms and flow behaviour in both conventional and fractured reservoirs (e.g., [1,32,33,34]). These models help to obtain an accurate result by avoiding the common simulation issues and artefacts such as upscaling problems and numerical convergence errors that frequently occurred at the reservoir scale modelling and simulation [3,35,36]. However, the detailed representation in the fine-scale model cannot be applied for the full field-scale due to the insufficient computational power resources and the simulation cost. Therefore, multiple scales were utilised in the current workflow to demonstrate the impact of low salinity waterflooding at each level, besides highlighting the upscaling challenges of the fine-scale model result into a larger scale, such as sector modelling or full-field modelling.
The impact of the low salinity waterflooding on the matrix recovery using the new proposed weighting factor formula has been evaluated using our previous fine-scale model and the Qamchuqa reservoir model. The details of both models can be found in Aljuboori et al. [2], where a linear relationship of the weighting factor has been used in their simulation. Meanwhile, we utilised the exponential weighting factor formula in the current workflow, and then we compared the results.

3.1. Single Matrix Block Models

Most often, the dual-porosity model is employed to simulate the fluid flow behaviour in naturally fractured reservoirs [1,33,37,38]. The model assumes two solution nodes in every gridblock with distinct properties that represent the matrix medium and fractures, respectively. The flow occurs between the gridblocks in the reservoir model through the fracture nodes only. Meanwhile, matrix nodes can solely exchange the fluid locally to the corresponding fracture node within the gridblock. Previous works have demonstrated that the dual-porosity model tends to overestimate the fluid exchange between the matrix and fractures and, hence, overestimate the matrix recovery (e.g., [1,39]). In contrast, fine-scale modelling can provide an accurate prediction of the fluid exchange rate between the matrix and fractures. Therefore, it is necessary to assess the impact of the low salinity waterflooding using both modelling scales.
A single matrix block with the dimensions 5 ft × 8 ft × 12 ft was used to construct two models to evaluate the effect of the low salinity waterflooding on the matrix recovery. The first model represents a discretised matrix scheme, with a 1 ft resolution in the x-, y-, and z-direction, respectively, with a total number of 480 cells in the simulation model. Meanwhile, the surrounding fractures have also been represented explicitly at the grid boundary in a discretised form, as illustrated in Figure 5A. The fine-scale model can be simulated using a single-porosity model. In the second model, the matrix and fractures were modelled in a dual-porosity scheme with two cells only, one for the matrix and the second for fractures, as shown in Figure 5B. The second model can be simulated using the dual-porosity model.
The single block model in Figure 5B represents the smaller element in the dual-porosity model. The dimensions of the matrix blocks are determined through fracture intensities, which determine their sizes, hence, oil recovery [40,41]. This model depicts the fluid flow behaviour using the dual-porosity model, which can be linked to the intermediate scale model or a full-field scale model. On the other hand, the fine-scale model can mimic the recovery mechanisms more precisely. It has the ability to capture the transient effect and saturation front, which the dual-porosity model has often failed to do.
In conventional waterflooding, it has been frequently observed that the injected water breaks through in the producers due to the high permeability of fractures leaving a significant amount of oil behind in the matrix cells. Similarly, the injected low salinity brine sweeps the oil from fractures much faster than sweeping the oil from the matrix cells. Then, a fluid exchange process occurs due to capillary imbibition, gravity drainage, and viscous forces. These conditions were mimicked by assuming fractures to be fully saturated with the injected low salinity brine. Meanwhile, the matrix is saturated with oil at connate formation water saturation. A straight-line relative permeability was assigned to fractures, and a relative permeability set of a fractured formation, as illustrated in Figure 1, has been used for the matrix. A prediction of 10 years was carried out to obtain the fluid exchange between the matrix and fractures under a low salinity waterflooding scheme. Further model input data and details can be found in Aljuboori et al. [2].

3.2. Intermediate Scale Modelling

An intermediate scale model was constructed to simulate the low salinity waterflooding using the same properties of the fine-scale model. The dimensions of the model are 200 ft × 200 ft × 60 ft with a gridblock resolution of 40 ft × 40 ft × 12 ft, and a total number of 250 cells for both matrix and fractures. Moreover, average matrix properties were used to identify the impact of low salinity injection water on oil recovery without heterogeneity effect. Appropriate fluid properties and relative permeability sets of Upper Qamchuqa formation were used to initialise the simulation model. The matrix block dimension in this model is the same as the dimension of the fine-scale model of the previous section (i.e., 5 ft × 8 ft × 12 ft) to allow for performance comparison.
Two recovery cases were simulated to evaluate the matrix recovery. The first model was designed to mimic the water advancement in fractures due to their high permeability leaving the matrix oil trapped behind, as shown in Figure 6A. In this case, it has been assumed that fractures were flooded with low salinity brine, while the matrix saturated with oil and connate formation water. Then, a fluid exchange process occurs under static conditions between the matrix and fractures due to capillary, gravity, and viscous forces. In the second case, the model comprises an injector and a producer located at the opposite corners of the grid, as illustrated in Figure 6B. In this case, fractures were assumed to be fully saturated with oil, while the matrix blocks saturated with oil and connate water. The low salinity brine was injected at a variable rate to maintain a constant bottom hole pressure, and the producer was opened to flow from the first layer only. The same prediction scenario of 10 years of the fine-scale model was adapted for simulating the above-discussed cases. The intermediate scale models have two objectives: (1) validate the simulation outcome of the single block model by comparing their results. (2) illustrate the differences between the injection scenario and the flooded scenario of the low salinity brine, which cannot be performed using the illustrated single block model in Figure 5B.

3.3. Field Scale Modelling—The Qamchuqa Reservoir

The Qamchuqa reservoir is one of the producing reservoirs of the Jambur field, which is located in the north of Iraq. The reservoir consists of two thick-bedded limestone formations with an average thickness of approximately 600 m in total; further geological description of the Qamchuqa reservoir can be found in Aljuboori et al. [1]. The history matching project has been carried out on the reservoir for four decades of production history, where consistent matching results were achieved for most of the producers [2]. Even recently, in 2020, the production of the Qamchuqa reservoir has continued using natural driving mechanisms, which resulted in a considerable pressure drop of (≈1200 psi) in the reservoir to the end of 2015 [2]. Therefore, the low salinity technique represents a competitive, affordable recovery method to support the reservoir pressure and to improve the sweep efficiency from the matrix medium as well as to sustain the production from the Qamchuqa reservoir. As highlighted previously, the same model of the Qamchuqa reservoir, which has been discussed in details in our previous work of Aljuboori et al. [2], has been used to assess the performance differences utilising linear relationship and exponential approach of the weighting factor. The same low salinity water injection scenario of 14 years was performed by injecting 45,000 STB/D through six injection wells. Furthermore, the salinity of the injected brine is 3000 ppm, where both injected brine salinity and the reservoir salinity represented as total dissolved solids (TDS) in the simulation model using ECLIPSE black oil simulator.

4. Results and Discussion

Before we delve into the results and discussions, it is essential to give a glimpse of the impact of low salinity brine injection on the fractional flow of the water phase and how this can increase the oil recovery. The generated relative permeability sets of the regions flooded by low salinity have been used to evaluate the fractional flow of water, as illustrated in Figure 7A. The calculated fractional flow profile demonstrated a systematic shifting towards the higher water saturation with preserving its shape during the change increment in wettability. This shift can improve the displacement of oil, which leads to a significant improvement in oil recovery and delay the water breakthrough remarkably. This example illustrates that the water breakthrough time has been slowed down considerably. At the maximum change in relative permeability, the water breakthrough has slowed down to 2613 days compared with 1777 days using the high salinity brine injection scenario (i.e., breakthrough time delayed by 47%), as indicated by the red and black colour in Figure 7B, respectively.

4.1. Single Matrix Block Model

An exponential weighting factor was used to simulate the low salinity waterflooding scenario, and the outcome was compared with the same model using a linear weighting factor. The results of the two weighting factor approaches were almost similar with a marginal difference of 0.1%, 0.2%, and 0.7% of the ultimate oil recovery for the minimum, average, and maximum alteration, respectively. However, the improvement in the oil recovery of the fine-scale model due to the impact of low salinity waterflooding is 2.3%, 7.5%, and 13% of STOIIP for the earlier mentioned alteration degree, respectively, as illustrated in Figure 8. These incremental percentages were calculated based on the result of the high salinity waterflooding scenario. In this fine-scale model, the single-porosity model was used in the simulation. Meanwhile, the same matrix block has been modelled using the dual-porosity model. The increment in oil recovery for the single block model using the dual-porosity option is 1.5%, 6.7%, and 10.7% of STOIIP for the minimum, average, and maximum cases, respectively, compared with the high salinity waterflooding scenario. The results of the above scenarios were compared to highlight the outcome differences, as illustrated in Figure 9 and summarised in Table 3.
The highlighted differences in the above table confirm the tendency of the dual-porosity model to overestimate the predicted oil recovery, the recovery speed, and the saturation profile, as concluded by [1,39]. Furthermore, the recovery speed and saturation front were significantly higher compared with the fine-scale modelling, as shown in Figure 9. The optimistic oil recovery of the dual-porosity model resulted from an exaggeration in the fluid exchange rate between the matrix and fractures. Moreover, the total height of the matrix block in the dual-porosity model (i.e., 12 ft) contributed to the oil exchange rate in the transfer rate calculations. Meanwhile, the height of a single cell in the fine-scale model (i.e., 1 ft based on the model resolution) has used in the fluid exchange calculations, which has a lower contribution. Moreover, the advancement of the low salinity waterfront took a longer time to sweep the same equivalent volume of the corresponding matrix block of the dual-porosity model. Figure 10 illustrates the saturation profile of the low salinity water phase in the fine-scale model and the single block model after 10 days. A single saturation value was computed for the matrix cell in the dual-porosity model, as illustrated in Figure 10A. Meanwhile, every cell in the fine-scale model show variations in saturation, see Figure 10B. Furthermore, the total exposed volume of the matrix for the fluid exchange with fractures are 480 ft3, and 300 ft3 for the dual-porosity model and the fine-scale model, respectively, which affected the matrix recovery in the early time. Therefore, the dual-porosity model exhibited a higher oil recovery due to the larger matrix volume exposed to the fluid exchange with fractures. The dominant force of the capillary pressure can be identified in Figure 10B, where the fluid transfer occurs in all directions of the matrix blocks.

4.2. Intermediate Scale Modelling

As stated earlier, these models were designed for comparison and validation purposes. The model simulated a low salinity waterflooding scenario. The results of this case were compared with the results of the single block model to validate the findings of the single block model against the intermediate scale model under the same condition using the dual-porosity solver. The simulation outcomes show almost identical oil recovery despite the differences in the scale of the two models, as shown in Figure 11.
On the other hand, the oil recoveries of both intermediate models (i.e., Case A and Case B in Figure 6) were compared to assess the matrix recovery under the sweep process by the injected low salinity brine scenario and the flooded scenario when the fluid exchange occurred under static condition. The three levels of uncertainty in the change to the relative permeability due to the impact of low salinity brine were simulated to evaluate oil recovery under the above-explained conditions. The simulation results indicate faster oil recovery in the flooded case compared with the injection case as all matrix blocks were in contact with the low salinity brine that filled fractures. Meanwhile, in the injection case, the matrix blocks were in contact with the oil in fractures. Then, the injected low salinity brine displace the oil from fractures and become in contact with additional matrix blocks as the low salinity waterfront advanced. The simulation results are illustrated in Figure 12.

4.3. Full Field Scale Modelling

A low salinity injection scenario has carried out in the north region of the Qamchuqa reservoir using six injection wells to sweep the oil towards the nearby producers, see Figure 13. The proposed locations of the injectors were designed to evaluate the impact of high salinity and low salinity brine injection scenarios on the performance of the Qamchuqa reservoir. Therefore, this injection design does not represent an optimal scenario of a secondary recovery project to develop the reservoir.
The dual-porosity model was used to simulate the effect of change in wettability on the performance of the Qamchuqa reservoir under a low salinity recovery scheme and included the exponential weighting factor. A secondary recovery mode of low salinity brine injection was carried out for 14 years. The uncertainty in the change to the Qamchuqa reservoir wettability has been addressed by using three levels of alteration to the relative permeability (minimum, average, and maximum).
Initially, the low salinity waterflooding scenarios gave a similar performance to the high salinity brine injection in the first six years of injection. Then, the change in wettability started to take effect and could be seen at the producers. This helped maintain the oil productivity for an extra two and a half years at a rate of 38,000 STB/D compared to an average rate of 32,500 STB/D for the high salinity waterflooding. Meanwhile, the oil flow rate maintained its high rate of 32,500 STB/D to the end of the prediction scenario of year 14 compared with 25,000 STB/D for the high salinity brine injection. At the same time, a reduction in the field water cut by 32.5% was also observed, as the injected low salinity brine exchanged with the oil in the matrix blocks. This process led to a reduced volume of injected water reaching the producer (i.e., less water cut) and increased oil production, as illustrated in Figure 14.
We can conclude from the above behaviour that the impact of the change in wettability required a longer time to sense its influence at the reservoir scale (six years were needed to show a significant effect on the Qamchuqa reservoir). Moreover, other impacts of the low salinity water injection on reservoir behaviour such as pressure maintenance. Pressure maintenance, in turn, has controlled the gas cusping in the producers; hence, preserving the drive mechanism energy of the dissolved gas or the gas cap in the Qamchuqa reservoir as well as less technical problems at the surface facilities due to excessive gas flow rate (e.g., field Gas-Oil Ratio (GOR) in Figure 15). That enabled the reduction of the workover cost, as most of the reported workover operations in the Qamchuqa reservoir were carried out due to the gas coning.
The advantage of low salinity brine injection is the ability to initiate a chemical reaction that changes reservoir wettability to be more water-wet. This reduces the mobility of the water phase in the flooded regions of the reservoir. Furthermore, this chemical interaction between the low salinity brine, crude oil, and the rock can also lead to a reduction in the residual oil saturation by displacing the oil from the small pore throats and rock surfaces towards the larger pores. As a result, the mobility of the oil phase increases in the flooded regions, which enhances the overall sweep efficiency compared with the conventional high salinity waterflooding. Moreover, the conventional water injection scenario often leads to an early water breakthrough.
Despite the positive simulation results of the Qamchuqa reservoir, an intensive evaluation program is necessary to determine the effective brine composition, and to assess the effect of the clay content, the composition of the crude oil, pH, and rock characteristics on the performance of the low salinity injection, as well as to reduce the uncertainty and to execute this technique effectively. Eventually, a significant improvement in the sweep efficiency and reservoir performance may require a longer time frame to observe its impact and to confirm the observed results of laboratory experiments.
It is worth mentioning that additional uncertainty due to the existence of bitumen in the Qamchuqa reservoir can impact the performance of the low salinity waterflooding significantly. Therefore, it is important to study the bitumen characteristics, its distribution, and how that can impact on the fluid flow in the Qamchuqa reservoir. The negative role of the bitumen might nullify the expected improvement in the oil recovery using a low salinity water injection scenario as the bitumen occasionally behaves like a flow barrier between the aquifer and the oil zone if it existed as a uniform layer. However, the bitumen may have less impact if it exists as scattered patches.

5. Conclusions

  • A new proposed approach of calculating the weighting factor in the mixing area has been presented to improve the simulation accuracy of the low salinity waterflooding. The weighting factor determines the salinity of the brine mixture, hence its properties, as well as it controls the fluid mobility through the calculated relative permeability in the mixing region. The impact of the exponential weighting factor might become significant in the small-scale modelling, which can be useful for modelling and simulation the laboratory coreflood experiments.
  • The results of the coreflood experiments have been integrated into the reservoir simulation to assess the improvement in the oil recovery due to the change in wettability to a more water-wet state. The current work highlighted the critical role of the laboratory results that should be performed carefully. Furthermore, the uncertainty in the reported coreflood results has been addressed by computing three levels of change to the relative permeability and simulate each level independently to evaluate their recoveries.
  • The remarkable delay in the water breakthrough can enhance the reservoir productivity and extend the production periods of the producers. However, accurate representation of the relative permeability in the mixing region between the high salinity set to the low salinity set is essential to avoid an overestimation of the sweep efficiency.
  • The relative permeability was successfully used as a proxy to simulate the physical and chemical interaction which have not been modelled explicitly in the simulation model during low salinity waterflooding. The relative permeability emulated the observed change in wettability and the reduction in the residual oil saturation, which enables to simulate the impact of the low salinity waterflooding flawlessly.
  • The study illustrated that there is an opportunity to increase oil recovery from the matrix using low salinity waterflooding even with a minor change in the relative permeability. This sensitive role to the generated relative permeability required accurate experimental results, which are essential for simulating the low salinity waterflooding scenario effectively and avoid any modelling or experimental artefact.
  • The accelerated recovery during low salinity waterflooding encourages a switch from the current conventional waterflooding projects to achieve a higher recovery within a reduced time frame. However, for a full field case, a longer time is required to observe the improvements as the low salinity brine required time to flood the reservoir rock and alter its wettability and reduce the residual oil saturation. Moreover, the fine-scale simulation demonstrated a precise estimation of oil recovery, and it should be used as a reference solution to calibrate the outcomes of the dual-porosity model and avoid any misestimation of the reservoir performance.

Author Contributions

Conceptualisation, F.A.A. and J.H.L.; methodology, K.A.E. and K.D.S.; software, F.A.A. and J.H.L.; validation, F.A.A., J.H.L., K.A.E. and K.D.S.; formal analysis, F.A.A.; investigation, F.A.A. and J.H.L.; permission acquisition of using field data, F.A.A.; writing—original draft preparation, F.A.A.; writing—review and editing F.A.A., K.D.S., J.H.L. and K.A.E.; visualisation, K.A.E. and K.D.S.; supervision, J.H.L., K.A.E. and K.D.S.; APC funding acquisition, J.H.L. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by Universiti Teknologi PETRONAS, YUTP grant (015LC0-109).

Acknowledgments

The authors would like to express their deepest gratitude to Universiti Teknologi PETRONAS for providing the required software license and creating the necessary working environment. The authors would also like to thank the Ministry of Oil of Iraq and the management of North Oil Company for their permission to use their data in this work.

Conflicts of Interest

The authors certify that they have no affiliations with or involvement in any organisation or entity with any financial interest.

Abbreviations

fwFractional flow of water, fraction
kroOil relative permeability, fraction
krwWater relative permeability, fraction
µoOil viscosity, cp
µwWater viscosity, cp
a, bCoeiffiecents
SwWater saturation, fraction
WiInjected water rate, bbl/D
tTime, day
DDistance from injection point, ft
ACross-sectional area, ft2
φporosity, fraction

Appendix A

In this Appendix, we present the analytical solution for the proposed approach of estimating the weighting factor in the mixing zone of the low salinity waterflooding scheme using 1D model. The model has already been discussed, where the simulation results illustrated by using saturation profiles of both tracers and water. Then, the weighting factor equation has been concluded based on the results of the tracer simulation model.
Initially, the water front of the 1D model has been calculated analytically using the Buckley–Leverett theory to validate the simulation results. The fractional flow of the injected water can be determined analytically using the following equation [42,43]:
f w = 1 1 + k r o k r w μ w μ o
where the relative permeability ratio in Equation (9) can be replaced with the exponential function of the plotted relationship between the water saturation and relative permeability ratio for the straight-line section, as illustrated in Figure A1A and expressed in Equation (10), where both a and b coefficients can be calculated [43]:
k r o k r w = a e b S w
Then, Equation (10) is substituted into Equation (9) to obtain:
f w = 1 1 + a e b S w μ w μ o
A tangent line was plotted from the initial water saturation to the maximum point in the water cut (fw) curve, as demonstrated in Figure A1B to calculate the saturation of the leading waterfront, as well as to determine the distance from the injector at the required time. An example is given in Figure A1C to illustrate the water saturation, the leading front, and the distance from the injector after 462 days. Moreover, this analytical solution was compared with the 1D tracer simulation scenario, which is illustrated in Figure 2. The analytical solution had a sharp waterfront at a distance of 260 ft from the injector. Meanwhile, the result of the numerical simulation exhibited a gradual mixing in the waterfront with an average equal or slightly higher than the analytical solution frontal distance. The distance from the injection point can be estimated using Equation (12), where the (dfw/dSw) can be calculated using the derivative form of Equation (11).
D =   5.615   W i   t   φ   A   d f w d S w
Figure A1. Water frontal in the 1D grid model. (A) The established exponential relationship between the relative permeability ratio (logarithmic value) and water saturation. (B) Water cut (fw) and the plotted tangent line. (C) Comparison of the analytical solution and numerical solution for the waterfront distance after 462 days. (D) Tracer concentration mixing curve after 462 days illustrating the tracer concentration of the injected water of 3000 and the tracer concentration of the connate water of 100,000 to emulate the low salinity water injection in the reservoir.
Figure A1. Water frontal in the 1D grid model. (A) The established exponential relationship between the relative permeability ratio (logarithmic value) and water saturation. (B) Water cut (fw) and the plotted tangent line. (C) Comparison of the analytical solution and numerical solution for the waterfront distance after 462 days. (D) Tracer concentration mixing curve after 462 days illustrating the tracer concentration of the injected water of 3000 and the tracer concentration of the connate water of 100,000 to emulate the low salinity water injection in the reservoir.
Applsci 10 04211 g0a1
Al-Ibadi [24] has illustrated that the high salinity water moving ahead to form the waterfront followed by low salinity water. The conclusion is the same even when the injected water has the same salinity as the formation of water, as illustrated in Figure A1D. In this example, both injected water and formation water have the same salinity while a tracer concentration of 3000 has used for the injected water compared with 100,000 for formation water. The simulation results demonstrated that a mixing zone of the tracer has formed, which starts with the maximum concentration in the waterfront and ends with the minimum concentration at a closer distance to the injection point. This conclusion is further illustrated using an overlapping plot of the subfigures shown in Figure A1C,D, as shown in Figure A2. In this plot, the leading frontal (i.e., beginning of the mixing zone) located at 305 ft from the injector and ending at 40 ft. Moreover, the mixing region of the brine mixture was situated within a 265 ft interval, which begins with the SCLoSal and ends with the SCMaxeff to the right. Therefore, the width of the mixing interval that starts altering the reservoir wettability and reduces the residual oil saturation depends on the weighting factor and the brine mixture salinity in this interval that determines the rate of change.
Figure A2. The location of the mixing zone within the leading waterfront after 462 days, illustrating the dilution curve of the tracer concentration when the injected water has significantly lower tracer concentration of 3000 compared to 100,000 of formation water.
Figure A2. The location of the mixing zone within the leading waterfront after 462 days, illustrating the dilution curve of the tracer concentration when the injected water has significantly lower tracer concentration of 3000 compared to 100,000 of formation water.
Applsci 10 04211 g0a2

References

  1. Aljuboori, F.A.; Lee, J.H.; Elraies, K.A.; Stephen, K.D. Gravity Drainage Mechanism in Naturally Fractured Carbonate Reservoirs; Review and Application. Energies 2019, 12, 3699. [Google Scholar] [CrossRef] [Green Version]
  2. Aljuboori, F.A.; Lee, J.H.; Elraies, K.A.; Stephen, K.D. The effectiveness of low salinity waterflooding in naturally fractured reservoirs. J. Pet. Sci. Eng. 2020, 191, 107167. [Google Scholar] [CrossRef]
  3. Elfeel, M.A.; Jamal, S.; Enemanna, C.; Arnold, D.; Geiger, S. Effect of DFN upscaling on history matching and prediction of naturally fractured reservoirs. In Proceedings of the 75th EAGE Annual Conference & Exhibition incorporating SPE Europec, London, UK, 10–13 June 2013. [Google Scholar]
  4. Aziz, R.; Niasar, V.; Martínez-Ferrer, P.J.; Godinez-Brizuela, O.E.; Theodoropoulos, C.; Mahani, H. Novel insights into pore-scale dynamics of wettability alteration during low salinity waterflooding. Sci. Rep. 2019, 9, 9257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Zhang, Y.; Xie, X.; Morrow, N.R. Waterflood Performance by Injection of Brine with Different Salinity for Reservoir Cores. In Proceedings of the SPE Annual Technical Conference and Exhibition, Anaheim, CA, USA, 11–14 November 2007; p. 12. [Google Scholar]
  6. Lager, A.; Webb, K.J.; Black, C.J.J.; Singleton, M.; Sorbie, K.S. Low salinity oil recovery—An experimental investigation. Petrophysics 2006, 49, 12. [Google Scholar]
  7. Zhang, Y.; Morrow, N.R. Comparison of Secondary and Tertiary Recovery with Change in Injection Brine Composition for Crude-Oil/Sandstone Combinations. In SPE/DOE Symposium on Improved Oil Recovery; Society of Petroleum Engineers: Tulsa, OK, USA, 2006; p. 14. [Google Scholar]
  8. Li, Y. Oil Recovery by Low Salinity Water Injection into a Reservoir: A New Study of Tertiary Oil Recovery Mechanism. Transp. Porous Media 2011, 90, 333–362. [Google Scholar] [CrossRef] [Green Version]
  9. Chen, Y.; Xie, Q.; Sari, A.; Brady, P.V.; Saeedi, A. Oil/water/rock wettability: Influencing factors and implications for low salinity water flooding in carbonate reservoirs. Fuel 2018, 215, 171–177. [Google Scholar] [CrossRef]
  10. Jackson, M.; Vinogradov, J.; Hamon, G.; Chamerois, M. Evidence, mechanisms and improved understanding of controlled salinity waterflooding part 1: Sandstones. Fuel 2016, 185, 772–793. [Google Scholar] [CrossRef] [Green Version]
  11. Bernard, G.G. Effect of Floodwater Salinity on Recovery of Oil from Cores Containing Clays. In SPE California Regional Meeting; Society of Petroleum Engineers: Los Angeles, CA, USA, 1967; p. 8. [Google Scholar]
  12. Jadhunandan, P.; Morrow, N. Effect of Wettability on Waterflood Recovery for Crude-Oil/Brine/Rock Systems. SPE Reserv. Eng. 1995, 10, 40–46. [Google Scholar] [CrossRef]
  13. Yildiz, H.; Valat, M.; Morrow, N. Effect of Brine Composition on Wettability and Oil Recovery of a Prudhoe Bay Crude Oil. J. Can. Pet. Technol. 1999, 38, 6. [Google Scholar] [CrossRef]
  14. Tang, G.-Q.; Morrow, N.R. Influence of brine composition and fines migration on crude oil/brine/rock interactions and oil recovery. J. Pet. Sci. Eng. 1999, 24, 99–111. [Google Scholar] [CrossRef]
  15. Wang, X.; Alvarado, V. Effects of Low-Salinity Waterflooding on Capillary Pressure Hysteresis. In SPE Improved Oil Recovery Conference; Society of Petroleum Engineers: Tulsa, OK, USA, 2016; p. 16. [Google Scholar]
  16. AlHammadi, M.; Mahzari, P.; Sohrabi, M. Fundamental investigation of underlying mechanisms behind improved oil recovery by low salinity water injection in carbonate rocks. Fuel 2018, 220, 345–357. [Google Scholar] [CrossRef]
  17. Pollen, E.N.; Berg, C.F. Experimental Investigation of Osmosis as a Mechanism for Low-Salinity EOR. In Abu Dhabi International Petroleum Exhibition & Conference; Society of Petroleum Engineers: Abu Dhabi, UAE, 2018; p. 20. [Google Scholar]
  18. Mahani, H.; Menezes, R.; Berg, S.; Fadili, A.; Nasralla, R.; Voskov, D.; Joekar-Niasar, V. Insights into the Impact of Temperature on the Wettability Alteration by Low Salinity in Carbonate Rocks. Energy Fuels 2017, 31, 7839–7853. [Google Scholar] [CrossRef]
  19. Shabaninejad, M.; Middleton, J.; Latham, S.; Fogden, A. Pore-Scale Analysis of Residual Oil in a Reservoir Sandstone and Its Dependence on Water Flood Salinity, Oil Composition, and Local Mineralogy. Energy Fuels 2017, 31, 13221–13232. [Google Scholar] [CrossRef]
  20. Nasralla, R.A.; Mahani, H.; Van Der Linde, H.A.; Marcelis, F.H.; Masalmeh, S.K.; Sergienko, E.; Brussee, N.J.; Pieterse, S.G.; Basu, S. Low salinity waterflooding for a carbonate reservoir: Experimental evaluation and numerical interpretation. J. Pet. Sci. Eng. 2018, 164, 640–654. [Google Scholar] [CrossRef]
  21. Sharma, H.; Mohanty, K.K. An experimental and modeling study to investigate brine-rock interactions during low salinity water flooding in carbonates. J. Pet. Sci. Eng. 2018, 165, 1021–1039. [Google Scholar] [CrossRef]
  22. Martínez-Mendoza, E.G.; Díaz-Viera, M.A.; Coronado, M.; Mendoza-Rosas, A.T. Capillary pressure and relative permeability estimation for low salinity waterflooding processes using pore network models. J. Pet. Sci. Eng. 2019, 182, 106253. [Google Scholar] [CrossRef]
  23. McGuire, P.L.; Chatham, J.R.; Paskvan, F.K.; Sommer, D.M.; Carini, F.H. Low Salinity Oil Recovery: An Exciting New EOR Opportunity for Alaska’s North Slope. In SPE Western Regional Meeting; Society of Petroleum Engineers: Irvine, CA, USA, 2005; p. 15. [Google Scholar]
  24. Al-Ibadi, H.; Stephen, K.D.; Mackay, E.J. Extended Fractional-Flow Model of Low-Salinity Waterflooding Accounting for Dispersion and Effective Salinity Range. SPE J. 2019, 24, 2874–2888. [Google Scholar] [CrossRef]
  25. Webb, K.; Lager, A.; Black, C. Comparison of high/low salinity water/oil relative permeability. In Proceedings of the International Symposium of the Society of Core Analysts, Abu Dhabi, UAE, 29 October–2 November 2008; Volume 29. [Google Scholar]
  26. Masalmeh, S.K.; Sorop, T.G.; Suijkerbuijk, B.M.; Vermolen, E.C.M.; Douma, S.; Van Del Linde, H.A.; Pieterse, S.G.J. Low Salinity Flooding: Experimental Evaluation and Numerical Interpretation. In Proceedings of the International Petroleum Technology Conference, Doha, Qatar, 19–22 January 2014; p. 13. [Google Scholar]
  27. Al-Ibadi, H.; Stephen, K.D.; Mackay, E.J. Insights into the fractional flow of low salinity water flooding in the light of solute dispersion and effective salinity interactions. J. Pet. Sci. Eng. 2019, 174, 1236–1248. [Google Scholar] [CrossRef]
  28. Jaber, A.K.; Awang, M.B.; Lenn, C.P. Box-Behnken design for assessment proxy model of miscible CO2-WAG in heterogeneous clastic reservoir. J. Nat. Gas Sci. Eng. 2017, 40, 236–248. [Google Scholar] [CrossRef]
  29. Schlumberger. Technical Description of Eclipse, 2014th ed.; Schlumberger: Houston, TX, USA, 2014. [Google Scholar]
  30. Al-Ibadi, H.; Stephen, K.; Mackay, E. Analytical Solution of Chemical Flooding in Heterogeneous Non-Communicating Layers with a Focus on Low Salinity Water Flooding. In Proceedings of the SPE Europec Featured at 81st EAGE Conference and Exhibition, London, UK, 3–6 June 2019; p. 27. [Google Scholar]
  31. Azim, R.A.; Faiz, S.; Rahman, S.; Elbagir, A.; Al Obaidi, N. Numerical Study of Low Salinity Water Flooding in Naturally Fractured Oil Reservoirs. In Recent Insights in Petroleum Science and Engineering; Zoveidavianpoor, M., Ed.; IntechOpen: Rijeka, Croatia, 2018. [Google Scholar]
  32. Ramirez, B.; Kazemi, H.; Al Kobaisi, M.; Özkan, E.; Atan, S. A Critical Review for Proper Use of Water/Oil/Gas Transfer Functions in Dual-Porosity Naturally Fractured Reservoirs: Part I. SPE Reserv. Eval. Eng. 2009, 12, 200–210. [Google Scholar] [CrossRef]
  33. Elfeel, M.A.; Al-Dhahli, A.; Geiger, S.; van Dijke, M.I.J. Multi-Scale Simulation of WAG Flooding in Naturally Fractured Reservoirs. In Proceedings of the SPE-164837-MS. Society of Petroleum Engineers, EAGE Annual Conference & Exhibition incorporating SPE Europec, London, UK, 10–13 June 2013. [Google Scholar]
  34. Aljuboori, F.; Corbett, P.; Bisdom, K.; Bertotti, G.; Geiger, S. Using outcrop data for geological well test modelling in fractured reservoirs. In Proceedings of the 77th EAGE Conference and Exhibition 2015: Earth Science for Energy and Environment, Madrid, Spain, 1–4 June 2015. [Google Scholar]
  35. Elfeel, M.A.; Geiger, S. Static and Dynamic Assessment of DFN Permeability Upscaling. In Proceedings of the SPE Europec/EAGE Annual Conference. Society of Petroleum Engineers, Copenhagen, Denmark, 4–7 June 2012; pp. 4–7. [Google Scholar]
  36. Sablok, R.; Aziz, K. Upscaling and Discretization Errors in Reservoir Simulation. Pet. Sci. Technol. 2008, 26, 1161–1186. [Google Scholar] [CrossRef]
  37. Wennberg, O.P.; Casini, G.; Jonoud, S.; Peacock, D.C. The characteristics of open fractures in carbonate reservoirs and their impact on fluid flow: A discussion. Pet. Geosci. 2016, 22, 91–104. [Google Scholar] [CrossRef]
  38. Kazemi, H.; Gilman, J.R. Multiphase flow in fractured petroleum reservoirs. In Flow and Contaminant Transport in Fractured Rock; Elsevier: Cambridge, MA, USA, 1993; pp. 267–323. [Google Scholar]
  39. Abushaikha, A.S.; Gosselin, O.R. Matrix-Fracture Transfer Function in Dual-Media Flow Simulation: Review, Comparison and Validation. In Proceedings of the Europec/EAGE Conference and Exhibition, Rome, Italy, 9–12 June 2008. [Google Scholar]
  40. Wennberg, O.P.; Svånå, T.; Azizzadeh, M.; Aqrawi, A.M.M.; Brockbank, P.; Lyslo, K.B.; Ogilvie, S. Fracture intensity vs. mechanical stratigraphy in platform top carbonates: The Aquitanian of the Asmari Formation, Khaviz Anticline, Zagros, SW Iran. Pet. Geosci. 2006, 12, 235–246. [Google Scholar] [CrossRef]
  41. Wennberg, O.P.; Azizzadeh, M.; Aqrawi, A.A.M.; Blanc, E.; Brockbank, P.; Lyslo, K.B.; Pickard, N.; Salem, L.D.; Svånå, T. The Khaviz Anticline: An outcrop analogue to giant fractured Asmari Formation reservoirs in SW Iran. In Fractured Reservoirs; The Geological Society: London, UK, 2007; Volume 270, pp. 23–42. [Google Scholar]
  42. Pieters, D.A.; Graves, R.M. Fracture Relative Permeability: Linear or Non-Linear Function of Saturation. In Proceedings of the International Petroleum Conference and Exhibition of Mexico, Veracruz, Mexico, 10–13 October 1994; Society of Petroleum Engineers: Veracruz, Mexico, 1994; p. 10. [Google Scholar]
  43. Ahmed, T. Reservoir Engineering Handbook, 3rd ed.; Gulf Professional Publishing, Elsevier Inc.: Houston, TX, USA, 2006. [Google Scholar]
Figure 1. The generated relative permeability sets for a fractured formation using the calculated ranges in Table 2 for the saturation exponents and saturation endpoints. The black curves represent the original relative permeability set, while the coloured curves represent the relative permeability of the flooded region by low salinity brine for three levels of alterations.
Figure 1. The generated relative permeability sets for a fractured formation using the calculated ranges in Table 2 for the saturation exponents and saturation endpoints. The black curves represent the original relative permeability set, while the coloured curves represent the relative permeability of the flooded region by low salinity brine for three levels of alterations.
Applsci 10 04211 g001
Figure 2. Numerical simulation results of the 1D model of water displacement illustrating (A) the profile of tracer concentration and (B) water saturation profile. The waterfront and the mixed zone were highlighted in both profiles.
Figure 2. Numerical simulation results of the 1D model of water displacement illustrating (A) the profile of tracer concentration and (B) water saturation profile. The waterfront and the mixed zone were highlighted in both profiles.
Applsci 10 04211 g002
Figure 3. The obtained change by the simulation results in the tracer concentration in the mixing zone. Region 1 and Region 3 represent the dominant regions by low and high tracer concentrations, respectively. Meanwhile, Region 2 represents the beginning of the interaction between the two tracers.
Figure 3. The obtained change by the simulation results in the tracer concentration in the mixing zone. Region 1 and Region 3 represent the dominant regions by low and high tracer concentrations, respectively. Meanwhile, Region 2 represents the beginning of the interaction between the two tracers.
Applsci 10 04211 g003
Figure 4. A comparison between the weighting factors of the mixing zone using the linear relationship in Equation (3) and the derived exponential relationship, as illustrated in Equation (4). The weighting factor value of one represents the beginning if the low salinity in the mixing zone. Meanwhile, a zero-weighting factor value represents the end of the mixing zone with high salinity concentration.
Figure 4. A comparison between the weighting factors of the mixing zone using the linear relationship in Equation (3) and the derived exponential relationship, as illustrated in Equation (4). The weighting factor value of one represents the beginning if the low salinity in the mixing zone. Meanwhile, a zero-weighting factor value represents the end of the mixing zone with high salinity concentration.
Applsci 10 04211 g004
Figure 5. An illustration of the single matrix block model. (A) A discretised matrix block (fine-scale model), which can be simulated using a single-porosity model, after Aljuboori et al. [2]. (B) Single block model, which can be simulated using the dual-porosity model.
Figure 5. An illustration of the single matrix block model. (A) A discretised matrix block (fine-scale model), which can be simulated using a single-porosity model, after Aljuboori et al. [2]. (B) Single block model, which can be simulated using the dual-porosity model.
Applsci 10 04211 g005
Figure 6. The simulated cases of the intermediate scale model, (A) fractures were saturated entirely with oil, and the model comprises a low salinity brine injector, and an oil producer (B) fractures were filled with water, and the fluid exchange occurred due to capillary, viscous, and gravity forces.
Figure 6. The simulated cases of the intermediate scale model, (A) fractures were saturated entirely with oil, and the model comprises a low salinity brine injector, and an oil producer (B) fractures were filled with water, and the fluid exchange occurred due to capillary, viscous, and gravity forces.
Applsci 10 04211 g006
Figure 7. The effect of low salinity waterflooding on the relative permeability curves. (A) Generated relative permeability sets due to wettability alteration and residual oil saturation. (B) Fractional flow of water for the generated relative permeability sets, where the water breakthrough time delayed by 47% when maximum alteration used (indicated by red legend) compared to the high salinity set (indicated by black legend).
Figure 7. The effect of low salinity waterflooding on the relative permeability curves. (A) Generated relative permeability sets due to wettability alteration and residual oil saturation. (B) Fractional flow of water for the generated relative permeability sets, where the water breakthrough time delayed by 47% when maximum alteration used (indicated by red legend) compared to the high salinity set (indicated by black legend).
Applsci 10 04211 g007
Figure 8. Oil recoveries of the fine-scale model using an exponential weighting factor and a single-porosity model for various levels of relative permeability alterations (minimum, average, and maximum), in addition to the high salinity waterflooding scenario. Meanwhile, the same scenarios were repeated using a linear weighting factor, as indicated by dashed lines.
Figure 8. Oil recoveries of the fine-scale model using an exponential weighting factor and a single-porosity model for various levels of relative permeability alterations (minimum, average, and maximum), in addition to the high salinity waterflooding scenario. Meanwhile, the same scenarios were repeated using a linear weighting factor, as indicated by dashed lines.
Applsci 10 04211 g008
Figure 9. A comparison of oil recoveries of the fine-scale model (solid lines) and the single block model (dashed lines) using an exponential weighting factor for various levels of relative permeability alterations (minimum, average, and maximum), in addition to high salinity waterflooding scenario.
Figure 9. A comparison of oil recoveries of the fine-scale model (solid lines) and the single block model (dashed lines) using an exponential weighting factor for various levels of relative permeability alterations (minimum, average, and maximum), in addition to high salinity waterflooding scenario.
Applsci 10 04211 g009
Figure 10. Water saturation profile in the matrix after 10 days of fluid exchange between the matrix and fractures (A) using the dual-porosity model for a single matrix block, and (B) using the single-porosity model for the same matrix block in (A) after discretisation. The corner grid cells in the fine-scale model have shown a higher water saturation compared to the remaining cells, as they have two faces in contact with fractures for fluid exchange.
Figure 10. Water saturation profile in the matrix after 10 days of fluid exchange between the matrix and fractures (A) using the dual-porosity model for a single matrix block, and (B) using the single-porosity model for the same matrix block in (A) after discretisation. The corner grid cells in the fine-scale model have shown a higher water saturation compared to the remaining cells, as they have two faces in contact with fractures for fluid exchange.
Applsci 10 04211 g010
Figure 11. Comparison of the simulation results between the single block model (solid line) and the intermediate scale model (dashed line). Almost identical recoveries were obtained which indicate that single block results can be used as a representative solution to a larger scale model under the same condition.
Figure 11. Comparison of the simulation results between the single block model (solid line) and the intermediate scale model (dashed line). Almost identical recoveries were obtained which indicate that single block results can be used as a representative solution to a larger scale model under the same condition.
Applsci 10 04211 g011
Figure 12. Comparison of the simulation results between the imbibition case (solid line) and injection case (dashed line) of the intermediate scale model. Faster oil recovery in the imbibition case was observed in the early time while higher ultimate recoveries were obtained by the injection case.
Figure 12. Comparison of the simulation results between the imbibition case (solid line) and injection case (dashed line) of the intermediate scale model. Faster oil recovery in the imbibition case was observed in the early time while higher ultimate recoveries were obtained by the injection case.
Applsci 10 04211 g012
Figure 13. Contour maps of Upper and Lower Qamchuqa formations illustrates the northern region. The filled black dots refer to the drilled wells in the area, while the empty dots with an arrow indicate the proposed injection wells.
Figure 13. Contour maps of Upper and Lower Qamchuqa formations illustrates the northern region. The filled black dots refer to the drilled wells in the area, while the empty dots with an arrow indicate the proposed injection wells.
Applsci 10 04211 g013
Figure 14. Comparison of the waterflooding scenarios using high salinity and low salinity schemes in the Qamchuqa reservoir, using the exponential weighting factor.
Figure 14. Comparison of the waterflooding scenarios using high salinity and low salinity schemes in the Qamchuqa reservoir, using the exponential weighting factor.
Applsci 10 04211 g014
Figure 15. Comparison of the pressure support to the Qamchuqa reservoir during high salinity and low salinity waterflooding, where a similar impact was observed for both flooding schemes on the reservoir subsurface pressure and the produced gas-oil ratio. The natural depletion was provided for comparison.
Figure 15. Comparison of the pressure support to the Qamchuqa reservoir during high salinity and low salinity waterflooding, where a similar impact was observed for both flooding schemes on the reservoir subsurface pressure and the produced gas-oil ratio. The natural depletion was provided for comparison.
Applsci 10 04211 g015
Table 1. An illustration of the brine properties using an increment step of 20% in the normalised salinity. The table is based on a maximum salinity of 100,000 ppm and a minimum of 3000 ppm, using a linear formula and the conditions of the Qamchuqa reservoir. This table (except Columns 1 and 2) is utilised in the simulation model to calculate the brine properties in the mixing region.
Table 1. An illustration of the brine properties using an increment step of 20% in the normalised salinity. The table is based on a maximum salinity of 100,000 ppm and a minimum of 3000 ppm, using a linear formula and the conditions of the Qamchuqa reservoir. This table (except Columns 1 and 2) is utilised in the simulation model to calculate the brine properties in the mixing region.
Normalised SalinitySalinity, PPMSalt Concentration, lb/bblFormation Volume Factor (Bw), rb/STBCompressibility (Cw), psi-1Viscosity (µw), cp
030001.0521.01532.98 × 10−60.352
0.222,4007.8511.01622.84 × 10−60.391
0.441,80014.6511.01692.73 × 10−60.422
0.661,20021.4511.01752.65 × 10−60.444
0.880,60028.2511.01812.57 × 10−60.460
1100,00035.0511.01872.49 × 10−60.478
Table 2. Quantitative summary of the impact of the low salinity on the oil recovery through relative permeability parameters which was calculated from the published results of [20,21,23,25,26]. The Minimum and Maximum parameter values were computed from the references’ results while the average was obtained arithmetically.
Table 2. Quantitative summary of the impact of the low salinity on the oil recovery through relative permeability parameters which was calculated from the published results of [20,21,23,25,26]. The Minimum and Maximum parameter values were computed from the references’ results while the average was obtained arithmetically.
The Changed Parameters during Low Salinity WaterfloodingNo. of SamplesChange %, (+ Increment, − Reduction)
MinimumAverageMaximum
Residual oil saturation (Sor_a)39−1.4−26.1−63.6
Oil saturation exponent (no_a)9−11.4−28.5−33.3
Water relative permeability (krw_a)3+9.3+32.3+48.4
Water saturation exponent (nw_a)9+12.5+36.2+56.7
Table 3. Simulation results of fine-scale and single block models using single and dual-porosity models, respectively, highlighting the recovery differences between their performance.
Table 3. Simulation results of fine-scale and single block models using single and dual-porosity models, respectively, highlighting the recovery differences between their performance.
ScenarioFine-Scale Model (WF-Exponential)Single Block ModelRecovery Difference % of STOIIP
High Salinity (Reference Model)0.4010.4868.5
LoSal—Minimum Alteration0.4240.5017.7
LoSal—Average Alteration0.4760.5537.7
LoSal—Maximum Alteration0.5310.5936.2

Share and Cite

MDPI and ACS Style

Aljuboori, F.A.; Lee, J.H.; Elraies, K.A.; Stephen, K.D. Using Low Salinity Waterflooding to Improve Oil Recovery in Naturally Fractured Reservoirs. Appl. Sci. 2020, 10, 4211. https://doi.org/10.3390/app10124211

AMA Style

Aljuboori FA, Lee JH, Elraies KA, Stephen KD. Using Low Salinity Waterflooding to Improve Oil Recovery in Naturally Fractured Reservoirs. Applied Sciences. 2020; 10(12):4211. https://doi.org/10.3390/app10124211

Chicago/Turabian Style

Aljuboori, Faisal Awad, Jang Hyun Lee, Khaled A. Elraies, and Karl D. Stephen. 2020. "Using Low Salinity Waterflooding to Improve Oil Recovery in Naturally Fractured Reservoirs" Applied Sciences 10, no. 12: 4211. https://doi.org/10.3390/app10124211

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop