Next Article in Journal
Preparation of Polyurethane Adhesives from Crude and Purified Liquefied Wood Sawdust
Next Article in Special Issue
Synthesis and Characterization of Nonwoven Cotton-Reinforced Cellulose Hydrogel for Wound Dressings
Previous Article in Journal
Recent Advances in Chitin and Chitosan/Graphene-Based Bio-Nanocomposites for Energetic Applications
Previous Article in Special Issue
UV-Protected Polyurethane/f-Oil Fly Ash-CeO2 Coating: Effect of Pre-Mixing f-Oil Fly Ash-CeO2 with Monomers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Green Enhanced Oil Recovery for Carbonate Reservoirs

Department of Petroleum Engineering, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
Submission received: 1 September 2021 / Revised: 13 September 2021 / Accepted: 15 September 2021 / Published: 25 September 2021

Abstract

:
Green enhanced oil recovery (GEOR) is an eco-friendly EOR technique involving the injection of specific green fluids to improve macroscopic and microscopic sweep efficiencies, boosting residual oil production. The environmentally friendly surfactant-polymer (SP) flood is successfully tested in a sandstone reservoir. However, the applicability of the SP method does not extend to carbonate reservoirs yet and requires comprehensive investigation. This work aims to explore the oil recovery competency of a green SP formulation in carbonate through experimental and modelling studies. Numerous formulations of SP with ketone, alcohol, and organic acid are selected based on phase behavior and interfacial tension (IFT) reduction capabilities to examine their potential for enhancing residual oil production from carbonate cores. A blending of nonionic green surfactant alkyl polyglucoside (APG), xanthan gum (XG) biopolymer, and butanone recovered 22% tertiary oil from the carbonate core. This formulation recovered more than double residual crude than that of the APG, XG, and acetone. Similarly, a combination of APG, XG, acrylic acid, and butanol increased significantly more oil than the APG, XG, and acrylic acid formulation. The APG, XG, and butanone mixture is efficient with regards to boosting tertiary oil recovery from the carbonate core.

1. Introduction

Oil recovery methods are grouped into three main categories [1]: primary, secondary, and tertiary processes. In the primary recovery, the oil is produced by the natural reservoir energy. After depletion of natural energy, the secondary stage is introduced. The secondary stage is water flooding, pressure maintenance, and gas injection. When the secondary method is no longer economically viable, the tertiary oil recovery, commonly known as the enhanced oil recovery (EOR), is implemented. According to Lake et al. (2014) [2], the EOR methods are chemical, gas, thermal, etc. Chemical EOR [1] processes involve injecting specific liquid chemicals that effectively displace oil because of their phase-behavior properties, decreasing the interfacial tension (IFT) between the displacing liquid and oil. The popular chemicals are surfactants, polymers, and alkaline. The primary displacing liquid slug is a complex chemical system called a micellar solution. The slug contains a surfactant, co-surfactant (alcohol), oil, electrolytes (i.e., NaCl), and water. Surfactant slug is followed by a mobility buffer called polymer solution. The primary mechanism of the chemical EOR is a reduction of IFT between oil and water/brine. The gas injection EOR method displaces crude with a miscible fluid and forms a single phase when mixed at all proportions with the oil at the existing condition. The process uses gases, for instance, carbon dioxide (CO2), nitrogen (N2), and natural gas, mainly methane (CH4) and flue gas. The injected gas expands in the reservoir and pushes additional oil to the production well or the gas dissolves in the oil to lower the viscosity and IFT and improves the recovery. Thermal methods include injecting hot water, steam, and other gas by conducting combustion in situ of the oil. The procedures apply heat to increase the temperature of the crude oil in the formation and reduce oil viscosity and/or vaporize part of the oil, thereby decreasing the mobility ratio and increasing oil recovery. Thermal EOR methods are generally applicable to heavy and dense crudes. Other processes [1] include microbial enhanced oil recovery (MEOR). MEOR [3] is an environmentally friendly enhanced oil recovery method that involves injecting microorganisms and produces surfactant, polymer, alcohol, ketone, acids, and gas in situ and enhances oil recovery. According to Haq [3,4], the green enhanced oil recovery (GEOR) is an eco-friendly EOR technique that involves the injection of specific green fluids, for example, surfactants, polymers, alcohols, acids, ketones, and gas (N2, CO2) that effectively improves macroscopic and microscopic sweep efficiencies as a result increases residual oil recovery. Classification of GEOR, surfactant, and polymer flooding mechanisms and challenges in carbonates, a summary of literature review, knowledge gaps, the objective of the work, and an outline of the research are discussed in the following sections.
GEOR processes [3] can be grouped into two categories: in situ and ex situ. Regarding in situ methods, bacteria with nutrients are injected into a reservoir, producing by-products such as surfactants, polymers, alcohols, ketones, acids, and gases (i.e., CO2 and CH4). These products increase microscopic and macroscopic displacement efficiencies and boost oil production. Microbial enhanced oil recovery (MEOR) is one example of an in situ GEOR method. Ex situ GEOR processes involve the injection of green gases and/or chemicals that improve microscopic and macroscopic displacement proficiencies and enhance oil recovery. Ex situ methods can be subdivided into three main categories: chemical, gas, and hybrid. Smart water, water alternating gas (WAG), foam, and surfactant-polymer (SP) flooding are all hybrid in nature. N2 and CO2 floods are grouped as gas flooding. Nature-friendly chemicals are used in green chemical enhanced oil recovery (GCEOR). GCEOR involves injecting specific eco-friendly chemicals such as surfactants, polymers, alcohols, ketones, and acids. In the GCEOR process, the primary displacing liquid slug is a complex chemical system called a micelle solution that contains nature-friendly surfactants, co-surfactants, oil, electrolytes, and water. The GEOR process is shown in Figure 1.
The surfactant flooding technique involves injecting surfactants to mobilize oil via IFT reduction and wettability alteration [5]. The surface-active agent, commonly known as surfactant, can change the rock’s wettability and decrease the IFT between oil and micellar solution [6]. This mechanism led to producing more oil from the formation. The governing mechanism of surfactant flooding is achieved by forming the microemulsion phase between the oil and water phases. The emulsion formed by surfactant at the fluid interface is dependent on the emulsion rheological properties, phase behavior, salinity, initial volume ratio, and internal structure. The primary mechanism of polymer flood is to improve the aqueous phase viscosity (μw) and the effective water permeability (kw) [7]. Mobility ratio (M) is calculated as
M = λ w λ o = k w μ w k o μ o
where λ w = brine mobility, λ o = oil mobility, k w = effective permeability of brine, k o = effective permeability of oil, μ w = brine viscosity, and μ o = viscosity.
The mobility ratio determines the flooding performance of the polymer. Low values of mobility ratio (M < 1) according to Lai (2007); Sorbie and Phil (1991) [8,9] indicate a piston-like displacement and good sweep efficiency while (M > 1) leads to more improvement in the sweep efficiency by the fluids.
There are many challenges of surfactant and polymer flooding in carbonates. The carbonate reservoirs are composed of dolomite, calcite, magnesite, anhydrite, and gypsum rocks. The carbonate formations typically show low porosity, and the mixed wet rock characteristics and fractured composition that leads to low oil recovery [10]. One of the main challenges of the surfactant flooding method is adsorption, which causes economic feasibility [10]. Many studies on a range of carbonate reservoirs with different mineralogy have been conducted to quantify adsorption phenomena and suitability for EOR [9,11,12,13,14]. The other challenges of surfactant flooding in carbonates are precipitation, gravity segregation, and phase trapping. The factors that affect the adsorption process are temperature, rock mineralogy, fluid saturations, salinity, and divalent cations. The surfactant adsorption is controlled by the structure, which must be accounted for in the design phase. Phase trapping involves the migration of the surfactant to either the oil or microemulsion phase. High salinity and temperature may cause the surfactant to transfer to the oil face, thus reducing the effectiveness of the flooding process and surfactant loss. On the other hand, the challenges of polymer flooding encountered in carbonate reservoirs are degradation, high temperature, salinity, and complexity. Degradation is a process that would break down the molecular structure of the macromolecules in the polymer solution resulting in a partial or complete loss of viscosity. Note that the polymers are expected to be stable throughout their propagation during polymer flooding. The carbonate reservoirs with high well spacing experience the major degradation challenge regardless of the kind of degradation, for example, chemical, thermal, mechanical, or biological [15]. It has been suggested that polymer flooding has the potential to increase oil recovery from the carbonates [16,17,18]. However, it becomes challenging to find suitable polymers for carbonates due to the complexity of the reservoir and high salinity and temperature. In addition, degradation of polymer is more in elevated temperature and salinity. According to Roehi and Choquette (1985) [19], the world’s known oil reserves in carbonate reservoirs is ~50%. The majority of the carbonate reservoirs are naturally fractured, poorly understood, and have lower recovery than an unfractured reservoir. Oil recovery in carbonates mainly depends on matrix permeability, wettability, fracture intensity, and fluid properties. The vital recovery controlling parameter is the wettability of the carbonate reservoir [20,21].
The challenges of previous studies in surfactant based chemical EOR can be summarized as follows: Most of the publications related to surfactant flooding found in literature deal with synthetic surfactants. These floodings are applied mainly in sandstone reservoirs. A relatively smaller number of surfactants flooding are found in the literature in carbonate, and these surfactants are synthetic. A significant portion of these surfactants is not environmentally friendly. A similar trend is observed in polymer flooding. Most of the polymers applied in chemical EOR are either HPAM or PAM, and they are synthetic. There are few publications on xanthan gum biopolymer, and the polymer is also used in a sandstone reservoir. The application of biopolymer in carbonates is restricted due to harsh reservoir conditions. Several publications of SP or ASP flooding in carbonate in tables deals with chemicals such as surfactant, polymer, alcohol, and alkali. These are synthetic, and many of them are not eco-friendly. There is a knowledge gap in environmentally friendly SP formulation. In addition, there is not enough information available on the combined effect of a green surfactant, alcohol, ketone, acid, and polymer on EOR in carbonates. It would be better to have a green SP formulation for carbonate reservoirs for sustainable development of the petroleum industry and requires step-by-step documentation to understand the role of green SP in carbonate.
The project aims to (a) determine optimum salinity through phase behavior modelling and (b) examine the combined impact of green surfactant, polymer, organic acids, and microbial products on reducing interfacial facial tension and boosting residual oil recovery from a carbonate core. The goals are achieved by meeting several research milestones, which are the results of several experiments and modeling. These are IFT measurement, phase behavior study, core flooding experiment, and mathematical modeling.

2. Literature Review

2.1. Surfactant Flooding in Carbonates

Surfactant flooding is an enhanced oil recovery (EOR) technique involving surfactant injection to mobilize oil via IFT reduction and wettability alteration. The main challenge in surfactant flooding is adsorption that leads to economic viability issues. The review focuses on the prime challenges.
The surfactant’s adsorption is caused by the electrostatic interaction affecting the mineralogy of the rock and structure of the surfactant [22]. Mannhardt et al. (1994) [23] found that adsorption could be minimized by adding cationic surfactant with low affinity with low to moderate brine. An experimental study was conducted by Trogus et al. (1977) [24], and it was found that the adsorption rises sharply with increasing nonionic surfactant concentration. It raised to a maximum point corresponding to the critical micelle concentration (CMC) and remained constant. It is also noted that the reduction of adsorption could be achieved by adding alcohol in the surfactant solution. They also conducted a modelling study of the adsorption process and found a minor dispersion effect in anionic surfactants. The adsorption was more in kaolin than sandstone. It was believed that the adsorption of cationic surfactant is less in carbonate core than sandstone [14]. Tabatabal et al. (1993) [25] found that the adsorption of cationic surfactant on carbonate core was relatively low because the mineral surface had a lattice charge.
Ahmadi and Sheng (2016) [11] carried out a surfactant flooding operation on carbonate reservoir samples to quantify the applicability of nano-surfactant extract from Ziziphus Spina-Christi leaves. Findings revealed that an increase in surfactant concentration and hydrophilic nano-silica concentration in the surfactant solution increases the ultimate recovery of the process.
Researchers found a number of ways to minimize adsorption: (1) low concentration surfactant injection, (2) pre-flush with a sacrificial chemical, and (3) co-injection of surfactant and co-surfactant. Polyacrylate works better as a sacrificial agent for carbonate and sandstone rocks and is independent of surfactant types. It is found that adsorption decreases with the increasing molecular, for example, sodium-based surfactants [26]. Krumrine et al. (1982) [27] performed an experimental study and found that adsorption could be reduced by adding sodium silicate, sodium carbonate, and sodium tripolyphosphate. Three minerals—dolomite, calcite, and kaoline—were selected to measure adsorption using CO2 injection at low and high pressure [12]. The results revealed that the tertiary amine Ethomeen C12 (anionic surfactant) converts to cation at high pressure (pH = 4) and reduces adsorption. At high pressure, divalent and trivalent cations decrease the dissolution of calcite and dolomite, leading to an increase in adsorption [13]. Tsau et al. (2000) [28] described that lignosulfonate as pre-flush gives more economic advantage when used as pre-flush than co-injection to reduce adsorption.
In summary, these challenges can be resolved by one of these techniques: (a) The salinity gradient technique involves injecting low salinity brine to change the high salinity state to the optimum salinity required for the successful implementation of surfactant flooding, followed by injection of surfactant slug [29]. (b) Use of sacrificial agents. The sacrificial agents have been proposed by many researchers to reduce adsorption and improve the flooding process’s effectiveness. Agents such as polystyrene sulfonate, sodium polyacrylate, calcium lignosulfonate, and polyacrylate bind to the active adsorption sites available prior to the surfactant injection. (c) Alkali-surfactant flooding. In this process, alkali is added as an adsorption reducing agent. Additionally, the alkali also generates in-situ surfactants from naphthenic acids in crude oil. (d) Surfactant foams. The method involves a surfactant aided CO2 flooding process. However, this has concerns about stability in the reservoir. Surfactant flooding is commonly seen in a sandstone reservoir. A list of surfactants applied in carbonate and sandstone is presented in Table A1 in Appendix A. It indicates that most of the fields use a synthetic surfactant. Two fields out of thirteen (~15%) use biosurfactants, and four are carbonate formation.

2.2. Polymer Flooding in Carbonates

Polymer flooding is a popular EOR method in oil industry [7] and is suitable for sandstone reservoirs. However, there are many challenges faced in carbonate reservoir due to high temperature and salinity, and low permeability. In recent years, researchers have been working toward extending this success to carbonate reservoirs in more challenging reservoir conditions in terms of temperature, salinity, and permeability. The review discusses some significant challenges: polymer degradation, salinity, and temperature.
Polymer degradation is a process that would break down the molecular structure of the macromolecules in the polymer solution resulting in a partial or complete loss of viscosity [7]. Note that the polymers are expected to be stable throughout their propagation irrespective of the type of degradation (chemical, thermal, mechanical or biological). Sandengen et al. (2017) [15] describe that carbonate reservoirs with high well spacing experience this major challenge. They cited an example of the giant carbonate offshore fields where the residence time inside the reservoir could be in years.
Traditionally, it is essential to mention that polymer flooding is sensitive to elevated temperature and salinity [30]. However, pre-flush of low salinity make brine may lead to favorable condition polymer flood which could reduce chemical degradation of polymers. Alfazazi et al. (2019) [30] described elaborately that HPAM might be applied to pre-flushed HTHS carbonate reservoir to improve oil recovery. A study by Diab and Al-Shalabi (2020) [7] shows that two conditions temperature and salinity. These are usually encountered in carbonate reservoirs worldwide, particularly in the Middle East, to apply polymer flooding.
A study on screening workflow for carbonate green field by Al-Shalabi et al. (2018) [31] mentioned how their screening ranking was performed based on pore-scale, compatibility of polymer with rock surface, macro-scale filtering, and industry guidance. Diab and Al-Shalabi (2020) [7] investigated the importance of the screening stage in carbonate reservoirs. They stated how vital this stage is in carbonate polymer flooding for EOR purposes. Screening tests, according to them, are inexpensive and provide enough knowledge about the solution rheology, thermal stability, and salinity tolerance. Some of the most important screening tests are viscosifying ability, long-term thermal stability, sensitivity for salinity and hardness, shear stability, and static adsorption. The field application of polymer is listed in Table A2, and it reveals that the field application of biopolymer is ~19% (3 out of 16), and all fields are sandstone reservoirs.

2.3. SP Flooding in Carbonates

SP is commonly known as surfactant-polymer. The SP flooding method involves injecting surfactant and polymer (SP) to alter EOR properties and increase residual oil production. The main issue in SP flooding is chemical losses due to adsorption, surfactant precipitation, and cost. These issues can be reduced by the injection of alkali, which is a cheap chemical. When SP blends are injected with alkali, the technique is referred alkaline surfactant polymer (ASP) flooding. The SP or ASP flooding technique applied in the carbonate fields is listed in Table A3. Experimental studies are summarized in Table A4. It shows that all SP and ASP formulations are synthetic. It would be better to have an eco-friendly formulation for carbonates.
In 2020, few publications on ASP gave exciting information, and these are discussed here. Ahmed and Shenj [32] conducted experiments and simulation studies of foam assisted ASP flooding. They examined the oil recovery performance of different injectivity scenarios of foaming agents and ASP formulation. The results revealed that ASP produced maximum incremental oil due to the gas phase. Fei Pan et al. [33] studied cationic and anionic surfactants’ role on IFT and oil recovery. They found that the anionic linear alkylbenzene sulfonic acid (LABSA) surfactant produced 15% more oil than the cationic Cetyl trimethyl ammonium bromide (CTAB) surfactant. LABSA has a higher IFT effect than CTAB. Afshin Davarpanah [34] conducted a parametric study of nanoparticles assisted polymer flooding in evaluating injectivity performance of two-phase flow. It was found that the higher mobility ratio and polymer concentration gave higher oil recovery.
The primary limitation of the earlier work on chemical EOR can be summarized as follows:
  • The majority of the publications of surfactant flooding in carbonates deals with synthetic chemical surfactants and co-surfactants. Many of them are not eco-friendly, either surfactant or co-surfactant. Corrosive acid is usually used to increase the productivity of the formulation. Several publications deal with environmentally friendly surfactants and co-surfactants in oil recovery. However, these formulations are applied in a sandstone reservoir. There is not enough information available on the combined effect of a green surfactant, alcohol, ketone, acid, and polymer on EOR in carbonates.
  • Polymer flooding, especially synthetic polymer, is a widely used chemical oil recovery method in sandstone reservoirs. Most of the polymers applied in chemical EOR are synthetic such as HPAM or PAM. There are few biopolymers, such as Xanthan gum, used in the sandstone reservoirs. The application of biopolymer in carbonates is restricted due to harsh reservoir conditions.
  • Surfactant-based polymer flooding is widespread in sandstone reservoirs but limited in carbonate reservoirs. Several publications of SP or ASP flooding in carbonate in tables deals with chemicals such as surfactant, polymer, alcohol, and alkali. These are synthetic, and many of them are not eco-friendly. There is a knowledge gap in environmentally friendly SP formulation. It requires a step-by-step documentation to understand the role of green SP in carbonate.

3. Carbonate Reservoirs in Saudi Arabia

The Kingdom of Saudi Arabia (KSA) is one of the largest countries in Asia. Since the first oil discovery in March 1938 at the Dammam oil well No. 7 at a depth of 1440 m, Saudi Arabia has become a major supplier of crude oil worldwide. The Saudi Arabian hydrocarbon accounts for ~50% of the GDP as well as 85% of export earnings [35]. Saudi Arabian oil reserves are the second-largest globally and are estimated to be at 268 billion barrels [36]. Some of the major oil fields in Saudi Arabia are as follows.

3.1. Abqaiq

Abqaiq is the northeast extension of the Ghawar field, the largest oil field globally. The main oil-producing section in the Abqaiq field is Arab-D, mainly composed of limestone and dolomitic mixture. The depth of Arab-D formation is 4400–4700 ft, and the temperatures in these reservoirs can be found to be around 196 °F. The current reservoir pressure in this formation has reached 2143 psi. The Arab-D section of the Abqaiq field has a relatively high permeability of about 80–150 md, and the porosities range from 5 to 30%. The crude is light crude with an API of about 30.5. The salinity of the reservoir fluid ranges between 38,000 and 240,000 ppm. There are currently 107 wells in operation, and the recovery mode is waterflooding [37,38,39]. Below the Arab-D formation in the Abqaiq field is the Hanifa formation, which can be found at a depth of 5200–7500 ft. The Arab-D reservoir separates it by about 450 ft of impermeable carbonates of the Jubaila formation. An engineering estimate can be made regarding the temperature of the formation to be around 237 °F. The porosity of the formation is similar to Arab-D and is approximately 5–30%. It has a very low permeability of approximately 0.1–10 md. The crude is Light oil with an API of ~30.5. The reservoir is mainly composed of organic-rich carbonate rocks and is recovered by waterflooding [40].

3.2. Haradh

Haradh is located in the eastern province of KSA and is nearly 80 km from the Arabian Gulf. It is situated in the southern part of the Ghawar field. The Haradh field is located at a 5200–7500 ft depth and is mainly composed of Carbonate rocks such as limestones and dolomites. One of the significant features of the Haradh field is the tilted oil-water contact which gets shallower from north to south. The temperature in the reservoir reaches 170 to 210 degrees Fahrenheit. The reservoir is maintained at a pressure of 2200 psi, and the bubble point pressure is about 1650 psi. Data for the permeability and porosity of the reservoir are not present, but it is known to have super permeability due to fractures. The crude oil is a light oil with an API of 33, and the salinity of the reservoir is around 37,000 to 150,000 ppm. There are 28 injections, 32 production, and 13 observation wells. Improved oil recovery (IOR) techniques include applying Intelligent wells and Maximum Reservoir Contact (MRC) wells. The field is using waterflooding techniques [40,41,42].

3.3. Khurais

The Khurais complex was discovered in 1970 and had the capacity to produce ~1.2 million barrels of crude oil per day. The complex consists of three fields: Khurais, Abu Jifan, and Mazalij. They are located approximately 100 miles east of Riyadh and consist of both Arab-D and Hanifa formations. Arab-D reservoir consists of clastic limestones and is a few hundred feet thick, and the Hanifa is located below it. The rock quality is poor, and the degree of fracturing is high. The pressure of the Khurais field is about 1800 psi, and the API is around 33–36. There are 50 wells in operation. Not a lot of other data for the Khurais field is present. Still, as it has both Arab-D and Hanifa formations, its properties would be similar to the Abqaiq field mentioned earlier in the report [40,41,42,43,44,45].

3.4. Khursaniyah

Khursaniyah is one of the oil fields situated on the coastal plain of Saudi Arabia and is a part of the Ghawar field. Khursaniyah was discovered in 1956, and regular production began when the Arab-C came online in 1960. The proven reserves of Khursaniyah are more than 1.5 billion barrels and can extend up to a depth of 7000 ft. Khursaniyah is mainly composed of carbonate rocks with a porosity of ~23% and permeabilities ranging between 80 and 250 md. Light crude oil is produced from this reservoir with an API of 31. The reservoir is currently being waterflooded to maintain pressure. Khursaniyah has also successfully implemented intelligent, latest systems and applications [46,47].

3.5. Qatif

In 1945, the Qatif field was discovered and started producing oil in 1946. It is situated on the western side of the Arabian Gulf. The field has seven oil-bearing reservoirs. These are Arab A to D, Hainfa, Upper Fadhili, and Lower Fadhili. Arab C and D are superior to others. Arab C produces medium crude, and D has light crude oils. The Qatif field is carbonate members of cycles that deposited marine carbonate sediments capped by regressive sabkhah anhydrite. The porosity of this field is around 12–35%, and the permeability ranges between 80 and 500 md. The produced fluid API varies between 28 and 42. The Qatif field is the first i-field in Saudi Aramco. The field is well equipped with asset measurement, communication, control, and field rate optimization [48,49].

3.6. Manifa

Manifa is the world’s fifth-largest oil field and is the world’s largest offshore crude oil production increment built in a single phase. The field was discovered in 1957, 200 km northwest of Dhahran, mainly offshore, with a Gulf water depth ranging from 6 to 36 feet. It is a carbonate reservoir and produces heavy oil in the range of 24 to 26 API. Not much data for the field is public, but the field’s depth is greater than 17,000 ft, and the longest well in Saudi Arabia, which is about 37,042 ft, has been drilled in this field. Extended reach wells are pretty common in this field [50].

4. Methodology

4.1. Materials

The Alkyl Polyglucoside (APG) 264 is a nonionic surfactant; the commercial name is the BASF supplied Glucopon 600 CSUP. The Xanthan gum, formic acid, acetic acid, acrylic acid, acetone, butanone, and alcohol were purchased from Sigma Aldrich (Dammam, Saudi Arabia). The physicochemical properties of acetone and butanone are given in Table 1. The acids and ketones were 99.98% pure. The NaCl was 99.87% pure and purchased from Sigma Aldrich. The Arabian light crude oil was used throughout this research and supplied by Saudi Aramco (Dhahran, Saudi Arabia). The crude is a light oil and black liquid. The density and viscosity of this crude oil were measured at 25 °C and 52 °C in our laboratory. The densities are 0.8699 g/cm3 and 0.8484 g/cm3 at 25 °C and 52 °C, respectively. Similarly, the viscosities are 19.8 cP and 8.2 cP at 25 °C and 52 °C, respectively.
Xanthan gum is a polysaccharide anionic polymer produced by fermentation. It is composed of a β-(1→4)-d-glucopyranose glucan backbone with side chains of (1→3)-α-d-mannopyranose-(2→1)-β-d-glucuronic acid-(4→1)-β-d-mannopyranose on alternating residues [51]. The chemical structure of XG is shown in Figure 2. APG is biodegradable and derived from sugar, glucose, and fatty alcohol. The raw materials of APG are starch and fat. The hydrophilic part of APG is typically sugar, and a hydrophobic end is an alkyl group with variable lengths [52]. The chemical structure is displayed in Figure 3 [52].
Table 1. Physicochemical properties of acetone and butanone [53,54,55,56,57].
Table 1. Physicochemical properties of acetone and butanone [53,54,55,56,57].
PropertiesFormic AcidAcetic AcidAcrylic AcidAcetoneButanone
Chemical FormulaH-COOHCH3-COOHCH2=CH-COOHCH3-CO-CH3CH3-CO-CH2CH3
Molar mass46.025 g/mol60.052 g/mol72.06 g/mol58.08 g/mol72.117 g/mol
AppearanceColorless liquidColorless liquidColorless liquidColorless liquidColorless liquid
Density1.22 g/cm31.049 g/cm31.05 g/cm30.78 g/cm3 at 25 °C0.80 g/cm3 at 25 °C
Boiling point100.8 °C118 °C141 °C56.05 °C79.64 °C
Viscosity1.57 cP at 20 °C1.22 cP at 20 °C1.3 cP at 20 °C0.29 cP at 25 °C0.43 cP at 25 °C

4.2. Selection of Chemical

In the microbial EOR process, tertiary oil can be recovered by reducing interfacial tension and the capillary trapping forces between oil and water with microbial products such as surfactants, polymers, alcohol, acid, and ketones. The effects of ketones and acids in green surfactant-polymer (SP) formulations in carbonates are not well understood yet require further analysis. For this reason, acetone, butanone, formic acid, acetic acids, and butanone are selected.
Alkyl polyglucoside (APG) 264 is chosen from the nonionic group because APG 264 is an eco-friendly synthetic surfactant made from coconut, palm oil, corn, potato, or wheat residues. It is entirely biodegradable. The toxicity of APG 264 is very low [52,58]. Xanthan gum (XG) is a polysaccharide, non-toxic, and biodegradable polymer. The XG is by a Xanthomonas campestris bacterium produced in a fermentation process. It is environmentally friendly and relatively cheaper than Schizophyllan and Sclerogucan [59]. That is why it is selected for this work.

4.3. Phase Behavior Modeling

The phase behavior study is an essential part for EOR application. It is very user-friendly method to predict surfactant performance. The main goals of most phase behavior studies are to identify regions of middle phase microemulsion and determine optimum salinity. The middle phase emulsion can be designed to have low Interfacial Tension (IFT) used in oil recovery processes.
The experiment was conducted in the laboratory using the concept of salinity scans and followed several steps. In this experiment, surfactant centration was fixed and NaCl concentration was varied from 1 to 7%. Each sample consists of an aqueous phase and the oil phase. It was determined that the sample size would be 9 mL, given the size of the vials: with water oil ratio (WOR) = 1, there would be 4.5 mL aqueous phases and 4.5 oil phases. The aqueous phase consisted of 3.5 mL brine and 1.0 mL surfactant. The concentration of APG 264 was 0.50%, and the NaCl concentration varied from 0 to 7%. The phase volumes were measured when they were stable and reached in equilibrium. The observed and measured phase volumes are presented in Table 2.
The development of a mathematical model for optimum salinity, interfacial tension, and residual oil calculations using experimental data and Healy and Reed correlation is shown in a process flow diagram. The workflow of the modelling is described in Figure 4. In the workflow, the solubilization parameters of oil and water in the micro-emulsion phases are first calculated using Equations (1) and (2). Then, the IFTs of oil and water in the microemulsion phases are calculated using Equations (3) and (4). After that, the solubilization parameters are plotted against salinity to obtain the optimum salinity. The optimum salinity is determined by the point of intersection between the oil and water solubilization parameter curves. Similarly, the IFTs of oil and water are plotted against salinity, and the intersecting point gives the IFT of the formulation.
P o = V o V s = Volume   of   oil   in   micro emulsion   phase Volume   of   surfactant   in   micro emusion   phase  
  P w = V w V s = Volume   of   brine   in   micro emulsion   phase Volume   of   surfactant   in   micro emusion   phase  
log σ mo = 7.058 + 6.285 0.04477   ( P o ) + 1
log σ mw = 12.856 + 12.167 0.01280   ( P w ) + 1
where
Po = Solubilization parameter of oil in the micro-emulsion;
Pw = Solubilization parameter f water in the micro-emulsion;
σ mo = IFT of oil in the micro-emulsion;
σ mw = IFT of water in the micro-emulsion.
Figure 4. Model workflow diagram.
Figure 4. Model workflow diagram.
Polymers 13 03269 g004

4.4. IFT Measurement

The Spinning Drop Tensiometer (STD) 100 from KRUSS shown in Figure 5 was used for the IFT measurements. The IFTs of the APG 264, acid, and ketone mixtures were determined to identify the optimum concentration for the core flood experiments. The density of each sample was measured at 25 °C to determine the IFT value at 25 °C. First, the cell is cleaned using hot water and soap. A 20 mL portion of each sample is required to complete a measurement. After loading the sample, the tool is calibrated before injecting the oil drop for each sample. Next, an oil drop is injected into the cell while it rotates at 500 RPM. Then, the rotation speed is set to 2800 RPM. The tool must be adjusted to centralize the oil drop. Then, adjustments are made to the temperature and camera on the tube. Then, the drop of oil is released. Finally, the IFT values are measured for at least two hours to obtain a constant IFT value. Seven samples of formic acid solution with various concentrations were prepared for IFT measurement and given in Table 7. The IFT of the 0.5% APG and 2% NaCl sample was measured when the value was closed to steady. It became nearly constant after two hours. Using this benchmark, samples 3 to 7 were measured and listed in Table 7.

4.5. Core Flood Experiment

Typically, a core flood experiment is performed in four phases. These are core preparation, water flood, chemical, and post floods. Five carbonate and two sandstone cores were cut and then dried for around four hours at 60 °C in an oven. After that, the dry weights of the cores were measured, and overburden pressure was applied to the core holder. Next, the plugs were placed in a vacuum chamber to remove the air. In the brine saturation process, 2% NaCl solution was applied to saturate each core, then flooded with Arabian light crude oil. Next, the oil-saturated plugs were put in aging for three days. Measured core properties such as length, diameter, pore-volume, dry weight, porosity, and permeability are shown in Table 3. After aging, the cores were ready for flooding.
First, a core plug was flooded with 2% NaCl solution about 2 to 4 PV when no oil was produced. Then, it was flooded by a chemical slug of about 2 to 3 PV. In this phase, seven formulations were flooded in seven core plugs. After that, the process was followed by a 1–2 PV post-flood with 2% NaCl solution to make sure all released oil inside the core and tube were accumulated in the measuring cylinder. Formulations are given in Tables 6–10. The flooding process is given in Figure 6. The schematic diagram of the core flood system is shown Figure 7. The equipment consists of four injection and three collection pumps and a temperature control chamber to simulate the required reservoir temperature. The core holder, separator, and injection and collection vessels are kept in the closed chamber to maintain the same temperature. There are three injection cylinders connected to the core holder through several valves to control the fluid flooding. These cylinders are crude oil, water, and gas. A manual hydraulic pump applies overburden pressure on the core to ensure no fluid passing outside the core. The pump is connected to two synthetic oil vessels to stabilize the pressure before the core inlet.
The detailed porosity, permeability, and oil saturation calculation and data are given in Table A5 and Table A6 and Figure A1 in Appendix B.

5. Results and Discussion

5.1. Optimum Salinity Determination

Samples were prepared with 4.5 mL of Alkyl Polygluside and NaCl formulations and 4.5 mL of crude oil. All samples were kept for three weeks to reach equilibrium at room temperature and pressure after mixing well. The aqueous, middle, and oleic phases of each system were measured and recorded.
Figure 8 indicates the phase volumes at a fixed surfactant concentration with an increasing NaCl concentration. The graph shows that the amount of micro-emulsion fluctuated with increasing NaCl. The figure demonstrates that the emulsion volume increases at the expense of oil phase and moves to the upper phase region. The maximum micro-emulsion is reached at 1% and 3% salinities. It can be concluded that optimum salinity can be found from a 2% NaCl concentration. The solubilization parameters for both oil and brine in the sample was calculated using the Healy and Reed correlation according to equations from 1 to 4. Figure 9 demonstrates the optimum salinity determination technique. After eliminating out of range data, the solubilization parameters are plotted against salinity, and best fit lines are drawn to obtain the optimum salinity. The optimum salinity for a pure APG-Crude oil system determined from Figure 9 is at 4.6% NaCl. Note that the optimum salinity will change if the formulation varies, for example, by adding alcohol, ketone, or polymer. IFT determination is not presented here because the IFT values of the surfactant and co-surfactant formulation are calculated using a spinning drop instrument.

5.2. IFT Measurement

IFT Values of Formic Acid with APG

The 0.50% of APG 264 with concentrations of formic acid ranging from 0.00% to 1.00% are measured and illustrated in Figure 10 and Table 4. The IFT declined sharply, approximately 22 dyne/cm from 23.00 dyne/cm to 0.19 dyne/cm, and then increased to 0.27 at 0.4% Formic acid. After that, it decreases to 0.20 dyne/cm at 1% formic acid concentration. The optimum concentration of formic acid was determined to be 0.60%.
IFT values of the acetone formulation (0.5% APG and 2% NaCl) and butanone solution (0.5% APG and 2 % NaCl) against Arabian light crude oil were measured using the same instrument, and the optimum point for acetone and butanone was selected. My previous paper titled “The Role of Microbial By-products in Green Enhanced Oil Recovery: Acetone and Butanone” [3] reported the detailed experimental process and result. In this article, a comparison of IFT value with formic was done and shown in Figure 11. Data are given in Table 5. It looks like the IFT behavior of formic acid, acetone, and butanone is similar.

5.3. Influence of Ketone Observed in Core Flood Experiment

5.3.1. Acetone, APG 264, and Xanthan Gum Formulation

Two core flood experiments were performed to observe the oil recovery performance of acetone. Acetone of 0.50% was blended with 0.50% APG and 1000 mg/L (1000 ppm) XG. The results are shown in Table 6 and Figure 12.
In the beginning, the core flooding instrument was calibrated with known results. First, the sandstone core was flooded with brine and oil and kept to age seven days. Next, the test sample was run. Brine (2% NaCl) was flooded; the oil recovery was 46% of OIIP. In the tertiary stage, an acetone and SP blend was injected. The oil production was 8.44% after 2.4 pore volumes of injection. The total oil recovery from the sandstone was 54.44% at the end of the brine, APG slug, and post-flood with brine process. Figure 12 demonstrates the total oil recovery from the carbonate core.

5.3.2. Butanone, APG 264, and Biopolymer Formulation

This experiment analyzed the effectiveness of butanone in the SP formulation in carbonate core. The formulation included 0.6% butanone, 0.5% APG 264, and 1000 mg/L XG. The core was vacuumed and saturated with brine (2% NaCl), then flooded with Arabian light crude oil and kept to age for seven days. First, the core was flooded with brine to recover residual oil saturation before the butanone-SP injection. The results of this flood are displayed in Figure 13 and Table 7. At the end of the waterflood, the oil recovery was 50%. Tertiary oil recovery from this formulation amounted to 23%. This was a significant amount of additional oil recovery. The total oil recovery of this formulation was 69% of OIIP. The APG, XG, and butanone combination are working effectively in the carbonate core.

5.4. Influence of Acrylic Acid in Carbonate

5.4.1. Acrylic Acid, APG 264, and Xanthan Gum Formulation

The influence of acrylic acid in the surfactant and polymer mixture was examined in this experiment. We used two formulations: the first is 0.4% acrylic acid, and the second is 0.5% acrylic acid. Both were mixed with 0.5% APG and 1000 ppm (mg/L) and flooded in two carbonate cores. The results are shown in Table 8 and Figure 14 and Figure 15.
In the secondary stage, brine (2% NaCl) was flooded; the oil recovery was 44.20% of OIIP. In the tertiary stage, an acrylic acid and SP blend was injected. The oil production was 11.86% after 3.4 pore volumes of injection. The total oil recovery from the sandstone was 58.10% at the end of the brine, acrylic acid, SP slug, and post-flood with brine process. Figure 14 demonstrates the total oil recovery from the sandstone core. Figure 15 shows the oil recovery performance of 0.4% acrylic acid and SP formulation as a function of the pore volume injected. At the end of the waterflood, the oil recovery was 47%. Tertiary oil recovery from this formulation amounted to 58%. It seems that both formulations can provide the same amount of tertiary oil.

5.4.2. Acrylic Acid, Butanol, APG 264, and Xanthan Gum Formulation

The experiment investigated the effect of butanol in acrylic acid and SP blend. In this formulation, 0.5% butanol was mixed with 0.5% acrylic acid, 0.5% APG and 1000 ppm (mg/L) and flooded in the carbonate core. The results are shown in Table 9 and Figure 16. In the secondary stage, the oil recovery was 41% of OIIP. In the tertiary stage, butanol, acrylic acid, and SP blend were injected, and the oil production was 18% after 3.0 pore volumes of injection. The total oil recovery from the sandstone was 59% at the end of the brine, butanol, acrylic acid, SP slug, and post-flood with brine process. Figure 16 demonstrates the total oil recovery from the sandstone core.

5.5. Comparison

Comparison of the Ketones, Acrylic Acids, Butanol, and SP Formulations

Five core floods were conducted to examine the oil recovery performance in carbonate cores. The core flood results for the formulations applying carbonate cores are given in Table 10, and Figure 17 and Figure 18.
Figure 17 and Figure 18 compare total oil recovery from the five carbonate cores applying the five formulations listed in Table 10. It can be seen from Figure 18 that the tertiary oil recoveries are around 8% from the acetone and SP solution and 23% from the butanone and SP solution. Similarly, from the 0.4% and 0.5% acrylic acid and SP solutions, tertiary oil recoveries are about 12% and 11%, respectively. However, the 0.5% acrylic acid, butanol and SP formulation produces approximately 18% tertiary oil from the carbonate core. It is clear from Figure 18 that the butanone and SP blend made 22% tertiary oil from the carbonate core. This formulation produced 14% higher than the acetone and SP mixture and 5% higher than the acrylic acid, butanol and SP blend. By comparing the five formulations, it can be concluded that the butanone and SP formulation is more efficient at recovering tertiary oil from the carbonate core.

5.6. Influence of Formic and Acetic Acids in Sandstone

The oil recovery potential of the green SP with formic and acetic acids formulations are also examined in the sandstone cores. The results of these combinations are listed in Table 11 and displayed in Figure 19 and Figure 20. Figure 19 demonstrates 0.5% formic acid, and SP produces ~13% tertiary oil from the sandstone core. The total oil recovery of this solution is ~55%.
However, the 0.5% acetic acid and SP mixture recovers about 15% residual oil and 59% total oil. This acetic acid and SP formulation are efficient in boosting tertiary oil from the sandstone core.

6. Conclusions

The work investigates the combined impact of green surfactant, polymer, organic acids, and ketones in boosting residual oil recovery from carbonate core via phase behavior modeling, IFT measurement, and core flood experiments. The main findings of this paper are highlighted below.
  • The sample’s solubilization parameters for oil and brine are calculated using the Healy and Reed correlation and plotted against salinity. The intersecting point of the two best-fit lines gives the optimum salinity. The optimum salinity for the APG–Crude oil system is determined at 4.6% NaCl.
  • A core flooding experiment in the carbonate core confirmed that a mixture of 0.5% APG, 1000 mg/L XG, and 0.6% acetone recovered ~8% residual oil. However, the same amount of APG and XG, blinding with 0.6% butanone, produced ~22% incremental oil, which was more than double crude oil. This formulation is efficient in recovering tertiary oil from carbonate core.
  • A combination of 0.5% APG, 1000 mg/L XG, and 0.5% acrylic acid gave ~11% incremental oil recovery. However, the same amount of APG, XG, and acrylic acid, mixing with 0.6% butanol, recovered ~18% tertiary oil. The SP, acrylic acid, and alcohol blend is effective in producing residual.
  • The blending of APG, XG with butanone is more effective in recovering residual crude than the mixture of APG, XG, acrylic acid, and butanol from the carbonate core.

7. Recommendations

  • It would be better to evaluate the oil recovery potential of formic acid and SP mixture and acetic acid and SP combination in carbonate core.
  • It recommended conducting wettability alteration analysis to understand this oil recovery controlling parameter [20,21].

Funding

The research was funded by the College of Petroleum and Geosciences (CPG) Start-Up Fund and grant number SF19005.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

I would like to thank the College of Petroleum and Geosciences (CPG) at KFUPM for providing the Start-Up Fund (SF19005) and BASF for giving the APG surfactant. I am thankful for the support of our group members at KFUPM and am also grateful to Assad Barri for his help in the lab. Last but not least, I thank my two beautiful boys, Iash Bashir and Isam Bashir, for their support in typing and editing, and my wife, Kamrun Naher, for mental support throughout the journey.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Surfactant and Polymer Field Application

Appendix A.1. Surfactant Flooding in Carbonates

Table A1. Field Application of Surfactant.
Table A1. Field Application of Surfactant.
S/LSurfactantSynthetic/Bio/GreenFormationFieldCountryRef.
1Petrostep-B100SyntheticcarbonateCretaceceous upper EdwardsUSA[60]
2Polyoxyethylene alcoholSyntheticcarbonateThe cottonwood creek[61,62]
3A combination of petroleum sulfonate and alkylaryl ether sulphateSynthetic Bob slaughter block[63]
4Non-ionic ethoxy alcoholGreencarbonateYates field[61]
5ORS-4lSyntheticsandstoneTanner[64]
6Petroleum sulfonate, lignosulfonate, alkyl benzene sulfonate (ABS), petroleum carboxylate, bio-surfactantSynthetic, Green and BiosandstoneDagingChina[65,66,67,68,69]
7Petroleum sulfonateSyntheticsandstoneKaramay[70,71]
8Petroleum sulfonateSyntheticsandstoneVirajIndia[72]
9Petrostep B-100sandstoneWest kiehl[73]
10Petroleum sulonateSyntheticsandstoneMinasIndonesia[74]
11carbonatesBaturaja[75]
12SS-6066Synthetic Chihuido de la sierra negraArgentina[76]
13Olefin sulfonateSyntheticsandstonebrambergeGermany[77]

Appendix A.2. Polymer Flooding in Carbonates

Table A2. Field Application of Polymer.
Table A2. Field Application of Polymer.
S/LPolymerSynthetic/BioFormationFieldCountryRef
1HPAMSyntheticSandstoneNorth BurbankUSA
2HPAMSyntheticSandstoneWest Kiehli[66]
3PAMSyntheticSandstoneCambridge Minnelusa[78]
4PAMSyntheticSandstoneTanner[65]
5HPAMSyntheticSandstoneTambaredjo[79]
6HPAMSyntheticSandstoneDaqingChina[61]
7HPAMSyntheticSandstoneGudong[60]
8HPAMSyntheticSandstoneXing Long Tai[63]
9XanthanBioSandstoneVoador Offshore FeildBrazil[80]
10HPAMSyntheticSandstoneVirajIndia[69]
11PAMSyntheticSandstoneSanand[70]
12HPAMSyntheticSandstoneMatzenAustria[73]
13HPAMSyntheticSandstonePelican LakeCanada[79]
14PAMSyntheticSandstoneDavid Pool[75]
15XanthanBioSandstoneEddesse-NordGermany[80]
16Hydroxyethylcellulose (HEC)BioSandstoneRussia Romashkino FieldRussia[76,80]

Appendix A.3. SP Flooding in Carbonates

Table A3. Application of the SP or ASP formulation in carbonate fields.
Table A3. Application of the SP or ASP formulation in carbonate fields.
SP or ASPSynthetic/Bio/GreenField and CountryFormationOil Recovery % Ref.
S: Petroleum sulfonate and alkyl ether sulfate
P: PAM
SyntheticWichita
County Regular Gunsight reservoir, Texas, USA
Carbonate22.0[81]
S: Petroleum sulfonates and Alkyl ether sulfate
P: PAM
SyntheticWesgum field
Smackover reservoir, Arkansas, USA
Carbonate26.7[82]
S: Alkyl ether sulfates and Witco petroleum sulfonateSyntheticBob slaughter block lease
San andres reservoir, Wyming, USA
Carbonate12.0[63]
Na2CO3 and anionic polymersSyntheticIsenhour, Wyming, USACarbonate26.4[83]
A: Na2CO3
S: Petrostep B-100
P: Alcoflood1175A
SyntheticCambridge Minnelusa, Wyming, USACarbonate36.0[78]
S: nonionic ethoxy alcohol (Shell 91-8) and Stepan CS-460 anionic ethoxy sulfate surfactantSyntheticYates field, san andres reservoir, Texas, USACarbonate15.0[84,85]
S: Nonionic polyoxyethylene alcohol (POA)SyntheticCottonwood Creek field, Wyming, USA Bighorn basin Carbonate10.4[31,32]
A: Sodium carbonate alkaline solution
S: 0.5% of surfactants in 200 gallons per ft of diesel oil and 0.5% of surfactants in 55 gallons per ft of xylene.
SyntheticMauddud carbonate reservoir, BahrainCarbonate10–15[86]
S: Nonionic surfactantsSyntheticSemoga field, baturaja Formation, IndonesiaCarbonate58,000 bbl per month[70]
Table A4. Laboratory studies of the SP and ASP method in carbonate.
Table A4. Laboratory studies of the SP and ASP method in carbonate.
SP or ASP FormulationSynthetic/Bio/GreenReservoirPermeability (mD)Oil Recovery % Ref.
Amphoteric petrostep B-100 surfactant and pusher 700E Polymer and sodium tripolyphosphate and sodium carbonate alkali.SyntheticCretaceous
Upper Edwards Reservoir
Carbonate Formations
From central Texas
7545[60]
Cationic surfactants of the type tetra alkyl ammonium
Anionic surfactants
SyntheticOutcrop chalk
From stevns Klint
Copenhagen
2–710–75[87]
Anionic (ethoxylated and propoxylated sulfate) surfactants and sodium carbonate alkali mixture.SyntheticDolomite cores40–12240–50[21]
Nonionic ethoxy alcohol surfactantsSyntheticDolomite cores -[88]
Anionic ethoxylated and propoxylated sulfate surfactants
Cationic (CTAB) surfactants
SyntheticCalcite, Lithographic
Limestone, Marble
Dolomite plates
35–55[89]
Cationic C12TAB surfactantsSyntheticOutcrop chalk from Stvens Klint Copenhagen1–350–90[90]
Cationic C12TAB surfactantsSyntheticOutcrop chalk2–320–60[91]
Five anionic (sulfonate, disulfonate and sulfate) surfactantsSyntheticCalcites plates limestones cores1560–75[92]
Anionic (sulfonate, disulfonate and sulfate) surfactantsSyntheticCalcites plates limestones cores1530–50[93]
Two anionic surfactants (ethoxylated sulfonate: AV-70, AV-150)
Three nonionic surfactants (NP ethoxylate, 15-s-ethoxylate, TDA 30EO)
Four cationic surfactants (CTAB, DTAB, Arquad C-50, Arquad T-50) surfactants
SyntheticLimestone 70–80[94]
Anionic surfactants: alkyl propoxy sulfates and their blends with internal olefin sulfonates, alkyl xylene sulfonates SyntheticSilurian dolomite outcrop cores19526–80[95]
Two anionic and two nonionic surfactants SyntheticSiliceous and carbonate shale cores -[96]
Anionic Guerbet alkoxy carboxylate (GAC) surfactantsSyntheticSilurian dolomite (478 mD)
Estaillade limestone core
478 and 18790–94.5[97]
Nonionic branched nonylphenol ethoxylates (Huntsman SURFONICS N-120 & Huntsman SURFONICS N-150) and branched isotridecyl ethoxylate (Huntsman SURFONICS TDA-9) surfactantsSyntheticSACROC carbonate cores13–16-[98]

Appendix B. Core Properties Measurement

Appendix B.1. Properties and Calculation Procedure of Core Sample 4

The Indiana limestone core of 6-inch length and 1.5-inch diameter was used in this experiment. First, the core plug was cut, dried, and vacuumed, and then properties were measured. Calculation only sample 4 is described here.

Appendix B.2. Pore Volume

Weight before saturation = 372.38 g
Weight after saturation = 404.6 g
Pore   Volume = W after W after density   = 404.6 372.38 1.0125 = 32   cm 3

Appendix B.3. Porosity

  V b = π 4 D 2 L = π 4 × ( 1.5 × 2.54 ) 2 × ( 6 × 2.54 ) = 173.75   c m 3  
= V p V b = 32 173.75 = 18.4 %  

Appendix B.4. Permeability

Table A5. Permeability variations with different flow rates.
Table A5. Permeability variations with different flow rates.
Flow Rate (cc/min)Inlet Pressure (psi)Outlet Pressure (psi)Permeability (mD)
41017.41012.3205.49
31040.41036178.67
21051.31048.8149.71
11099.71098.7262.00
From Darcy’s law
K = 1000 q μ L A P
K: permeability, md
q: flow rate, cc/s
μ: viscosity, cp
L: length of the core, cm
A: area of the core, cm2
∆P: pressure drop, atm
Figure A1. Permeability calculations.
Figure A1. Permeability calculations.
Polymers 13 03269 g0a1
Figure A2 shows that our core sample has an average permeability of 185.966 md.
Figure A2. Average permeability calculation.
Figure A2. Average permeability calculation.
Polymers 13 03269 g0a2

Appendix B.5. Saturation

Table A6. Saturation calculations.
Table A6. Saturation calculations.
PV (mL)Brine Produced (mL)Tube Volume (mL)Brine Produced from Core (mL)SoiSwi
322811170.530.47
Volume of oil = 17 mL
oil   saturation =   V o V p = 17 32 = 0.53125
water   saturation = 1 oil   saturation = 1 0.53125 = 0.46875

References

  1. Green, D.W.; Willhite, G.P. Enhanced Oil Recovery; SPE: Richardson, TX, USA, 1998. [Google Scholar]
  2. Lake, L.W.; Johns, R.; Rossen, B.; Pope, G. Fundamental of Enhanced Oil Recovery; SPE: Richardson, TX, USA, 2014. [Google Scholar]
  3. Haq, B. The role of microbial products in green enhanced oil recovery: Acetone and butanone. Polymers 2021, 13, 1946. [Google Scholar] [CrossRef]
  4. Haq, B.; Liu, J.; Liu, K. Green enhanced oil recovery (GEOR). APPEA J. 2017, 57, 150–170. [Google Scholar] [CrossRef]
  5. Ahmadi, M.A.; Shadizadeh, S.R. Nano-surfactant flooding in carbonate reservoirs: A mechanistic study. Eur. Phys. J. Plus 2017, 132, 246. [Google Scholar] [CrossRef]
  6. Zhou, X.; Han, M.; Fuseni, A.B.; Yousef, A.A. Adsorption-desorption of an amphoteric surfactant onto permeable carbonate rocks. In Proceedings of the SPE Improved Oil Recovery Symposium, Tulsa, OK, USA, 14–18 April 2012; Volume 1, pp. 552–572. [Google Scholar] [CrossRef]
  7. Diab, W.N.; Al-Shalabi, E.W. Recent developments in polymer flooding for carbonate reservoirs under harsh conditions. In Proceedings of the Offshore Technology Conference Brasil, Rio de Janeiro, Brazil, 29–31 October 2019. [Google Scholar] [CrossRef]
  8. Lai, H. Defining Enhanced Oil Recovery 1–414. Sci. Eng. 2007, 158, 680–692. [Google Scholar] [CrossRef]
  9. Sorbie, K.S.; Phil, D. Polymer-Improved Oil Recovery; Springer Science & Business Media: Berlin/Heldelberg, Germany, 1991. [Google Scholar]
  10. Manrique, E.J.; Muci, V.E.; Gurfinkel, M.E. EOR Field Experiences in Carbonate Reservoirs in the United States. In Proceedings of the SPE/DOE Symposium on Improved Oil Recovery, Tulsa, OK, USA, 22–26 April 2006; pp. 667–686. [Google Scholar]
  11. Ahmadi, M.A.; Sheng, J. Performance improvement of ionic surfactant flooding in carbonate rock samples by the use of nanoparticles. Pet. Sci. 2016, 13, 725–736. [Google Scholar] [CrossRef] [Green Version]
  12. Cui, L.; Ma, K.; Abdala, A.A.; Lu, L.J.; Tanakov, I.; Biswal, S.L.; Hirasaki, G.J. Adsorption of a Switchable Cationic Surfactant on Natural Carbonate Minerals. SPE J. 2015, 20, 70–78. [Google Scholar] [CrossRef] [Green Version]
  13. Meyers, K.O.; Oil, A.; Co, G.; Salter, S.; Oil, A. The Effect of Oil/Brine Ratio on Surfactant Adsorption from Microemulsion. Soc. Pet. Eng. J. 1981, 21, 500–512. [Google Scholar] [CrossRef]
  14. Celik, M.S.; Somasundaran, P. Wettability of reservoir minerals by flotation and correlation with surfactant adsorption. In Proceedings of the SPE Oilfield and Geothermal Symposium, Standford, CA, USA, 28–30 May 1980. [Google Scholar] [CrossRef]
  15. Sandengen, K.; Widerøe, H.C.; Nurmi, L.; Hanski, S. Hydrolysis kinetics of ATBS polymers at elevated temperature, via 13C NMR spectroscopy as the basis for accelerated aging tests. J. Pet. 2017, 158, 680–692. [Google Scholar]
  16. Bennetzen, M.V.; Gilani, S.F.; Mogensen, K.; Ghozali, M.; Bounoua, N. Successful Polymer Flooding of Low-Permeability, Oil-Wet, Carbonate Reservoir Cores. In Proceedings of the Abu Dhabi International Petroleum Exhibition and Conference, Abu Dhabi, United Arab Emirates, 10–13 November 2014; pp. 1–28. [Google Scholar]
  17. Gao, C. The viscosity of partially hydrolyzed polyacrylamide under shearing and heat. J. Pet. Explor. Prod. Technol. 2013, 3, 203–206. [Google Scholar] [CrossRef] [Green Version]
  18. Vermolen, E.C.M.; Van Haasterecht, M.J.T.; Masalmeh, S.K.; Faber, M.J.; Boersma, D.M.; Gruenenfelder, M. Pushing the envelope for polymer flooding towards high-temperature and high-salinity reservoirs with polyacrylamide based terpolymers. In Proceedings of the SPE Middle East Oil and Gas Show and Conference, Manama, Bahrain, 25–28 September 2011; Volume 2, pp. 1001–1009. [Google Scholar] [CrossRef]
  19. Roehl, P.O.; Choquette, P.W. Carbonate Petroleum Reservoirs; Springer: New York, NY, USA, 1985. [Google Scholar]
  20. Allan, J.; Sun, S.Q. Controls on Recovery Factor in Fractured Reservoirs: Lessons Learned from 100 Fractured Fields. In Proceedings of the SPE Annual Technical Conference and Exhibition of the Denver, Denver, CO, USA, 5–8 October 2003. [Google Scholar]
  21. Hirasaki, G.; Zhang, D.L. Surface Chemistry of Oil Recovery from Fractural, Oil-Wet, Carbonate Formations. In Proceedings of the International Symposium on Oilfield Chemistry, Houston, TX, USA, 5–7 February 2003. [Google Scholar]
  22. Zhong, X.; Pu, H.; Zhou, Y.; Petroleum, N.; Zhao, J.X. Static Adsorption of Surfactants on Bakken Rock Surfaces in High Temperature, High Salinity Conditions. In Proceedings of the SPE International Conference on Oilfield Chemistry, Galveston, TX, USA, 8–9 April 2019; pp. 1–10. [Google Scholar]
  23. Mannhardt, K.; Novosad, J.J.; Jha, K.N.N. Adsorption of Foam-forming Surfactants in Berea Sandstone. J. Can. Pet. Technol. 1994, 33, PETSOC-94-02-04. [Google Scholar] [CrossRef]
  24. Trogus, F.J.; Sophany, T.; Schechter, R.S.; Wade, W.H. Static and Dynamic Adsorption of Anionic and Nonionic Surfactants. Soc. Pet. Eng. AIME J. 1977, 17, 337–344. [Google Scholar] [CrossRef]
  25. Tabatabal, A.; Gonzalez, M.V.; Harwell, J.H.; Scamehorn, J.F. Reducing Surfactant Adsorption in Carbonate Reservoirs. SPE Reserv. Eng. 1993, 8, 117–122. [Google Scholar]
  26. Shamsijazeyi, H.; Hirasaki, G.J.; Verduzco, R. SPE 164061 Sacrificial Agent for Reducing Adsorption of Anionic Surfactants. In Proceedings of the SPE International Symposium on Oilfield Chemistry, The Woodlands, TX, USA, 8–10 April 2013. [Google Scholar]
  27. Krumrine, P.H.; Falcone, J.S.; Campbell, T.C. Surfactant Flooding 2*: The Effect of Alkaline Additives on Permeability and Sweep Efficiency. Soc. Pet. Eng. J. 1982, 22, 983–992. [Google Scholar] [CrossRef]
  28. Tsau, J.; Syahputra, A.E.; Grigg, R.B.; Recovery, P. SPE 59365 Economic Evaluation of Surfactant Adsorption in CO2 Foam Application. In Proceedings of the SPE/DOE Improved Oil Recovery Symposium, Tulsa, OK, USA, 3–5 April 2000. [Google Scholar]
  29. Pal, S.; Mushtaq, M.; Banat, F.; Al Sumaiti, A.M. Review of surfactant-assisted chemical enhanced oil recovery for carbonate reservoirs: Challenges and future perspectives. Pet. Sci. 2018, 15, 77–102. [Google Scholar] [CrossRef] [Green Version]
  30. Alfazazi, U.; AlAmeri, W.; Hashmet, M.R. Experimental investigation of polymer flooding with low-salinity preconditioning of high temperature–high-salinity carbonate reservoir. J. Pet. Explor. Prod. Technol. 2019, 9, 1517–1530. [Google Scholar] [CrossRef] [Green Version]
  31. Al-Shalabi, E.W.; Diab, W.D. SPE-192700-MS Robust EOR Screening Workflow for a Carbonate Green Field: A Case Study. In Proceedings of the Offshore Technology Conference Brasil, Rio de Janeiro, Brazil, 29–31 October 2018. [Google Scholar]
  32. Ahmadi, M.A.; Sheng, J. Numerical simulation and laboratory evaluation of alkali–surfactant–polymer and foam flooding. Int. J. Environ. Sci. Technol. 2016, 17, 1123–1136. [Google Scholar]
  33. Pan, F.; Zhang, Z.; Zhang, X.; Davarnanah, A. Impact of anionic and cationic surfactants interfacial tension on the oil recovery enhancement. Power Technol. 2020, 373, 93–98. [Google Scholar] [CrossRef]
  34. Davarpanah, A. Parametric Study of Polymer-Nanoparticles-Assisted Injectivity Performance for Axisymmetric Two-Phase Flow in EOR Processes. Nanomaterials 2020, 10, 1818. [Google Scholar] [CrossRef]
  35. OPEC. The Organization of the Petroleum Exporting Countries (OPEC); Infobase Publishing: Vienna, Austria, 2016. [Google Scholar]
  36. Learsy, R.J. Oil and Finance: The Epic Corruption; iUniverse: Bloomington, IN, USA, 2011; p. 89. [Google Scholar]
  37. Murty, C.R.K.; Fathalla, M.; Al-Mahroos, F.M. Reservoir Performance Analysis of Arab “D” Zone (Bahrain Field). In Proceedings of the Middle East Oil Technical Conference and Exhibition, Manama, Bahrain, 14–17 March 1983. [Google Scholar] [CrossRef]
  38. Ma, S.M.; Al-Muthana, A.S.; Dennis, R.N. Use of Core Data in Log Lithology Calibration: Arab-D Reservoir, Abqaiq and Ghawar Fields. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 29 September–2 October 2002. [Google Scholar] [CrossRef]
  39. Okasha, T.M.; Alshiwaish, A. Effect of Brine Salinity on Interfacial Tension in Arab-D Carbonate Reservoir, Saudi Arabia. In Proceedings of the SPE Middle East Oil and Gas Show and Conference, Manama, Bahrain, 15–18 March 2009. [Google Scholar] [CrossRef]
  40. Grover, G.A. Abqaiq Hajifa Reservoir: Geologic Attributes Controlling Hydrocarbon Production and Water Injection. In Proceedings of the Middle East Oil Show, Manama, Bahrain, 3–6 April 1993. [Google Scholar] [CrossRef]
  41. Stenger, B.A.; Pham, T.R.; Al-Sahhaf, A.A.; Al-Muhaish, A.S. Assessing the Oil Water Contact in Haradh Arab-D. In Proceedings of the SPE Annual Technical Conference and Exhibition, New Orleans, LA, USA, 30 September–3 October 2001. [Google Scholar] [CrossRef]
  42. Stenger, B.A.; Ameen, M.S.; Al-Qahtani, S.; Pham, T.R. Pore Pressure Control of Fracture Reactivation in the Ghawar Field, Saudi Arabia. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 29 September–2 October 2002. [Google Scholar] [CrossRef]
  43. Jacob, S.; Baim, A.S.; Al-Bani, F.A. Drilling and Completing Intelligent Multilateral MRC wells in Haradh Inc-3. In Proceedings of the SPE/IADC Drilling Conference, Amsterdam, The Netherlands, 20–22 February 2007. [Google Scholar] [CrossRef]
  44. Al-Afaleg, N.I.; Al-Garni, S.; BasemAhmed, R.; Al-Towailib, A. Successful Integration of Sparsely Distributed Core and Well test Derived Permeability Data in a Viable Model of a Giant Carbonate Reservoir. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 29 September–2 October 2002. [Google Scholar] [CrossRef]
  45. Al-Mulhim, W.A.; Al-Ajmi, F.A.; Al-Shehab, M.A.; Pham, T.R. Best Practices in Khurais Complex Development. In Proceedings of the SPE/DGS Saudi Arabia Section Technical Symposium and Exhibition, Al-Khobar, Saudi Arabia, 4–7 April 2010. [Google Scholar] [CrossRef]
  46. Alotaibi, A.Z.; Alyahya, A.M. Offshore Gas Program Acceleration (Karan Gas Program). In Proceedings of the Offshore Technology Conference (OTC), Houston, TX, USA, 4–7 May 2015. [Google Scholar] [CrossRef]
  47. Almadi, S.M.; Alaidarous, O.; Al-Dhubaib, T.A.; AhmadHusain, H.A.; Al-Amri, A.D. I-Field Data Acquisition and Delivery Infrastructure: Khursaniyah Field Best in Class Practices. In Proceedings of the SPE Intelligent Energy Conference and Exhibition, Utrecht, The Netherlands, 23–25 March 2010. [Google Scholar] [CrossRef]
  48. Wilson, A.O. Jurassic Arab-C and -D Carbonate Petroleum Reservoirs, Qatif Field, Saudi Arabia. In Proceedings of the Middle East Technical Conference and Exhibition, Manama, Bahrain, 9–12 March 1981. [Google Scholar] [CrossRef]
  49. Al-Dossary, F.M.; Al-Ghamdi, A.A. The Operation of the Multiphase Flow Meters on the Sour Environment of Qatif Field: Case Study. In Proceedings of the SPE Intelligent Energy Conference and Exhibition, Utrecht, The Netherlands, 23–25 March 2010. [Google Scholar] [CrossRef]
  50. Arukhe, J.O.; Duthie, L.S. Technology Improvements in the Giant Manifa Field Maturation Project. In Proceedings of the SPE Middle East Oil & Gas Show and Conference, Manama, Bahrain, 8–11 March 2015. [Google Scholar] [CrossRef]
  51. Wever, D.A.Z.; Picchioni, F.; Broekhuis, A.A. Polymers for enhanced oil recovery: A paradigm for structure-property relationship in aqueous solution. Prog. Polym. Sci. 2011, 36, 1558–1628. [Google Scholar] [CrossRef]
  52. Wikipedia. The Free Encyclopedia. Available online: https://en.wikipedia.org/wiki/Alkyl_polyglycoside (accessed on 1 September 2021).
  53. Wikipedia. The Free Encyclopedia. Available online: https://en.wikipedia.org/wiki/Acetone (accessed on 1 September 2021).
  54. Wikipedia. The Free Encyclopedia. Available online: https://en.wikipedia.org/wiki/Butanone (accessed on 1 September 2021).
  55. Wikipedia. The Free Encyclopedia. Available online: https://en.wikipedia.org/wiki/Acetic_acid (accessed on 1 September 2021).
  56. Wikipedia. The Free Encyclopedia. Available online: https://en.wikipedia.org/wiki/Formic_acid (accessed on 1 September 2021).
  57. Wikipedia. The Free Encyclopedia. Available online: https://en.wikipedia.org/wiki/Acrylic_acid (accessed on 1 September 2021).
  58. Haq, B. The Role of Green Surfactants in Microbial Enhanced Oil Recovery. Ph.D. Thesis, University of Western Australia, Perth, Australia, 2020; p. 189. [Google Scholar]
  59. Muhammed, N.S.; Al Shehri, D.; Rahaman, M.M.; Keshavarz, A.; Hossain, S.M.Z. Comparative study of green and synthetic polymers for Enhanced Oil Recovery. Polymers 2020, 12, 2429. [Google Scholar] [CrossRef] [PubMed]
  60. Olsen, D.; Hicks, M.; Hurd, B.; Sinnokrot, A.; Sweigart, C. Design of a novel flooding system for an oil-wet Central Texas carbonate reservoir. In Proceedings of the SPE/DOE Enhanced Oil Recovery Symposium, Tulsa, OK, USA, 22–25 April 1990. [Google Scholar]
  61. Xie, X.; Weiss, W.W.; Tong, Z.; Morrow, N.R. Improved oil recovery from carbonate reservoirs by chemical stimulation. In Proceedings of the SPE/DOE Symposium on Improved Oil Recovery, Tulsa, OK, USA, 17–21 April 2004. [Google Scholar]
  62. Weiss, W.W.; Xie, X.; Weiss, J.; Subramanian, V.; Taylor, A.R.; Edens, F.J. Artificial intelligence used to evaluate 23 single-well surfactant-soak treatments. SPE Reserv. Eval. Eng. 2006, 9, 209–216. [Google Scholar] [CrossRef]
  63. Adams, W.T.; Schievelbein, V.H. Surfactant Flooding Carbonate Reservoirs. SPE Reserv. Eng. 1987, 2, 619–626. [Google Scholar] [CrossRef]
  64. Pitts, M.; Dowling, P.; Wyatt, K.; Surkalo, H.; Adams, K. Alkaline-surfactant-polymer flood of the Tanner Field. In Proceedings of the SPE/DOE Symposium on Improved Oil Recovery, Tulsa, OK, USA, 22–26 April 2006. [Google Scholar]
  65. Chang, H.L.; Zhang, Z.Q.; Wang, Q.M.; Xu, Z.S.; Guo, Z.D.; Sun, H.Q.; Cao, X.L.; Qiao, Q. Advances in Polymer Flooding and Alkaline/Surfactant/Polymer Processes as Developed and Applied in the People’s Republic of China. J. Pet. Technol. 2006, 58, 84–89. [Google Scholar] [CrossRef]
  66. Demin, W.; Jiecheng, C.; Junzheng, W.; Zhenyu, Y.; Yuming, Y. Summary of ASP pilots in Daqing oil field. In Proceedings of the SPE Asia Pacific Improved Oil Recovery Conference, Kuala Lumpur, Malaysia, 25–26 October 1999. [Google Scholar]
  67. Li, H.; Liao, G.; Han, P.; Yang, Z.; Wu, X.; Chen, G.; Xu, D.; Jin, P. Alkline/Surfactant/Polymer (ASP) Commercial Flooding Test in Central Xing2 Area of Daqing Oilfield. In Proceedings of the SPE International Improved Oil Recovery Conference in Asia Pacific, Kuala Lumpur, Malaysia, 20–21 October 2003. [Google Scholar]
  68. Li, D.; Shi, M.-Y.; Wang, D.; Li, Z. Chromatographic separation of chemicals in alkaline surfactant polymer flooding in reservoir rocks in the Daqing oil field. In Proceedings of the SPE International Symposium on Oilfield Chemistry, The Woodlands, TX, USA, 20–22 April 2009. [Google Scholar]
  69. Pu, H. An Update and Perspective on Field-Scale Chemical Floods in Daqing Oilfield China. In Proceedings of the SPE Middle East Oil and Gas Show and Conference, Manama, Bahrain, 15–18 March 2009. [Google Scholar]
  70. Qi, Q.; Gu, H.; Li, D.; Dong, L. The pilot test of ASP combination flooding in Karamay oil field. In Proceedings of the International Oil and Gas Conference and Exhibition in China, Beijing, China, 7–10 November 2000. [Google Scholar]
  71. Gu, H.; Yang, R.; Guo, S.; Guan, W.; Yue, X.; Pan, Q. Study on reservoir engineering: ASP flooding pilot test in Karamay Oilfield. In Proceedings of the SPE International Oil and Gas Conference and Exhibition in China, Beijing, China, 2–6 November 1998. [Google Scholar]
  72. Pratap, M.; Gauma, M. Field Implementation of Alkaline-Surfactant-Polymer (ASP) Flooding: A maiden effort in India. In Proceedings of the SPE Asia Pacific Oil and Gas Conference and Exhibition, Perth, Australia, 18–20 October 2004. [Google Scholar]
  73. Meyers, J.J.; Pitts, M.J.; Wyatt, K. Alkaline-Surfactant-Polymer Flood of the West Kiehl, Minnelusa Unit. In Proceedings of the SPE/DOE Enhanced Oil Recovery Symposium, Tulsa, OK, USA, 22–24 April 1992; pp. 423–435. [Google Scholar]
  74. Bou-Mikael, S.; Asmadi, F.; Marwoto, D.; Cease, C. Minas surfactant field trial tests two newly designed surfactants with high EOR potential. In Proceedings of the SPE Asia Pacific Oil and Gas Conference and Exhibition, Brisbane, Australia, 16–18 October 2000. [Google Scholar]
  75. Rilian, N.A.; Sumestry, M.; Wahyuningsih, W. Surfactant Stimulation to Increase Reserves in Carbonate Reservoir: “A Case Study in Semoga Field”. In Proceedings of the SPE EUROPEC/EAGE Annual Conference and Exhibition, Barcelona, Spain, 14–17 June 2010. [Google Scholar]
  76. Zaitoun, A.; Fonseca, C.; Berger, P.; Bazin, B.; Monin, N. New surfactant for chemical flood in high-salinity reservoir. In Proceedings of the International Symposium on Oilfield Chemistry, Houston, TX, USA, 5–7 February 2003. [Google Scholar]
  77. Tabary, R.; Douarche, F. Design of a surfactant/polymer process in a hard brine context: A case study applied to Bramberge reservoir. In Proceedings of the SPE EOR Conference at Oil and Gas West Asia, Muscat, Oman, 16–18 April 2012; pp. 1–13. [Google Scholar]
  78. Jay, V.; Jim, T.; Bob, V.; Pitts, M.; Kon, W.; Harry, S.; David, P. Alkaline-surfactant-polymer flooding of the Cambridge Minnelusa Field. SPE Reserv. Eval. Eng. 2000, 3, 552–558. [Google Scholar]
  79. Mogollon, J.L.; Lokhandwala, T. Rejuvenating Viscous Oil Reservoirs by Polymer Injection: Lessons Learned in the Field. In Proceedings of the 2013 SPE Enhanced Oil Recovery Conference, Kuala Lumpur, Malaysia, 2–4 July 2013. [Google Scholar]
  80. Abbas, S.; Donovan, J.; Sanders, A. Applicability of Hydroxyethylcellulose Polymers for Chemical EOR. In Proceedings of the SPE Enhanced Oil Recovery Conference, Kuala Lumpur, Malaysia, 2–4 July 2013. [Google Scholar]
  81. Leonard, J. Annual production report. Enhanced oil recovery. Oil Gas J. 1984, 82, 100–101. [Google Scholar]
  82. Leonard, J. Production/enhanced oil recovery report. Oil Gas J. 1986, 84, 94–95. [Google Scholar]
  83. Doll, T.E. An update of the polymer-augmented alkaline flood at the Isenhour Unit, Sublette County, Wyoming. SPE Reserv. Eng. 1988, 3, 604–608. [Google Scholar] [CrossRef]
  84. Yang, H.D.; Wadleigh, E.E. Dilute surfactant IOR—Design improvement for massive, fractured carbonate applications. In Proceedings of the SPE International Petroleum Conference and Exhibition in Mexico, Villahermosa, Mexico, 1–3 February 2000. [Google Scholar] [CrossRef]
  85. Mathiassen, O.M. CO2 as Injection Gas for Enhanced Oil Recovery and Estimation of the Potential on the Norwegian Continental Shelf. Ph.D. Thesis, Department Petroleum Engineering and Applied Geophysics, Norwegian University of Science and Technology, Trondheim, Norway, 2003. [Google Scholar]
  86. Zubari, H.K.; Sivakumar, V.C.B. Single well tests to determine the efficiency of alkaline-surfactant injection in a highly oil-wet limestone reservoir. In Proceedings of the Middle East Oil Show, Manama, Bahrain, 9–12 June 2003. [Google Scholar] [CrossRef]
  87. Standnes, D.C.; Austad, T. Wettability alteration in carbonates: Interaction between cationic surfactant and carboxylates as a key factor in wettability alteration from oil-wet to water-wet conditions. Colloids Surf. A 2003, 216, 243–259. [Google Scholar] [CrossRef]
  88. Vijapurapu, C.S.; Rao, D.N. Compositional effects of fluids on spreading, adhesion and wettability in porous media. Colloids Surf. A 2004, 241, 335–342. [Google Scholar] [CrossRef]
  89. Seethepalli, A.; Adibhatla, B.; Mohanty, K.K. Physicochemical interactions during surfactant flooding of fractured carbonate reservoirs. SPE J. 2004, 9, 411–418. [Google Scholar] [CrossRef]
  90. Hognesen, E.J.; Olsen, M.; Austad, T. Capillary and gravity dominated flow regimes in displacement of oil from an oil-wet chalk using cationic surfactant. Energy Fuels 2006, 20, 1118–1122. [Google Scholar] [CrossRef]
  91. Strand, S.; Hognesen, E.J.; Austad, T. Wettability alteration of carbonates: Effects of potential determining ions (Ca2+ and SO42−) and temperature. Colloids Surf. A 2006, 275, 1–10. [Google Scholar] [CrossRef]
  92. Gupta, R.; Mohanty, K. Temperature effects on surfactant-aided imbibition into fractured carbonates. SPE J. 2010, 15, 587–597. [Google Scholar] [CrossRef]
  93. Gupta, R.; Mohanty, K.K. Wettability alteration mechanism for oil recovery from fractured carbonate rocks. Transp. Porous Media 2011, 87, 635–652. [Google Scholar] [CrossRef]
  94. Sharma, G.; Mohanty, K.K. Wettability alteration in high temperature and high salinity carbonate reservoirs. SPE J. 2013, 18, 646–655. [Google Scholar] [CrossRef]
  95. Sagi, A.R.; Puerto, M.C.; Bian, Y.; Miller, C.A.; Hirasaki, G.J.; Salehi, M.; Thomas, C.P.; Kwan, J.T. Laboratory studies for surfactant flood in low-temperature, low-salinity fractured carbonate reservoir. In Proceedings of the SPE International Symposium on Oilfield Chemistry, The Woodlands, TX, USA, 8–10 April 2013. [Google Scholar] [CrossRef]
  96. Alvarez, J.O.; Neog, A.; Jais, A.; Schechter, D.S. Impact of surfactants for wettability alteration in stimulation fluids and the potential for surfactant EOR in unconventional liquid reservoirs. In Proceedings of the SPE Unconventional Resources Conference, The Woodlands, TX, USA, 1–3 April 2014. [Google Scholar] [CrossRef]
  97. Lu, J.; Liyanage, P.J.; Solairaj, S.; Adkins, S.; Arachchilage, G.P.; Kim, D.H.; Britton, C.; Weerasooriya, U.; Pope, G.A. New surfactant developments for chemical enhanced oil recovery. J. Pet. Sci. Eng. 2014, 120, 94–101. [Google Scholar] [CrossRef]
  98. McLendon, W.J.; Koronaios, P.; Enick, R.M.; Biesmans, G.; Salazar, L.; Miller, A.; Soong, Y.; McLendon, T.; Romanov, V.; Crandall, D. Assessment of CO2 soluble nonionic surfactants for mobility reduction using mobility measurements and CT imaging. J. Pet. Sci. Eng. 2014, 119, 196–209. [Google Scholar] [CrossRef]
Figure 1. GEOR classification.
Figure 1. GEOR classification.
Polymers 13 03269 g001
Figure 2. Chemical structure of xanthan gum [51].
Figure 2. Chemical structure of xanthan gum [51].
Polymers 13 03269 g002
Figure 3. Chemical structure of alkyl polyglucoside [52].
Figure 3. Chemical structure of alkyl polyglucoside [52].
Polymers 13 03269 g003
Figure 5. Spinning drop tensiometer (STD 100).
Figure 5. Spinning drop tensiometer (STD 100).
Polymers 13 03269 g005
Figure 6. Core flooding procedure.
Figure 6. Core flooding procedure.
Polymers 13 03269 g006
Figure 7. A schematic diagram of a core flood system.
Figure 7. A schematic diagram of a core flood system.
Polymers 13 03269 g007
Figure 8. Phase diagram of 0.5% APG, crude oil, and 0–7% NaCl mixture.
Figure 8. Phase diagram of 0.5% APG, crude oil, and 0–7% NaCl mixture.
Polymers 13 03269 g008
Figure 9. Optimum salinity determination.
Figure 9. Optimum salinity determination.
Polymers 13 03269 g009
Figure 10. IFT measurements as a function of formic acid concentration.
Figure 10. IFT measurements as a function of formic acid concentration.
Polymers 13 03269 g010
Figure 11. Comparison of IFT values.
Figure 11. Comparison of IFT values.
Polymers 13 03269 g011
Figure 12. Total oil production from the carbonate core after the brine, acetone, and SP blend, followed by a post-flood with brine.
Figure 12. Total oil production from the carbonate core after the brine, acetone, and SP blend, followed by a post-flood with brine.
Polymers 13 03269 g012
Figure 13. Total oil production from the carbonate core after brine, butanone, and SP mixture, followed by a post-flood with brine.
Figure 13. Total oil production from the carbonate core after brine, butanone, and SP mixture, followed by a post-flood with brine.
Polymers 13 03269 g013
Figure 14. Total oil production from the carbonate core after brine, 0.4% Acrylic acid, and SP mixture, followed by a post-flood with brine.
Figure 14. Total oil production from the carbonate core after brine, 0.4% Acrylic acid, and SP mixture, followed by a post-flood with brine.
Polymers 13 03269 g014
Figure 15. Total oil production from the carbonate core after brine, 0.5% Acrylic acid, and SP mixture, followed by a post-flood with brine.
Figure 15. Total oil production from the carbonate core after brine, 0.5% Acrylic acid, and SP mixture, followed by a post-flood with brine.
Polymers 13 03269 g015
Figure 16. Total oil production from the carbonate core after brine, 0.5% Acrylic acid, butanol and SP mixture, followed by a post-flood with brine.
Figure 16. Total oil production from the carbonate core after brine, 0.5% Acrylic acid, butanol and SP mixture, followed by a post-flood with brine.
Polymers 13 03269 g016
Figure 17. Total oil recovery observed from the five formulations.
Figure 17. Total oil recovery observed from the five formulations.
Polymers 13 03269 g017
Figure 18. Comparison of total oil recovery observed from the five formulations.
Figure 18. Comparison of total oil recovery observed from the five formulations.
Polymers 13 03269 g018
Figure 19. Total oil production from the carbonate core after brine, 0.5% formic acid, and SP mixture, followed by a post-flood with brine.
Figure 19. Total oil production from the carbonate core after brine, 0.5% formic acid, and SP mixture, followed by a post-flood with brine.
Polymers 13 03269 g019
Figure 20. Total oil production from the carbonate core after brine, 0.5% acetic acid, and SP mixture, followed by a post-flood with brine.
Figure 20. Total oil production from the carbonate core after brine, 0.5% acetic acid, and SP mixture, followed by a post-flood with brine.
Polymers 13 03269 g020
Table 2. Phase behavior study.
Table 2. Phase behavior study.
Salinity (%)Initial Vol of Oil (mL)Initial Vol of Slug (mL)Final Oil Volume (mL)Final Vol of Slug (mL)Vol of Microemulsion (mL)Phase Position
0.004.504.505.5003.50Upper
1.004.504.501.503.004.50Upper
2.004.504.503.801.004.20Upper
3.004.504.501.103.404.50Upper
4.004.504.503.103.002.90Middle
5.004.504.501.603.703.70Middle
6.004.504.501.004.503.50Lower
7.004.504.503.204.301.50Lower
Table 3. Core properties.
Table 3. Core properties.
Core PlugLength (cm)Diameter (cm)Pore Volume (CC)Dry Weight (gm)Porosity (%)Permeability (mD) Ka
C115.093.7932398.8018.91102.00
C215.53.8033.24400.1619.9491.16
C315.243.8133.2368.6019.10121.58
C415.243.8132372.3818.40185.97
C515.023.7830.32-17.80190.00
S615.053.7935.04355.4520.17124.2
S715.183.8035.14358.8420.22129.02
Table 4. APG surfactant and different concentration of formic acid.
Table 4. APG surfactant and different concentration of formic acid.
Sample No.Concentration (%)IFT (dyne/cm)
12% NaCl23.00
20.5% APG + 2% NaCl0.285
30.5% APG + 0.2% Formic Acid + 2% NaCl0.190
40.5% APG + 0.4% Formic Acid + 2% NaCl0.266
50.5% APG + 0.6% Formic Acid + 2% NaCl0.220
60.5% APG + 0.8% Formic Acid + 2% NaCl0.250
70.5% APG + 1.0% Formic Acid + 2% NaCl0.200
Table 5. IFT values of formic acid, acetone, and butanone with 0.5% APG and 2% NaCl.
Table 5. IFT values of formic acid, acetone, and butanone with 0.5% APG and 2% NaCl.
Sample No.Concentration of Acetone/Butanone/Formic Acid (%)IFT (dyne/cm)
AcetoneButanoneFormic Acid
10232323
20.10.310.310.29
30.20.30.250.19
40.40.370.240.27
50.60.20.250.22
60.80.250.250.25
710.250.250.2
Table 6. Flood results for acetone, APG 264, and Xanthan gum blend.
Table 6. Flood results for acetone, APG 264, and Xanthan gum blend.
Formulation 1: 0.5% APG + 0.6% Acetone + 1000 mg/L XG + 2% NaCl Water
CorePV (cc)Oil Volume (cc)Soi (%)Water Flood RecoveryS.P. Flood RecoveryTotal
cc%cc%%
C1321752.67.8461.438.4454.44
Table 7. Flood results for butanone, APG 264, and Xanthan gum mixture.
Table 7. Flood results for butanone, APG 264, and Xanthan gum mixture.
Formulation 2: 0.5% APG + 0.6% Butanone + 1000 mg/L XG + 2% NaCl Water
CorePV (cc)Oil Volume (cc)Soi (%)Water Flood RecoveryS.P. Flood RecoveryTotal
cc%cc%%
C233.241751.147.9646.827.2122.5969.41
Table 8. Flood results for acrylic acid, APG 264, and Xanthan gum blend.
Table 8. Flood results for acrylic acid, APG 264, and Xanthan gum blend.
Formulation 3: 0.4% Acrylic Acid + 0.5% APG + 1000 mg/L XG + 2% NaCl Water
Formulation 4: 0.5% Acrylic Acid + 0.5% APG + 1000 mg/L XG + 2% NaCl Water
CorePV (cc)Oil Volume (cc)Soi (%)Water Flood RecoveryS.P. Flood RecoveryTotal
cc%cc%%
C333.241751.27.51444.202.0211.8656.06
C4331753.17.9546.751.9311.3558.10
Table 9. Flood results for acrylic acid, butanol, APG 264, and Xanthan gum blend.
Table 9. Flood results for acrylic acid, butanol, APG 264, and Xanthan gum blend.
Formulation 5: 0.5% Acrylic Acid + 0.5% Butanol + 0.5% APG + 1000 mg/L XG + 2% NaCl
CorePV (cc)Oil Volume (cc)Soi (%)Water Flood RecoveryS.P. Flood RecoveryTotal
cc%cc%%
C530.418.4617.54413.221859
Table 10. Summary flood results of the five formulations.
Table 10. Summary flood results of the five formulations.
CoreFormulationWater Flood Recovery (%)S.P. Flood Recovery (%)Total (%)
C10.6% Acetone + 0.5% APG + 1000 mg/L XG + 2% NaCl 468.4454.44
C20.6% Butanone + 0.5% APG + 1000 mg/L XG + 2% NaCl46.8222.5969.41
C30.4% Acrylic Acid + 0.5% APG + 1000 mg/L XG + 2% NaCl44.2011.8656.06
C40.5% Acrylic Acid + 0.5% APG + 1000 mg/L XG + 2% NaCl 46.7511.3558.10%
C50.5% Acrylic Acid + 0.5% Butanol + 0.5% APG + 1000 mg/L XG + 2% NaCl411859
Table 11. Flood results for acrylic acid, APG 264, and Xanthan gum blend.
Table 11. Flood results for acrylic acid, APG 264, and Xanthan gum blend.
Formulation 6: 0.5% Formic Acid + 0.5% APG + 1000 mg/L XG + 2% NaCl Water
Formulation 7: 0.5% Acetic Acid + 0.5% APG + 1000 mg/L XG + 2% NaCl Water
CorePV (cc)Oil Volume (cc)Soi (%)Water Flood RecoveryS.P. Flood RecoveryTotal
cc%cc%%
S635.04 59.93 42.2 12.955
S735.14 59.75 42.7 15.758.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Haq, B. Green Enhanced Oil Recovery for Carbonate Reservoirs. Polymers 2021, 13, 3269. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13193269

AMA Style

Haq B. Green Enhanced Oil Recovery for Carbonate Reservoirs. Polymers. 2021; 13(19):3269. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13193269

Chicago/Turabian Style

Haq, Bashirul. 2021. "Green Enhanced Oil Recovery for Carbonate Reservoirs" Polymers 13, no. 19: 3269. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13193269

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