Next Article in Journal
A Study of the Chemical Composition, Antioxidant Potential, and Acute Toxicity of Bulgarian Tanacetum vulgare L. Essential Oil
Previous Article in Journal
Characteristic Photoprotective Molecules from the Sphagnum World: A Solution-Phase Ultrafast Study of Sphagnic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Displacement Mechanism of Supercritical CO2 in Low-Permeability Heavy Oil Reservoir and Improvement Mechanism of Displacement Agents

1
College of Petroleum Engineering, Liaoning Petrochemical University, Fushun 113001, China
2
Baikouquan Oil Production Plant of Petrochina Xinjiang Oilfield Branch, Karamay 834000, China
3
D&P Technology Research Institute, Petrochina Liaohe Oilfield Company, Panjin 124000, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(16), 6154; https://doi.org/10.3390/molecules28166154
Submission received: 30 July 2023 / Revised: 9 August 2023 / Accepted: 15 August 2023 / Published: 21 August 2023

Abstract

:
With the continuous growth of global energy demand and the late stage of conventional oilfield exploitation, the demand for developing and utilizing low-permeability heavy oil reservoirs is becoming increasingly urgent. However, the exploitation of low-permeability heavy oil reservoirs faces many challenges due to their high viscosity, low permeability, and complex geological conditions. To overcome these challenges, researchers have gradually introduced SC-CO2 as an oil displacement agent in the exploitation of heavy oil reservoirs. However, the oil displacement mechanism of SC-CO2 in low-permeability heavy oil reservoirs and its improvement mechanism are still not completely understood. The article provides a detailed study and understanding of the oil displacement mechanism of SC-CO2, which involves the expansion of heavy oil volume through SC-CO2 dissolution. This mechanism reduces the capillary resistance and flow resistance during the oil flow process. The permeation of CO2 disrupts the internal structure and arrangement of heavy oil, reducing its viscosity. CO2 extracts both light and heavy components from the heavy oil, reducing the residual oil saturation. In addition, the mechanism of improving the effect of oil displacement agents such as nanoparticles, polymers, and surfactants on SC-CO2 displacement was also explored. By further exploring the mechanisms and improvement mechanisms of SC-CO2 displacement for heavy oil, it can guide the selection and optimization of oil displacement agents. Furthermore, understanding the mechanism can also provide a theoretical basis for engineering practice and technical innovation. While the research on CO2 flooding is analyzed and evaluated, the obstacles and challenges that still exist at this stage are indicated, and future research work on CO2 in low-permeability heavy oil reservoirs is proposed.

1. Introduction

In the 21st century, the increasing global oil and gas demand and the gradual decline in conventional oil and gas production have not met the needs of consumption growth. The effective use of unconventional oil and gas is of key strategic significance in alleviating the contradiction between oil and gas supply and demand, ensuring energy security, promoting a low-carbon transformation of the energy mix, and promoting carbon sequestration [1,2,3]. Chemical flooding has achieved great success in conventional medium-high-permeability reservoirs. However, it has entered a late stage of development after a long period of development. The production gradually decreases, and the reservoir development target gradually shifts from medium-high-permeability reservoirs to low-permeability heavy oil reservoirs.
Improving the recovery efficiency of low-permeability heavy oil reservoirs has become one of the important challenges in the petroleum industry [4]. The main reason is that there is more gum, asphaltene, petroleum wax, and organic acids in heavy oil, which results in crude oil having higher density and viscosity. This leads to insufficient fluidity and makes it unable to be extracted using conventional methods [5]. The other reason is that the pore size of low-permeability reservoirs is small, resulting in a large specific surface area and strong heterogeneity. This makes it difficult for the oil displacement agent to be injected smoothly [6]. The adsorption and retention losses of HPAM, alkali, and surfactant increase, leading to an increase in displacement cost [7,8]. The traditional waterflood mining cannot effectively improve the flow of heavy oil due to its high viscosity. Moreover, there are issues of high initial pressure and high residual oil content in waterflood mining [9]. In addition, the low permeability of reservoirs leads to high displacement pressure, which further increases the difficulty in water flooding [10]. CO2 flooding is an oil recovery technology that uses the pressure difference generated after CO2 gas injection to drive the crude oil in the formation, making it more fluid and easier to recover. Therefore, CO2 flooding has become an effective method to address the poor fluidity of heavy oil and the low injection capacity of oil displacement agents, ultimately improving the oil recovery efficiency of low-permeability heavy oil reservoirs [11]. The field tests show that the oil recovery rate of CO2 flooding is 10% higher than that of water flooding. Simultaneously, the reservoir can also partially store CO2 to alleviate the greenhouse effect [12]. In the future, CO2 flooding technology will link CO2 capture and transmission technologies to form a green and sustainable CO2 industry chain, providing a two-way drive for environmental protection and oil exploitation.
At present, most of the existing oilfields belong to continental sedimentary reservoirs with high oil viscosity, serious heterogeneity, low water drive recovery, and high recombination content. Most of the newly discovered crude oil is in low-permeability reservoirs, and it is difficult to achieve a high oil recovery rate under conventional production technology. As an oil displacement agent to enhance oil recovery, CO2 has attracted extensive attention in scientific research and engineering applications [13,14,15]. Existing studies have shown that CO2 in the supercritical state exhibits the characteristics of gas-like viscosity and liquid-like solubility [16]. In most reservoir conditions, CO2 exists in a supercritical state, which can reduce the viscosity of crude oil, improve the oil–water mobility ratio, and cause crude oil to expand. At the same time, it can also extract the light hydrocarbon components from crude oil and effectively reduce the interfacial tension between oil and water, thereby increasing the oil washing efficiency and significantly increasing the crude oil recovery [17,18,19,20,21].
An important mechanism of CO2 flooding to enhance oil recovery is that CO2 can be easily dissolved into crude oil and cause a swelling effect. CO2 is more soluble in oil than common hydrocarbon gases, and its solubility in crude oil is greater than its solubility in water, thus supplementing reservoir pressure and increasing oil saturation. The dissolution of CO2 expands the volume of crude oil, which can increase the energy of elasticity of the formation and thus increase the internal kinetic energy of crude oil [22]. The viscosity of crude oil will decrease significantly when CO2 is dissolved in crude oil, which mainly depends on pressure and temperature of the reservoir, and the viscosity of the non-carbonated crude oil. It effectively expands the swept area of crude oil and improves the mobility ratio of crude oil. The flow of crude oil is improved, thus improving the efficiency of oil displacement [23,24,25]. CO2 extracts or vaporizes the light components in crude oil due to which the surface tension at the oil–water interface is reduced, and the remaining oil saturation is also reduced, thus improving oil recovery. CO2 can preferentially extract and vaporize light hydrocarbons in crude oil, and then heavier hydrocarbons will also be vaporized. After the hydrocarbon components in crude oil are extracted, the oil–water interfacial tension decreases after continuous enrichment [26,27,28,29]. CO2 occupies a certain gap in the oil reservoir, which increases the pressure of the oil reservoir and increases the crude oil production. During the production process, the pressure drop causes the CO2 dissolved in the crude oil to escape, which has the effect of dissolved gas flooding and improves the oil displacement effect [30,31].
Due to the limitation of experimental conditions, the relevant studies on the mechanism of swelling and viscosity reduction of heavy oil by SC-CO2 are still relatively few, and the theoretical basis is relatively weak, and the microscopic mechanism of fluid interaction is still unclear. In view of the above reasons, the characteristics and microscopic mechanism of SC-CO2 displacement in low-permeability heavy oil reservoirs are discussed. The aim is to further understand the mechanism of action of CO2 and effectively pave the way for further improvements in oil recovery. Asphaltene deposition is an inevitable problem during CO2 flooding in low-permeability reservoirs, which is affected by many factors. Asphalt has a complex structure and contains heteroatoms such as nitrogen, sulfur, or oxygen, making it the most polar and surfactant-like component in crude oil [32]. The main mechanisms by which nanomaterials enhance oil recovery include changes in wettability, reduction in interfacial tension between oil and water, and alteration in oil–water viscosity ratio [33]. The strong polarity of asphaltene can be used to inhibit the deposition of asphaltene and stably disperse asphaltene in the crude oil system under the action of nanoparticles [34,35,36,37]. CO2 foams are commonly used in reservoirs where CO2 is supercritical. When CO2 is supercritical, it forms a weak and unstable foam [38,39,40]. In addition, in low-permeability heavy oil reservoirs, gas channeling and low oil recovery are inevitable under SC-CO2 displacement [41,42,43]. HPAM enters high-permeability layers by flowing along large channels and by stretching deformations, thereby reducing the permeability of the high-permeability layers. It improves the heterogeneity of the reservoir, expands the swept volume, and thereby increases the EOR. Most oil reservoirs have excessively high miscibility pressure between crude oil and CO2, which prevents successful miscible displacement. During enhanced oil recovery, CO2 is utilized to carry surfactants into the formation and act on the oil–gas interface. It reduces the interfacial tension between oil and water, thereby lowering the minimum miscibility pressure of the system and compensating for the drawbacks of existing methods [44,45]. Therefore, this paper studies and discusses the effect and interaction mechanism of oil displacement agents, such as nanoparticles and polymers, and surface activity on CO2 flooding, highlighting the improvement technology of CO2 flooding. While the research on CO2 flooding is analyzed and evaluated, the obstacles and challenges that still exist at this stage are indicated, and future research work on CO2 in low-permeability heavy oil reservoirs is proposed.

2. Overview of SC-CO2 and Heavy Oil

2.1. Basic Properties of SC-CO2

A substance can be present in three states, i.e., solid, liquid, and gas, with the change in ambient pressure and temperature. Any pressure can convert the substance from a gaseous state to a liquid state. When the temperature of a substance is above a specific value, the temperature at that point is called the critical temperature Tc. The lowest pressure at which a gaseous state can be converted to a liquid state is called the critical pressure Pc at the critical temperature. A supercritical fluid (SCF) is a fluid whose temperature and pressure are at the critical temperature (Tc) and critical pressure (Pc), respectively. SC-CO2 refers to a fluid in which the temperature and pressure exceed its critical point. The critical temperature (Tc) of CO2 is about 31.7 °C, and the critical pressure (Pc) is about 7.37 MPa [46]. The phase diagram of SC-CO2 is shown in Figure 1.
The green area in the diagram is the supercritical state. Currently, the most studied supercritical fluid is CO2. SC-CO2 is used in a wide range of industrial manufacturing applications including electrochemical deposition, preparation of nanoparticles, processing of hard-to-cut materials, cleaning of precision instruments, and centrifugal compression of petroleum [40,47,48,49]. CO2 at its critical temperature and pressure has a higher density, and as the temperature increases, the diffusion coefficient of CO2 gradually increases while the viscosity decreases. As the pressure increases, the diffusion coefficient of CO2 gradually decreases, while the viscosity increases. Under normal conditions, the density of SC-CO2 is generally similar to that of crude oil, and the low viscosity and high diffusivity of SC-CO2 enable it to rapidly diffuse into the narrow pores of the reservoir, thereby displacing resident oil and gas. In addition, the conditions required for CO2 to reach a supercritical state are low, making these conditions less demanding on production equipment and highly operable. As a non-toxic and odorless environmentally friendly gas, CO2 can be recovered and reused from industrial waste gases, enabling industrial production and improving economic efficiency. At the same time, it can alleviate environmental problems such as the greenhouse effect faced by the world to a certain extent. Furthermore, CO2 exhibits strong solubility for substances with lower molecular weights or weaker polarity, such as alkanes below C18.

2.2. Composition and Viscous Mechanism of Heavy Oil

Heavy oil is a complex product composed of light components such as saturated hydrocarbons and aromatic hydrocarbons, as well as heavy components such as resins and asphaltenes in a certain proportion. The reason why heavy oil is “thick” is because of its high proportion of gum and asphaltene. There are many studies and assumptions about the structure and adhesion mechanism of gum and asphaltene, and it is generally believed that asphaltene is the main influencing factor.
The individual molecules of asphaltene exhibit strong polarity and contain more aromatic heterocyclic structures. These aromatic heterocyclic structures are stacked with each other by π-π conjugation, and this stacking structure can be further stabilized by hydrogen bonding of oxygen, nitrogen, sulfur, and other heteroatoms in asphaltene or the chelation of metal heteroatoms such as Ni and V in heavy oil, forming asphaltene supramolecular structures [50,51]. Asphaltenes are encapsulated by colloidal molecules with polar groups that form a surrounding coating, resulting in asphaltene particles encapsulated by colloids that are also highly polar and further leading to their aggregation through hydrogen bonding and eventually forming large molecular aggregates. This phenomenon contributes to the high viscosity of heavy oil. Through hydrogen bonding, they further accumulate and eventually form macromolecular aggregates, resulting in high viscosity of heavy oil. Figure 2 shows a supramolecular assembly structure of asphaltene aggregates proposed by Gray et al. They believe that the asphaltene aggregates are supramolecular structures formed by hydrogen bonding, metal chelation, van der Waals force, π-π interaction, and acid–base interaction [52].
Additionally, heavy oil easily forms stable W/O type emulsions with small amounts of water, resulting in increased viscosity. Processing and treating this stable W/O type emulsion of heavy oil is relatively difficult and challenging in the petroleum industry. The presence of asphaltene is conducive to the stability of W/O emulsions because the asphaltene molecule itself is a natural surfactant. Its polar functional groups such as metal heteroatoms have strong hydrophilicity, while the aromatic ring structure and hydrocarbon chain in the molecule are hydrophobic. The amphiphilicity of asphaltene molecules enables them to adsorb at the oil–water interface, forming an interfacial film with a certain mechanical strength, thus preventing water droplets from gathering and enhancing the stability of W/O emulsions [53].

2.3. Geological Characteristics of Low-Permeability Reservoirs and EOR Techniques—CO2 Flooding

With the continuous development of onshore oilfields, old oilfields have entered the late stage of high water cut, and the focus of petroleum exploration and development has gradually shifted to low-permeability reservoirs. Low-permeability reservoirs have abundant reserves both domestically and internationally, with recoverable reserves reaching 483 × 108 t [54]. Low-permeability oil and gas reservoirs refer to reservoirs with low porosity, poor fluid flow capacity, and low natural productivity, making it difficult to effectively develop them using conventional extraction methods. Currently, there is no unified classification standard for low-permeability reservoirs both domestically and internationally. In the US, reservoirs with permeability below 10 mD are classified as medium to poor reservoirs. In China, low-permeability reservoirs are divided into low-permeability (10–50 mD), ultra-low-permeability (1–10 mD), and super-low-permeability reservoirs (0.1–1 mD) [55,56,57,58,59]. Low-permeability oil reservoirs are predominantly lithologic reservoirs, characterized by a single type, mainly elastic drive reservoir. Its reservoir physical properties are poor, with low porosity and permeability, small pore throats, and strong heterogeneity. However, it has a high oil saturation and favorable crude oil properties. Therefore, the low-permeability oil reservoir has enormous potential and high exploitation value [60,61,62,63].
The United States is the country with the fastest development and the highest number of projects in CO2 flooding technology. In the 121 projects implemented in 2012, 75.8% of the reservoirs were classified as low-permeability oil reservoirs, indicating the significant importance of CO2 flooding in EOR for low-permeability and ultra-low-permeability reservoirs [64]. Compared to other gases, CO2 has a lower critical temperature and critical pressure, generating a more favorable miscibility with crude oil. Even without miscibility, CO2 still exhibits excellent oil displacement effects. After dissolving in crude oil, CO2 causes the oil to expand and reduces its viscosity, making it easier to transport. Under certain pressure, CO2 can extract both light and heavy components, causing the composition of the crude oil to approach equilibrium. CO2 dissolved in crude oil can also reduce interfacial tension. The weak acidity of CO2 can play a role in plugging in acidification to some extent.
The injection of external fluids complicates the flow relationship between oil, water, and gas in low-permeability reservoirs. During the oil displacement process, physical and chemical reactions such as convection, diffusion, adsorption, and ion exchange often occur. Studying the oil displacement mechanism of carbon dioxide in low-permeability reservoirs is crucial for injection-production parameters optimization and effect forecasting. Therefore, this study elaborates on the oil displacement mechanism of carbon dioxide in low-permeability heavy oil reservoirs and the improvement mechanism in Section 3 and Section 4. The ecological environment in low-permeability oilfields is fragile, and the contradiction between efficient exploitation of petroleum resources and environmental protection is becoming increasingly prominent. CO2 flooding can achieve the storage of greenhouse gases and improve the oil recovery rate, making it a highly promising technology for enhanced oil recovery in low-permeability reservoirs.

3. Main Mechanism of SC-CO2 Flooding of Heavy Oil

CO2 flooding is one of the oil displacement methods in tertiary oil recovery, and SC-CO2 miscible drive is one of them. After reaching the reservoir, CO2 becomes supercritical. It displaces oil from the large pores and mixes with it, forming an oil bank ahead of the miscible zone. The SC-CO2 zone is followed by the water lock, which drives oil from the reservoir to the bottom of the production well, ultimately allowing oil to be recovered. The working principle diagram of SC-CO2 miscible flooding technology is shown in Figure 3.
The use of SC-CO2 miscible flooding is due to its excellent fluidity in the formation and its ability to expand the volume of crude oil when dissolved in it, thereby reducing its viscosity and lowering the interfacial tension of the crude oil [65]. In particular, heavy oil reservoirs with high viscosity and density have a good mutual solubility with SC-CO2, enabling a significant reduction in crude oil viscosity, which is very beneficial for the replacement of heavy oil [66]. At the same time, CO2 has the effect of a dissolving gas after dissolving in crude oil. CO2 separates from the crude oil and creates bubbles in the crude oil, pushing the crude oil to flow and improving the oil drive efficiency, when the formation pressure drops below the saturation pressure [67].

3.1. The Swelling of Heavy Oil by SC-CO2

The solubility of CO2 in crude oil is an important parameter for EOR and CCUS, while the volume expansion of crude oil is a mechanism for CO2-EOR. Consequently, it is necessary to investigate the mechanism of CO2 volume expansion under near-critical and supercritical conditions for enhanced recovery, which will provide us a better understanding of the chemical and physical reactions in the CO2-EOR process. The important mechanism of crude oil expansion involves CO2 easily dissolving into crude oil and causing it to expand, thus supplementing reservoir pressure and increasing oil saturation [68]. Volume expansion increases the internal kinetic energy of crude oil and reduces the capillary resistance and flow resistance in the process of crude oil flow, thus improving the flow capacity of crude oil [69]. The effect of expansion on heavy oil is mainly manifested in three aspects. The residual oil in the reservoir after water flooding is inversely proportional to the expansion coefficient, that is, the greater the expansion, the lesser the residual oil in the reservoir [70]. The dissolution of CO2 in oil droplets displaces water from the pore space, resulting in a process where the water-wet system becomes a drainage rather than water absorption. The relative permeability curve of oil discharge is higher than that of oil suction, creating an expedient flow environment under any given saturation condition [71]. On the one hand, volume expansion can significantly increase the oil saturation and formation elastic energy of the reservoir. On the other hand, the remaining oil after expansion can be separated or partially separated from the formation water and become mobile oil [72].
Crude oil is a mixture composed of compounds such as n-alkanes, cycloalkanes, aromatic hydrocarbons, resins, and asphaltenes, making the interaction between CO2 and the constituents of crude oil exceedingly intricate. Since Whorton et al. applied for the patent of CO2 flooding in 1950, many experts and scholars have studied the phenomenon of volumetric expansion effect of dissolved CO2 in crude oil [73]. The alterations in the properties of crude oil are primarily as a result of CO2 dissolving in crude oil and are affected by the dispersion of CO2 in crude oil. Yang et al. analyzed the swelling behavior of oil by qualitatively discussing the dispersion properties of CO2 in the oil. The results show that the dispersion of CO2 molecules in the alkane phase plays a dominant role in the increase in the volume of CO2–alkane system. The solubility of CO2 in alkanes and the volume of the CO2–alkane system decrease with increasing alkane molecular length [74]. However, the expansion mechanism of CO2 dissolved in crude oil cannot be fully characterized under experimental conditions. In recent years, some researchers have investigated the interfacial properties and phase equilibria of the CO2–crude oil system by using molecular simulation techniques to understand the dynamic behavior of alkane solid surfaces. Zhang et al. investigated the solubility of CO2 in octane and its effect on the expansion of octane (n-octane) by performing configurational-bias Monte Carlo simulations in an osmotic ensemble. The results show that the interaction between octane and CO2 strengthens with increasing CO2 solubility in the pressure range of 2–10 MPa and at temperatures ranging from 323–353 K. The interaction between octane and CO2 is the main cause of octane dissolution [75]. Similar conclusions were reached by Liu et al., who found that the interaction between CO2 and decane molecules is the main reason for the volume swelling of the CO2–alkane system. The dispersive interaction between CO2 and n-decane molecules is the essence of the volume swelling of the CO2–n-decane system. At the same time, they noted that the volume swelling rate of decane was accelerated by the increase in the average spacing between decane molecules and the stretching of decane molecules. Figure 4 illustrates the microscopic process of volume expansion in the CO2–octane system [76].
As the solubility of CO2 in oil increases, it leads to an increase in the distance between the oil molecules, and the force between oil molecules decreases, resulting in an expansion of the oil volume [77]. Figure 5 shows the process of dissolution and volume expansion of CO2 in n-alkane. As can be seen from the figure, when the molar fraction of CO2 is 25%, the molecules fill the space between the hydrocarbon molecules, forming a dispersed form. When the molar fraction reaches 50%, the overall volume of the system expands to a certain extent, and the CO2 and hydrocarbon molecules interpenetrate each other. When the CO2 mole fraction reaches 75%, the number of CO2 molecules is absolutely dominant, and the total volume of the system is greatly increased due to the dramatic increase in the number of both molecules. In addition, the longer the carbon chain and the larger the molecule of a hydrocarbon, the harder it is to dissolve CO2 [78].
Normal n-alkanes with shorter chains have smaller negative intermolecular interaction energies, which means that the molecules are more easily separated from each other, and therefore a higher swelling factor can be observed when octane is saturated by CO2. The interaction between alkane chains decreases with increasing pressure, indicating that the diffusion capacity of oil molecules increases when more CO2 molecules are dissolved by heavy oil [79]. In order to observe the CO2 dissolution process, Wang et al. observed the CO2–octane system by using molecular dynamics simulation methods. At the initial stage, some CO2 molecules begin to penetrate into the octane phase, resulting in a slight expansion of the octane volume. As a result, the CO2–octane system expands further, providing more space for the diffusion of CO2. The swelling process of the CO2–octane binary system is shown in Figure 6 [80]. It was also found that when CO2 was injected into the system, the alkane chain was slightly extended, meaning that the presence of CO2 molecules caused the alkane molecules to expand, which further led to the expansion of the systemic volume.
Regarding the study on the swelling behavior of CO2 with cycloalkanes, Sima et al. studied the vapor–liquid equilibrium of binary systems of CO2 with cycloalkanes (cyclopentane and cyclohexane) and showed that CO2 is more soluble and miscible in n-alkanes than in cycloalkanes [81]. Similar conclusions were obtained by Zhang et al. in their study of the solubility and volume expansion factors of CO2 in cycloalkanes. They believe that as the number of carbon atoms increases, the dispersibility between alkane molecules increase, the interactions between alkane molecules are enhanced, and the solubility of CO2 in alkanes decreases [82]. A large number of research works have investigated the solubility and volume expansion mechanism of CO2 in single alkane, brine, and alkane–water binary systems, while the microscopic mechanism of CO2 in hydrocarbon mixtures is not sufficiently clear. Therefore, the study of CO2 in hydrocarbon mixtures has received much attention. Wang et al. measured the volume expansion of CO2 and n-dodecane in fused silica capillaries using Raman spectroscopy. The results show that the expansion coefficient increases with the increase in pressure and decreases with the increase in temperature [83]. CO2 is less soluble in cycloalkanes and aromatics than in n-alkanes with the same number of carbon atoms. It is concluded that the solubility of CO2 in hydrocarbon mixture is mainly affected by high-carbon hydrocarbons through the above study, i.e., the higher the proportion of high-carbon hydrocarbons in the mixture, the lower the solubility of CO2 [84].

3.2. SC-CO2 Viscosity Reduction Characteristics in Heavy Oil

SC-CO2 is characterized by a large diffusion coefficient and a strong solvency capacity, and it can rapidly penetrate into the mixture [85,86]. When CO2 is dissolved in crude oil, the attraction force between oil molecules becomes smaller and the internal friction during flow is reduced due to the carboxylation reaction, thus the viscosity of the oil is reduced and the fluidity of the crude oil is improved. For heavy oil reservoirs, when a certain concentration of CO2 is injected into heavy oil, the internal intermolecular force changes from the liquid–liquid intermolecular force to the liquid–gas intermolecular force, and the intermolecular force is greatly reduced. Furthermore, the gum asphalt macromolecular structure of heavy oil will also be destroyed after dissolving CO2, the volume will be expanded, the density will be reduced, and the viscosity will be rapidly reduced [87].
The high content of asphaltene and gum in heavy oil leads to adsorption of micelles on gum to form a solvation layer. Therefore, the dispersed phase occupy a large volume fraction, the attractive interaction between the particles of the dispersed phase is strong, and the internal friction is large, so it has a high viscosity. The stronger is the interaction between colloidal particles and solvent, the higher is the degree of solvation, the higher is the bound solvent, the greater is the viscosity of the system. In recent years, researchers have conducted a lot of research on the effect of CO2 flooding on viscosity reduction in heavy oil. Seyyedsar et al. suggested that CO2 can cause methane in heavy oil to be released in the form of microbubbles. The release of methane from heavy oil results in an increase in oil viscosity, and this increase is significant for heavy oil systems. However, the dissolution of CO2 in the oil will compensate for this viscosity increase. Experimental studies have shown that the saturation of CO2 in heavy oil is generally not linearly related to the viscosity reduction rate of heavy oil. In other words, as long as a small amount of CO2 is added to the oil, the viscosity of the heavy oil can be greatly reduced [88,89]. In order to further prove the viscosity reduction effect of CO2, Sun et al. compared the viscosity reduction effect of CO2 and nitrogen on heavy oil. The experimental results show that the viscosity reduction effect of CO2 on heavy oil is much better than that of N2 under the same conditions. Meanwhile, with the increase in CO2 pressure, the shear thinning feature becomes more and more obvious. With the decrease in temperature, the viscosity reduction rate of CO2 on heavy oil increases. The main mechanism is that when CO2 is dissolved in crude oil, the dispersion effect increases the distance between crude oil molecules, and ultimately reduces the viscosity of crude oil [90]. Regarding the effect of pressure on the viscosity reduction in CO2, Wang investigated the viscosity reduction in heavy oil by CO2 flooding through a combination of software simulation and experiment. The main findings are shown in Figure 7 below. With the increase in CO2 injection, the critical pressure of CO2–heavy oil system gradually increases, and the saturation pressure at reservoir temperature also gradually increases, and the viscosity of crude oil rapidly decreases as CO2 is injected [91]. We can conclude that the affinity of CO2 to asphaltene is greater than that of gum, and it can be adsorbed on asphaltene in the state of small molecules, replacing the original gum adsorption layer (solvated layer). On one hand, CO2 releases small molecules that are bound in the solvation layer, reducing the attraction between molecules, allowing them to move freely, resulting in a decrease in system viscosity. On the other hand, CO2 molecules are too small to form a thick adsorption layer, so that the asphaltene micelle are exposed. The volume fraction of the dispersed phase in the system is greatly reduced, and it is not easy for the micelles to interact with each other. The above factors cause a decrease in the viscosity of the heavy oil system.
The dissolution of CO2 can significantly reduce oil viscosity has been reported in the literature, especially for heavy oil. The solubility of CO2 in three heavy oils with different densities and the corresponding crude oil viscosity and density have been studied by Miller et al. at different temperatures and pressures. When there is no CO2, the viscosity of crude oil increases with the increase in pressure, while when there is CO2, the viscosity of crude oil decreases with the increase in pressure, and the viscosity of crude oil containing CO2 is significantly reduced with the increase in temperature [92]. Li et al. studied the effect of CO2 on the expansion and viscosity reduction in crude oil through experiments. The experimental results show that the viscosity of the dissolved CO2 containing oil decreases with the increase in saturation pressure. When the formation pressure is 23 MPa, the viscosity decreases by 44% [93]. Zhao et al. studied the effect of CO2 on the physical properties of high pour-point oil. The experimental results show that the solubility of CO2 in high pour-point oil decreases with the increase in temperature. At the same temperature, the viscosity of high pour-point oil decreases as the amount of dissolved CO2 increases, and the more dissolved CO2 there is, the greater the decrease in viscosity. The viscosity of pour-point oil is 27.0 mPa·s at the formation temperature of 55 °C, and the viscosity of crude oil fully saturated with CO2 decreases to 10.2 mPa·s, with a decrease of 62.2% [94]. Xin et al. measured the viscosity reducing effect of CO2 on crude oil in the North China NB reservoir under an experimental pressure of 15.9 MPa and a temperature of 94 °C. The experimental results show that the gas–oil ratio of the crude oil saturated with CO2 increased from 34 m3/t to 114 m3/t, and the viscosity of the crude oil decreased from 3.41 mPa·s to 1.92 mPa·s, a reduction of about 34% [95]. Zhao et al. carried out a feasibility study on CO2 flooding in ultra-low-permeability reservoirs using laboratory experiments and numerical simulation. Similar to the results of many of the above researchers, it is observed that the amount of dissolved CO2 in oil increases with the increase in pressure, and the viscosity of the heavy oil rapidly decreases. Under the formation pressure (18.18 MPa), the dissolved oil–gas ratio is 80.07 m3/t, and the crude oil viscosity is 1.745 mPa·s. When CO2 is injected under this pressure, the formation oil viscosity is 1.104 mPa·s, and the crude oil viscosity is reduced by 36.73% [96]. Liu et al. studied the viscosity reduction effect of CO2 flooding on ultra-low-permeability reservoir in Daqing Oilfield by using thin tubes and natural core. The experimental results show that under the formation pressure (18.2 MPa), the dissolved oil–gas ratio is 80.07 m3/t, and the crude oil viscosity is 1.745 mPa·s [97]. After the CO2 was injected, the oil–gas ratio is increased to 236.06 m3/t, and the viscosity of formation oil is 1.104 mPa·s, the viscosity of crude oil is reduced by 36.73%, and the fluidity of crude oil is increased. CO2 dissolved in crude oil significantly reduces the viscosity of crude oil, and the higher the viscosity of crude oil, the greater is the reduction [98]. From the above data, we conclude that CO2 has high solubility in heavy oil, and the complete miscibility of CO2 with heavy oil can dilute heavy oil, which is the main reason that the viscosity of heavy oil decreases greatly and the recovery efficiency increases.

3.3. SC-CO2 Extraction

The principle of SC-CO2 extraction and separation is to use pressure and temperature to control the dissolution capacity of supercritical fluid, and then achieve the extraction and separation of substances. Hydrocarbons and various lipid organic compounds in the reservoir medium can be extracted when CO2 reaches supercritical state, and the material inside the matrix are dissolved in the process and flow out of the pores with the gas [99,100]. The pore is widened and the crude oil adsorbed on the matrix is resolved due to the extraction of SC-CO2, which changes the heavy oil from the adsorbed state to the free state, thereby causing the crude oil to flow out of the pore channels. This process also expands the passage of crude oil, allowing gas to enter the matrix more easily and also allows the crude oil to be extracted quickly.
Kesavan et al. investigated the recovery of oil from oil shale using supercritical extraction technology with CO2 as the solvent. Their work successfully demonstrated the feasibility of using CO2 as a solvent for extracting oil from oil shale, completing groundbreaking research on the utilization of SC-CO2 for oil shale extraction [101]. Deo et al. found that two solvents, CO2 and propane, exhibited strong extraction capacity for crude oils from light-chain alkanes to heavy-chain alkanes around their respective critical temperatures [102]. With further studies on crude oil extraction, Guiliano et al. found that only n-alkanes between C14 and C24 were present in the extract solution at a pressure of 130 bar, with few branched chain alkanes and cycloparaffins present. Normal alkanes were extracted up to C32 at a pressure of 300 bar. The results showed that with the increase in pressure (130–300 bar), a greater variety of branched chain alkanes and cycloparaffins were extracted, along with heavier aromatic compounds [103]. A similar study was performed by Mihkel Koel et al., which was different from Guiliano as they studied the application results of SC-CO2 extraction at different temperatures. They found that pure CO2 preferentially extracted n-alkanes [104]. Rudolf et al. studied the extraction of aliphatic and aromatic hydrocarbons from shale rock using SC-CO2. SC-CO2 was found to be effective in extracting “free” hydrocarbons from samples in a lower temperature range [105].
Studies have shown that SC-CO2 mainly extracts components below C20, and the existence of a pressure critical point in this extraction process ensures that the extraction proceeds smoothly [106]. As can be seen in Figure 8a, the carbon fraction of the oil production remains essentially unchanged before the gas breakthrough. From Figure 8a,b, it can be observed that at the same pressure of 25 MPa, the fractions of components below C20 are 56.3% and 65.3%, respectively, while the ratio of components below C20 in crude oil is 46.3% [107]. The extraction capacity of CO2 increases with the increase in its density. The appropriate increase in pressure increases the density of CO2, and the extraction is easier to carry out. Factors affecting the optimum extraction temperature include the increase in diffusion coefficient due to warming and the decrease in CO2 density, which together influence the magnitude of the optimum extraction temperature [108]. Peng et al. conducted a study on the content and viscosity of the components in the extracted material and concluded that the extraction of light hydrocarbon components from heavy oils is enhanced with increasing extraction pressure and temperature in CO2 extraction process. As can be seen from Figure 9, with the increase in extraction pressure, the percentage of light components in the residue after CO2 extraction is reduced by about 56% compared with the original oil sample. The results show that with the increase in pressure, the extraction effect of CO2 on super heavy oil is enhanced, and the light components are continuously extracted [109]. When the pressure reaches a high level, according to the colloidal system theory of heavy oil, the original stable colloidal structure of heavy oil is disrupted, and the heavy asphaltenes agglomerate, leading to asphaltene precipitation [110].
Allawzi et al. concluded that SC-CO2 extraction of oil is selective and preferentially extracts non-polar substances. It was found that CO2 mainly extracted saturated hydrocarbon, olefins, and some aromatics in crude oil [111]. Small aromatic hydrocarbons (such as methylbenzene, methyl naphthalene, etc.) exist in the form of micelles in crude oil, while paraffin exists in the form of colloids. The original equilibrium state of small saturated hydrocarbons and aromatic hydrocarbons is disrupted by the extraction of SC-CO2, leading to the deposition of macromolecular compound [112]. Ni et al. further studied the selective extraction of SC-CO2. It was found that SC-CO2 cannot completely extract all components from the oil, and it preferentially extracts saturates and leaves behind asphaltene. The results show that low-boiling and non-polar or low-polar petroleum components can be quickly extracted from rocks using SC-CO2 [113]. The selectivity of SC-CO2 extraction of crude oil molecules under different pressures and temperatures was studied by Cao et al. [114]. SC-CO2 preferentially extracts small molecules with relatively low polarity, while large molecules with relatively high aromaticity and polarity are not easily extracted using SC-CO2, and compounds containing multiple heteroatoms are more difficult to be extracted than hydrocarbons [115]. This is because non-polar saturated hydrocarbon molecules are less affected by the adsorption of other components, so they are easily extracted using SC-CO2. Aromatic hydrocarbons are less affected by the circumflex network structure of asphaltene and gum due to their smaller molecular size. In addition, SC-CO2 molecules have weak polarity, making it easier to dissolve aromatic hydrocarbon molecules and enhance the diffusion coefficient of aromatic hydrocarbons [116].

4. Research Progress on the Mechanism of CO2 Flooding Improved by Typical Oil Flooding Agents

4.1. Nanoparticles Inhibit Asphaltene Deposition during CO2 Flooding

When CO2 gas is injected into the reservoir for oil displacement, the temperature and pressure in the reservoir change after CO2 molecules contact with crude oil, breaking the dynamic stability of the dispersion system. As shown in Figure 10, CO2 molecules compete with gum that coats the asphaltene surface for spatial positions after entering the crude oil. As a result, the force of forming the gum solvation layer is weakened, causing the gums that stabilize the asphaltene to continuously detach from the surface of the asphaltene particles [117]. Asphaltene particles that have lost their solvation layer coating may have attractive collision and aggregation under the interaction of electrostatic, dipole, and hydrogen bonding, which leads to asphaltene deposition [118]. The surface activity of asphaltene molecules is caused by the existence of polar groups, which will lead to self-association between asphaltene molecules, causing asphaltene aggregation and even deposition. This results in many problems such as blockage of reservoir pore throat, wettability reversal, and permeability decrease [119]. Therefore, solving or mitigating the issue of asphaltene deposition during CO2 flooding processes is crucial for enhancing oil recovery in low-permeability reservoirs.

4.1.1. Adsorption of Nanoparticles

Nanoparticles are considered to be one of the most promising solutions for asphaltene deposition due to their large specific surface area and adsorption capacity, better suspension and catalytic performance, and less susceptibility to reservoir damage compared to chemical agents. The strong polarity of the surface of asphaltenes is the fundamental reason for self-association and even deposition between asphaltene molecules. In the presence of nanoparticles, the strong polarity of asphaltenes can be exploited to inhibit asphaltene deposition and stabilize their dispersion in the crude oil system. By modifying the nanoparticles, on the one hand, the charge distribution on the surface of the particles can be changed, so that the particles will adsorb asphaltene after entering the reservoir, and asphaltene aggregation and deposition can be avoided. On the other hand, nanoparticles can be surface grafted so that the surface of the particles carries the corresponding polar groups, which can interact with asphaltene molecules after entering the reservoir and promote the stable dispersion of asphaltene. According to the DLVO theory, the strong polarity of the asphaltene and the surface charge of the nanoparticles facilitate an interaction between the two [120]. For example, silica nanoparticles containing hydroxyl groups and electronegativity can adsorb with heavy components that are interfacial activity. Moreover, silica nanoparticles can destroy the self-assembled structure of the heavy components of crude oil [36]. From Figure 11, it can be seen that silica nanoparticles are adsorbed on the collapsed layer, which proves that there is an interaction between silica nanoparticles and heavy components [121].
The surface charge of the nanoparticles can be adjusted by changing the pH of the nanofluid. Moreover, nanoparticles have strong fluidity in porous media and possess high surface activity and specific surface area. Therefore, they can adsorb the asphaltenes in crude oil, thereby inhibiting asphaltene deposition [122,123]. A Fe3O4–shell polyose nanocomposite was prepared by Panahi et al. This composite material exhibits a stronger capacity for adsorbing asphaltenes compared to single Fe3O4 nanoparticles. Its asphaltene adsorption isotherm is consistent with the Freundlich model, which further confirms the existence of self-association and multilayer adsorption of asphaltene on the nanoparticles [124]. Asphaltenes in crude oil tend to rapidly aggregate and deposit in low-permeability reservoirs during CO2 flooding. When the nanofluid is injected, the asphaltene will be adsorbed by the nanoparticles, making it stably dispersed in the crude oil system, as shown in Figure 12.

4.1.2. Dispersion of Nanoparticles

The self-association between asphaltene molecules will accelerate the flocculation and deposition of asphaltene. The dispersion of asphaltene by nanoparticles is mainly based on the formation of stable interaction force or space steric hindrance between the nanoparticles and asphaltene molecules. The dispersion effect disrupts the self-aggregation of asphaltene molecules, thereby inhibiting asphaltene deposition and removing asphaltene aggregates [125]. The surface-modified nanoparticles are prepared into nanofluid suspension and injected into the reservoir. When the nanoparticles enter the space between the gum and asphaltene, they can bind to asphaltene through dipole interaction, charge-transfer complexing interaction, and hydrogen-bonding interaction. An alkyl space stabilizing layer is established on the surface of asphaltene molecules, thereby dispersing the asphaltene and preventing its self-aggregation, and thus stabilizing the asphaltene and preventing it from being deposited [126]. As shown in Figure 13, the surface-modified nanoparticles are prepared into a nanofluid suspension and injected into the reservoir. When the nanoparticles enter the space between the gum and asphaltene, they bind together with the asphaltene through the effects of dipole interactions, charge transfer, and hydrogen bonding. The asphaltene molecules form a stable layer on the surface, which in turn disperses the asphaltene, avoiding self-aggregation and stabilizing the asphaltene to prevent deposition. Setoodeh found that the presence of multiple sulfur heterocycles in the structure of PT, which produce higher polarity in PT and can form stronger interactions with heteroatoms such as N, S, and O in asphaltene molecules. This creates more effective PT-coated nanoparticles than Fe3O4 nanoparticles in inhibiting asphaltene deposition [127].
Nanofluids can further strip the heavy oil in the pores where CO2 does not enter through the CO2 displacement channel. Nanofluids can change the wettability of the pore surface, so that the oil on the surface of the micropore is stripped into oil droplets during the CO2 flooding process. The advantage of the synergistic effect of the two is that nanofluids can not only change the reservoir wettability, but also increase the water injection capacity. Moreover, it can inhibit the pointing phenomenon and provide a larger sweep area for subsequent CO2 flooding [128,129]. In fact, nanoparticles are considered to be effective additives for increasing sweep efficiency, and can also increase the viscosity and density of SC-CO2. In Cao et al.’s experiment on water-based nanofluid-alternating-CO2 injection, it was found that the mass content of C10~C16 intermediate components in the crude oil extracted using NWAG was 9.55%, which was much lower than that of water-alternating-gas (WAG). The results show that after emulsification with nanofluids, CO2 can extract more intermediate components from crude oil, which improves the interaction between crude oil and CO2 [121]. The above studies focused on the development of mathematical models for water-based nanofluids, so SC-CO2-based nanofluids were studied by Dezfuli et al. The results indicate that for heavy oil reservoirs, the maximum ultimate recovery of heavy oil obtained by injection of SC-CO2 is 27.82% under the condition of 5.0 vol. % nano-particle concentration [130].

4.2. Polymer Enhanced CO2 Flooding

4.2.1. Polymer Improves CO2 Channeling

CO2 flooding has the disadvantage of low sweep efficiency, which is mainly caused by gas flow in the high-permeability area of heterogeneous reservoir, viscous fingering caused by viscosity difference between oil and gas, and gravity override caused by density difference between oil and gas [131]. The main reason for CO2 gas channeling is that the viscosity of CO2 is much smaller than that of formation water and crude oil, and the viscosity pointing caused by the unfavorable mobility ratio, which makes the injected CO2 bypass the crude oil and crossflow occurs, and the sweep efficiency decreases. The heterogeneity of the reservoir causes injected CO2 to form channeling paths along high-permeability zones, which will affect the sweep efficiency [132].
PAM gel is a conventional crosslinked gel polymer, which is mainly composed of polyacrylamide macromolecules and a crosslinking agent. It has the advantages of low adsorption loss, insensitivity to bacterial attack, and easy dissolution [133]. As a weak gel, HPAM preferentially enters the high-permeability region through porous media by flowing and undergoing stretching deformation along the large pores. This behavior can reduce the permeability of high-permeability layers, improve the intralayer heterogeneity of reservoir, expand the swept volume, and thus enhance oil recovery [134]. HPAM contains a large number of hydrophilic groups such as carboxyl and amide groups, and the electrostatic repulsion between carboxyl groups can unfold its main chain. The hydration and electrostatic repulsion of hydrophilic groups increase the volume of gel particles, showing a larger hydrodynamics radius. Simultaneously, the three-dimensional network structure formed by crosslinking allows it to possess a certain degree of strength, enabling it to create blockages in the deeper layers of the reservoir and alter the flow direction of the fluid [135,136]. As shown in Figure 14, Al-Ali et al. found that low concentration of HPAM gel has weaker strength and would flow with the produced oil, and only a small part played the role of blocking CO2. However, it is effective to inject high concentration HPAM gel to block the permeation channel or crack [137]. The study conducted by Sun et al. revealed that the polyacrylamide-chromium(III) polymer gel would be degraded when it is in contact with CO2. Therefore, after the gel in the long-term CO2 environment is degraded, a channel will be formed inside it, that is, the gel breaks through, thereby reducing the plugging efficiency of the polymer gel [138]. In summary, delayed crosslinked PAM can effectively block channeling and improve sweep efficiency, which is suitable for high temperature and high salt reservoirs. However, it has disadvantages such as uncontrollable gel time, easy blockage, and damage to the formation.
Preformed particle gel (PPG) is typically formed by PAM or HPAM with crosslinking agents on the ground to form gel particles. Both retarding crosslinking HPAM and pre-crosslinking gel particles diverting CO2 to the low-permeability zone by plugging high-permeability zones to improve swept efficiency. Zhou et al. found that AR-PPG exhibited better swelling and shear resistance in acidic environments. This is mainly due to the fact that the electrostatic repulsion between the cationic group and H+ and the steric hindrance effect between the sulfonic acid groups make the polymer chain difficult to be compressed, thereby increasing the acid resistance and shear resistance of AR-PPG [139]. The large presence of CO2 leads to an acidic environment in the reservoir during CO2 flooding. The development of PPG gel has resolved the issue of reduced plugging efficiency caused by the degradation of PAM in acidic reservoirs. IPN-ASSAP polymer gels were synthesized by Pu et al., which exhibit excellent CO2 responsiveness. CO2 will change the particle size, strength, and other properties of the particles, thereby improving the plugging capacity [140]. In conclusion, the research on PPG polymer gels has addressed the shortcomings of traditional PAM gels, which cannot remain stable in acidic environments for a long time. However, preformed particle gel still have the disadvantages of being too large in size, having a fast water absorption rate, and being difficult to inject.

4.2.2. Polymer Enhanced Stability of CO2 Foam

In the subsurface environment, SC-CO2 has lower viscosity and the density of SC-CO2 is lower than that of water and heavy oil, which leads to the phenomenon of viscous fingering and gravity override in the process of reservoir seepage. The above phenomenon will lead to a reduction in its wave coefficient and seriously weaken the effect of the recovery rate. It has been demonstrated that CO2 foam can effectively adjust the section plane of the displacement front and reduce the fluidity of CO2 [141]. The generating ability and stability of foam can effectively improve the sweep coefficient of foam flooding. However, the high solubility of CO2 in water accelerates the diffusion rate of CO2 through the liquid film, resulting in CO2 foam generally being less stable than nitrogen foam [142,143]. Therefore, how to effectively improve the stability of CO2 foam has attracted much attention in recent years [144].
Polymer-reinforced foam is a kind of oil displacement system in which a polymer is added to the foam as a foam stabilizer. In recent years, polymer-reinforced foam systems have been extensively reported in CO2 foam flooding. Lv et al. found that with the addition of HPAM, the drainage rate of CO2 foam film decreased. The liquid inside the foam film is not easily lost, and the permeability of the liquid film is reduced, inhibiting the diffusion of gas within the foam. In the experimental range, when the concentration of HPAM is 0.025%, the CO2 foam half-life more than doubled and reached a maximum. If the polymer is added in excess, the foamability will be greatly reduced [145]. This is mainly because foam stability is primarily related to film drainage rate, Marangoni repair effect for the film, foam disproportionation rate, and liquid-sheet disintegration rate. The viscosity of the solution increases substantially with the addition of polymer. According to the film drainage rate formula, the film drainage rate decreases and the liquid film thickness increases when the solution viscosity increases. It also reduces the gas diffusion rate between the bubbles and decreases the foam disproportionation rate. On a macroscopic level, the foam exhibits higher stability [146].
V R E = d h f d t = 2 h f 3 3 η c r f i l m Δ P f i l m
V R E : film drainage rate, m/s, h f : liquid film thickness, m, t: time, s, r f i l m : liquid film curvature radius, m, η c : solution viscosity, Pa·s, and P f i l m : pressure difference on both sides of liquid film, Pa.
Within a certain mass concentration range, polymers can increase the viscosity of the foam system, increase the thickness of the foam film, reduce the film drainage rate and diffusion rate of the foam, and enhance the stability of the foam [147]. This view was also confirmed in the experiments of Zhang et al. Hydroxypropyl methyl cellulose (HPMC) can be used as a foam stabilizer to effectively improve the stability of the foam, forming a polymer-reinforced CO2 foam system. However, the foaming ability of the foaming agent system will be weakened by excessively high polymer quality fraction [148]. In addition, the addition of PAM to the foam solution also improves the strength of the liquid film surface, thereby reducing the emission of liquid in the foam and minimizing the diffusion of the gas phase [149].
The above studies only introduce the stability mechanism of polymer on CO2 foam from the macro aspect, and some researchers have carried out relevant discussions on the micro scale. Polymer-reinforced properties of CO2 foam depend on the type of polymer. Pu et al. found that the CHSB-nonionic polyacrylamide system was more stable than the CHSB-anionic polyacrylamide system. Specifically, the foam reinforced with xanthan gum exhibited the best stability. As can be seen from Figure 15, the shape of the CHSB foam changes to a polyhedral shape over time, while the polymer-reinforced foam has a smaller change in size [150]. Yang et al. found that the three-dimensional network spatial structure of acid-resistant hydrophobic polymer nanospheres prompts them to form a “skeleton” at the gas–liquid interface. It increases the strength of the liquid film and regulates the viscoelasticity and liquid storage capacity of the liquid film. Polymer nanospheres swell and release free water, which prevents the rupture of liquid films and increases foam stability. Figure 16 shows the mechanism of stabilizing CO2 foam with nanospheres [41].

4.3. Surfactants Improve CO2 Flooding

In order to improve the effect of heavy oil mining, various additives are often used to assist CO2 to further increase the EOR. After the addition of a surfactant, the solubility of CO2 in crude oil under high pressure conditions increases, but the increase is not significant. The main reason is that CO2 is a non-polar substance, while surfactants have the capability to alter the properties of CO2, thereby increasing its solubility in crude oil [151,152]. In general, the carbon atoms on the carbon chain of surfactants (C12–C18) belong to the light components of crude oil. When the surfactant is added, it disrupts the initial oil–gas phase equilibrium of crude oil, so that CO2 can extract more light components in the crude oil [153]. The high content of macromolecular gums and asphaltenes in heavy oil is the main factor leading to a high value of the MMP. With the addition of surfactants, the internal structure of the molecules of the heavy oil is disrupted and their content is reduced, thus allowing the properties of the crude oil to change. Surfactants with polar groups such as ester groups and hydroxyl groups have the ability to form strong hydrogen bonds with hydroxyl or amino groups in gums and asphaltenes. Hydrogen bonding causes the molecular structure to be branched and the cohesive energy of the crude oil to be reduced, which plays a role in viscosity reduction [154,155]. The surfactant forms reverse micelle or microemulsion structure with CO2 and crude oil. The polar groups of surfactants form a polar core with the polar components in crude oil through hydrogen bonding. The polar groups of surfactants form a polar core with the polar components in crude oil through hydrogen bonding, and polar macromolecules such as asphaltenes and gums are encapsulated within it. The highly branched hydrophobic alkyl chains extend into CO2 and aggregate with each other by van der Waals force and thus increasing the extent of CO2 miscibility in crude oil [156,157,158]. When the surfactant spontaneously accumulates at the two-phase interface, the interaction between the CO2-philic end and CO2 is close to the interaction between the oil-wet end and crude oil. The pressure difference between the two phases decreases, causing the interfacial tension between the oil and gas phases to decrease.

4.3.1. Surfactants on Foaming Properties and Foam Stability of CO2 Foam

The surfactant is injected before the CO2 injection, and foam is generated when the surfactant comes into contact with the CO2. It can selectively block the high-permeability channeling paths, improve the heterogeneity of the formation, and expand the swept volume, which effectively delays the time of CO2 channeling and prolongs the contact time between CO2 and the remaining oil [159]. Foaming ability and foam stability are two important properties to consider in CO2 foam flooding. Surfactants, with their characteristics of fast foaming speed, high foam output, and high foaming ability at lower concentrations, are typical chemical agents used in assisting CO2 flooding. Therefore, the foaming ability of surfactants has been extensively studied by researchers. Jones et al. suggest that as the kinetic energy of surfactant molecules increases, they are more likely to adsorb at the gas-liquid interface. A greater adsorption amount indicates a reduction in surface tension, which benefits the foaming ability of EAB surfactants [160]. This is because the generation of bubbles is related to surface tension, as shown in the following equation.
W e = G μ S R γ
G : shearing stress, μ S :viscosity of continuous phase, R : bubble radius, γ : IFT [161]. However, the link between interfacial tension and foamability has been viewed differently by researchers. Some scholars believe that the diffusion rate of surfactants increases with the maximum rate of decrease in interfacial tension. As a result, the diffusion rate of foaming properties also increases [162]. Another group of researchers believed that the excessive adsorption of surfactants leads to the disappearance of the interfacial tension differential value, which may result in damage of foaming or the cessation of the Marangoni effect [163]. The different perspectives arise from the different methods of forming CO2 foam [164,165]. Foam stability is determined by the interaction of interfacial tension, elasticity, temperature, and pressure. For example, when the bubbles expand, the decrease in surfactant concentration at the expansion site causes a change in interfacial tension. At this time, the surfactant migrates to the surface of the thin film, compensating for the thinning of the film [166].
CO2 dissolves in water, causing the aqueous solution to become acidic, which will accelerate the diffusion rate of CO2 through the liquid film. Therefore, it is difficult for conventional surfactants to maintain the stability of CO2 foam for a long time. Nonionic surfactant with an affinity for CO2 have attracted the attention of researchers [167,168,169]. Niu et al. discovered that EAB surfactants form worm-like micelles in foam solutions, increasing the diffusion resistance of CO2 between bubbles. At the same time, the worm-like micelles increase the interfacial viscosity and reduce the drainage rate of the liquid film. The mechanism of EAB surfactant stabilizing CO2 foam is shown in Figure 17 [170].

4.3.2. Surfactants Reduce the Minimum Miscible Pressure (MMP) of CO2 Flooding

CO2 is a non-polar molecule, which has poor solubility with polar macromolecules such as gum and asphaltene in crude oil. In order to improve their miscibility, it is an effective method to add suitable surfactants to reduce the interfacial tension between CO2 and crude oil. Table 1 summarizes the experimental data of surfactants reducing the MMP between CO2 and crude oil.
CO2 is a weak solvent with low dielectric constant and low polarity, which causes most surfactants to be insoluble or slightly soluble in SC-CO2. The hydrophobic tail chain of the gas-soluble surfactants contain functional groups have an affinity for CO2, thereby enhancing the compatibility between surfactants and CO2. In addition to this, it has a low cohesive energy density, reducing the interaction between CO2 and surfactants. The branched chain on the hydrophilic group increases the space steric hindrance, and the more flexible the molecular chain is, the better the compatibility with CO2. Luo et al. showed that the surfactant CiPOj significantly reduced the IFT between CO2 and crude oil, and the extent of CO2 miscibility with crude oil could be directly observed. In the scenario where the dosage of CiPOj is 0.6% and the temperature is 60 °C, the MMP and first contact miscibility pressure of the CO2 and crude oil systems decreased from 19.1 MPa and 43.0 MPa to 13.8 MPa and 19.0 MPa, respectively [176]. The study by Sun et al. found that diisobutyl citrate can reduce the MMP between CO2 and oil. Injecting surfactants through the slug method can improve the EOR compared to CO2 flooding [181]. Currently, research on reducing the MMP between CO2 and oil systems using gas-soluble surfactants is in its preliminary stage, with shortcomings such as low degradability and high toxicity, which limits their application.
Based on the above reasons, oil-soluble surfactants have attracted the attention of researchers. The study conducted by Wu et al. demonstrates that adding a concentration of 0.1% to 0.5% of surfactant (composed of a mixture of five surfactants) in crude oil can effectively reduce the oil–gas interfacial tension. Among them, the addition of 0.5% mix-benzene (aromatic hydrocarbon mixture) achieved the highest reduction, reaching 25.51%. The addition of mixed benzene increases the light hydrocarbon component of the crude oil and reduces the viscosity of the crude oil, which is conducive to the dissolution of CO2 in the crude oil and reduces the interfacial tension between the CO2 and the crude oil [171]. The oil displacement mechanism of oil-soluble surfactants and CO2-soluble surfactants are compared in Figure 18, highlighting the limitations of oil-soluble surfactants. The development of amphiphilic surfactants is the main research direction at present [179]. Yang et al. found that acetyl glucose is a CO2-philic end and aliphatic hydrocarbon is oil-wet end. Therefore, the amphiphilic acetyl glucose dodecyl ester can reduce the interfacial tension, and the addition of a 1% concentration of the reagent reduced the MMP by 27.47% [155].

5. Conclusions and Prospect

Low-permeability heavy oil reservoirs typically have low permeability and high porosity, which pose challenges to the flow of CO2 and its oil displacement efficiency. CO2 may encounter greater resistance in low-permeability reservoirs, resulting in poor mobility and limiting its diffusion capacity within the reservoir, leading to lower recovery. Heavy oil has higher viscosity and temperature, which increases the complexity of the interaction between CO2 and oil, and the solubility and diffusion rate of CO2 are influenced by temperature and viscosity. In addition, in low-permeability heavy oil reservoirs, it is a challenge for SC-CO2 to maintain the required temperature and pressure throughout the reservoir. Low-permeability heavy oil reservoir typically exhibit reservoir homogeneity, resulting in differences in rock properties and fluid distribution. This heterogeneity will affect the sweep efficiency and the distribution of CO2 in the reservoir, resulting in uneven recovery. Due to the low permeability of the reservoir, the transfer rate of CO2 is slow, and it is easy to take shortcuts, which will also lead to uneven distribution of CO2. Meanwhile, the migration path and dose control of CO2 are the key issues. In addition, CO2 can affect wellbore integrity, lead to cement degradation or casing corrosion, and can cause reservoir damage such as oil well plugging and pore collapse, thus impacting the long-term stability of the well. Ensuring the safe capture, transport, and storage of CO2 and reducing its impact on the environment is also a major challenge.
The article provides a detailed description of the main oil recovery mechanisms of SC-CO2 in low-permeability heavy oil reservoirs, as well as the improvement effect of nanoparticles, polymers, and surfactants on the defects of CO2 flooding. The dissolution of CO2 expands the volume of heavy oil, which can increase the elastic strain energy of formation, and thus increase the internal kinetic energy of crude oil. Capillary resistance and flow resistance are greatly reduced, the volume of formation void is increased, the flow of oil flow in rock pores is promoted, and the oil recovery rate is improved. CO2 interacts with heavy oil molecules through physical adsorption and chemical reactions. The infiltration of CO2 disrupts the internal structure and arrangement of heavy oil, breaking the adhesion force between oil molecules and reducing the viscosity of heavy oil, thereby improving its flowability. CO2 extracts light and heavy components in heavy oil or vaporizes them, thereby reducing the surface tension of oil–water interface, reducing residual oil saturation, and thereby increasing crude oil recovery. Asphaltene deposition, gas channeling, stability of CO2 foam, and reduction in minimum reverberation pressure are all problems to be considered in CO2 flooding. Therefore, this paper focuses on the improvement mechanism of nanoparticles, polymers, and surfactants on CO2 flooding. In response to the above challenges and the progress of current research, more laboratory studies and numerical simulations are needed to understand the interaction and flow mechanisms between CO2 and heavy oil. In addition, new dosage control methods need to be further developed to enhance the fluidity and solubility of CO2, and advanced monitoring and evaluation techniques should be employed to optimize the SC-CO2 flooding process.

Author Contributions

Y.S. was responsible for the conceptualization; visualization; methodology; formal analysis; and editing, and review of the original draft. W.Z. completed the literature research and drafted this manuscript. J.T. and Y.M. are responsible for the field investigation and analysis of oil displacement data. L.Z. is responsible for the supervision and reasonableness of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Liaoning Provincial Education Department grant number Grant No.LJKZ0420.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Chen, S.; Liu, Y.; Zhang, J.; Li, P.; Tang, X.; Li, Z.; Dong, Z.; Xu, L.; Zhao, X. Formation conditions and evolution of fractures in multiple tight rocks: Implications for unconventional reservoir exploitation. J. Pet. Sci. Eng. 2021, 200, 108354. [Google Scholar] [CrossRef]
  2. Dong, K.; Jiang, M.; Li, J.; Zhang, D. Research progresses in formation mechanism of complex fracture network for unconventional reservoir. Arab. J. Geosci. 2020, 13, 750. [Google Scholar] [CrossRef]
  3. Pei, Y.; Zhou, H.; Li, D.; Zhang, S.; Tian, F.; Zhao, L. A Novel Unconventional Reservoir Fracturing Method-Liquid Self-Propping Fracturing Technology. In Proceedings of the International Petroleum Technology Conference, Dhahran, Kingdom of Saudi Arabia, 13–15 January 2020. [Google Scholar]
  4. Zhou, X.; Yuan, Q.; Peng, X.; Zeng, F.; Zhang, L. A critical review of the CO2 huff ‘n’ puff process for enhanced heavy oil recovery. Fuel 2018, 215, 813–824. [Google Scholar] [CrossRef]
  5. Luo, P.; Zhang, Y.; Huang, S. A promising chemical-augmented WAG process for enhanced heavy oil recovery. Fuel 2013, 104, 333–341. [Google Scholar] [CrossRef]
  6. Luo, T.; Yao, Z.; Duan, J.; Li, J.; Zhao, Y. Research on Geologic Characterization and 3D Geological Modeling for Low Permeability Reservoir. J. Liaoning Petrochem. Univ. 2020, 40, 40–44. [Google Scholar]
  7. Fang, Y.; Yang, E.; Yin, D.; Gan, Y. Study on distribution characteristics of microscopic residual oil in low permeability reservoirs. J. Dispers. Sci. Technol. 2019, 41, 575–584. [Google Scholar] [CrossRef]
  8. Shi, Y.; Jia, Y.; Pan, W.; Huang, L.; Yan, J.; Zheng, R. Potential evaluation on CO2-EGR in tight and low-permeability reservoirs. Nat. Gas Ind. B 2017, 4, 311–318. [Google Scholar] [CrossRef]
  9. Salmo, I.C.; Pettersen, Ø.; Skauge, A. Polymer Flooding at Adverse Mobility Ratio -Acceleration of Oil Production by Crossflow into Water Channels. Energy Fuels 2017, 31, 5948–5958. [Google Scholar] [CrossRef]
  10. Li, L.; Su, Y.; Hao, Y.; Zhan, S.; Lv, Y.; Zhao, Q.; Wang, H. A comparative study of CO2 and N2 huff-n-puff EOR performance in shale oil production. J. Pet. Sci. Eng. 2019, 181, 106174. [Google Scholar] [CrossRef]
  11. Wan, T.; Wang, X.; Jing, Z.; Gao, Y. Gas injection assisted steam huff-n-puff process for oil recovery from deep heavy oil reservoirs with low-permeability. J. Pet. Sci. Eng. 2020, 185, 106613. [Google Scholar] [CrossRef]
  12. Cheng, Q.; Li, Z.; Zhu, G.; Zhang, H. Research and Application of CO2 Flooding Enhanced Oil Recovery in Low Permeability Oilfield. Open J. Geol. 2017, 7, 1435–1440. [Google Scholar] [CrossRef]
  13. Fang, T.; Wang, M.; Gao, Y.; Zhang, Y.; Yan, Y.; Zhang, J. Enhanced oil recovery with CO2/N2 slug in low permeability reservoir: Molecular dynamics simulation. Chem. Eng. Sci. 2019, 197, 204–211. [Google Scholar] [CrossRef]
  14. Wei, J.; Zhou, J.; Li, J.; Zhou, X.; Dong, W.; Cheng, Z. Experimental study on oil recovery mechanism of CO2 associated enhancing oil recovery methods in low permeability reservoirs. J. Pet. Sci. Eng. 2021, 197, 108047. [Google Scholar] [CrossRef]
  15. Zhao, F.; Hao, H.; Hou, J.; Hou, L.; Song, Z. CO2 mobility control and sweep efficiency improvement using starch gel or ethylenediamine in ultra-low permeability oil layers with different types of heterogeneity. J. Pet. Sci. Eng. 2015, 133, 52–65. [Google Scholar] [CrossRef]
  16. Du, Y.; Liu, L.; Liao, G.; Zhang, F.; Jiaqiang, E. Density fluctuations and transport properties of supercritical carbon dioxide calculated by molecular dynamics simulation near the Widom line. J. Supercrit. Fluids 2023, 200, 106003. [Google Scholar] [CrossRef]
  17. Wang, S.; Chen, C.; Shiau, B.; Harwell, J.H. In-situ CO2 generation for EOR by using urea as a gas generation agent. Fuel 2018, 217, 499–507. [Google Scholar] [CrossRef]
  18. Huang, L.; Liu, X.; Xiong, J.; Liang, L. Experimental on the pore structure characteristics of Longmaxi Formation shale in southern Sichuan Basin, China. Petroleum 2021, 7, 135–141. [Google Scholar] [CrossRef]
  19. Su, X.; Yue, X. Mechanism study of the relation between the performance of CO2 immiscible flooding and rock permeability. J. Pet. Sci. Eng. 2020, 195, 107891. [Google Scholar] [CrossRef]
  20. Xiao, P.; Yang, Z.; Wang, X.; Xiao, H.; Wang, X. Experimental investigation on CO2 injection in the Daqing extra/ultra-low permeability reservoir. J. Pet. Sci. Eng. 2017, 149, 765–771. [Google Scholar] [CrossRef]
  21. Yang, H.; Zhou, B.; Zhu, T.; Wang, P.; Zhang, X.; Wang, T.; Wu, F.; Zhang, L.; Kang, W.; Ketova, Y.A.; et al. Conformance control mechanism of low elastic polymer microspheres in porous medium. J. Pet. Sci. Eng. 2021, 196, 107708. [Google Scholar] [CrossRef]
  22. Fakher, S.; Imqam, A. An experimental investigation of immiscible carbon dioxide interactions with crude oil: Oil swelling and asphaltene agitation. Fuel 2020, 269, 117380. [Google Scholar] [CrossRef]
  23. Rangriz Shokri, A.; Babadagli, T. Feasibility assessment of heavy-oil recovery by CO2 injection after cold production with sands: Lab-to-field scale modeling considering non-equilibrium foamy oil behavior. Appl. Energy 2017, 205, 615–625. [Google Scholar] [CrossRef]
  24. Huang, T.; Zhou, X.; Yang, H.; Liao, G.; Zeng, F. CO2 flooding strategy to enhance heavy oil recovery. Petroleum 2017, 3, 68–78. [Google Scholar] [CrossRef]
  25. Song, Z.; Zhu, W.; Wang, X.; Guo, S. 2-D Pore-Scale Experimental Investigations of Asphaltene Deposition and Heavy Oil Recovery by CO2 Flooding. Energy Fuels 2018, 32, 3194–3201. [Google Scholar] [CrossRef]
  26. Rezk, M.G.; Foroozesh, J. Effect of CO2 mass transfer on rate of oil properties changes: Application to CO2-EOR projects. J. Pet. Sci. Eng. 2019, 180, 298–309. [Google Scholar] [CrossRef]
  27. Wanat, E.; Teletzke, G.; Newhouse, D.; Lawrence, J.; Shehhi, R.A.; Willingham, T.; Jawhari, M.O. Quantification of Oil Recovery Mechanisms during Gas Injection EOR Coreflood Experiments. In Proceedings of the Abu Dhabi International Petroleum Exhibition & Conference, Abu Dhabi, United Arab Emirates, 13–16 November 2017. [Google Scholar]
  28. Al-Shargabi, M.; Davoodi, S.; Wood, D.A.; Rukavishnikov, V.S.; Minaev, K.M. Carbon Dioxide Applications for Enhanced Oil Recovery Assisted by Nanoparticles: Recent Developments. ACS Omega 2022, 7, 9984–9994. [Google Scholar] [CrossRef]
  29. Qian, K.; Yang, S.; Dou, H.; Wang, Q.; Wang, L.; Huang, Y. Experimental Investigation on Microscopic Residual Oil Distribution During CO2 Huff-and-Puff Process in Tight Oil Reservoirs. Energies 2018, 11, 2843. [Google Scholar] [CrossRef]
  30. Shi, Y.; Yang, D. Quantification of Gas Exsolution and Mass Transfer of Foamy Oil in Bulk Phase under Solution Gas Drive Conditions. In Proceedings of the SPE Western Regional Meeting, Garden Grove, CA, USA, 22–26 April 2018. [Google Scholar]
  31. Bakhshi, P.; Kharrat, R.; Hashemi, A.; Zallaghi, M. Experimental evaluation of carbonated waterflooding: A practical process for enhanced oil recovery and geological CO2 storage. Greenh. Gases Sci. Technol. 2018, 8, 238–256. [Google Scholar] [CrossRef]
  32. Fakher, S.; Imqam, A. Asphaltene precipitation and deposition during CO2 injection in nano shale pore structure and its impact on oil recovery. Fuel 2019, 237, 1029–1039. [Google Scholar] [CrossRef]
  33. Sun, Y.; Zhang, W.; Li, J.; Han, R.; Lu, C. Mechanism and Performance Analysis of Nanoparticle-Polymer Fluid for Enhanced Oil Recovery: A Review. Molecules 2023, 28, 4331. [Google Scholar] [CrossRef]
  34. Li, N.; Mao, G.; Liu, Y. Effect of the Evaluation and Mechanism Analysis of a Novel Nanohybrid Pour Point Depressant on Facilitating Flow Properties of Crude Oil. Energy Fuels 2018, 32, 10563–10570. [Google Scholar] [CrossRef]
  35. Kazemzadeh, Y.; Shojaei, S.; Riazi, M.; Sharifi, M. Review on application of nanoparticles for EOR purposes: A critical review of the opportunities and challenges. Chin. J. Chem. Eng. 2019, 27, 237–246. [Google Scholar] [CrossRef]
  36. Ke, H.; Yuan, M.; Xia, S. A review of nanomaterials as viscosity reducer for heavy oil. J. Dispers. Sci. Technol. 2020, 43, 1271–1282. [Google Scholar] [CrossRef]
  37. Montes, D.; Henao, J.; Taborda, E.A.; Gallego, J.; Cortes, F.B.; Franco, C.A. Effect of Textural Properties and Surface Chemical Nature of Silica Nanoparticles from Different Silicon Sources on the Viscosity Reduction of Heavy Crude Oil. ACS Omega 2020, 5, 5085–5097. [Google Scholar] [CrossRef]
  38. Rezaeiakmal, F.; Parsaei, R.; Shafiabadi, A.; Rezaei, A. Insights into the flow behaviour of the pre-generated polymer enhanced foam in heterogeneous porous media during tertiary oil recovery: Effect of gravitational forces. J. Pet. Sci. Eng. 2022, 213, 110385. [Google Scholar] [CrossRef]
  39. Yang, H.; Lv, Z.; Wang, L.; Feng, C.; Wang, J.; Xu, Z.; Huang, Y.; Li, Z.; Kang, W. Stability mechanism of controlled acid-resistant hydrophobic polymer nanospheres on CO2 foam. Fuel 2023, 346, 128332. [Google Scholar] [CrossRef]
  40. Abdelaal, A.; Gajbhiye, R.; AlShehri, D.; Mahmoud, M.; Patil, S.; Zhou, X. Improvement of supercritical carbon dioxide foam performance for EOR in sandstone reservoirs: An experimental and optimization study. Gas Sci. Eng. 2023, 110, 204859. [Google Scholar] [CrossRef]
  41. Shen, H.; Yang, Z.; Li, X.; Peng, Y.; Lin, M.; Zhang, J.; Dong, Z. CO2-responsive agent for restraining gas channeling during CO2 flooding in low permeability reservoirs. Fuel 2021, 292, 120306. [Google Scholar] [CrossRef]
  42. Wang, J.; Zhou, F.; Fan, F.; Zhang, L.; Yao, E. Study on the Influence of CO2 Finger-Channeling Flooding on Oil Displacement Efficiency and Anti-Channeling Method. In Proceedings of the ARMA-CUPB Geothermal International Conference, Beijing, China, 5–8 August 2019. [Google Scholar]
  43. Zhang, L.; Zhao, F.; Hou, J. Experimental Study of Improving the CO2 Flooding Development Effect in Ultra-Low Permeability Reservoir. Adv. Mater. Res. 2012, 594, 2470–2474. [Google Scholar]
  44. Li, Q.; Zhao, D.; Yin, J.; Zhou, X.; Li, Y.; Chi, P.; Han, Y.; Ansari, U.; Cheng, Y. Sediment Instability Caused by Gas Production from Hydrate-bearing Sediment in Northern South China Sea by Horizontal Wellbore: Evolution and Mechanism. Nat. Resour. Res. 2023, 321, 1595–1620. [Google Scholar] [CrossRef]
  45. Li, Q.; Zhang, C.; Yang, Y.; Ansari, U.; Han, Y.; Li, X.; Cheng, Y. Preliminary experimental investigation on long-term fracture conductivity for evaluating the feasibility and efficiency of fracturing operation in offshore hydrate-bearing sediments. Ocean Eng. 2023, 281, 114949. [Google Scholar] [CrossRef]
  46. Delgado-García, Y.I.; Luna, S.; López-Malo, A.; Morales-Camacho, J.I. Effect of supercritical carbon dioxide on physicochemical and techno-functional properties of amaranth flour. Chem. Eng. Process. 2022, 178, 109031. [Google Scholar] [CrossRef]
  47. Daş, E.; Gürsel, S.A.; Yurtcan, A.B. Simultaneously deposited Pt-alloy nanoparticles over graphene nanoplatelets via supercritical carbon dioxide deposition for PEM fuel cells. J. Alloys Compd. 2021, 874, 159919. [Google Scholar] [CrossRef]
  48. Senyay-Oncel, D.; Kimiz-Gebologlu, I.; Yesil-Celiktas, O. New developments in supercritical fluids as a green technology: Processing of β-glucanase with sub- and supercritical carbon dioxide. Biochem. Eng. J. 2023, 195, 108934. [Google Scholar] [CrossRef]
  49. Jung, B.; Im, S.; Westerhoff, P.; Park, K. Supercritical carbon dioxide fluid cleaning of extracellular polymeric substance surrogates on seawater reverse osmosis membranes. Desalination 2023, 559, 116637. [Google Scholar] [CrossRef]
  50. Cai, X.; Long, J.; Ren, Q.; Dong, M.; Wang, W.; Tian, S.; Liu, Z. Aggregation Mechanism of Asphaltene Molecular Aggregates. Acta Pet. Sin. (Pet. Process.) 2019, 35, 920–928. [Google Scholar]
  51. Liu, G.; Yang, J.; Song, J.; Xu, X. Inhibition of asphaltene precipitation in blended crude oil using novel oil-soluble maleimide polymers. Energy Sources Part A Recovery Util. Environ. Effects 2019, 41, 2460–2470. [Google Scholar] [CrossRef]
  52. Gray, M.R.; Tykwinski, R.R.; Stryker, J.M.; Tan, X. Supramolecular Assembly Model for Aggregation of Petroleum Asphaltenes. Energy Fuels 2011, 25, 3125–3134. [Google Scholar] [CrossRef]
  53. Liu, J.; Zhao, Y.; Ren, S. Molecular Dynamics Simulation of Self-Aggregation of Asphaltenes at an Oil/Water Interface: Formation and Destruction of the Asphaltene Protective Film. Energy Fuels 2015, 29, 1233–1242. [Google Scholar] [CrossRef]
  54. U.S. Energy Information Administration. Technically Recoverable Shale Oil and Shale Gas Resources: An Assessment of 137 Shale Formations in 41 Countries Outside the United States; U.S. Energy Information Administration: Washington, DC, USA, 2013.
  55. Ji, B.; Fang, J. An overview of efficient development practices at low permeability sandstone reservoirs in China. Energy Geosci. 2023, 4, 100179. [Google Scholar] [CrossRef]
  56. Cong, S.; Li, J.; Liu, W.; Shi, Y.; Li, Y.; Zheng, K.; Luo, X.; Luo, W. EOR mechanism of fracture oil displacement agent for ultra-low permeability reservoir. Energy Rep. 2023, 9, 4893–4904. [Google Scholar] [CrossRef]
  57. Wang, T.; Wang, L.; Meng, X.; Chen, Y.; Song, W.; Yuan, C. Key parameters and dominant EOR mechanism of CO2 miscible flooding applied in low-permeability oil reservoirs. Geoenergy Sci. Eng. 2023, 225, 211724. [Google Scholar] [CrossRef]
  58. Song, S.; Ghanizadeh, A.; Younis, A.; Clarkson, C.R. Experimental evaluation of the impact of induced fractures on cyclic solvent injection performance in low-permeability hydrocarbon reservoirs. Fuel 2023, 354, 129327. [Google Scholar] [CrossRef]
  59. Wang, W.; Guo, X.; Duan, P.; Kang, B.; Zheng, D.; Zafar, A. Investigation of plugging performance and enhanced oil recovery of multi-scale polymer microspheres in low-permeability reservoirs. Nat. Gas Ind. B 2023, 10, 223–232. [Google Scholar] [CrossRef]
  60. Li, W.; Ji, Z. Finite element analysis of temporary plugging and fracturing mechanism. Fault-Block Oil Gas Field 2016, 23, 514–517. [Google Scholar]
  61. Zhou, D.; Xiong, X.; He, J.; Dong, B.; He, Y. Multi-Stage Deflective Fracturing Technology for Low Permeability Reservoir. Pet. Drill. Technol. 2020, 48, 85–89. [Google Scholar]
  62. Li, H.; Li, J.; Li, W.; Hu, J.; Yuan, G.; Liang, Y.; Zhang, X.; Guo, Y. Synthesis of associative polymer for flooding in low permeability reservoir and its reservoir adaptability. Pet. Reserv. Eval. Dev. 2020, 10, 24–32. [Google Scholar]
  63. Wang, G. Microscopic residual oil utilization mechanism of low viscosity polymer flooding in low permeability reservoir. Fault-Block Oil Gas Field 2018, 25, 776–780. [Google Scholar]
  64. Peng, X.; Wang, Y.; Diao, Y.; Zhang, L.; Yazid, I.M.; Ren, S. Experimental investigation on the operation parameters of carbon dioxide huff-n-puff process in ultra low permeability oil reservoirs. J. Pet. Sci. Eng. 2019, 174, 903–912. [Google Scholar] [CrossRef]
  65. Wang, G. Carbon dioxide capture, enhanced-oil recovery and storage technology and engineering practice in Jilin Oilfield, NE China. Pet. Explor. Dev. 2023, 50, 245–254. [Google Scholar] [CrossRef]
  66. Sun, X.; Du, M.; Han, B.; Sun, Y.; Zhao, M.; Guan, B.; Dai, C. Review on Carbon Dioxide Fracturing Technology. Oilfield Chem. 2017, 34, 374–380. [Google Scholar]
  67. Zhang, L.; Wang, S.; Zhang, L.; Ren, S.; Guo, Q. Assessment of CO2 EOR and its geo-storage potential in mature oil reservoirs, Shengli Oilfield, China. Pet. Explor. Dev. 2009, 36, 737–742. [Google Scholar]
  68. Pedersen, K.S.; Christensen, P.L.; Shaikh, J.A. Phase Behavior of Petroleum Reservoir Fluids; CRC Press: Boca Raton, FL, USA, 2014. [Google Scholar]
  69. Cai, X.; Zhou, Z.; Du, F. Increase production mechanism and application of CO2 huff and puff technology. Oil Drill. Prod. Technol. 2002, 24, 45–46. [Google Scholar]
  70. Zhang, T.; Ming, T.; Yuan, L.; Zhu, G.; Zhang, C.; Liu, Y.; Li, Y.; Wang, W.; Yang, X. Experimental study on stress-dependent multiphase flow in ultra-low permeability sandstone during CO2 flooding based on LF-NMR. Energy 2023, 278, 127874. [Google Scholar] [CrossRef]
  71. Sambo, C.; Liu, N.; Shaibu, R.; Ahmed, A.A.; Hashish, R.G. A Technical Review of CO2 for Enhanced Oil Recovery in Unconventional Oil Reservoirs. Geoenergy Sci. Eng. 2023, 221, 111185. [Google Scholar] [CrossRef]
  72. Sharafi, M.S.; Ghasemi, M.; Ahmadi, M.; Kazemi, A. An experimental approach for measuring carbon dioxide diffusion coefficient in water and oil under supercritical conditions. Chin. J. Chem. Eng. 2021, 34, 160–170. [Google Scholar] [CrossRef]
  73. Whorton, L.P.; Brownscombe, E.R.; Dyes, A.B. Method for Producing Oil by Means of Carbon Dioxide. U.S. Patent No. 2,623,596; U.S. Patent and Trademark Office, Washington, DC, USA, 30 December 1950. [Google Scholar]
  74. Yang, Z.; Li, M.; Peng, B.; Lin, M.; Dong, Z. Dispersion Property of CO2 in Oil. 1. Volume Expansion of CO2+ Alkane at near Critical and Supercritical Condition of CO2. J. Chem. Eng. Data 2012, 57, 1305–1311. [Google Scholar] [CrossRef]
  75. Zhang, J.; Pan, Z.; Liu, K.; Burke, N. Molecular Simulation of CO2 Solubility and Its Effect on Octane Swelling. Energy Fuels 2013, 27, 2741–2747. [Google Scholar] [CrossRef]
  76. Liu, B.; Shi, J.; Sun, B.; Shen, Y.; Zhang, J.; Chen, X.; Wang, M. Molecular dynamics simulation on volume swelling of CO2–alkane system. Fuel 2015, 143, 194–201. [Google Scholar] [CrossRef]
  77. Yang, Z.; Li, M.; Peng, B.; Lin, M.; Dong, Z. Volume expansion of CO2+oil at near critical and supercritical conditions of CO2. Fuel 2013, 112, 283–288. [Google Scholar] [CrossRef]
  78. Han, H.; Yuan, S.; Li, S.; Liu, X.; Chen, X. Dissolving capacity and volume expansion of carbon dioxide in chain n-alkanes. Pet. Explor. Dev. 2015, 42, 97–103. [Google Scholar] [CrossRef]
  79. Li, C.; Pu, H.; Zhao, J. Molecular Simulation Study on the Volume Swelling and the Viscosity Reduction of n-Alkane/CO2 Systems. Ind. Eng. Chem. Res. 2019, 58, 8871–8877. [Google Scholar] [CrossRef]
  80. Wang, Y.; Chen, Y.; Wang, J.; Pan, Z.; Liu, J. Molecular Simulation Study on the Density Behavior of n-Alkane/CO2 Systems. ACS Omega 2021, 6, 29618–29628. [Google Scholar] [CrossRef] [PubMed]
  81. Sima, S.; Milanesio, J.M.; Ramello, J.I.; Cismondi, M.; Secuianu, C.; Feroiu, V.; Geană, D. The effect of the naphthenic ring on the VLE of (carbon dioxide + alkane) mixtures. J. Chem. Thermodyn. 2016, 93, 374–385. [Google Scholar] [CrossRef]
  82. Zhang, W.; Guo, P.; Ge, X.; Du, J.; Wang, Z. Influence factors on CO2 solubility in cycloalkanes and cycloalkane volume expansion: Temperature, pressure and molecular structure. J. Mol. Liq. 2021, 332, 115859. [Google Scholar] [CrossRef]
  83. Wang, J.; Zhou, S.; Bei, K.; Chou, I.M.; Lin, C.; Pan, Z. A new approach for the measurement of the volume expansion of a CO2+ n-dodecane mixture in a fused silica capillary cell by Raman spectroscopy. Fuel 2017, 203, 113–119. [Google Scholar] [CrossRef]
  84. Gao, Y.; Li, C.; Xia, S.; Ma, P. The solubility of CO2 in (hexane + cyclohexane) and (cyclopentane + ethylbenzene) and (toluene + undecane) systems at high pressures. J. Chem. Thermodyn. 2021, 154, 106324. [Google Scholar] [CrossRef]
  85. Souza, J.L.; Fabri, F.; Buffon, R.; Schuchardt, U. Reduced anatase on silica as a support for a Ziegler catalyst. Appl. Catal. A Gen. 2007, 323, 234–241. [Google Scholar] [CrossRef]
  86. He, F.; Zhang, L. Organo-modified ZnAl layered double hydroxide as new catalyst support for the ethylene polymerization. J. Colloid Interface Sci. 2007, 315, 439–444. [Google Scholar] [CrossRef]
  87. Li, Y.; Zhang, J.; Li, M. Research on Supercritical CO2-Heavy Oil Viscosity Reduction Characteristics and Calculation Model. Sci. Technol. Eng. 2013, 13, 294–298. [Google Scholar]
  88. Seyyedsar, S.M.; Sohrabi, M. Intermittent CO2 and viscosity-reducing gas (VRG) injection for enhanced heavy oil recovery. Fuel Process. Technol. 2017, 164, 1–12. [Google Scholar] [CrossRef]
  89. Seyyedsar, S.M.; Farzaneh, S.A.; Sohrabi, M. Experimental investigation of tertiary CO2 injection for enhanced heavy oil recovery. J. Nat. Gas Sci. Eng. 2016, 34, 1205–1214. [Google Scholar] [CrossRef]
  90. Sun, G.; Li, C.; Wei, G.; Yang, F. Characterization of the viscosity reducing efficiency of CO2 on heavy oil by a newly developed pressurized stirring-viscometric apparatus. J. Pet. Sci. Eng. 2017, 156, 299–306. [Google Scholar] [CrossRef]
  91. Wang, L.; Zhao, Q.; Li, Z. Experimental investigation of carbon dioxide flooding in heavy oil reservoirs for enhanced oil recovery. Energy Rep. 2022, 8, 10754–10761. [Google Scholar] [CrossRef]
  92. Miller, J.S.; Jones, R.A. A Laboratory Study To Determine Physical Characteristics Of Heavy Oil After CO2 Saturation. In Proceedings of the SPE/DOE Enhanced Oil Recovery Symposium, Tulsa, OK, USA, 5–8 April 1981. [Google Scholar]
  93. Li, Z.; Li, X.; Yuan, M.; Huang, D.; Zhang, G. Study on laboratory experiments of CO2 drive in Shang 13–22 unit. Pet. Geol. Recovery Effic. 2000, 3, 005. [Google Scholar]
  94. Zhao, G.; Liu, R.; Zhang, W.; Fu, M. Experimental Research on Affection of CO2 to the Physical Property of High Solidification Point Oil. J. Jianghan Pet. Univ. Staff. 2003, 4, 35–37. [Google Scholar]
  95. Xin, S.; Luo, T.; Zhang, Y.; Zhang, L.; He, Y.; Zhang, H. Laboratory test evaluation of CO2 injection miscible flooding in BN reservoir, North China. Pet. Geol. Eng. 2004, 18, 30–32. [Google Scholar]
  96. Zhao, M. Laboratory and Numerical Simulation Study on CO2 Flooding in Ultra Low Permeability Reservoir; Northeast Petroleum University: Daqing, China, 2008; Volume 11, pp. 1469–1472. [Google Scholar]
  97. Liu, S. Rsearch on Laboratory experiment of CO2 drive in ultra-low permeability reservoir. Sci. Technol. Ed. 2011, 33, 133–136. [Google Scholar]
  98. Xiang, P. Research and application of carbon dioxide oil displacement technology. Petrochem. Ind. Technol. 2021, 28, 33–34. [Google Scholar]
  99. Jaffé, R.; Diaz, D.; Hajje, N.; Chen, L.; Eckardt, C.; Furton, K.G. Hydrocarbon speciation in ancient sediments studied by stepwise high-temperature supercritical carbon dioxide extraction. Org. Geochem. 1997, 26, 59–65. [Google Scholar] [CrossRef]
  100. Akinlua, A.; Torto, N.; Ajayi, T.R. Supercritical fluid extraction of aliphatic hydrocarbons from Niger Delta sedimentary rock. J. Supercrit. Fluids 2008, 45, 57–63. [Google Scholar] [CrossRef]
  101. Kesavan, S.K.; Ghosh, A.; Polasky, M.E.; Parameswaran, V.; Lee, S. Supercritical Extraction of Stuart Oil Shale. Fuel Sci. Technol. Int. 1988, 6, 505–523. [Google Scholar] [CrossRef]
  102. Deo, M.D.; Hwang, J.; Hanson, F.V. Supercritical fluid extraction of a crude oil, bitumen-derived liquid and bitumen by carbon dioxide and propane. Fuel 1992, 71, 1519–1526. [Google Scholar] [CrossRef]
  103. Guiliano, M.; Boukir, A.; Doumenq, P.; Mille, G.; Crampon, C.; Badens, E.; Charbit, G. Supercritical Fluid Extraction of Bal 150 Crude Oil Asphaltenes. Energy Fuels 2000, 14, 89–94. [Google Scholar] [CrossRef]
  104. Koel, M.; Ljovin, S.; Bondar, Y. Supercritical carbon dioxide extraction of Estonian oil shale. Oil Shale 2000, 17, 225–235. [Google Scholar] [CrossRef]
  105. Jaffé, R.; Diaz, D.; Furton, K.G.; Lafargue, E. High temperature supercritical carbon dioxide extractions of geological samples: Effects and contributions from the sample matrix. Appl. Geochem. 2000, 15, 79–89. [Google Scholar] [CrossRef]
  106. Ma, Q.; Yang, S.; Chen, H.; Wang, L.; Qian, K.; Meng, Z.; Lei, H.; Wang, Z. Effect and influencing factors of CO2 huff and puff in a tight oil reservoir-Taking the Lucaogou formation in the Xinjiang Jimsar sag as an example. Pet. Sci. Bull. 2018, 3, 434–445. [Google Scholar]
  107. Li, M.; Shan, W.; Liu, X.; Shang, G. Laboratory study on miscible oil displacement mechanism of supercritical carbon dioxide. Acta Pet. Sinica 2006, 27, 80. [Google Scholar]
  108. Wang, Q.; Zhang, N.; Wu, Q.; Zhang, P.; Zhao, J. Research progress on the solubility of supercritical CO2. Refin. Chem. Ind. 2011, 22, 1–5. [Google Scholar]
  109. Peng, B. Study On the Mechanism of Carbon Dioxide Auxiliary Steam Thermal Recovery; China University of Petroleum: Beijing, China, 2012. [Google Scholar]
  110. Fan, P.; Zhu, W.; Lin, j.; Gao, T.; Wang, Y. Factors and Regularity of Supercritical cOExtraction of Heavy Oil lmpact. Sci. Technol. Eng. 2017, 17, 31–36. [Google Scholar]
  111. Allawzi, M.; Al-Otoom, A.; Allaboun, H.; Ajlouni, A.; Al Nseirat, F. CO2 supercritical fluid extraction of Jordanian oil shale utilizing different co-solvents. Fuel Process. Technol. 2011, 92, 2016–2023. [Google Scholar] [CrossRef]
  112. Priyanto, S.; Mansoori, G.A.; Suwono, A. Measurement of property relationships of nano-structure micelles and coacervates of asphaltene in a pure solvent. Chem. Eng. Sci. 2001, 56, 6933–6939. [Google Scholar] [CrossRef]
  113. Ni, H.; Hsu, C.S.; Lee, P.; Wright, J.; Chen, R.; Xu, C.; Shi, Q. Supercritical carbon dioxide extraction of petroleum on kieselguhr. Fuel 2015, 141, 74–81. [Google Scholar] [CrossRef]
  114. Cao, X.; Peng, B.; Ma, S.; Ni, H.; Zhang, L.; Zhang, W.; Li, M.; Hsu, C.S.; Shi, Q. Molecular Selectivity in Supercritical CO2 Extraction of a Crude Oil. Energy Fuels 2017, 31, 4996–5002. [Google Scholar] [CrossRef]
  115. Wang, Q.; Li, Z.; Qian, M.; Jiang, Q.; Cao, T.; Liu, P.; Bao, Y.; Tao, G. Movable oil extraction from shale with supercritical carbon dioxide. Pet. Geol. Exp. 2020, 42, 646–652. [Google Scholar]
  116. Li, H. Molecular Dynamics Simulation for Supercritical Carbon Dioxide Extraction of Heavy Oil Components in Deep Reservoir. Sci. Technol. Eng. 2021, 21, 12543–12550. [Google Scholar]
  117. Ding, C. Research and Analysis of the Influencement of Carbon Dioxide Flooding on Asphaltene Precipitation; China University of Petroleum: Beijing, China, 2013. [Google Scholar]
  118. Chen, L.; Yv, H.; Tang, R.; Wang, W.; Yao, Z. Effect of Asphalt Deposition on CO2 Flooding in Light Oil Reservoir. Oilfield Chem. 2017, 34, 87–91. [Google Scholar]
  119. Kazemzadeh, Y.; Sourani, S.; Doryani, H.; Reyhani, M.; Shabani, A.; Fallah, H. Recovery of Asphaltenic Oil During Nano Fluid Injection. Pet. Sci. Technol. 2014, 33, 139–146. [Google Scholar] [CrossRef]
  120. Mohammadi, M.; Akbari, M.; Fakhroueian, Z.; Bahramian, A.; Azin, R.; Arya, S. Inhibition of Asphaltene Precipitation by TiO2, SiO2, and ZrO2 Nanofluids. Energy Fuels 2011, 25, 3150–3156. [Google Scholar] [CrossRef]
  121. Cao, C.; Song, Z.; Su, S.; Tang, Z.; Xie, Z.; Chang, X.; Shen, P. Water-based nanofluid-alternating-CO2 injection for enhancing heavy oil recovery: Considering oil-nanofluid emulsification. J. Pet. Sci. Eng. 2021, 205, 108934. [Google Scholar] [CrossRef]
  122. Hassanpour, S.; Malayeri, M.R.; Riazi, M.; Mousavi, S.M.; Negahban, Z. On the impact of Co3O4 nanoparticles on interaction of heavy oil and brine mixtures. J. Pet. Sci. Eng. 2018, 171, 680–686. [Google Scholar] [CrossRef]
  123. Mohammadi, M.; Dadvar, M.; Dabir, B. Application of response surface methodology for optimization of the stability of asphaltene particles in crude oil by TiO2/SiO2 nanofluids under static and dynamic conditions. J. Dispers. Sci. Technol. 2017, 39, 431–442. [Google Scholar] [CrossRef]
  124. Panahi, S.; Sardarian, A.R.; Esmaeilzadeh, F.; Mowla, D. Synthesize and characterization of chitosan and silica supported on Fe3O4 nanoparticles for the adsorption and removal of asphaltene molecules from crude oil. Mater. Res. Express 2018, 5, 095022. [Google Scholar] [CrossRef]
  125. Li, A.; Zhang, W.; Shu, F.; Li, L.; Li, M.; Lei, B. Properties of asphaltene in BA crude oil and mechanism of deposition inhibitor. Chem. Ind. Eng. Prog. 2019, 38, 110–115. [Google Scholar]
  126. Wang, X. Unlocking the Mechanism of Asphaltene Dispersion in Petroleum Fluids; China University of Petroleum: Beijing, China, 2014. [Google Scholar]
  127. Setoodeh, N.; Darvishi, P.; Lashanizadegan, A. A comparative study to evaluate the performance of coated Fe3O4nanoparticles for adsorption of asphaltene from crude oil in bench scale. J. Dispers. Sci. Technol. 2017, 39, 711–720. [Google Scholar] [CrossRef]
  128. Xu, L.; Li, Q.; Myers, M.; Cao, X. Investigation of the enhanced oil recovery mechanism of CO2 synergistically with nanofluid in tight glutenite. Energy 2023, 273, 127275. [Google Scholar] [CrossRef]
  129. Ju, B.; Fan, T.; Li, Z. Improving water injectivity and enhancing oil recovery by wettability control using nanopowders. J. Pet. Sci. Eng. 2012, 86, 206–216. [Google Scholar] [CrossRef]
  130. Dezfuli, M.G.; Jafari, A.; Gharibshahi, R. Optimum volume fraction of nanoparticles for enhancing oil recovery by nanosilica/supercritical CO2 flooding in porous medium. J. Pet. Sci. Eng. 2020, 185, 106599. [Google Scholar] [CrossRef]
  131. Farajzadeh, R.; Andrianov, A.; Krastev, R.; Hirasaki, G.J.; Rossen, W.R. Foam–oil interaction in porous media: Implications for foam assisted enhanced oil recovery. In Proceedings of the SPE EOR Conference at Oil and Gas West Asia, Muscat, Oman, 16–18 April 2012; Volume 183, pp. 1–13. [Google Scholar]
  132. Zhang, S. Study on identification method for gas channeling of CO2 flooding in low permeability reservoirs. Pet. Geol. Recovery Effic. 2020, 27, 101–106. [Google Scholar]
  133. Cui, C.; Zhou, Z.; He, Z. Enhance oil recovery in low permeability reservoirs: Optimization and evaluation of ultra-high molecular weight HPAM/phenolic weak gel system. J. Pet. Sci. Eng. 2020, 195, 107908. [Google Scholar] [CrossRef]
  134. Li, H.; Li, Q.; Wang, J.; Xu, F. Flow regularity and new knowledge of enhancing recovery mechanism of weak gel. Pet. Geol. Recovery Effic. 2006, 13, 85–87. [Google Scholar]
  135. Ma, Y.; Zhang, H.; Yuan, S. Hydration Structure of Partially Hydrolyzed Preformed Particle Gel. Chem. J. Chin. Univ. 2015, 36, 386–394. [Google Scholar]
  136. Peng, H. Research on Methods of Anti-channeling of CO2-Flooding in Low-permeability Reservoirs; China University of Petroleum: Beijing, China, 2018. [Google Scholar]
  137. Al-Ali, A.H.; Schechter, D.S.; Lane, R.H. Application of Polymer Gels as Conformance Control Agents for Carbon Dioxide EOR WAG Floods. In Proceedings of the SPE International Symposium on Oilfield Chemistry, The Woodlands, TX, USA, 8–10 April 2013. [Google Scholar]
  138. Sun, X.; Bai, B.; Alhuraishawy, A.K.; Zhu, D. Understanding the Plugging Performance of HPAM-Cr (III) Polymer Gel for CO2 Conformance Control. SPE J. 2021, 26, 3109–3118. [Google Scholar] [CrossRef]
  139. Zhou, B.; Kang, W.; Yang, H.; Zhu, T.; Zhang, H.; Li, X.; Sarsenbekuly, B.; Sarsenbek, T. Preparation and properties of an acid-resistant preformed particle gel for conformance control. J. Pet. Sci. Eng. 2021, 197, 107964. [Google Scholar] [CrossRef]
  140. Pu, W.; Du, D.; Fan, H.; Chen, B.; Yuan, C.; Varfolomeev, M.A. CO2-responsive preformed gel particles with interpenetrating networks for controlling CO2 breakthrough in tight reservoirs. Colloids Surf. A Physicochem. Eng. Asp. 2021, 613, 126065. [Google Scholar] [CrossRef]
  141. Zhang, X.; Zheng, W.; Zhang, T.; Ge, J.; Jiang, P.; Zhang, G. CO2 in water foam stabilized with CO2-dissolved surfactant at high pressure and high temperature. J. Pet. Sci. Eng. 2019, 178, 930–936. [Google Scholar] [CrossRef]
  142. Du, D.; Beni, A.N.; Farajzadeh, R.; Zitha, P.L.J. Effect of Water Solubility on Carbon Dioxide Foam Flow in Porous Media: An X-ray Computed Tomography Study. J. Ind. Eng. Chem. 2008, 47, 6298–6306. [Google Scholar] [CrossRef]
  143. Farajzadeh, R.; Andrianov, A.; Bruining, H.; Zitha, P.L.J. Comparative Study of CO2 and N2 Foams in Porous Media at Low and High Pressure-Temperatures. Ind. Eng. Chem. Res. 2009, 48, 4542–4552. [Google Scholar] [CrossRef]
  144. Wang, Y.; Zhang, Y.; Liu, Y.; Zhang, L.; Ren, S.; Lu, J.; Wang, X.; Fan, N. The stability study of CO2 foams at high pressure and high temperature. J. Pet. Sci. Eng. 2017, 154, 234–243. [Google Scholar] [CrossRef]
  145. Lv, M.; Wang, S. Stability of carbon dioxide foam and effect of polymer on its foam properties. CIESC J. 2014, 65, 2219–2224. [Google Scholar]
  146. Adkins, S.S.; Chen, X.; Chan, I.; Torino, E.; Nguyen, Q.P.; Sanders, A.W.; Johnston, K.P. Morphology and stability of CO2-in-water foams with nonionic hydrocarbon surfactants. Langmuir 2010, 26, 5335–5348. [Google Scholar] [CrossRef] [PubMed]
  147. Zhao, Z.; Zheng, J. Effect of Polymer on Foam Stability of CO2 Flooding Anti-Channeling System. Shandong Chem. Ind. 2022, 51, 9–11. [Google Scholar]
  148. Zhang, X.; Zhang, G.; Ge, J.; Jiang, J. Research on the stability of polymer enhanced CO2 foam. Fault-Block Oil Gas Field 2020, 27, 799–802. [Google Scholar]
  149. Davarpanah, A.; Shirmohammadi, R.; Mirshekari, B. Experimental evaluation of polymer-enhanced foam transportation on the foam stabilization in the porous media. Int. J. Environ. Sci. Technol. 2019, 16, 8107–8116. [Google Scholar] [CrossRef]
  150. Pu, W.; Wei, P.; Suna, L.; Wang, S. Stability, CO2 sensitivity, oil tolerance and displacement efficiency of polymer enhanced foam. RSC Adv. 2017, 7, 6251–6258. [Google Scholar] [CrossRef]
  151. Qi, G.; Zhou, L.; Long, M.; Wu, S.; Zhou, X. Experimental Study of the CO2 Miscible Modulator Oil-displacement Mechanism. Sci. Technol. Eng. 2016, 16, 180–183. [Google Scholar]
  152. Lv, W.; Gong, H.; Li, Y.; Li, Z.; Dong, M. The potential and mechanism of nonionic polyether surfactants dissolved in CO2 to improve the miscibility of CO2-hydrocarbon systems. Fuel 2022, 326, 125012. [Google Scholar] [CrossRef]
  153. Sun, L. Experimental Study on Reducing CO2 Miscibility Pressure by Modifiers; China University of Petroleum: Beijing, China, 2016. [Google Scholar]
  154. Liu, Y. Technology Research of Improving the Oil Displacement Efficiency for CO2 Flooding; Northeast Petroleum University: Daqing, China, 2017. [Google Scholar]
  155. Yang, S.; Lian, L.; Yang, Y.; Li, S.; Tang, J.; Ji, Z.; Zhang, Y. Molecular Optimization Design and Evaluation of Miscible Processing Aids Applied to CO2 Flooding. Xinjiang Pet. Geol. 2015, 36, 555–559. [Google Scholar]
  156. Wang, F.; Luo, H.; Ren, Y.; Fan, W.; Liang, M.; Zhang, C. Progress of Miscibility Pressure Reduction of Carbon Dioxide Flooding. J. Petrochem. Univ. 2015, 28, 93–97. [Google Scholar]
  157. Wu, C.; Shen, Z.; Li, Y.; Sha, O. Methods of Lowering the Minimum Miscibility Pressure of CO2 Flooding. Chem. World 2016, 57, 451–456. [Google Scholar]
  158. Lv, W.; Dong, M.; Sarma, H.; Li, Y.; Li, Z.; Sun, J.; Gong, H. Effects of CO2-philic nonionic polyether surfactants on miscibility behaviors of CO2-hydrocarbon systems: Experimental and simulation approach. Chem. Eng. J. 2023, 464, 142701. [Google Scholar] [CrossRef]
  159. Wang, X. Performance evaluation and oil displacement experiment study of CO2 foam system. Pet. Geol. Recovery Effic. 2020, 27, 69–74. [Google Scholar]
  160. Jones, S.A.; Laskaris, G.; Vincent-Bonnieu, S.; Farajzadeh, R.; Rossen, W.R. Effect of surfactant concentration on foam: From coreflood experiments to implicit-texture foam-model parameters. J. Ind. Eng. Chem. 2016, 37, 268–276. [Google Scholar] [CrossRef]
  161. Wang, J.; Nguyen, A.V.; Farrokhpay, S. A critical review of the growth, drainage and collapse of foams. Adv. Colloid Interface Sci. 2016, 228, 55–70. [Google Scholar] [CrossRef]
  162. Kanokkarn, P.; Shiina, T.; Santikunaporn, M.; Chavadej, S. Equilibrium and dynamic surface tension in relation to diffusivity and foaming properties: Effects of surfactant type and structure. Colloids Surf. A Physicochem. Eng. Asp. 2017, 524, 135–142. [Google Scholar] [CrossRef]
  163. Tamura, T.; Ohyama, M.; Ohyama, M. Dynamic Surface Tension and Foaming Properties of Aqueous Polyoxyethylene n-Dodecyl Ether Solutions. J. Colloid Interface Sci. 1995, 173, 493–499. [Google Scholar] [CrossRef]
  164. AlQuaimi, B.I.; Rossen, W.R. Study of foam generation and propagation in fully characterized physical-model fracture. J. Pet. Sci. Eng. 2019, 175, 1169–1181. [Google Scholar] [CrossRef]
  165. Fu, T.; Ma, Y. Bubble formation and breakup dynamics in microfluidic devices: A review. Chem. Eng. Sci. 2015, 135, 343–372. [Google Scholar] [CrossRef]
  166. Wei, P.; Pu, W.; Sun, L.; Pu, Y.; Li, D.; Chen, Y. Role of water-soluble polymer on foam-injection process for enhancing oil recovery. J. Ind. Eng. Chem. 2018, 65, 280–289. [Google Scholar] [CrossRef]
  167. Adkins, S.S.; Chen, X.; Nguyen, Q.P.; Sanders, A.W.; Johnston, K.P. Effect of branching on the interfacial properties of nonionic hydrocarbon surfactants at the air-water and carbon dioxide-water interfaces. J. Colloid Interface Sci. 2010, 346, 455–463. [Google Scholar] [CrossRef]
  168. Xing, D.; Wei, B.; McLendon, W.; Enick, R.; McNulty, S.; Trickett, K.; Mohamed, A.; Cummings, S.; Eastoe, J.; Rogers, S.; et al. CO2-Soluble, Nonionic, Water-Soluble Surfactants That Stabilize CO2-in-Brine Foams. SPE J. 2012, 17, 1172–1185. [Google Scholar] [CrossRef]
  169. Ren, G.; Sanders, A.W.; Nguyen, Q.P. New method for the determination of surfactant solubility and partitioning between CO2 and brine. J. Supercrit. Fluids 2014, 91, 77–83. [Google Scholar] [CrossRef]
  170. Niu, Q.; Dong, Z.; Lv, Q.; Zhang, F.; Shen, H.; Yang, Z.; Lin, M.; Zhang, J.; Xiao, K. Role of interfacial and bulk properties of long-chain viscoelastic surfactant in stabilization mechanism of CO2 foam for CCUS. J. CO2 Util. 2022, 66, 102297. [Google Scholar] [CrossRef]
  171. Wu, S.; He, J.; Li, Z.; Lin, W.; Wang, Q.; Wu, F.; Zhao, B.; Zhang, J. Studies on the chemical agent system for reducing in the minimal miscibility pressure of CO2 flooding. China Sci. 2015, 10, 2161–2164. [Google Scholar]
  172. Almobarak, M.; Wu, Z.; Zhou, D.; Fan, K.; Liu, Y.; Xie, Q. A review of chemical-assisted minimum miscibility pressure reduction in CO2 injection for enhanced oil recovery. Petroleum 2021, 7, 245–253. [Google Scholar] [CrossRef]
  173. Rommerskirchen, R.; Nijssen, P.; Bilgili, H.; Sottmann, T. Reducing the Miscibility Pressure in Gas Injection Oil Recovery Processes. In Proceedings of the Abu Dhabi International Petroleum Exhibition & Conference, Abu Dhabi, United Arab Emirates, 7–10 November 2016. [Google Scholar]
  174. Wang, F.; Luo, H.; Ren, Y.; Fan, W.; Liang, M.; Nan, Z. Influences of fatty alcohol polyoxypropylene ether on the minimum miscibility pressure of carbon dioxide flooding. Pet. Geol. Oilfield Dev. Daqing 2016, 35, 118–122. [Google Scholar]
  175. Guo, P.; Hu, Y.; Qin, J.; Li, S.; Jiao, S.; Chen, F.; He, J. Use of oil-soluble surfactant to reduce minimum miscibility pressure. Pet. Sci. Technol. 2017, 35, 345–350. [Google Scholar] [CrossRef]
  176. Luo, H.; Zhang, Y.; Fan, W.; Nan, G.; Li, Z. Effects of the Non-ionic Surfactant (CiPOj) on the Interfacial Tension Behavior between CO2 and Crude Oil. Energy Fuels 2018, 32, 6708–6712. [Google Scholar] [CrossRef]
  177. Liu, J.; Sun, L.; Li, Z.; Wu, X. Experimental Study on Reducing CO2-Oil Minimum Miscibility Pressure with Hydrocarbon Agents. Energies 2019, 12, 1975. [Google Scholar] [CrossRef]
  178. Choubineh, A.; Helalizadeh, A.; Wood, D.A. The impacts of gas impurities on the minimum miscibility pressure of injected CO2-rich gas–crude oil systems and enhanced oil recovery potential. Pet. Sci. 2018, 16, 117–126. [Google Scholar] [CrossRef]
  179. Zhang, C.; Xi, L.; Wu, P.; Li, Z. A novel system for reducing CO2-crude oil minimum miscibility pressure with CO2-soluble surfactants. Fuel 2020, 281, 118690. [Google Scholar] [CrossRef]
  180. Kuang, N.; Yang, S.; Yuan, Z.; Wang, M.; Zhang, Z.; Zhang, X.; Wang, M.; Zhang, Y.; Li, S.; Wu, J.; et al. Study on Oil and Gas Amphiphilic Surfactants Promoting the Miscibility of CO2 and Crude Oil. Engineering 2021, 6, 27170–27182. [Google Scholar] [CrossRef] [PubMed]
  181. Sun, N.; Fu, B. Study on reducing the minimum miscible pressure (MMP) of CO2 flooding by miscible solvent method. Liaoning Chem. Ind. 2017, 46, 3. [Google Scholar]
Figure 1. Phase change in CO2.
Figure 1. Phase change in CO2.
Molecules 28 06154 g001
Figure 2. Supramolecular assembly structure of asphaltenes: acid-base interactions and hydrogenbonding (blue), metal coordination complex (red), a hydrophobic pocket (orange), π-π stacking (face-to-face dark green; within a clathrate containing toluene, light green) [52].
Figure 2. Supramolecular assembly structure of asphaltenes: acid-base interactions and hydrogenbonding (blue), metal coordination complex (red), a hydrophobic pocket (orange), π-π stacking (face-to-face dark green; within a clathrate containing toluene, light green) [52].
Molecules 28 06154 g002
Figure 3. The principle of SC-CO2 miscible flooding technology.
Figure 3. The principle of SC-CO2 miscible flooding technology.
Molecules 28 06154 g003
Figure 4. The microscopic process of the volume swelling for CO2–decane system. Color code: gray, carbon; red, oxygen; white, hydrogen [76].
Figure 4. The microscopic process of the volume swelling for CO2–decane system. Color code: gray, carbon; red, oxygen; white, hydrogen [76].
Molecules 28 06154 g004
Figure 5. Diagram of volume expansion caused by CO2 dissolved in n-alkane (the mole fraction of CO2 is 25%, 50%, and 75%).
Figure 5. Diagram of volume expansion caused by CO2 dissolved in n-alkane (the mole fraction of CO2 is 25%, 50%, and 75%).
Molecules 28 06154 g005
Figure 6. Simulation snapshot of swelling of CO2–octane system at P = 7.69 MPa and T = 333 K: (a) t = 0, (b) 20, (c) 100, (d) 300, (e) 1000, and (f) 1500 ps. Color code:C8H18 (black) and CO2 (red) [80].
Figure 6. Simulation snapshot of swelling of CO2–octane system at P = 7.69 MPa and T = 333 K: (a) t = 0, (b) 20, (c) 100, (d) 300, (e) 1000, and (f) 1500 ps. Color code:C8H18 (black) and CO2 (red) [80].
Molecules 28 06154 g006
Figure 7. Effect of CO2 mole fraction on oil expansion coefficient and oil viscosity [91].
Figure 7. Effect of CO2 mole fraction on oil expansion coefficient and oil viscosity [91].
Molecules 28 06154 g007
Figure 8. (a) Carbon component under continuously ascending pressure. (b) Carbon component under individual pressure after CO2 breakthrough [107].
Figure 8. (a) Carbon component under continuously ascending pressure. (b) Carbon component under individual pressure after CO2 breakthrough [107].
Molecules 28 06154 g008
Figure 9. The percentage of extraction residue component [109].
Figure 9. The percentage of extraction residue component [109].
Molecules 28 06154 g009
Figure 10. Asphaltene deposition during CO2 flooding, asphaltene (blue), aromatic hydrocarbon (yellow), gum (red), CO2 (green).
Figure 10. Asphaltene deposition during CO2 flooding, asphaltene (blue), aromatic hydrocarbon (yellow), gum (red), CO2 (green).
Molecules 28 06154 g010
Figure 11. SEM image after carbonization at 800 °C: (a) oil–water emulsion and (b) oil–nanofluid emulsion [121].
Figure 11. SEM image after carbonization at 800 °C: (a) oil–water emulsion and (b) oil–nanofluid emulsion [121].
Molecules 28 06154 g011
Figure 12. Adsorption inhibition of asphaltene deposition by nanoparticles: (a) CO2 flooding and (b) CO2–nanofluid flooding.
Figure 12. Adsorption inhibition of asphaltene deposition by nanoparticles: (a) CO2 flooding and (b) CO2–nanofluid flooding.
Molecules 28 06154 g012
Figure 13. Schematic diagram illustrating the mechanism of nanoparticle dispersion inhibiting asphaltene deposition.
Figure 13. Schematic diagram illustrating the mechanism of nanoparticle dispersion inhibiting asphaltene deposition.
Molecules 28 06154 g013
Figure 14. Schematic diagram illustrating channeling blocking with high-strength HPAM gel.
Figure 14. Schematic diagram illustrating channeling blocking with high-strength HPAM gel.
Molecules 28 06154 g014
Figure 15. CHSB foam: (a) polymer-reinforced foam and (b) measurement of the size of foam over time [150].
Figure 15. CHSB foam: (a) polymer-reinforced foam and (b) measurement of the size of foam over time [150].
Molecules 28 06154 g015
Figure 16. Mechanism diagram of polymer nanospheres stabilizing CO2 foam [41].
Figure 16. Mechanism diagram of polymer nanospheres stabilizing CO2 foam [41].
Molecules 28 06154 g016
Figure 17. Schematic diagram of the CO2 foam stabilization mechanism and instability process for EAB [170].
Figure 17. Schematic diagram of the CO2 foam stabilization mechanism and instability process for EAB [170].
Molecules 28 06154 g017
Figure 18. Reducing the MMP of CO2 and crude oil using (a) oil-soluble surfactants and (b) CO2-soluble surfactants, especially in the heterogeneous formation using (c) oil-soluble surfactants and (d) CO2-soluble surfactants [179].
Figure 18. Reducing the MMP of CO2 and crude oil using (a) oil-soluble surfactants and (b) CO2-soluble surfactants, especially in the heterogeneous formation using (c) oil-soluble surfactants and (d) CO2-soluble surfactants [179].
Molecules 28 06154 g018
Table 1. Experimental data of surfactant reducing MMP of between CO2 and crude oil.
Table 1. Experimental data of surfactant reducing MMP of between CO2 and crude oil.
SurfactantGasMMP
Reduction/%
MethodReference
Acetyl glucose dodecyl esterCO227.47IFT method[155]
MultipleCO217.86IFT method[171]
Oil-solubleCO217.86VIT[172]
SurfactantsCO216.0–22.0Observation cell[173]
QDM, QHBCO215.0IFT method[151]
Lauryl alcohol polyoxypropylene etherCO225.6-[174]
Surfactant CAECO220.0Slim-tube[175]
PropoxylatedCO227.7VIT[176]
Oil-solubleCO28.28IFT method[177]
C1, C2, C3CO213.2Slim-tube[178]
NP-9CO25.72IFT method[179]
2EH-PO5-EO9CO28.03IFT method[179]
AmphiphilicCO220.0Slim-tube[180]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sun, Y.; Zhang, W.; Tian, J.; Meng, Y.; Zhang, L. Research Progress on Displacement Mechanism of Supercritical CO2 in Low-Permeability Heavy Oil Reservoir and Improvement Mechanism of Displacement Agents. Molecules 2023, 28, 6154. https://doi.org/10.3390/molecules28166154

AMA Style

Sun Y, Zhang W, Tian J, Meng Y, Zhang L. Research Progress on Displacement Mechanism of Supercritical CO2 in Low-Permeability Heavy Oil Reservoir and Improvement Mechanism of Displacement Agents. Molecules. 2023; 28(16):6154. https://doi.org/10.3390/molecules28166154

Chicago/Turabian Style

Sun, Yuanxiu, Weijie Zhang, Jinlong Tian, Yanzhao Meng, and Liping Zhang. 2023. "Research Progress on Displacement Mechanism of Supercritical CO2 in Low-Permeability Heavy Oil Reservoir and Improvement Mechanism of Displacement Agents" Molecules 28, no. 16: 6154. https://doi.org/10.3390/molecules28166154

Article Metrics

Back to TopTop