Next Article in Journal
Sensing Based on Plasmon-Induced Transparency in H-Shaped Graphene-Based Metamaterials
Previous Article in Journal
Effect of Calcination Temperature on the Microstructure, Composition and Properties of Agglomerated Nanometer CeO2-Y2O3-ZrO2 Powders for Plasma Spray–Physical Vapor Deposition (PS-PVD) and Coatings Thereof
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Systematic Review of Solubility, Thickening Properties and Mechanisms of Thickener for Supercritical Carbon Dioxide

by
Xiaohui Wang
1,2,
Qihong Zhang
1,
Shiwei Liang
1 and
Songqing Zhao
3,*
1
Beijing Key Laboratory of Optical Detection Technology for Oil and Gas, China University of Petroleum (Beijing), Beijing 102249, China
2
National Key Laboratory of Petroleum Resources and Engineering, China University of Petroleum (Beijing), Beijing 102249, China
3
State Key Laboratory of Heavy Oil Processing, China University of Petroleum (Beijing), Beijing 102249, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2024, 14(12), 996; https://doi.org/10.3390/nano14120996
Submission received: 12 March 2024 / Revised: 29 May 2024 / Accepted: 5 June 2024 / Published: 7 June 2024
(This article belongs to the Topic Carbon Capture Science & Technology (CCST))

Abstract

:
Supercritical carbon dioxide (CO2) has extremely important applications in the extraction of unconventional oil and gas, especially in fracturing and enhanced oil recovery (EOR) technologies. It can not only relieve water resource wastage and environmental pollution caused by traditional mining methods, but also effectively store CO2 and mitigate the greenhouse effect. However, the low viscosity nature of supercritical CO2 gives rise to challenges such as viscosity fingering, limited sand–carrying capacity, high filtration loss, low oil and gas recovery efficiency, and potential rock adsorption. To overcome these challenges, low–rock–adsorption thickeners are required to enhance the viscosity of supercritical CO2. Through research into the literature, this article reviews the solubility and thickening characteristics of four types of polymer thickeners, namely surfactants, hydrocarbons, fluorinated polymers, and silicone polymers in supercritical CO2. The thickening mechanisms of polymer thickeners were also analyzed, including intermolecular interactions, LA–LB interactions, hydrogen bonding, and functionalized polymers, and so on.

1. Introduction

In recent years, carbon dioxide (CO2) has become the focus of carbon emission reduction. As a major greenhouse gas, the utilization and storage of CO2 (as a buffer gas when producing hydrogen) are required to be environmentally friendly and economically sustainable [1,2,3,4,5,6]. For the chemical properties of CO2, the carbon atoms in the CO2 molecule are hybridized in sp mode, and the electrons form two mutually perpendicular π bonds. The bond length of the carbon–oxygen double bond (O=C=O) is shorter than that of the carbonyl bond (C=O). Thus, the CO2 structure is stable and the chemical properties are not very reactive [7], and the phase diagram of CO2 can be seen in Figure 3 of Ref. [8]. The critical points of CO2 are at 31 °C and 74 bar [8]. When the temperature and pressure exceed the critical point, CO2 undergoes a transition into the supercritical state, resulting in the formation of supercritical CO2.
Supercritical CO2 is a substance that exhibits properties intermediate between those of a gas and a liquid. It possesses advantageous characteristics such as high diffusivity, low viscosity, low surface tension, and controllable solubility. This unique nature of supercritical CO2 finds extensive applications in oil displacement technology and fracturing technology [9], effectively addressing the limitations associated with hydraulic fracturing [10,11]. These limitations include excessive water consumption, clay swelling, reservoir damage caused by residual working fluids, and inadequate flowback leading to groundwater pollution [12,13]. However, the low viscosity characteristics of pure supercritical CO2, with a viscosity of only 0.02–0.05 mPa·s, will cause a series of problems including viscous fingering, limited sand–loading capacity, and filter damage. The viscosity fingering problem stems from the viscosity difference between supercritical CO2 and crude oil, which causes supercritical CO2 to form finger–like flows in the reservoir, bypassing the oil layer and reducing recovery efficiency [14]. High filtration loss means that part of the fracturing fluid is adsorbed, retained, and permeated into the formation during the fracturing process. The problem of limited sand–carrying capacity is due to the low viscosity and high diffusivity of supercritical CO2, which limits its ability to carry sand particles and affects the effective support of fractures [10,15,16,17].
Therefore, understanding how to increase the viscosity of supercritical CO2 has become extremely important. The most direct and effective method is to add a thickener to supercritical CO2. The ideal supercritical CO2 thickeners should be effective in increasing viscosity at low doses. From a quantitative point of view, the thickening ratio is sufficient when it can thicken supercritical CO2 200–300 times, which is achieved in fluorinated thickeners [18], yet fluorinated thickeners are toxic, limiting their on–site application. In principle, the higher the viscosity of supercritical CO2, the better it would fit within the range to be achieved. In oilfield applications, the ideal efficiency of thickening depends on the actual application needs. For a broad range of applications, the increase of 5 to 10 times will serve the purpose. For oil extraction in the Middle East, a viscosity increase of 10 times is more than adequate. And in the context of supercritical CO2 fracturing technology, the viscosity enhancement should range from 20 to 30 times the original value. In addition to an increase in viscosity, an ideal thickener should also be cheap, environmentally friendly, safe, high efficiency, and soluble in supercritical CO2 but insoluble in water [19]. At present, supercritical CO2 thickeners are generally divided into the following four categories: surfactants, hydrocarbon polymers, fluorinated polymers, and silicone polymers [20]. In particular, fluorine–containing thickeners have the best solubility and thickening effects, but their use is restricted as they are expensive and not environmentally friendly [21,22]. Furthermore, they are not effective at low concentrations and also adsorb to rock [23]. This article mainly reviews the research progress of the four types of thickeners in terms of solubility, thickening properties, and mechanisms.

2. Characterization Parameters of Solubility and Thickening Properties

Through much literature research, it can be seen that the ideal supercritical CO2 thickener requires both efficient solubility and thickening properties in CO2 without any additional cosolvents. The characterization parameter of solubility and thickening properties are summarized below.

2.1. Solubility Properties

In terms of solubility, the main influencing factors are the interaction between the polymer and CO2 molecules, solvent CO2 density, solute relative molecular mass and molecular polarity, and especially the density factor. Solubility increases exponentially with the density of a supercritical CO2 system [24]. For thickening properties, it is mainly influenced by the spatial network structure formed by the interaction between thickener molecules. This structure effectively impedes the flow of CO2 molecules, resulting in the thickening of supercritical CO2.
The dissolving properties of thickeners in supercritical CO2 can be described by the solubility parameter δ [25], which is equal to the arithmetic square root of the cohesive energy density. The closer the polymer solubility parameter is to the CO2 solubility parameter, the better the solubility in CO2. The addition of cosolvents can reduce differences in the solubility parameters [25].
According to Enick’s research on thermodynamics, in order to be dissolved in supercritical CO2, Gibbs free energy (Gmix) must be reduced, that is ΔGmix < 0 [26], and its related expression is as follows:
G m i x = H m i x T S m i x
Here, ΔHmix, ΔSmix, and T are the mixing enthalpy change, mixing entropy change, and absolute temperature, respectively. Therefore, the above problem is transformed into the problem of how to improve ΔHmix and ΔSmix.
For Hmix, the main influencing factors are the density of the CO2 mixed solution, the interaction between CO2 molecules, between the polymer thickener molecules, and between the polymer thickener and CO2 molecules. However, the interaction between the polymer thickener and the CO2 molecules is critical in determining whether the thickener can be dissolved in liquid CO2. This interaction effectively promotes enthalpy reduction, which in turn reduces the free energy, allowing the thickener to dissolve in supercritical CO2 [27]. The strength of the interaction between molecules can be described by cohesive energy density [25]. For CO2 molecules, the electron distribution is near oxygen atoms and the CO2 molecule has a zero dipole moment; however, a large quadrupole moment and low polarizability [28] can cause weak interactions with non–polar covalently bonded fragments like C–C motifs, but reasonably strong interactions with non–hydrogen–bonded polar functional groups like esters, ethers, C–F groups, or aromatic structures. Therefore, in order to be dissolved in CO2, it should be weakly polar and have low cohesive energy density or contain a certain number of CO2–philic groups.
As for ΔSmix, the solubility of the polymer can be improved by increasing entropy of the system. For example, increasing the free volume of the polymer, improving the flexibility of the polymer chain, and lowering the glass transition temperature can all help reduce the interaction between polymer molecules and thus promote dissolution. Additionally, increasing the degree of branching of the polymer thickener can have a similar effect [29].
The interaction between the polymer and CO2 is one of the key factors determining the solubility of the polymer. If the interaction energy between the polymer and CO2 is strong, it usually means that the solubility of the polymer in supercritical CO2 is better because this strong interaction helps to overcome the attraction between the polymer molecules or between the CO2 molecules such that the thickener can be dispersed into the solvent. The interaction energy between the polymer and the CO2 molecules provides a measure of the strength of this interaction and can be calculated via Formula (2) [30].
E i n t e r = E p o l y m e r C O 2 E p o l y m e r + E C O 2
The total energy ECO2 of supercritical CO2, the total energy Epolymer, and the total energy Epolymer–CO2 of the mixed system can be calculated through molecular dynamics (MD). The larger the absolute value of Einter, the stronger the interaction. Hu et al. used MD methods to study poly(vinyl acetate–alt–maleate) copolymers. The results show that this type of polymer reduces the interaction energy between the polymer and CO2 after copolymerizing vinyl acetate, but it remains higher than the interaction energy between PVAc and CO2 [31].
In addition, the adsorption of polymers in supercritical CO2 can be reflected by the potential of mean force (PMF), which is calculated as shown in (3) [32]. The change in PMF can reflect the positional preference of polymers in CO2. If the PMF value is negative, it indicates that the polymer is more stable at a certain position in supercritical CO2, which contributes to polymer solubilization.
E ( r ) = k B T ln g ( r )
where E(r) is the mean force potential, kB is Boltzmann’s constant, T is the absolute temperature, g(r) is the radial distribution function between the polymer and CO2. The physical significance of g(r) can be expressed as the ratio of the local density of the B atom to the intrinsic density of the A atom at a distance r from the central atom A. This indicates that CO2 molecules need to overcome a certain barrier to approach the thickener. Moreover, the solubility of polymers in CO2 is also affected by polymer–polymer interactions to the same degree. The weaker the polymer–polymer interaction and the stronger the polymer–CO2 interaction, the more favorable the solubility.
Among the interaction between the polymer and CO2, Lewis acid–base (LA–LB) interaction can effectively promote dissolution of polymer [25]. In an LA–LB interaction, the Lewis acid acts as an electron pair acceptor and can accept electron pairs, while the Lewis base acts as an electron pair donor and can donate electron pairs. For example, Gong et al. [25] found that the LA–LB interaction between O atoms in PVAc and C atoms in CO2 can enhance the solubility of PVAc in supercritical CO2. This interaction helps more CO2 molecules distribute around the carbonyl groups in the PVAc molecular chain, thereby increasing the solubility of PVAc in supercritical CO2.

2.2. Thickening Properties

The viscosity of the mixed solution can be calculated by the following formula [33].
η = τ w γ w = D Δ p / 4 L 8 ν / D
where η(Pa·s) is the viscosity, τw and γw are the wall shear stress, and apparent shear rate, relatively. D is the capillary diameter, ∆p is the capillary pressure difference, L is the capillary length, and ν is the flow rate of CO2 in the thickened liquid. Research shows that the thickened supercritical CO2 system is a non–Newtonian fluid and the viscosity has a non–linear relationship with the shear rate [33]. Moreover, the viscosity can also be acquired by fitting the following transverse current autocorrelation function below:
C ( k , t ) ~ e η k 2 ρ t
which can be obtained through MD simulation [32].

3. Supercritical CO2 Thickeners

3.1. Surfactants

Surfactant thickeners are polymers composed of nonpolar hydrophilic groups and polar hydrophobic groups [34,35]. In supercritical CO2, the hydrophilic groups of the surfactant undergo physical interactions with a small amount of water, while the hydrophobic groups are exposed to and interact with CO2, forming a reverse micelle structure that further develops into a spatial network structure [36,37]. This network structure can restrict the mobility of CO2 molecules, thereby serving a thickening function.
The solubility of the polymer ZCJ–01 (copolymers of styrene with modified sulfonated fluorinated acrylates), surfactant thickener APRF–2 (consists of sodium succinate (–2–ethyl) sulfonate, ethanol, and H2O, etc.) and surfactant thickener SC–T–18 in supercritical CO2 was investigated by Zhai et al. [38]. The main component of SC–T–18 is a comb copolymer with polydimethylsiloxane as the main chain and amino groups as side chains. The research results show that SC–T–18 has the highest solubility among them, and the experimental values are in good agreement with the theoretical values [38]. SC–T–18 and supercritical CO2 form a single, stable, homogeneous emulsion micelle after sufficient mixing, which is due to the side–chain amino group in SC–T–18 effectively increasing its solubility in CO2 through an LA–LB interaction. At 25 °C and 6.894 MPa, Enick [39] conducted a study on the use of a 1 wt% surfactant tributyltin fluoride thickener and a 40–45 wt% pentane cosolvent to increase the viscosity of the supercritical CO2 system. The results showed that this combination can thicken the viscosity by 10–100 times. However, the flaw of this thickener is the necessity to use a considerable number of cosolvents to facilitate its dissolution in supercritical CO2, which makes the process very inefficient. Considering that the addition of cosolvents will bring more serious environmental problems, Shi et al. [40,41,42] introduced a fluoroalkyl group into the trialkyltin fluoride polymer molecule to obtain a semi–fluorinated trialkyltin fluoride thickener. Research shows that this thickener has high solubility in supercritical CO2 systems without the use of any cosolvents. Under 10–18 MPa, 4% mass fraction of semi–fluorinated trialkyltin fluoride can increase the system viscosity by up to 3.3 times. Enick et al. [43] also used perfluoropolyether glycol and fluorinated diisocyanate to react to synthesize a fluorinated polyurethane thickener. At 25 °C and 25 MPa, 4% fluorinated polyurethane can increase the viscosity of the system by 1.7 times.
Trickett et al. [44] designed a surfactant that can not only dissolve in CO2, but also form rod–shaped micelles with enhanced viscosity when a small amount of water is added. They changed Na+ in dialkyl sulfosuccinic acid into M2+ (Co2+ or Ni2+), as shown in Figure 1, turning the spherical micelles into rod–shaped micelles, and then forming reverse micelles in supercritical CO2. Then, through the interaction between these micelles, the viscosity of the CO2 mixed system increased. Research results show that Co(di–HCF4)2 and Ni(di–HCF4)2 with 6–10 wt% can increase the viscosity of the CO2 mixed system by 20–90 wt%. For some surfactant thickeners that are difficult to dissolve in supercritical CO2, CO2–philic groups can be introduced, such as fluorinated amine– and oxygen–containing surfactants [45]. Semi–fluorinated and fluorinated surfactant thickeners are found to be soluble in CO2 liquids and can also increase CO2 viscosity through the addition of a small amount of water [46]. By studying the solubility of oxygenated hydrocarbon surfactants in CO2, it was found that this thickener has a similar level to that of fluorinated surfactants, and both show high solubility properties, indicating that the O atoms in oxygenated hydrocarbon surfactants can increase solubility [47].
The principle of surfactant thickening of CO2 involves the formation of reverse micelles by surfactants. These reverse micelles overlap and entangle with each other, creating a spatial network structure. This structure restricts the flow of CO2 molecules, resulting in the thickening effect of the system. In addition, hydrocarbons, polar, or ionic groups can also be introduced into surfactants to increase the viscosity of supercritical CO2 systems [48]. The interaction between ion charges and water dipoles, as well as the Van der Waals force between alkyl chains, are also important factors in the formation of ionic surfactant micelles. The former is required to be stronger than the interaction between ions and CO2, and as for the latter, specific functional groups can be introduced to enhance the Van der Waals forces. In order to enhance the solubility of thickener molecules in supercritical CO2, CO2–philic groups, such as fluoroalkyl groups, carbonyl groups, and oxygen atoms, can be introduced into surfactant thickener molecules.

3.2. Hydrocarbon Polymers

Generally, this type of thickener contains carbon (C), hydrogen (H) elements, and may contain oxygen (O) elements. Hydrocarbon polymer thickeners usually have low solubility and weak thickening properties. Heller et al. [49] conducted a study on commercially available hydrocarbon polymer thickeners and found that the viscosity of the supercritical CO2 system did not significantly increase through the introduction of hydrocarbon polymer thickeners. They discovered that only a portion of the thickeners was soluble in supercritical CO2. The solubility of this portion was attributed to the polymers’ amorphous and irregular structure, which lacked a compact crystalline arrangement. This structural characteristic allowed for greater space, facilitating the penetration and solubilization of CO2 molecules. Not only that, the solubility of this type of polymer in supercritical CO2 was influenced by the interactions between the polymers, as well as between the polymers and CO2. The solubility is higher when the interaction between the polymer and CO2 is stronger.
Sarbu et al. [50] believed that polymers soluble in CO2 should have monomer units of LA–LB interaction with CO2 and monomer units with high free volume and high flexibility. Shen et al. [51] confirmed that among all the hydrocarbon thickeners, polyvinyl acetate (PVAc, Mw = 125,000) has the best solubility in supercritical CO2. The reason is that PVAc contains acetic acid groups, which can effectively increase solubility, but its ability to thicken CO2 is relatively weak. Because of this, PVAc has become an ideal supercritical CO2 thickener design material. Shen et al. [52] used azobisisobutyronitrile (AIBN) as a catalyst to synthesize a polyvinyl acetate telomer through free radical reaction, and then polymerized it with styrene to form a binary copolymer, namely styrene vinyl acetate binary copolymer, and the reaction mechanism is shown in Ref. [52] (1), (2) and (3). This copolymer molecule has CO2–philic groups and thickening groups, namely acetic acid groups and styrene groups, respectively, and is expected to become an economical and environmentally friendly thickener. Zhang Jian [53] used AIBN as an initiator to synthesize four–arm PVAc through the reversible addition–fragmentation–chain transfer (RAFT) polymerization method. In the case of adding cosolvent ethanol, at 35 °C and 15 MPa, adding four–arm PVAc with a concentration of 1 wt% and ethanol with a concentration of 5 wt% can increase the viscosity of the supercritical CO2 system by 31–55%. However, this still does not meet the standards for actual use.
In recent years, MD simulations have been used to study the solubility of supercritical CO2 and the thickening properties of thickeners. Xue et al. [32] used MD simulation methods to calculate the thickening mechanism of supercritical CO2 thickener and found that polyvinyl acetate–copolyvinyl ether (PVAEE) molecular chains formed a spatial network structure by intertwining with each other. Due to the interaction between CO2 molecules and polymers (including electrostatic interactions and Van der Waals interactions), the polymer restricted CO2 molecules to the network structure formed by PVAEE (degree of polymerization of each chain was N = 50) molecular chains. This restriction reduced the supercritical CO2 molecules flow, thereby increasing the viscosity of the supercritical CO2 fluid. They also obtained PMF by calculating the radial distribution function (RDF) in the results. The contact minimum (CM) and the second solvent separation minimum (SSM) are 0.4 nm and 0.85 nm, respectively. Their corresponding energy values determine the binding stability of CO2 and the polymer in the first and second solvent layers, respectively. A higher energy barrier exists between the CM and the SSM, which is the barrier of the solvent layer (BS). The equations for calculating the binding energy barrier and dissociation energy barrier between CO2 and polymer groups are ∆E+ = EBSESSM and ∆E = EBSECM, respectively. The corresponding binding and dissociation energy barriers for the ether (ester) groups are 700 kJ/mol and 300 kJ/mol (540 kJ/mol and 190 kJ/mol), respectively. These values indicate that the ester group is more readily bonded to CO2. Goicochea et al. [54] also used MD simulation to study the interaction between polymers and CO2 molecules. Research shows that both intermolecular interactions and branching can improve the viscosity of supercritical CO2. In particular, intermolecular π–π stacking plays a crucial role in the thickening effect of supercritical CO2. These studies show that MD simulation is a very effective method to study supercritical CO2 systems at the molecular level.
Double–chain polyether carbonate (TPA–PEC, Mw = 2168, N = 30), tri–chain polyether carbonate (TMA–PEC, Mw = 2211, N = 30), and four–chain polyether carbonate (TFA–PEC, Mw = 2254, N = 30) were synthesized as CO2 thickeners by Chen et al. [55]. The respective dissolution properties were studied using the MD simulation method at 24.85 °C and 20 MPa. The results show that both TPA–PEC and TMA–PEC have better solubility than TFA–PEC due to their stronger interaction with the CO2 molecules, but their thickening effect is poor. TFA–PEC has the highest viscosity and only needs an addition of 0.95 wt% to thicken the supercritical CO2 viscosity by 11 times, while TMA–PEC needs to be added at 0.72 wt% to thicken the CO2 viscosity by 3.9 times. Among the three, TPA–PEC has the worst CO2 thickening ability. From the perspective of solubility and thickening properties, the multi–chain structure is beneficial to the thickening ability but not to solubility. Polyether carbonate is also a polymer thickener that is easily degraded under natural conditions and has the advantage of being environmentally friendly.
Afra et al. [23] investigated the supercritical CO2 system of Poly–1–decene (P1D, Mw = 2950) in sandstone rock properties. It was shown that 1.5 wt% P1D increased the viscosity of the supercritical CO2 system by a factor of six at 24.13 MPa and 35 °C. When the temperature was increased to 90 °C, the viscosity increased by a factor of 4.8. It was also shown that the large number of methyl groups in the P1D molecule contributes to its solubility in CO2, while the branched structure of the molecule positively affects the thickening effect as well. In addition, poly–1–decene (an oligomer of about 20 repeating units) is not only very effective in CO2 viscosification, but also reduces the remaining water saturation from 40 to about 27% at 24.13 MPa and 90 degrees celsius, which can improve the storage efficiency of CO2.
Sun et al. [56] used AIBN as the initiator and synthesized a series of copolymers P(HFDA–co–MMA) and P(HFDA–co–EAL) using HFDA (1H,1H,2H,2H–perfluorodecyl acrylate), EAL (ethyl acrylate), and MMA (methyl methacrylate), as shown in Ref. [56] Scheme 1. The microstructure and intermolecular interactions in the supercritical CO2 system were studied through MD simulations. According to their research, an increase in the content of the EAL group enhances the interaction between copolymer chains and reduces their flexibility, leading to a decrease in solubility. Moreover, the intermolecular association of the copolymer is strengthened, resulting in an increased thickening ability. At 35.05 °C and 30 MPa, the P(HFDA0.19–co–EAL0.81, Mw = 3576) copolymer with the highest EAL content increases the viscosity of supercritical CO2 by 96 times at 5 wt% concentration and has the best thickening property among all copolymers. The solubility and thickening properties of P(HFDA0.37–co–EAL0.63) are higher than those of P(HFDA0.39–co–MMA0.61). P(HFDA0.37–co–EAL0.63, Mw = 3394) increases the viscosity of supercritical CO2 by 70 times, while P(HFDA0.39–co–MMA0.61) only increases it by 40 times. It shows that although EAL and MMA are isomers, the differences in their structures and compositions make huge differences in the intermolecular interactions of copolymer–CO2 and the association between copolymer chains. The presence of methyl groups in the main chain of P(HFDA–co–MMA) increases steric hindrance, which reduces intermolecular association, free volume, and chain flexibility.
Furthermore, at 344.3 K and 25–45 MPa, a coarse–grained molecular modeling study optimized by Kazuya [18,57] via the particle swarm optimization algorithm showed that branched hydrocarbon poly–1–decene oligomers (especially the model with six repeating units and Mw = 1000) showed a significant increase in solubility in supercritical CO2 compared to straight–chained alkanes with the same molecular weight, up to a factor of 270 times. This increase is attributed to an increase in the number of branches in the molecular structure, especially structural edges (methyl groups), which have enhanced interactions with CO2 and thus increase solubility. The branched structure of the thickener not only increases its solubility in CO2 but also reduces the adsorption of the thickener to the rock as compared to the change in chemical composition [57]. These findings provide important molecular design principles for the development of thickeners with high solubility in supercritical CO2. Ding et al. [58] conducted a study on the solubility and thickening properties of oligomers of 1–decene (O1D) with six repeat units and oligomers with branches of 1–dodecene and 1–hexadecene (O1D1H). The research confirmed that branches and methyl groups can promote solubility. At approximately 13.8 MPa and 35 °C, the solubility of O1D in supercritical CO2 is 0.6 wt%. While at 24.1 MPa and 35 °C, the solubility of O1D1H is 0.3 wt%. However, in terms of relative viscosity, the 0.3 wt% concentration of O1D1H provides better viscosity performance than the 0.6 wt% concentration of O1D.
In general, the biggest problem with hydrocarbon polymer thickeners is their low solubility and difficulty in completely dissolving in supercritical CO2. At present, the main way to quickly solve the problem of solubility is to add a large amount of cosolvent, but this also has the implication of environmental problems caused by the cosolvent, which is neither economically nor environmentally friendly. Therefore, it is still necessary to modify and design the hydrocarbon thickener molecules themselves to find thickeners with high solubility and high thickening properties.

3.3. Fluorinated Polymers

Compared with hydrocarbon polymer thickeners, fluorinated polymer thickeners obtained after fluorination have stronger CO2–philic properties and can be effectively dissolved in liquid CO2 without adding cosolvents. At the same time, the polymer has a better thickening effect.
DeSimone et al. [59] demonstrated for the first time that fluoropolymers can be strongly dissolved in supercritical CO2 without the assistance of cosolvents and show good thickening properties. Research shows that at 50 °C and 300 bar, 3.7 wt% poly(1,1–dihydroperfluorooctyl acrylate) (PFOA, Mw = 1,400,000) in supercritical CO2 can increase the viscosity of the system from 0.08 cP to 0.20–0.25 cP. However, PFOA is toxic to aquatic organisms, which may cause disruption to aquatic ecosystems.
Huang et al. [60] synthesized copolymers (PolyFAST) using a fluorinated acrylate and styrene copolymer. Among them, the fluorocarbon group is CO2–philic and can improve the solubility of PolyFAST in supercritical CO2, while the styrene group is CO2–phobic and can thicken CO2 but can also reduce the solubility of PolyFAST in supercritical CO2. The ratio of fluorinated acrylate to styrene is crucial in determining the thickening effect. Through multiple experiments, it has been found that the most significant increase in viscosity occurs when using a ratio of 71 mol% fluorinated acrylate and 29 mol% styrene [60]. Additionally, incorporating 1–5 wt% PolyFAST in supercritical CO2 can result in viscosity increasing up to 5–400 times. However, the production cost of this polymer is relatively high, and it is not environmentally friendly.
Heller et al. [49] studied telechelic polymer thickeners which have corresponding ionic groups at each end and form a network structure in the form of ion pair aggregation. Enick et al. [61] synthesized poly–sulfonated polyurethane. Fluorinated telechelic ionic polymers have good solubility in CO2 and the addition of 4 wt% the polymer can increase the viscosity of CO2 by 2.7 times at 25 °C and 25 MPa.
Shi et al. [42] synthesized a series of semi–fluorinated trialkyl tin fluorides. Among them, 4 wt% tris(1,1,2,2–tetrahydroperfluorohexyl)tin fluoride is soluble in CO2 and can increase the thickening of the supercritical CO2 system by 3.3 times. The mechanism is that the positively charged Sn atoms and the negatively charged F atoms form Sn–F bridges to create transient polymer chains.
Sun et al. [62] used an all–atom MD to simulate the molecular model of the polymer–CO2 system and studied the solubility and thickening properties of the copolymer in supercritical CO2. Research shows that 5 wt% of P(HFDA0.49–co–VPc0.51, Mw = 3023) can increase the viscosity of the CO2 system by 62 times. The thickening performance of 1.5 wt% of P (HFDA0.31–co–VAc0.69, Mw = 3539) is higher than that of P (HFDA0.49–co–VAc0.51) under the same conditions, but it is not easily dissolved at higher concentrations. The main reason is the high concentration of VAc which increases the number of methyl groups in the polymer chain, resulting in a decrease in chain flexibility. Therefore, the length and composition of polymer side chains can greatly affect the thickening performance.
Huang zhou [27] synthesized a CO2 thickener containing a fluoro–urea group. Research shows that DCT ([1,6–Bis(1,3–diperfluorooctanoic acid propyl–2–ureido)]heptane, double–chain thickener) begins to thermally degrade at around 200 °C, and SCT ((1,6–Difluorooctanoic acid ethyl ester urethyl)hexane, single–chain thickener) begins to thermally degrade at 170 °C. The optimal mass fraction obtained through single–factor experiments is 2 wt%, and DCT makes the viscosity of supercritical carbon dioxide is increased to 1.54 mPa·s, while the SCT is 1.46 mPa·s.
Kilic et al. [63] synthesized a series of aromatic acrylate–fluoroacrylate copolymer supercritical CO2 thickeners, and studied their structure and mechanism of thickening supercritical CO2. Research results show that the thickening property of this copolymer increases firstly and then decreases as the content of aromatic acrylate groups increases. The best solution is a 29% phenyl acrylate–71% fluoroacrylate copolymer. In the supercritical CO2 system, the copolymer only needs 5 wt% to increase the viscosity of the system by 205 times at 21.85 °C and 41.4 MPa. At the same time, it was also found that 26% phenyl acrylate (PHA)–74% fluoroacrylate (FA) has a better thickening effect than 27% CHA (cyclohexyl acrylate)–74% FA (fluoroacrylate). This proves that π–π stacking between aromatic rings plays a crucial role in thickening supercritical CO2.
In addition, Goicochea et al. [54], also used molecular simulation to study the thickening properties of polymer HFDA. Research shows that the thickening principle of fluorinated polymers mainly has two aspects. On the one hand, the fluorocarbon groups in the molecules can effectively enhance the CO2–philic properties of the polymers; on the other hand, the coupling mechanism between polymer molecules, which is the π–π association between styrenes, is stronger than the intramolecular interaction, making it difficult for polymer molecules to diffuse and aggregate, hindering the flow of CO2 molecules. This further enhances the thickening property of the supercritical CO2 system.
At present, according to the above research results, it can be seen that fluorine–containing polymer thickeners have impressive characteristics in terms of CO2–philic properties and CO2–thickening properties. However, the economic cost of such fluorinated polymers is too high, and they cannot be metabolized by organisms in the ecosystem. At the same time, they can also cause varying degrees of damage to organisms, such as weakening germ cell activity, interfering with enzyme activity, and damaging cell membrane structures, and so on [22]. Nevertheless, the research on this type of polymer provides theoretical guidance for the future design of economical, environmentally friendly, and pollution–free supercritical CO2 thickeners.

3.4. Silicone Polymer

Silicone polymers show reliable performance in thickening and are also pollution–free [64], and thus they can be an ideal potential supercritical CO2 thickener.
Bae et al. [65,66] used polydimethylsiloxane (PDMS) as a thickener to thicken supercritical CO2. Research shows that at 54 °C and 17.2 MPa, the viscosity of the 4% PDMS thickener + 20% toluene cosolvent + 76% liquid CO2 system increases to a maximum of 1.2 mPa·s. Compared with pure supercritical CO2, the viscosity increased by 30 times. But the disadvantage is that a large amount of cosolvent needs to be added. Zhao et al. [67] also used PDMS to thicken supercritical CO2, and the difference was that kerosene was used as a cosolvent because kerosene has a better solubilizing effect than toluene. Research results show that at 51.85 °C, the viscosity of the 5% PDMS thickener + 5% kerosene cosolvent + 90% liquid CO2 system increases to 4.67 mPa·s, which makes an increase of 54 times, while the amount of cosolvent is reduced.
Fink et al. [68] studied the feasibility of side–chain functionalization to improve the thickening properties of silicone polymers. The results show that silicone polymers with the appropriate amounts of side–chain functionalization act similarly to fluorinated polyether materials in supercritical CO2. Kilic et al. [68,69] enhanced the solubility of PDMS in supercritical CO2 through functionalization with propyldimethylamine. O’Brien et al. [70] synthesized a series of aromatic amidated polydimethylsiloxane (PDMS), as shown in Figure 5 of Ref. [70], and studied their solubility and thickening properties in supercritical CO2. Research results show that PDMS with anthraquinone–2–carboxamide (AQCA) end groups can thicken supercritical CO2 with hexane as a cosolvent, as shown in Figure 6 of Ref. [70]. The reason is that the content of CO2–philic groups and benzene ring groups in PDMS containing AQCA is low, and hexane is needed to thicken supercritical CO2.
Li et al. [71] synthesized a silicone terpolymer using 0.09 g tetramethylammonium hydroxide catalyst and a molar ratio of aminopropyltriethoxysilane and methyltriethoxysilane of 2:1. At 35 °C and 12 MPa, 3 wt% silicone terpolymer and 7 wt% toluene can thicken the viscosity of the supercritical CO2 system by 5.7 times. The mechanisms of silicone terpolymer and toluene are shown in Figure 2. CO2 interacts with amino groups. Specifically, N in the amino groups donates electrons to C in the CO2 and CO2 is located above N. Hydroxyl enhances the stability of the spatial network structure formed by siloxane and CO2 molecules. The reason why this type of polymer can thicken supercritical CO2 is that the hydroxyl group enhances the spatial network structure. Additionally, the chain structure generated by intermolecular interactions also plays a certain binding role, thereby increasing the flow resistance of CO2.
Wang et al. [72] synthesized epoxy–terminated polydimethylsiloxane, as shown in Figure 1 of Ref. [72], and studied its thickening performance in supercritical CO2. Research results show that when the shear rate increases, the polymer network structure will be destroyed by shear, and the viscosity of the system also decreases, that is, shear thinning. When the temperature rises, the activity and migration rate of various molecules in the system will be enhanced, which will weaken the intermolecular interaction, thereby destroying the network structure of the polymer and resulting in a decrease in the viscosity of the system. When the pressure increases in the range of 8–14 MPa, the degree of damage to the polymer’s spatial network structure will decrease and the viscosity will increase.
Shen et al. [6] used benzoyl peroxide as the initiator and synthesized a graft copolymer of methylsilsesquioxane and vinyl acetate through graft polymerization, as shown in Figure 3 of Ref. [6]. The thickener does not contain fluorine. Studies have shown that the grafting of PVAc enhances the solubility of siloxane polymers in supercritical CO2, and what plays a thickening role would be the network structure generated by polymethylsilsesquioxane. This research provides ideas for solving the solubility problem of polymers in supercritical CO2.

4. Thickening Mechanism

To obtain an ideal thickener, it should have a certain amount of CO2–philic groups (ether groups, carbonyl groups, acetate groups, acetyl groups, sugar ester groups, etc.) and CO2–phobic groups in the molecule. CO2–philic groups contribute to improve the solubility of the thickener, while CO2–phobic groups enhance the viscosity of supercritical CO2 through intermolecular association.
The introduced chain–like CO2–philic groups should have good flexibility, low cohesive energy, and high free volume. CO2–phobic groups can associate or their chains can cross and entangle with each other to form a spatial network structure to restrict the flow of CO2 molecules [32]. According to the results of Sagisaka et al. [73], at a certain concentration, the surfactant self–assembles to form linear or rod–like micelles that would intertwine with each other, forming a network structure and increasing the viscosity of CO2. It was also observed that rod–like reverse micelles, with different length–to–diameter ratios, exhibit varying thickening effects on supercritical CO2 at the same temperature and pressure. For instance, at 45 °C and 350 bar, anisotropic reverse micelles of about 5 to 7 wt% with rod lengths of approximately 166 Å and 583 Å increase the viscosity by 24% and 200%, respectively. Meanwhile, these two groups cannot be too many or too few. If there are too few CO2–phobic groups, the solubility of the thickener would be insufficient to achieve the desired thickening effect; while too many CO2–phobic groups would also affect the solubility of the thickener [27]. Kilic et al. [63] showed that the thickening properties of aromatic acrylate–fluoroacrylate copolymers exhibited an increase and then a decrease with the content of aromatic acrylate groups. Copolymers containing 29% phenyl acrylate and 71% fluoroacrylate were found to be the most desirable. The addition of only 5 wt% of the copolymer could increase the viscosities of supercritical CO2 by up to a factor of 205. Thus, research seeking an optimal ratio or dosage is still needed.
Generally, for surfactant thickeners, one end should be soluble in CO2 and the other end should be soluble in water or organic solvents to reduce the surface tension of water or organic solvents in CO2. At the same time, to form reverse micelles, two conditions should be satisfied, one is the multiple branched non–polar tail chain and low cohesive energy density, and the other one is a hydrogen–bonding interaction between the polar head group and water [74].
Hydrocarbon thickeners should meet two characteristics. On the one hand, they require large free volume, high chain flexibility, small steric resistance, weak interaction, low glass transition temperature and small steric hindrance, which help the polymer dissolve in CO2. On the other hand, polymer chains can cross and entangle with each other to form a spatial network structure, which hinders the flow of CO2 molecules and thicken CO2 [27].
Fluorine–containing polymer thickeners are obtained by fluorination of hydrocarbon polymers. They are usually weakly polar and have dipole–quadrupole interactions with CO2 molecules. At the same time, molecular chains can cross and entangle with each other to form a spatial network structure.
Silicone thickeners generally require cosolvents to improve solubility and thickening effects. The π–π stacking between phenyl groups produces intermolecular interactions, which has the thickening effect of supercritical CO2.
For thickening properties of thicker in supercritical CO2, in addition to the molecular structure, ratio of CO2–philic groups and CO2–phobic groups mentioned above, temperature, pressure, thickener molecular weight, and so on are also important influencing factors. The temperature and pressure conditions vary across different reservoir depths, leading to differences in the corresponding properties. The viscosity of supercritical CO2 exhibits changes in response to temperature variations. Much research has been conducted [18,75] on the effects of temperature and pressure on CO2 viscosity, and the results reveal a decreasing trend in CO2 viscosity with increasing temperature, while an there is an increasing trend with increasing pressure.

5. Thickening Supercritical CO2 in Porous Media

5.1. The Flow of CO2 in Porous Media

The displacement process of CO2 in heterogeneous porous media is one of the most important mechanisms [76]. Fluid physical parameters will cause phase flow instability during the CO2 displacement process. Research shows that when CO2 is injected into deep salt–water layers, it will displace the pores in a supercritical state. The dominant force in the displacement process is viscous force, and it will affect the form and distribution of fluid flow during the displacement process [77]. The simulation results show that under the condition of low–viscosity enhancement, the displacement process is obviously unstable, and the whole process has a relatively obvious fingering phenomenon; on the contrary, under the condition of higher viscosity enhancement, the displacement process is more stable, and no obvious fingering phenomenon occur [78].

5.2. Adsorption in Porous Media

During CO2 fracturing, CO2 thickeners may remain in the shale reservoir, and these chemicals may pollute the reservoir environment. Therefore, there is the need for the low adsorption of CO2 thickeners in porous media in practical industrial applications. If the adsorption within the porous medium is excessive, it may lead to the blockage of the pores [58]. Afra et al. [23] conducted experiments on a variety of thickeners currently available on the market, and the results showed that several thickeners containing the element fluorine showed significant adsorption to the surface of porous media [23]. They also used MD simulations to study the adsorption problem and came up with the agreement between the simulation and experimental results. They proposed an effective theoretical approach to study the adsorption of thickeners in porous media [23]. Li et al. [79] modified the thickener and prepared a new type of PDMS, and then investigated its contact angle. The results showed that the contact angle of PDMS decreased from 138° to 99° with increasing temperature, with a significant decreasing trend, while the contact angle of the prepared novel PDMS decreased from 135° to 127° [79]. Compared with the two, the novel PDMS has less adsorption on the reservoir surface, which is more favorable to reduce the contamination of the thickener on the reservoir.
In conclusion, the thickener should not exhibit excessive adsorption on the surface of porous media, as it would lead to a poorer process, economic problems, and excessively large reductions in permeability due to wettability alteration. If the thickener is brine–soluble (which is unlikely, given the low mutual solubility of CO2 and water), the thickener may separate out into the brine within the porous media. If the thickener is crude–oil–soluble, a portion of the thickener may ultimately contaminate the crude oil product and potentially cause problems in downstream processing equipment within refineries [23].

6. Summary and Outlook

This article provides a comprehensive review of four types of polymer thickeners, namely surfactants, hydrocarbons, fluorine–containing polymers, and silicones. We focused on analyzing their solubility and thickening characteristics in supercritical CO2 systems, and also explained the thickening mechanisms. Furthermore, we discussed the flow and adsorption of thickeners in porous media.
For surfactants, the thickening property is adequate, while solubility is far from satisfactory [39,43]. For example, a 1 wt% surfactant tributyltin fluoride thickener and a 40–45 wt% pentane cosolvent can thicken the viscosity by 10–100 times [39]. However, the solubility of the thickener is poor, requiring a significant amount of cosolvent or CO2–philic groups. In addition, silicones show similar solubility and thickening characteristics to surfactants, where 5% PDMS thickener with a small amount of cosolvent increases the viscosity to 4.67 mPa·s, which is an increase of up to 54 times [67]. The economic cost and environmental problems of cosolvents have become an urgent issue to be addressed.
Among the four thickeners, fluorinated thickeners have the most outstanding thickening properties, brought about by the interactions between the fluorine element and CO2. According to our research, the addition of 5 wt% PolyFast could increase the viscosity of supercritical CO2 by up to 400 times [67], which is the highest on record so far, according to our knowledge. Meanwhile, this copolymer also has fantastic solubility under reservoir conditions. However, this type of thickener is not commonly used, mainly because of its economic cost and biological toxicity. At present, most other polymer thickeners still require cosolvents to thicken liquid CO2; however, it is not environment friendly.
It was found that PVAc is one of the optimal CO2–philic hydrocarbon homopolymers because of the acetic acid group [73], yet its thickening properties are not ideal at present. However, PVAc is an ideal economical and environmentally friendly thickener, and an abundance of research has been conducted to improve the viscosity with it [32,52,53], such as forming a binary copolymer or spatial network structure, etc. This has made the PVAc–based system a mainstream thickener in work sites.
In recent years, the thickening mechanism and promotion of thickeners have been investigated through molecular modeling of polymer–CO2 systems. Regarding the thickening mechanism, it has been recognized that CO2–soluble polymers may have a moderately branched structure, high free volume, low solubility parameter, and contain Lewis acid–base groups. By introducing CO2–philic groups, the interaction between the thickener molecules and CO2 can be facilitated, thereby increasing the solubility of the thickener in supercritical CO2. The polymers should also contain CO2–phobic groups, which can combine with neighboring CO2–phobic groups to form a viscosity–enhancing network structure. Furthermore, the thickeners should exhibit low adsorption onto rock to minimize the blockage of rock pores, maintain the fluidity of the fracturing fluid, and reducing pollution and damage to the rock environment. Therefore, further research may focus on these aspects, addressing economic and technological barriers, as well as environmental concerns. The development of efficient, environmentally friendly, and cost–controllable thickeners can help promote engineering site applications.

Author Contributions

Writing—review and editing, X.W. and Q.Z.; visualization, S.L.; supervision, S.Z.; funding acquisition, X.W. and S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 11804028 and 12175023, and by the National Key Laboratory of Petroleum Resources and Engineering, China University of Petroleum (Beijing), grant number 2462024YJRC005.

Acknowledgments

The authors would like to acknowledge the National Key Laboratory of Petroleum Resources and Engineering, China University of Petroleum (Beijing), for supporting the research project.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Muhammad, A.; Nurudeen, Y.; Nilanjan, P.; Alireza, K.; Stefan, I.; Hussein, H. Influence of pressure, temperature and organic surface concentration on hydrogen wettability of caprock; implications for hydrogen geo–storage. Energy Rep. 2021, 7, 5988–5996. [Google Scholar] [CrossRef]
  2. Muhammad, A.; Nurudeen, Y.; Nilanjan, P.; Alireza, K.; Stefan, I.; Hussein, H. Influence of organic molecules on wetting characteristics of mica/H2/brine systems: Implications for hydrogen structural trapping capacities. J. Colloid Interface Sci. 2022, 608, 1739–1749. [Google Scholar] [CrossRef] [PubMed]
  3. Sherif, F.; Ahmed, E.-T.; Hesham, A.; Youssef, E.; Abdulmohsin, I. Increasing Oil Recovery from Unconventional Shale Reservoirs Using Cyclic Carbon Dioxide Injection. In Proceedings of the SPE Europec, EAGE Conference and Exhibition, Virtual, 1–3 December 2020. [Google Scholar] [CrossRef]
  4. Sean, S.; Patricia, C.; Barbara, K.; Sittichai, N. Characterizing Pore–Scale Geochemical Alterations in Eagle Ford and Barnett Shale from Exposure to Hydraulic Fracturing Fluid and CO2/H2O. Energy Fuels 2020, 35, 583–598. [Google Scholar] [CrossRef]
  5. Zhang, Q.; Zuo, L.; Wu, C.; Sun, C.; Zhu, X. Effects of crude oil characteristics on foaming and defoaming behavior at separator during CO2 flooding. Colloids Surf. A Physicochem. Eng. Asp. 2020, 608, 125562. [Google Scholar] [CrossRef]
  6. Aiguo, S.; Jinbo, L.; Yuehui, S.; Fuchang, S.; Zhengliang, W. Synthesis of the Copolymer of Vinylacete–Methylsilsesquioxane as a Potential Carbon Dioxide Thickener. Polym. Mater. Sci. Eng. 2011, 27, 157–159. (In Chinese) [Google Scholar] [CrossRef]
  7. Zhao, Z.; Li, X.; Zhang, B.; Gan, B.; Li, G. Experimental Study on Supercritical CO2 Fracturing. Nat. Gas Explor. Dev. 2016, 39, 58–63+14. (In Chinese) [Google Scholar]
  8. Etienne, G.; Thierry, T.; Jean-Daniel, M.; Mathias, D. Structure–Property Relationships in CO2–philic (Co)polymers: Phase Behavior, Self–Assembly, and Stabilization of Water/CO2 Emulsions. Chem. Rev. 2016, 116, 4125–4169. [Google Scholar] [CrossRef] [PubMed]
  9. XueSong, X.; Guo, Y.; Zhang, J.; Sun, N.; Shen, G.; Chang, X.; Yu, W.; Tang, Z.; Chen, W.; Wei, W.; et al. Fracturing with Carbon Dioxide: From Microscopic Mechanism to Reservoir Application. Joule 2019, 3, 1913–1926. [Google Scholar] [CrossRef]
  10. Liu, W. Research advance in supercritical CO2 thickeners. Fault-Block Oil Gas Field 2019, 19, 658–661. (In Chinese) [Google Scholar] [CrossRef]
  11. Whorton, L.P.; Brownscombe, E.R.; Dyes, A.B. Method for producing oil by means of carbon dioxide. US. Patent US2623596A, 30 December 1952. [Google Scholar]
  12. Birgit, C.G.; Ewers, U.; Fritz, H.F. Hydraulic fracturing: A toxicological threat for groundwater and drinking–water? Environ. Earth Sci. 2013, 70, 3875–3893. [Google Scholar] [CrossRef]
  13. Min, L.; Mehdi, S.; Peyman, M. Impact of mineralogical heterogeneity on reactive transport modelling. Comput. Geosci. 2017, 104, 12–19. [Google Scholar] [CrossRef]
  14. Zhang, Y. Study on the Influencing Factors of Viscous Fingering of CO2 in Oil Reservoirs; China University of Petroleum: Beijing, China, 2010. [Google Scholar]
  15. Shen, Z.; Wang, H.; Li, G. Numerical simulation of the cutting–carrying ability of supercritical carbon dioxide drilling at horizontal section. Pet. Explor. Dev. 2011, 38, 233–236. (In Chinese) [Google Scholar] [CrossRef]
  16. Li, Q.; Chen, M.; Jin, Y.; Wang, M.; Jiang, H. Application of New Fracturing Technologies in Shale Gas Development. Spec. Oil Gas Reserv. 2012, 19, 1–7+141. (In Chinese) [Google Scholar] [CrossRef]
  17. Wang, H.; Li, G.; Zhu, B.; Sepehrnoori, K.; Shi, L.; Zheng, Y.; Shi, X. Key problems and solutions in supercritical CO2 fracturing technology. Front. Energy 2019, 13, 667–672. [Google Scholar] [CrossRef]
  18. Pal, N.; Zhang, X.; Ali, M.; Mandal, A.; Hoteit, H. Carbon dioxide thickening: A review of technological aspects, advances and challenges for oilfield application. Fuel 2022, 315, 122947. [Google Scholar] [CrossRef]
  19. Xie, W.; Chen, S.; Wang, M.; Yu, Z.; Wang, H. Progress and Prospects of Supercritical CO2 Application in the Exploitation of Shale Gas Reservoirs. Energy Fuels 2021, 35, 18370–18384. [Google Scholar] [CrossRef]
  20. LI, Q.; WANG, Y.; LI, Q.; WANG, F.; YUAN, L.; BAI, H. Thickening Performance and Thickening Mechanism of a Viscosifier for CO2 Fracturing Fluid. Drill. Fluid Complet. Fluid 2019, 36, 102–108. (In Chinese) [Google Scholar] [CrossRef]
  21. Xue, W.; Wenge, C.; Qiuyuan, Y.; Hongyun, N.; Qian, L.; Yun, L.; Ming, G.; Ming, X.; An, X.; Sijin, L.; et al. Preliminary investigation on cytotoxicity of fluorinated polymer nanoparticles. J. Environ. Sci. China 2018, 69, 217–226. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, Y.; Dong, W.; Ye, L.; Ren, P. Progress in Research on Pollution Status and Hazards of Perfluorinated Organic Compounds (Pfcs) in Surface Water. Water Pollut. Control 2015, 33, 43–47. (In Chinese) [Google Scholar] [CrossRef]
  23. Salar, A.; Mohamed, H.A.; Abbas, F. Improvement in CO2 geo–sequestration in saline aquifers by viscosification: From molecular scale to core scale. Int. J. Greenh. Gas Control 2023, 125, 103888. [Google Scholar] [CrossRef]
  24. Wang, J.; Zhang, N.; Wu, Q.; Wang, S.; Zhang, P.; Zhao, J. Research progress on the solubility of supercritical CO2. Refin. Chem. Ind. 2011, 22, 1–5+83. (In Chinese) [Google Scholar] [CrossRef]
  25. Gong, H.; Zhang, H.; Xu, L.; Li, Y.; Dong, M. Effects of cosolvent on dissolution behaviors of PVAc in supercritical CO2: A molecular dynamics study. Chem. Eng. Sci. 2019, 206, 22–30. [Google Scholar] [CrossRef]
  26. Robert, M.E.; Eric, J.B.; Hamilton, A. Novel CO2–Thickeners for Improved Mobility Control; UNT Digital Library: Denton, TX, USA, 2001. [Google Scholar] [CrossRef]
  27. Huang, Z. Preparation of CO2 Thickening Agent and Evaluation of Its Fracturing Performance; SouthWest Petroleum University: Chengdu, China, 2017. [Google Scholar]
  28. Poovathinthodiyil, R.; Yutaka, I.; Scott, L.W. Polar Attributes of Supercritical Carbon Dioxide. Acc. Chem. Res. 2005, 38, 478–485. [Google Scholar] [CrossRef] [PubMed]
  29. Gérard, C.; Romain, D.; Delmas, G. Thermodynamic properties of polyolefin solutions at high temperature: 2. Lower critical solubility temperatures for polybutene–1, polypentene–1 and poly(4–methylpentene–1) in hydrocarbon solvents and determination of the polymer–solvent interaction parameter for PB1 and one ethylene–propylene copolymer. Polymer 1981, 22, 1190–1198. [Google Scholar] [CrossRef]
  30. Liu, B.; Wang, Y.; Liang, L.; Zeng, Y. Achieving solubility alteration with functionalized polydimethylsiloxane for improving the viscosity of supercritical CO2 fracturing fluids. RSC Adv. 2021, 11, 17197–17205. [Google Scholar] [CrossRef]
  31. Hu, D.; Sun, S.; Yuan, P.-Q.; Zhao, L.; Liu, T. Exploration of CO2–Philicity of Poly(vinyl acetate–co–alkyl vinyl ether) through Molecular Modeling and Dissolution Behavior Measurement. J. Phys. Chem. B 2015, 119, 12490–12501. [Google Scholar] [CrossRef] [PubMed]
  32. Xue, P.; Shi, J.; Cao, X.; Yuan, S. Molecular dynamics simulation of thickening mechanism of supercritical CO2 thickener. Chem. Phys. Lett. 2018, 706, 658–664. [Google Scholar] [CrossRef]
  33. Sui, S.; Xiao, P.; CUI, M. Effect of Modified Silicone CO2 Thickener on Fluid Rheology and Oil Displacement Efficiency. Oilfield Chem. 2023, 40, 229–234. (In Chinese) [Google Scholar] [CrossRef]
  34. Li, S.; Wang, J. Recent Developments of the Surfactants Using in Supercritical Carbon Dioxide. Chem. World 2007, 496–499+510. (In Chinese) [Google Scholar] [CrossRef]
  35. Liu, J.; Li, G.L.; Han, B. Research on the Aggregation and Microenvironment of Fluorine-Free and Silicon-Free Nonionic Surfactants in Supercritical Carbon Dioxide; Shandong University: Jinan, China, 2022. [Google Scholar]
  36. Du, M. Investigation on Supercritical Carbon Dioxide Fracturing Fluid System; China University of Petroleum (East China): Qingdao, China, 2016. [Google Scholar]
  37. Masanobu, S.; Yuuki, S.; Sajad, K.; Shirin, A.; Tretya, A.; Azmi, M.; Robert, M.E.; Sarah, E.R.; Christopher, H.; Julian, E. Thickening supercritical CO2 at high temperatures with rod-like reverse micelles. Colloids Surf. A Physicochem. Eng. Asp. 2024, 686, 133302. [Google Scholar] [CrossRef]
  38. Zhai, H.; Zhang, J.; Dong, J.; Wang, J.; Zhang, F.; Xiaofeng, L. Solubility Evaluation of Supercritical Carbon Dioxide Thickener. Oilfield Chem. 2021, 38, 422–426. (In Chinese) [Google Scholar] [CrossRef]
  39. Enick, M.R. A Literature Review of Attempts to Increase the Viscosity of Dense Carbon Dioxide. 1998. Available online: https://www.semanticscholar.org/paper/A-Literature-Review-of-Attempts-to-Increase-the-of-Enick/899dccfc1e36981b94973f48d5bf62064ae56e14 (accessed on 11 March 2024).
  40. Joel, F.; Naiping, H. The molecular basis of CO2 interaction with polymers containing fluorinated groups: Computational chemistry of model compounds and molecular simulation of poly bis(2,2,2–trifluoroethoxy)phosphazene. Polymer 2003, 44, 4363–4372. [Google Scholar] [CrossRef]
  41. Zhang, Y. Design, Synthesis and Properties of Polyethers as Carbon Dioxide Thickening Agent; Jilin University: Changchun City, China, 2017. [Google Scholar]
  42. Chao, S.; Zhihua, H.; Eric, J.B.; Robert, M.E.; Sun-Young, K.; Dennis, P.C. Semi-Fluorinated Trialkyltin Fluorides and Fluorinated Telechelic Ionomers as Viscosity–Enhancing Agents for Carbon Dioxide. Ind. Eng. Chem. Res. 2001, 40, 908–913. [Google Scholar] [CrossRef]
  43. Robert, M.E.; Eric, J.B.; Chen, S.; Eddy, K. Formation of fluoroether polyurethanes in CO2. In Proceedings of the 4th International Symposium on Supercritical Fluids (ISSF 97), Sendai, Japan, 11–14 May 1998. [Google Scholar]
  44. Kieran, T.; Dazun, X.; Robert, M.E.; Julian, E.; Martin, J.H.; Kevin, J.M.; Sarah, E.R.; Richard, K.H.; David, C.S. Rod-Like Micelles Thicken CO2. Langmuir 2009, 26, 83–88. [Google Scholar] [CrossRef]
  45. Hoefling, T.A.; Newman, D.A.; Robert, M.E.; Eric, J.B. Effect of structure on the cloud-point curves of silicone-based amphiphiles in supercritical carbon dioxide. J. Supercrit. Fluids 1993, 6, 165–171. [Google Scholar] [CrossRef]
  46. Stephen, C.; Robert, M.E.; Sarah, E.R.; Richard, K.H.; Julian, E. Amphiphiles for supercritical CO2. Biochimie 2012, 94, 94–100. [Google Scholar] [CrossRef]
  47. Xin, F.; Vijay, P.; Michael, C.M.; Yan, W.; Juncheng, L.; Robert, M.E.; Andrew, D.H.; Christopher, B.R.; Johnson, J.K.; Eric, J.B. Oxygenated Hydrocarbon Ionic Surfactants Exhibit CO2 Solubility. J. Am. Chem. Soc. 2005, 127, 11754–11762. [Google Scholar] [CrossRef] [PubMed]
  48. Cui, W.; Wang, C.; Xu, J.; Duan, Y. Application and research of liquid carbon dioxide based gel fracturing fluid in tight gas reservoir. In Proceedings of the 2016 Natural Gas Academic Annual Conference, Yinchuan, China, 28 September 2016. [Google Scholar]
  49. Heller, J.P.; Dandge, D.K.; Roger, J.C.; Donaruma, L.G. Direct Thickeners for Mobility Control of CO2 Floods. Soc. Pet. Eng. J. 1985, 25, 679–686. [Google Scholar] [CrossRef]
  50. Traian, S.; Thomas, S.; Eric, J.B. Non-fluorous polymers with very high solubility in supercritical CO2 down to low pressures. Nature 2000, 405, 165–168. [Google Scholar] [CrossRef] [PubMed]
  51. Shen, Z.; McHugh, M.; Xu, J.; Belardi, J.; Kilic, S.; Mesiano, A.; Bane, S.; Karnikas, C.; Beckman, E.; Enick, R. CO2–solubility of oligomers and polymers that contain the carbonyl group. Polymer 2003, 44, 1491–1498. [Google Scholar] [CrossRef]
  52. Shen, A.; Liu, J.; She, Y.; Shu, F.; Wang, Z. Design and Synthesis of Styrene Vinyl-acetate Copolymer as Potential CO2 Thickener. J. Oil Gas Technol. 2011, 33, 131–134+168. (In Chinese) [Google Scholar]
  53. Zhang, J. Carbon Dioxide Thickening Using Poly(Vinyl Acetate) and Amphiphilic Surfactant; East China University of Science and Technology: Shanghai, China, 2017. [Google Scholar]
  54. Goicochea, A.G.; Abbas, F. CO2 Viscosification by Functional Molecules from Mesoscale Simulations. J. Phys. Chem. C 2019, 123, 29461–29467. [Google Scholar] [CrossRef]
  55. Chen, R.; Zheng, J.; Ma, Z.; Zhang, X.; Fan, H.; Bittencourt, C. Evaluation of CO2–philicity and thickening capability of multichain poly(ether–carbonate) with assistance of molecular simulations. J. Appl. Polym. Sci. 2020, 138, 49700. [Google Scholar] [CrossRef]
  56. Sun, W.; Wang, H.; Zha, Y.; Yu, J.; Zhang, J.; Ge, Y.; Sun, B.; Zhang, Y.; Gao, C. Experimental and microscopic investigations of the performance of copolymer thickeners in supercritical CO2. Chem. Eng. Sci. 2020, 226, 115857. [Google Scholar] [CrossRef]
  57. Kazuya, K.; Abbas, F. Branching in molecular structure enhancement of solubility in CO2. PNAS Nexus 2023, 2, pgad393. [Google Scholar] [CrossRef] [PubMed]
  58. Ding, B.; Kantzas, A.; Firoozabadi, A. Spatiotemporal X-ray Imaging of Neat and Viscosified CO2 in Displacement of Brine-Saturated Porous Media. SPE J. 2024, 1–16. [Google Scholar] [CrossRef]
  59. Joseph, M.D.; Zihibin, G.; Elsbernd, C.S. Synthesis of Fluoropolymers in Supercritical Carbon Dioxide. Science 1992, 257, 945–947. [Google Scholar] [CrossRef] [PubMed]
  60. Huang, Z.; Shi, C.; Xu, J.; Kilic, S.; Enick, R.M.; Beckman, E.J. Enhancement of the Viscosity of Carbon Dioxide Using Styrene/Fluoroacrylate Copolymers. Macromolecules 2000, 33, 5437–5442. [Google Scholar] [CrossRef]
  61. Robert, M.E.; Eric, J.B.; Ali Vaziri, Y.; Val, J.K.; Hans, S.; Jon, L.H. Phase behavior of CO2–perfluoropolyether oil mixtures and CO2–perfluoropolyether chelating agent mixtures. J. Supercrit. Fluids 1998, 13, 121–126. [Google Scholar] [CrossRef]
  62. Baojiang, S.; Wenchao, S.; Haige, W.; Ying, L.; Haiming, F.; Hao, L.; Xiuping, C. Molecular simulation aided design of copolymer thickeners for supercritical CO2 as non-aqueous fracturing fluid. J. CO2 Util. 2018, 28, 107–116. [Google Scholar] [CrossRef]
  63. Sevgi, K.; Robert, M.E.; Eric, J.B. Fluoroacrylate–aromatic acrylate copolymers for viscosity enhancement of carbon dioxide. J. Supercrit. Fluids 2019, 146, 38–46. [Google Scholar] [CrossRef]
  64. Li, Q.; Wang, Y.; Li, Q.; Wang, F.; Li, Y.; Tang, L. Synthesis and performance evaluation of supercritical CO2 thickener for fracturing. Fault-Block Oil Gas Field 2018, 25, 541–544. (In Chinese) [Google Scholar]
  65. Jae-Heum, B.; Irani, C.A. A Laboratory Investigation of Viscosified CO2 Process. SPE Adv. Technol. Ser. 1993, 1, 166–171. [Google Scholar] [CrossRef]
  66. Jae-Heum, B. Viscosified CO2 Process: Chemical Transport and Other Issues. In Proceedings of the SPE International Symposium on Oilfield Chemistry, San Antonio, TX, USA, 14–17 February 1995. [Google Scholar] [CrossRef]
  67. Zhao, M.; Li, Y.; Gao, M.; Wang, T.; Dai, C.; Wang, X.; Guan, B.; Liu, P. Formulation and performance evaluation of polymer-thickened supercritical CO2 fracturing fluid. J. Pet. Sci. Eng. 2021, 201, 108474. [Google Scholar] [CrossRef]
  68. Fink, R.; Hancu, D.; Valentine, R.; Beckman, E.J. Toward the Development of “CO2–philic” Hydrocarbons. 1. Use of Side–Chain Functionalization to Lower the Miscibility Pressure of Polydimethylsiloxanes in CO2. J. Phys. Chem. B 1999, 103, 6441–6444. [Google Scholar] [CrossRef]
  69. Sevgi, K.; Yang, W.; Johnson, J.K.; Eric, J.B.; Robert, M.E. Influence of tert–amine groups on the solubility of polymers in CO2. Polymer 2009, 50, 2436–2444. [Google Scholar] [CrossRef]
  70. Michael, J.O.B.; Robert, J.P.; Mark, D.D.; Jason, J.L.; Aman, D.; Eric, J.B.; Robert, M.E. Anthraquinone Siloxanes as Thickening Agents for Supercritical CO2. Energy Fuels 2016, 30, 5990–5998. [Google Scholar] [CrossRef]
  71. Li, Q.; Wang, Y.; Li, Q.; Foster, G.; Lei, C. Study on the optimization of silicone copolymer synthesis and the evaluation of its thickening performance. RSC Adv. 2018, 8, 8770–8778. [Google Scholar] [CrossRef] [PubMed]
  72. Wang, Y.; Li, Q.; Dong, W.; Li, Q.; Wang, F.; Bai, H.; Zhang, R.; Owusu, A.B. Effect of different factors on the yield of epoxy-terminated polydimethylsiloxane and evaluation of CO2 thickening. RSC Adv. 2018, 8, 39787–39796. [Google Scholar] [CrossRef] [PubMed]
  73. Masanobu, S.; Shotaro, O.; Craig, J.; Atsushi, Y.; Azmi, M.; Frédéric, G.; Robert, M.E.; Sarah, E.R.; Adam, C.; Christopher, H.; et al. Anisotropic reversed micelles with fluorocarbon–hydrocarbon hybrid surfactants in supercritical CO2. Colloids Surf. B Biointerfaces 2018, 168, 201–210. [Google Scholar] [CrossRef] [PubMed]
  74. Hoefling, T.A.; David, S.; Reid, M.D.; Eric, J.B.; Robert, M.E. The incorporation of a fluorinated ether functionality into a polymer or surfactant to enhance CO2–solubility. J. Supercrit. Fluids 1992, 5, 237–241. [Google Scholar] [CrossRef]
  75. Liu, S.; Yuan, L.; Zhao, C.; Zhang, Y.; Song, Y. A review of research on the dispersion process and CO2 enhanced natural gas recovery in depleted gas reservoir. J. Pet. Sci. Eng. 2021, 208, 109682. [Google Scholar] [CrossRef]
  76. Miri, R. Effects of CO2–Brine–Rock Interactions on CO2 Injectivity–Implications for CCS; University of Oslo: Oslo, Norway, 2015. [Google Scholar]
  77. Lenormand, R.; Éric, T.; Zarcone, C. Numerical models and experiments on immiscible displacements in porous media. J. Fluid Mech. 1988, 189, 165–187. [Google Scholar] [CrossRef]
  78. Li, Y. The Numerical Simulation Study on CO2–Water Two Phase Flow in Porous Media Based on CFD; Dalian University of Technology: Dalian, China, 2015. [Google Scholar]
  79. Li, Q.; Wang, Y.; Owusu, A.B. A modified Ester-branched thickener for rheology and wettability during CO2 fracturing for improved fracturing property. Environ. Sci. Pollut. Res. 2019, 26, 20787–20797. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of Na+(di–HCF4) and M2+(di–HCF4). Reprinted (adapted) with permission from [44]. Copyright© 2010, American Chemical Society.
Figure 1. Structures of Na+(di–HCF4) and M2+(di–HCF4). Reprinted (adapted) with permission from [44]. Copyright© 2010, American Chemical Society.
Nanomaterials 14 00996 g001
Figure 2. Thickening mechanism of silicone terpolymer [71].
Figure 2. Thickening mechanism of silicone terpolymer [71].
Nanomaterials 14 00996 g002
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

Wang, X.; Zhang, Q.; Liang, S.; Zhao, S. Systematic Review of Solubility, Thickening Properties and Mechanisms of Thickener for Supercritical Carbon Dioxide. Nanomaterials 2024, 14, 996. https://doi.org/10.3390/nano14120996

AMA Style

Wang X, Zhang Q, Liang S, Zhao S. Systematic Review of Solubility, Thickening Properties and Mechanisms of Thickener for Supercritical Carbon Dioxide. Nanomaterials. 2024; 14(12):996. https://doi.org/10.3390/nano14120996

Chicago/Turabian Style

Wang, Xiaohui, Qihong Zhang, Shiwei Liang, and Songqing Zhao. 2024. "Systematic Review of Solubility, Thickening Properties and Mechanisms of Thickener for Supercritical Carbon Dioxide" Nanomaterials 14, no. 12: 996. https://doi.org/10.3390/nano14120996

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

Article Metrics

Back to TopTop