Next Article in Journal
Multifunctional Polymer-Modified P-CaO2@Au@OVA@Cu@DHPs Nanoparticles Enhance SARS-CoV-2 mRNA Vaccine-Induced Immunity via the cGAS–STING Signaling Pathway
Previous Article in Journal
Transparent Polyurethane Elastomers with Excellent Foamability and Self-Healing Property via Molecular Design and Dynamic Covalent Bond Regulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Linear Modified Siloxane-Based Thickeners and Study of Their Phase Behavior and Thickening Mechanism in Supercritical Carbon Dioxide

1
Research Institute of Natural Gas Technology, PetroChina Southwest Oil & Gas Field Company, Chengdu 610500, China
2
Shale Gas Evaluation and Exploitation Key Laboratory of Sichuan Province, Chengdu 610213, China
3
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Southwest Petroleum University, Chengdu 610500, China
*
Authors to whom correspondence should be addressed.
Polymers 2025, 17(19), 2640; https://doi.org/10.3390/polym17192640
Submission received: 28 August 2025 / Revised: 16 September 2025 / Accepted: 22 September 2025 / Published: 30 September 2025
(This article belongs to the Special Issue Application of Polymers in Enhanced Oil Recovery)

Abstract

To address critical limitations of ultra-low viscosity supercritical CO2 fracturing fluids, including excessive fluid loss and inadequate proppant transport capacity, a series of thickeners designed to significantly enhance CO2 viscosity were synthesized. Initially, FT-IR and 1H NMR characterization confirmed successful chemical reactions and incorporation of both solvation-enhancing and -thickening functional groups. Subsequently, dissolution and thickening performance were evaluated using a custom-designed high-pressure vessel featuring visual observation capability, in-line viscosity monitoring, and high-temperature operation. All thickener systems exhibited excellent solubility, with 5 wt% loading elevating CO2 viscosity to 3.68 mPa·s. Ultimately, molecular simulations performed in Materials Studio elucidated the mechanistic basis, electrostatic potential (ESP) mapping, cohesive energy density analysis, intermolecular interaction energy, and radial distribution function comparisons. These computational approaches revealed dissolution and thickening mechanisms of polymeric thickeners in CO2.

Graphical Abstract

1. Introduction

With the rapid development of the modern industrial society worldwide, the use of fossil fuels such as crude oil and natural gas has gradually increased, resulting in a large amount of CO2 emissions [1]. In response to the rapid growth of CO2 emissions, more attention has been paid to CO2 capture, utilization, and storage (CCUS) research [2,3]. The global CCUS technologies have been developing for more than 40 years and has been slow to develop due to high investment costs and policy influences [4]. In recent years, there have been developments that can re-create value after CO2 capture and storage, such as using CO2 to flood oil [5].
With the advancement of oil and gas field development technology, the development of unconventional oil and gas resources is receiving increasing attention [6]. Compared with the problems of reservoir permeability damage exposed by traditional hydraulic fracturing technology, supercritical CO2 fracturing technology, which has the advantages of avoiding clay swelling and water locking, enhancing fracturing expansion, easy backflow, and reducing water consumption, is regarded as an ideal fracturing technology with great application potential [7,8]. However, the widespread application of supercritical CO2 fracturing is severely limited due to issues such as gravity separation, viscous fingering, and poor sand-carrying capacity caused by its extremely low viscosity during the fracturing process [9].
To solve the above series of problems caused by supercritical CO2 fracturing, many scholars have conducted research on CO2 thickening; thickening technology refers to the process of dissolving thickening agent such as fluoropolymer, siloxane polymer, hydrocarbon polymer, and surfactant dissolved in CO2, and forming a transparent and stable homogeneous phase, so as to achieve the purpose of thickening CO2 [6,10,11]. Among them, the solubility and thickening ability of thickeners are of paramount importance in the study of thickeners [9]. Researchers typically seek structural balance between CO2-philic and CO2-phobic group to achieve a dynamic balance of solubility and thickening ability [10]. Because of its strong interaction with CO2 and good solubility, fluoropolymers have become the main CO2 solubilizing structure and CO2-philic substance [12,13]. Dai et al. [4] prepared a copolymer of heptafluorodecyl acrylate and styrene [P(HFDA/STY)] as a thickener for supercritical CO2. The unique binding mode between phenyl groups on the copolymer chains and CO2, as well as the molecular simulation radial distribution function (RDF), indicate that phenyl groups have higher aggregation and thickening abilities compared to fluorinated groups. Kilic et al. [14] studied the influence of the composition of copolymer of aromatic acrylate–fluoroacrylate on the viscosity of CO2. It was found that the copolymer composed of 29% aromatic acrylate and 71% fluoroacrylate was the most effective CO2 thickener. At a pressure of 15 MPa and temperature of 295 K, the thickener is miscible with CO2, and the viscosity increased to a certain extent. While achieving a balance between the thickening and solubilizing structure of polymer thickeners, less attention has been paid to the mechanisms of thickening and solubilizing polymer and the interaction between various functional groups and CO2.
On the basis of modified siloxane as a thickening group, this study introduces fluorinated groups as solubilizing groups to synthesize fluorinated modified siloxane thickeners. The solubility and thickening ability of fluorinated groups, modified siloxanes, and their copolymers were tested and characterized by infrared and nuclear magnetic resonance, and the microscopic behavior of fluorinated groups, modified siloxanes, and their copolymers in supercritical CO2 was studied using full atomic molecular dynamics simulation. The mechanism of action of different groups was explored, providing ideas for the molecular design of CO2 thickeners in the future.

2. Chemicals and Methods

2.1. Chemicals

Concentrated sulfuric acid, sodium carbonate, trimethylolpropane trimethacrylate, isocyanuric acid triallyl ester, tetramethyldisiloxane (HMM), octamethylcyclotetrasiloxane (D4), azodiisobutyronitrile, and chloroplatinic acid were purchased from Chengdu Kelong Co., Ltd., Chengdu, China. All the chemicals were used as received without further purification. CO2 and N2 were sourced from Chengdu Xindu Jinnengda Gas Co., Ltd., Chengdu, China.

2.2. Synthesis of Linear Modified Silicone-Based Supercritical CO2 Thickener

2.2.1. Preparation of Silicon Hydrogen-Terminated Linear Polysiloxane

Under the conditions of 90 °C and concentrated sulfuric acid as a catalyst, 1 g of HMM and 50 g of D4 were polymerized in a N2 atmosphere for 12 h to obtain silicon hydrogen-terminated linear siloxane. Subsequently, sodium carbonate was added to filter out the sulfuric acid catalyst, followed by vacuum distillation to remove impurities, resulting in a silicon hydrogen-terminated linear polysiloxane, labeled HMM-D4. The synthesis route is shown in Figure 1.

2.2.2. Synthesis of Supercritical CO2 Copolymer Thickeners

A quantity of trimethylolpropane triacrylate or triallyl isocyanurate was added to a three-necked flask, stirred and heated to 70 °C, and then chloroplatinic acid was added for activation for 2 h. Subsequently, a certain amount of HMM-D4 was added to the flask (the molar ratio of siloxane to acrylic ester was 1:2) under stirring conditions. After 4 h, chloroplatinic acid was separated by precipitation with anhydrous methanol, and the low boiling impurities were removed by vacuum distillation on a rotary evaporator to obtain a yellow liquid, which was the final copolymer thickener, named SHTT or SHTA. The synthesis route is shown in Figure 2.

2.2.3. Modification of SHTT and SHTA

An amount of 0.001 g of azobisisobutyronitrile was added to 25 g of SHTT or SHTA. After activation for 2 h, 3.885 g of fluorinated acrylate was added under stirring conditions. The modified SHTT and SHTA product was obtained by condensation reflux reaction at a temperature of 80 °C for 6 h and named SHTTF and SHTAF. The modification route is shown in Figure 3.

2.3. Characterization

The chemical structure of copolymers and modified copolymers was characterized by FT-IR and 1H NMR. The FT-IR spectrum of the copolymers was recorded using a Nicolet Nexus 170SX FT-IR (Madison, WI, USA) on KBr tablets. The 1H NMR spectrum of copolymers and modified copolymers were investigated using a Bruker 400 MHZ NMR spectrometer (Japan Electronics Corporation, Tokyo, Japan) using deuterated chloroform as the solvent.

2.4. Solubility and Thickening Properties of Copolymers and Modified Copolymers

The solubility and thickening properties of copolymers and modified copolymers were investigated using a visualization supercritical CO2 fracturing fluid evaluation device. The schematic diagram of the experimental device is shown in Figure 4. The experimental steps were as follows.
(1)
Accurately weigh the required thickener and add it to the intermediate container.
(2)
Seal the visual reaction vessel and heat it to the measurement temperature.
(3)
Inject CO2 to 0.5 MPa and then reduce the pressure. Repeat this cycle 5 times to expel the air inside the visual reaction vessel.
(4)
Use an ISCO pump to inject the weighed thickener in the intermediate container into the visual reaction vessel.
(5)
Use a CO2 gas booster pump to pressurize CO2 to the pressure required for the experiment. The physical property parameters of CO2 were referenced from the software REFPROP (version 9.1).
(6)
Start mechanical stirring at the bottom of the visual reaction vessel (stirring speed range: 0–1350 rpm). Stir at a constant speed for at least 40 min to promote the dissolution of the thickener and form a transparent and homogeneous solution. (The dissolution of the thickener in supercritical CO2 was observed through the front viewing window using an external LED light source placed behind the rear viewing window. High solubility of the thickener was indicated when the mixed fluid inside the visual reaction vessel became clear (the reference object was clearly visible).)
(7)
A circulating water vacuum pump was used to evacuate the Cambridge online viscometer.
(8)
Open the intake valve of the online viscometer to allow the completely dissolved thickener solution from step (6) to enter. Meanwhile, manually and slowly adjust the volume of the visual reaction vessel to maintain a stable system pressure. Use a data acquisition system to obtain viscosity data.
(9)
After the test is completed, direct the system into the waste liquid collection device. Simultaneously, clean the equipment and wait for the next experiment.

2.5. Molecular Simulation Study

Molecular simulation techniques were employed to investigate the thickening mechanisms of supercritical CO2 thickener, utilizing the Material Studio (version 2023) software package. The Compass force field, which is widely applied in recent years and known for its high accuracy, was selected to calculate the interparticle interactions [5,15,16,17]. Within the Amorphous Cell module, simulation systems containing CO2 molecules and modified silicone-based thickener molecules were constructed. Three-dimensional periodic boundary conditions were applied to simulate the supercritical state. The system underwent sequential structural optimization and annealing to obtain the minimum energy conformation. Subsequently, equilibrium molecular dynamics (MD) calculations were performed under the NPT ensemble for both CO2 and the thickener polymer, resulting in stable structures for subsequent studies [5,18]. Table 1 lists the detailed information and optimized structures of the different monomers and modified silicone-based thickeners used in the simulations. The simulation process is listed below:
(1)
A model was constructed based on the molecular structures presented in Table 1. Force field charges were assigned, and the configuration with the lowest energy was selected for geometric optimization using the Forcite module.
(2)
The model obtained from the initial geometric optimization was subjected to annealing using the Forcite module. The system went through 10 annealing cycles, starting from 313.15 K, ramping up to 500 K, and then returning to 313.15 K to reach system equilibrium.
(3)
Under the NPT ensemble, the desired temperature and pressure conditions (305.15 K, 10 MPa) were set. The Andersen thermostat and Berendsen barostat were employed, respectively. The time step was set to 1 fs, and the total simulation duration was 500 ps. In the aforementioned molecular dynamics process, one frame was output every 5 ps, and the data from the last 100 ps were used for subsequent data analysis.

2.5.1. Cohesive Energy Density (CED)

It is defined as the energy required to overcome the intermolecular force in 1 mole of a condensed matter substance quantitatively characterized in the interactions between polymer molecules [19]. And the solubility parameter (δ) is defined as the square root of CED [20].
E = E i n t e r = E v a n + E e l e c t + E o t h e r
C E D = E V
δ = C E D = E V
where the CED is the cohesive energy density, J/m3; E is the energy of the substance, kJ/mol; V is the molar volume of the substance, L/mol; δ is the solubility parameter, (J/m3)1/2; Einter is the total intermolecular energy of the system (kJ/mol); Evan is the energy from van der Waals interactions (kJ/mol); and Eelect is the energy from electrostatic interactions (kJ/mol).

2.5.2. Interaction Energy

The interaction energy serves to quantitatively characterize the strength of interactions between the polymer and CO2. A larger absolute magnitude of the interaction energy indicates stronger interactions between the polymer and CO2 [1].
E i n t = E p o l y m e r + C O 2 ( E p o l y m e r + E C O 2 )
where Eint is the polymer–CO2 interaction energy, (kJ/mol); Epolymer+CO2 is the total energy of the polymer–CO2 system (kJ/mol); Epolymer is the energy of the polymer (kJ/mol); and ECO2 is the energy of CO2, kJ/mol.

2.5.3. Radial Distribution Function (RDF)

The RDF is commonly used to investigate the microscopic structure of systems and evaluate intermolecular interactions. It typically represents the probability density of finding a molecule (or atom) of A at a specific distance from a molecule (or atom) of B [21] calculated as
g A B ( r ) = 1 ρ A B · 4 π r · r j = 1 N A B N A B ( r r + r ) N A B
where g A B ( r ) is the radial distribution function value; N A B is the number of species A and B (molecules or atoms) in the system; r is the distance interval width; N A B is the number of B particles (or A particles) found within the distance range r~r + Δr from a central A particle (or B particle); and ρ A B is the density of system.

3. Results and Discussion

3.1. Characterization of Supercritical CO2 Thickener

3.1.1. FT-IR

Figure 5 shows the FT-IR spectra of HMM-D4, SHTT, and SHTA. The FT-IR spectra of HMM-D4 primarily exhibit characteristic peaks including the Si-CH3 stretching vibration at 1257 cm−1, Si-O stretching vibration at 1064 cm−1 and 1052 cm−1, and the Si-C stretching vibration at 797 cm−1. As the hydrosilylation reaction progresses, the characteristic peaks of HMM-D4 are still retained in the infrared spectra of the obtained thickeners SHTT and SHTA. However, the intensities of the peaks corresponding to the Si-CH3 stretching vibration at 1257 cm−1 and the Si-O stretching vibration at 1052 cm−1 are significantly reduced [22]. Additionally, new peaks emerge at 1680 cm−1 (C=C stretching vibration of CH2=CH–), 1719 cm−1 (C=O stretching vibration), and 1447 cm−1 (bending vibration of saturated C–H bonds in –CH2– and –CH3 groups) [23,24]. Based on the above analysis, the appearance of new characteristic peaks in the infrared spectra of the thickeners SHTT and SHTA corresponds to the changes in functional groups during the hydrosilylation reaction [17]. This conclusion is consistent with the previous research results of Tang et al. [22] and Liu et al. [17,19].

3.1.2. 1H NMR

Figure 6 shows the 1H NMR spectra of the thickener copolymers. The peak at 7.26 ppm corresponds to the residual protons in the solvent CDCl3. For SHTT (Figure 6a), key signals include alkenyl protons (CH2=C(CH3)C=O–) at 5.5–6.2 ppm, methylene protons (–(CH2)3CCH2CH3) at 3.9–4.2 ppm, methyl protons (CH2=C(CH3)C=O–) at 1.94 ppm, methylene protons of the ethyl group (–(CH2)3CCH2CH3) at 1.2 ppm, methylene and methyl protons (O=C–CH(CH3)CH2Si–) at 0.89–0.98 ppm, and methyl protons bonded to silicon (Si–CH3) at 0.1–0.2 ppm. For SHTA (Figure 6b), key signals include methine proton (–NCH2CH=CH2) at 5.8–5.93 ppm, alkenyl protons (–NCH2CH=CH2) at 5.2–5.34 ppm, methylene protons of the allyl group (–CH2–) at 4.4–4.6 ppm, methylene protons adjacent to silicon (Si–CH2CH2–) at 1.2–1.6 ppm, and methyl protons bonded to silicon (Si–CH3) at 0–0.2 ppm [17,19].
Compared with SHTT (Figure 6a), in SHTTF (Figure 6c), peaks corresponding to the hydrogen atoms of the methylene group adjacent to the oxygen atom in –O–CH2–CH2–C8F17 appear at δ = 3.9–4.2 ppm, and peaks corresponding to the hydrogen atoms of the methylene group adjacent to the fluorine atoms in –O–CH2–CH2–C8F17 appear at δ = 1.7–1.94 ppm. Similarly, in SHTAF (Figure 6d), the peak corresponding to the hydrogen atoms of the methylene group adjacent to the oxygen atom in –O–CH2–CH2–C8F17 appears at δ = 4.2 ppm, and the peak corresponding to the hydrogen atoms of the methylene group adjacent to the fluorine atoms in –O–CH2–CH2–C8F17 appears at δ = 1.7–1.96 ppm. Similar to the findings of Dai et al. [4] and Sun et al. [13], the changes in the values of these characteristic peaks correspond to the changes in the corresponding functional groups after the introduction of fluorides. Combining the previous infrared analysis and the above NMR results, it indicates the successful preparation of the thickener copolymers.

3.2. Phase Behavior

The phase behavior of SHTT and SHTTF in ScCO2 was investigated using visualized reaction vessel as shown in Figure 4, with results presented in Figure 7 and Table 2. The observation revealed that the system remained clear and transparent under stirred conditions at 25 °C and 7.5 MPa, allowing distinct visibility of the reference wire behind the observation window (Figure 7(a1,b1)). When stirring was initiated at 32° C and 7.5 MPa, the system turned turbid as temperature increase intensified molecular thermal motion, promoting gradual dissolution of the thickener and completely obscuring the reference wire (Figure 7(a2,b2)). Upon maintaining the temperature while slowly increasing pressure to 10 MPa, the thickener became fully dissolved in ScCO2, accompanied by brightening of the solubility phase diagram and restored visibility of the reference wire (Figure 7(a3,b3)). When pressure was reduced back to 7.5 MPa, the system appeared more turbid than in Figure 7(a3,b3) at 32 °C and 10 MPa but retained better clarity than in Figure 7(a2,b2), suggesting that although pressure decreased, the polymeric thickener remained dissolved in supercritical CO2 without complete precipitation [15]. During this process, the dissolution pressures of SHTT-CO2 and SHTTF-CO2 were measured to be 9.2 MPa and 8.6 MPa, respectively. The dissolution pressure is defined as the pressure at which the thin-line reference changes from visible to invisible during the pressure reduction process starting from 10 MPa (the measurement is repeated three times, and the average value is taken) [17,19].

3.3. Thickening Ability

The thickening ability of SHTT, SHTA, SHTTF, and SHTAF were investigated at 32 °C and 10 MPa, and the results are shown in Figure 8 and Table 3. All ScCO2 fracturing fluid systems with different thickeners exhibited the same trend. Based on the literature review, a good thickener should achieve a thickening ratio of 10~100 times (compared to pure CO2) at a low concentration (0~5 wt%) and achieve excellent solubility while maintaining thickening performance, cost-effectiveness, and safety [9,10]. Based on this, in this study, the authors will investigate the concentration of the additive (thickener) in the range of 0~5 wt%. The viscosity of ScCO2 increased with higher thickener content. However, fluorinated modified copolymers (SHTTF, SHTAF) demonstrated a significantly stronger impact on ScCO2 viscosity than SHTT and SHTA, highlighting its superior thickening capacity. For instance, adding 5 wt% SHTTF increased the viscosity of CO2 to 3.68 mPa·s. At the initial stage of adding SHTT or SHTA, the system contained only a small number of thickener molecules, resulting in limited van der Waals interactions and hydrogen bonding between CO2 molecules and the thickener. This formed a sparse spatial network structure. As the thickener content increased—particularly with fluorinated thickeners like SHTTF and SHTAF—more modified siloxane polymers dissolved in CO2. This enhanced the interactions between CO2 molecules and thickener molecules, leading to the formation of a denser microscopic spatial network structure [16,19]. Consequently, the apparent viscosity of CO2 increased macroscopically.
Under experimental conditions of 32 °C and 5 wt% thickener dosage, the effect of pressures on the viscosity of CO2 thickened by various agents was studied, with results shown in Figure 9 and Table 4. The viscosity of thickened CO2 increased with rising pressure, with SHTTF demonstrating the optimal thickening performance. The likely reasons are (1) the incorporation of fluoride groups enhances the solubility of the modified siloxane polymer in CO2; (2) during pressurization, decreasing intermolecular distances facilitate the formation of hydrogen bonding through Lewis acid–base pairing. These two effects work synergistically, manifesting macroscopically as an increase in CO2 viscosity [19,25].
Figure 10 presents the curves depicting the influence of temperature on the thickening capacity of various thickeners for the CO2 system (Table 5 shows the original data corresponding to Figure 10). The results indicate that the viscosity of CO2 thickened by SHTT, SHTA, SHTTF, and SHTAF declined progressively with increasing temperature. This occurs because, at lower temperatures, the three-dimensional network structures formed by molecular intertwining are insufficient to counteract the disruption of the network structure induced by the temperature rise, as well as the detrimental effect on viscosity caused by the breakdown of Lewis acid–base pairs [12,26].
To further clarify the gap between the thickener prepared in this study and similar thickeners in the current literature, Table 6 and Figure 11 present the detailed viscosity data of similar thickeners under different experimental conditions in the literature and a 3D intuitive comparison diagram, respectively. Overall, the SHTTF thickener prepared in our study exhibits relatively excellent CO2 thickening ability, which is superior to the reported HBD-1 thickener (32 °C/8 MPa) [17], HS series thickeners (32 °C/10 MPa) [19], PDMS (42 °C/20 MPa) [27], EEPDMS (30 °C/8 MPa) [28,29], and Ester-branched polydimethylsiloxane (30 °C/12 MPa) [30]. In addition, it is worth mentioning that, in comparison with the thickeners in these references, no cosolvent was added in this study. However, there is still a significant gap in the thickening ability between the thickener in this study and the reported HBD-2 [17] and SiO2/HBD-2 thickeners [16] in the literature, which is exactly the direction of our future efforts (structure optimization, compounding of different types of thickeners, or introduction of nanomaterials).

3.4. Thickening Mechanisms

3.4.1. Molecular Electrostatic Potential Map

Figure 12 shows the electrostatic potential distribution of CO2 molecules and modified siloxane thickeners. In CO2, negative charges concentrate on oxygen atoms, whereas positive charges in both modified siloxane and fluorinated modified siloxane thickeners distribute evenly across carbon-containing groups [32]. Simultaneously, the electrostatic potential at polymerization sites between modified siloxane and fluorides (Figure 12d,e) appears slightly higher than that of surrounding functional groups [15]. Consequently, both modified and fluorinated modified siloxane thickeners can bind with CO2 molecules through electrostatic attraction [1,15].

3.4.2. CED and Δ

CED and δ are commonly employed to characterize intermolecular interactions in polymers. CED is defined as the energy required to vaporize one mole of a condensed substance per unit volume, overcoming intermolecular forces, which primarily reflects inter-group interactions [33]. Relevant studies demonstrate that polymers with lower CED exhibit higher solubility in CO2. Furthermore, smaller differences between polymer and CO2 solubility parameters correlate with enhanced dissolution performance of the polymer in CO2 [34]. The CED and δ of CO2 and thickened CO2 system are tabulated in Table 7, all four thickener polymers show minor δ differences with CO2. Among these, SHTTF exhibits the smallest δ difference with CO2, confirming its optimal solubility in the CO2 system.

3.4.3. Intermolecular Interaction Energy

Interaction energy (Eint) quantitatively characterizes the strength of interactions between polymers and CO2 molecules (calculated via Equation (4)) [18,35]. A larger absolute value of Eint indicates stronger polymer–CO2 interactions, which enhances polymer solubility in CO2 [33,36]. As shown in Table 8, in modified siloxane systems, the SHTT-CO2 system exhibits a greater absolute Eint than the SHTA-CO2 system, confirming superior solubility of thickener polymer SHTT in CO2. After fluorination modification, thickener polymer SHTTF maintains stronger solubility in CO2 than SHTAF. Based on the above interaction energy data between different thickeners and CO2, the solubility order of different thickeners in CO2 is as follows: SHTTF-CO2 > SHTAF-CO2 > SHTT-CO2 > SHTA-CO2. Among them, the maximum interaction energy between the thickener SHTTF and CO2 indicates that it has the strongest polymer–CO2 interaction with CO2, which improves its solubility in CO2 and thickening ability. This result corresponds to the result in the thickening ability test in Section 3.3. In addition, the reason for the enhanced polymer–CO2 interaction may be that the interaction between the polar sites on the polymer chain and CO2 breaks the intermolecular interaction between polymer–polymer molecules [36].

3.4.4. Radial Distribution Function (RDF)

Calculation of radial distribution functions between different thickener molecules and carbon dioxide molecules by means of trajectory files of molecular dynamics optimized system models (Figure 13a–d). Figure 14 shows the radial distribution functions between different thickeners and CO2: (a) four repeating units of the thickener + 1000 CO2 molecules; (b) thickener polymer chains consisting of four chains with 6, 8, and 10 repeating units of the thickener (SHTT/SHTA) + 1000 CO2 molecules, respectively; (c) thickener polymer chains consisting of four chains with 3, 4, and 5 repeating units of the thickener (SHTTF/SHTAF) + 1000 CO2 molecules, respectively; (d) a combination of Figure 13b,c. From the RDFs (Radial Distribution Functions) of different thickeners and CO2 in Figure 14, it can be seen that a higher RDF value indicates better miscibility between the thickener and CO2. From Figure 14a,d, it can be obtained that the dissolution abilities of thickeners with different types, different chain lengths, and different molecular weights in CO2 are in the following order: SHTTF-CO2 > SHTAF-CO2 > SHTT-CO2 > SHTA-CO2. That is, the introduction of fluorocarbon chains weakens the intermolecular interactions between the thickener polymer molecules, enhancing their solubility and thickening ability [4,36]. From the RDF data in Figure 14b,c, it can be seen that as the chain length and molecular weight of the thickeners (SHTT/SHTA, SHTTF/SHTAF) increase, their dissolution abilities in CO2 decrease. The reason for this phenomenon may be that while the increase in the chain length and molecular weight of the thickener polymer introduces more polar sites, the intermolecular interactions between polymer–polymer molecules also increase, and the interaction between the polymer and CO2 weakens, resulting in a decrease in solubility and thickening ability [37].

4. Conclusions

(1)
Four modified silicone-based thickeners for supercritical CO2 fracturing fluids were successfully synthesized. FT-IR and NMR characterization confirmed the successful incorporation of CO2-philic groups into thickener molecular structures, achieving the synthesis of fluorinated silicone thickeners.
(2)
The fluorinated silicone thickeners demonstrated exceptional thickening performance and favorable solubility. At 5 wt% concentration under 10 MPa and 32 °C, they formed homogeneous mixtures with CO2, elevating the viscosity of CO2 to 3.68 mPa·s. Systematic experiments established positive correlations between CO2 viscosity and both thickener concentration and pressure, while revealing an inverse relationship with temperature. The limitation of this study is that it only investigated the thickener concentrations within the range of 0~5 wt%.
(3)
Integrating experimental results with molecular simulations, the thickening mechanism was elucidated: The introduction of fluorinated compounds with carbon dioxide-affinitive moieties significantly enhances the dissolution of thickening functional-modified silicone segments in CO2, thereby substantially increasing the viscosity of CO2 fracturing fluids.

Author Contributions

Conceptualization, data curation and visualization, P.C.; conceptualization, methodology and funding acquisition, Y.X.; writing—original draft and formal analysis, D.D.; data curation and investigation, R.J.; conceptualization, methodology, writing—original draft, and writing—review and editing, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support generously provided by China Petroleum Southwest Oil and Gas Field Company School Enterprise Joint Research Project (Grant No. 20230302-33) for this study.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

Authors Pengfei Chen, Ying Xiong, and Rui Jiang were employed by PetroChina Southwest Oil and Gas Field Company. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The authors declare that this study received funding from China Petroleum Southwest Oil and Gas Field Company School Enterprise Joint Research Project. The funding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
CCUSCO2 capture, utilization, and storage
FT-IRfourier transform infrared spectrum
1H NMR1H nuclear magnetic resonance spectroscopy
ESPelectrostatic potential
HMMtetramethyldisiloxane
D4octamethylcyclotetrasiloxane
MDmolecular dynamics
CEDcohesive energy density
Ethe energy of the substance
Vthe molar volume of the substance
δthe solubility parameter
Einterthe total intermolecular energy of the system
Evanthe energy from van der Waals interactions
Eelectthe energy from electrostatic interactions
Eintthe polymer–CO2 interaction energy
Epolymer+CO2the total energy of the polymer–CO2 system
Epolymerthe energy of the polymer
ECO2the energy of CO2, kJ/mol
RDFradial distribution function
g A B ( r ) the radial distribution function value
N A B the number of species A and B (molecules or atoms) in the system
r the distance interval width
N A B the number of B particles (or A particles) found within the distance range r~r + Δr from a central A particle (or B particle)
ρ A B the density of system

References

  1. Li, J.; Pu, W.; Du, D.; Wu, T.; Shen, C.; Liao, Z. Molecular dynamics simulation of solvation behavior and thickening properties of modified siloxanes in supercritical carbon dioxide under different pressure and temperature. J. Mol. Liq. 2024, 414, 126082. [Google Scholar] [CrossRef]
  2. Song, Y.; Zhang, Q.; Zou, X.; Fan, J.; Wang, S.; Zhu, Y. Research progress on and outlook of direct CO2 thickeners for enhanced oil recovery. RSC Adv. 2025, 15, 714–731. [Google Scholar] [CrossRef]
  3. Prasad, S.K.; Sangwai, J.S.; Byun, H.-S. A review of the supercritical CO2 fluid applications for improved oil and gas production and associated carbon storage. J. CO2 Util. 2023, 72, 102479. [Google Scholar] [CrossRef]
  4. Dai, C.; Liu, P.; Gao, M.; Liu, Z.; Liu, C.; Wu, Y.; Wang, X.; Liu, S.; Zhao, M.; Yan, H. Preparation and thickening mechanism of copolymer fluorinated thickeners in supercritical CO2. J. Mol. Liq. 2022, 360, 119563. [Google Scholar] [CrossRef]
  5. Zhang, G.; Wu, T.; Li, J.; Pang, Q.; Yang, H.; Liu, G.; Huang, H.; Zhu, Y. Dynamics simulation of the effect of cosolvent on the solubility and tackifying behavior of PDMS tackifier in supercritical CO2 fracturing fluid. Colloids Surf. A Physicochem. Eng. Asp. 2023, 662, 130985. [Google Scholar] [CrossRef]
  6. 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. [Google Scholar] [CrossRef] [PubMed]
  7. Zheng, B.; Qi, S.; Lu, W.; Guo, S.; Wang, Z.; Yu, X.; Zhang, Y. Experimental Research on Supercritical Carbon Dioxide Fracturing of Sedimentary Rock: A Critical Review. Acta Geol. Sin.–Engl. Ed. 2023, 97, 925–945. [Google Scholar] [CrossRef]
  8. Gupta, N.; Verma, A. Supercritical Carbon Dioxide Utilization for Hydraulic Fracturing of Shale Reservoir, and Geo-Storage: A Review. Energy Fuels 2023, 37, 14604–14621. [Google Scholar] [CrossRef]
  9. 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]
  10. Li, N.; Zhang, H.; Ren, X.; Wang, J.; Yu, J.; Jiang, C.; Zhang, H.; Li, Y. Development status of supercritical carbon dioxide thickeners in oil and gas production: A review and prospects. Gas Sci. Eng. 2024, 125, 205312. [Google Scholar] [CrossRef]
  11. Zhou, M.; Ni, R.; Zhao, Y.; Huang, J.; Deng, X. Research progress on supercritical CO2 thickeners. Soft Matter 2021, 17, 5107–5115. [Google Scholar] [CrossRef] [PubMed]
  12. Gandomkar, A.; Reza Nasriani, H.; Enick, R.M.; Torabi, F. The effect of CO2-philic thickeners on gravity drainage mechanism in gas invaded zone. Fuel 2023, 331, 125760. [Google Scholar] [CrossRef]
  13. 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]
  14. Kilic, S.; Enick, R.M.; Beckman, E.J. Fluoroacrylate-aromatic acrylate copolymers for viscosity enhancement of carbon dioxide. J. Supercrit. Fluids 2019, 146, 38–46. [Google Scholar] [CrossRef]
  15. Zhao, M.; Yan, R.; Li, Y.; Wu, Y.; Dai, C.; Yan, H.; Liu, Z.; Cheng, Y.; Guo, X. Study on the thickening behavior and mechanism of supercritical CO2 by modified polysiloxane. Fuel 2022, 323, 124358. [Google Scholar] [CrossRef]
  16. Wang, Y.; Liu, B.; Li, D.; Liang, L. SiO2 synergistic modification of siloxane thickener to improve the viscosity of supercritical CO2 fracturing fluid. Energy Sources Part A Recovery Util. Environ. Eff. 2022, 44, 6602–6617. [Google Scholar] [CrossRef]
  17. Liu, B.; Wang, Y.; Liang, L. Preparation and Performance of Supercritical Carbon Dioxide Thickener. Polymers 2021, 13, 78. [Google Scholar] [CrossRef]
  18. Wang, X.; Liang, S.; Zhang, Q.; Wang, T.; Zhang, X. Molecular Dynamics Simulation on Thickening and Solubility Properties of Novel Thickener in Supercritical Carbon Dioxide. Molecules 2024, 29, 2529. [Google Scholar] [CrossRef]
  19. 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]
  20. Chang, K.-S.; Chung, Y.-C.; Yang, T.-H.; Lue, S.J.; Tung, K.-L.; Lin, Y.-F. Free volume and alcohol transport properties of PDMS membranes: Insights of nano-structure and interfacial affinity from molecular modeling. J. Membr. Sci. 2012, 417–418, 119–130. [Google Scholar] [CrossRef]
  21. 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]
  22. Tang, M.; Fang, X.; Li, B.; Xu, M.; Wang, H.; Cai, S. Hydrophobic Modification of Wood Using Tetramethylcyclotetrasiloxane. Polymers 2022, 14, 2077. [Google Scholar] [CrossRef] [PubMed]
  23. Mali, P.; Sonawane, N.; Patil, V.; Mawale, R.; Pawar, N. Designing efficient UV-curable flame-retardant and antimicrobial wood coating: Phosphorodiamidate reactive diluent with epoxy acrylate. J. Polym. Res. 2021, 28, 376. [Google Scholar] [CrossRef]
  24. Huang, M.; Peng, Z.-h.; Zhang, W.; Xiang, Y.; Cao, F. Influence of β-Si3N4 whiskers on sintering and crystallization of fused silica ceramics. J. Am. Ceram. Soc. 2023, 106, 967–976. [Google Scholar] [CrossRef]
  25. Li, Q.; Wang, Y.; Wang, X.; Yu, H.; Li, Q.; Wang, F.; Bai, H.; Kobina, F. An application of thickener to increase viscosity of liquid CO2 and the assessment of the reservoir geological damage and CO2 utilization. Energy Sources Part A Recovery Util. Environ. Eff. 2019, 41, 368–377. [Google Scholar]
  26. Dai, C.; Wang, T.; Zhao, M.; Sun, X.; Gao, M.; Xu, Z.; Guan, B.; Liu, P. Impairment mechanism of thickened supercritical carbon dioxide fracturing fluid in tight sandstone gas reservoirs. Fuel 2018, 211, 60–66. [Google Scholar] [CrossRef]
  27. Du, M.; Sun, X.; Dai, C.; Li, H.; Wang, T.; Xu, Z.; Zhao, M.; Guan, B.; Liu, P. Laboratory experiment on a toluene-polydimethyl silicone thickened supercritical carbon dioxide fracturing fluid. J. Pet. Sci. Eng. 2018, 166, 369–374. [Google Scholar] [CrossRef]
  28. Li, Q.; Wang, Y.; Wang, F.; Li, Q.; Kobina, F.; Bai, H.; Yuan, L. Effect of a Modified Silicone as a Thickener on Rheology of Liquid CO2 and Its Fracturing Capacity. Polymers 2019, 11, 540. [Google Scholar] [CrossRef]
  29. 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]
  30. 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]
  31. Pu, W.; Li, J.; Du, D.; Zhao, J.; Wu, T.; Xiong, Y.; Chen, P.; Jiang, R. Research progress of thickener for supercritical carbon dioxide fracturing fluid. Petroleum, 2025; in press. [Google Scholar] [CrossRef]
  32. Nelson, P.N. A theoretical study of the interactions between carbon dioxide and some Group(III) trihalides: Implications in carbon dioxide sequestration. J. Mol. Struct. 2021, 1223, 129212. [Google Scholar] [CrossRef]
  33. Sun, W.; Sun, B.; Li, Y.; Fan, H.; Gao, Y.; Sun, H.; Li, G. Microcosmic understanding on thickening capability of copolymers in supercritical carbon dioxide: The key role of π–π stacking. RSC Adv. 2017, 7, 34567–34573. [Google Scholar] [CrossRef]
  34. Li, M.; Zhang, J.; Zou, Y.; Wang, F.; Chen, B.; Guan, L.; Wu, Y. Models for the solubility calculation of a CO2/polymer system: A review. Mater. Today Commun. 2020, 25, 101277. [Google Scholar] [CrossRef]
  35. Wang, J.; Ding, W.; Zhang, B.; Jin, H. Polycyclic aromatic hydrocarbons dissolution in supercritical carbon dioxide by molecular dynamics simulation. J. Mol. Liq. 2023, 391, 123358. [Google Scholar] [CrossRef]
  36. Sun, B.; Sun, W.; Wang, H.; Li, Y.; Fan, H.; Li, H.; Chen, X. 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]
  37. Tan, J.; Wang, Z.; Chen, S.; Hu, H. Progress and Outlook of Supercritical CO2–Heavy Oil Viscosity Reduction Technology: A Minireview. Energy Fuels 2023, 37, 11567–11583. [Google Scholar] [CrossRef]
Figure 1. Synthetic routes of HMM-D4.
Figure 1. Synthetic routes of HMM-D4.
Polymers 17 02640 g001
Figure 2. Synthetic routes of copolymer thickeners.
Figure 2. Synthetic routes of copolymer thickeners.
Polymers 17 02640 g002
Figure 3. Modification routes of copolymer thickeners.
Figure 3. Modification routes of copolymer thickeners.
Polymers 17 02640 g003
Figure 4. Schematic diagram of visualization supercritical CO2 fracturing fluid evaluation device.
Figure 4. Schematic diagram of visualization supercritical CO2 fracturing fluid evaluation device.
Polymers 17 02640 g004
Figure 5. FTIR spectra of HMM-D4, SHTT and SHTA.
Figure 5. FTIR spectra of HMM-D4, SHTT and SHTA.
Polymers 17 02640 g005
Figure 6. 1H NMR spectra of SHTT (a), SHTA (b), SHTTF (c), and SHTAF (d).
Figure 6. 1H NMR spectra of SHTT (a), SHTA (b), SHTTF (c), and SHTAF (d).
Polymers 17 02640 g006
Figure 7. Solubility performance under different temperatures and pressures: (a) SHTT-CO2 (a1: 25 °C, 7.5 MPa; a2: 32 °C, 7.5 MPa; a3: 32 °C, 10 MPa; a4: 32 °C, 7.5 MPa); (b) SHTTF-CO2 (b1: 25 °C, 7.5 MPa; b2: 32 °C, 7.5 MPa; b3: 32 °C, 10 MPa; b4: 32 °C, 7.5 MPa).
Figure 7. Solubility performance under different temperatures and pressures: (a) SHTT-CO2 (a1: 25 °C, 7.5 MPa; a2: 32 °C, 7.5 MPa; a3: 32 °C, 10 MPa; a4: 32 °C, 7.5 MPa); (b) SHTTF-CO2 (b1: 25 °C, 7.5 MPa; b2: 32 °C, 7.5 MPa; b3: 32 °C, 10 MPa; b4: 32 °C, 7.5 MPa).
Polymers 17 02640 g007
Figure 8. The influence of different contents and types of thickeners on the viscosity of thickened supercritical CO2 at 32 °C and 10 MPa.
Figure 8. The influence of different contents and types of thickeners on the viscosity of thickened supercritical CO2 at 32 °C and 10 MPa.
Polymers 17 02640 g008
Figure 9. Effect of pressure on thickened CO2 viscosity.
Figure 9. Effect of pressure on thickened CO2 viscosity.
Polymers 17 02640 g009
Figure 10. Effect of temperature on thickened CO2 viscosity.
Figure 10. Effect of temperature on thickened CO2 viscosity.
Polymers 17 02640 g010
Figure 11. A 3D diagram comparing the viscosity data of this study with those of similar thickeners in the existing literature. The corresponding references in the figure are as follows: Wang, et al., 2022 [16], Liu, et al., 2021a [17], Liu, et al., 2021b [19], Du, et al., 2018 [27], Li, et al., 2019a [28], Wang, et al., 2018 [29], Li, et al., 2019b [30].
Figure 11. A 3D diagram comparing the viscosity data of this study with those of similar thickeners in the existing literature. The corresponding references in the figure are as follows: Wang, et al., 2022 [16], Liu, et al., 2021a [17], Liu, et al., 2021b [19], Du, et al., 2018 [27], Li, et al., 2019a [28], Wang, et al., 2018 [29], Li, et al., 2019b [30].
Polymers 17 02640 g011
Figure 12. Electrostatic potential distribution of CO2 molecules and modified siloxane thickeners ((a), CO2; (b), SHTT; (c), SHTA; (d), SHTTF; (e), SHTAF).
Figure 12. Electrostatic potential distribution of CO2 molecules and modified siloxane thickeners ((a), CO2; (b), SHTT; (c), SHTA; (d), SHTTF; (e), SHTAF).
Polymers 17 02640 g012
Figure 13. The system model after molecular dynamics optimization ((a) SHTT-CO2, (b) SHTA-CO2, (c) SHTTF-CO2, (d) SHTAF-CO2)).
Figure 13. The system model after molecular dynamics optimization ((a) SHTT-CO2, (b) SHTA-CO2, (c) SHTTF-CO2, (d) SHTAF-CO2)).
Polymers 17 02640 g013
Figure 14. The radial distribution functions of thickeners with different chain lengths and molecular weights and CO2 molecules. The compositions of each system are as follows: (a) four repeating units of the thickener + 1000 CO2 molecules; (b) thickener polymer chains consisting of four chains with 6, 8, and 10 repeating units of the thickener (SHTT/SHTA) + 1000 CO2 molecules, respectively; (c) thickener polymer chains consisting of four chains with 3, 4, and 5 repeating units of the thickener (SHTTF/SHTAF) + 1000 CO2 molecules, respectively; (d) A combination of the Figure (b,c).
Figure 14. The radial distribution functions of thickeners with different chain lengths and molecular weights and CO2 molecules. The compositions of each system are as follows: (a) four repeating units of the thickener + 1000 CO2 molecules; (b) thickener polymer chains consisting of four chains with 6, 8, and 10 repeating units of the thickener (SHTT/SHTA) + 1000 CO2 molecules, respectively; (c) thickener polymer chains consisting of four chains with 3, 4, and 5 repeating units of the thickener (SHTTF/SHTAF) + 1000 CO2 molecules, respectively; (d) A combination of the Figure (b,c).
Polymers 17 02640 g014
Table 1. The repetitive unit structures of different thickeners (yellow, white, gray, red, blue, and purple denote silicon, hydrogen, carbon, oxygen, fluorine, and nitrogen atoms, respectively).
Table 1. The repetitive unit structures of different thickeners (yellow, white, gray, red, blue, and purple denote silicon, hydrogen, carbon, oxygen, fluorine, and nitrogen atoms, respectively).
NameMolecular Structure
CO2Polymers 17 02640 i001
SHTTPolymers 17 02640 i002
SHTAPolymers 17 02640 i003
SHTTFPolymers 17 02640 i004
SHTAFPolymers 17 02640 i005
Table 2. Phase behavior and dissolution pressure of thickeners (SHTT and SHTTF) in CO2 under different temperatures and pressures.
Table 2. Phase behavior and dissolution pressure of thickeners (SHTT and SHTTF) in CO2 under different temperatures and pressures.
SystemTemperature (°C)/Pressure (MPa)Dissolution
Pressure (MPa)
25 °C, 7.5 MPa32 °C, 7.5 MPa32 °C, 10 MPa32 °C, 7.5 MPa
SHTT-CO2a1a2a3a49.2
ClarifiedTurbid, reference object not visibleClear and transparentTurbid, reference object visible
SHTTF-CO2b1b2b3b48.6
ClarifiedTurbid, reference object not visibleClear and transparentTurbid, reference object visible
Table 3. The original viscosity data of supercritical CO2 thickened by different contents and types of thickeners at 32 °C and 10 MPa.
Table 3. The original viscosity data of supercritical CO2 thickened by different contents and types of thickeners at 32 °C and 10 MPa.
Temperature (°C)/Pressure (MPa) SystemSHTT-CO2SHTA-CO2SHTTF-CO2SHTAF-CO2
Content
(wt%)
Viscosity (mPa·s)
32 °C,
10 MPa
00.06335 (REFPROP)
10.8230.3552.0261.564
21.0270.5392.6121.636
31.2340.662.9511.986
41.3780.853.3962.47
51.5121.1623.682.676
Table 4. The original viscosity data of supercritical CO2 thickened by different types of thickeners under different pressures at 32 °C with a thickener content of 5 wt%.
Table 4. The original viscosity data of supercritical CO2 thickened by different types of thickeners under different pressures at 32 °C with a thickener content of 5 wt%.
Temperature (°C)/Content (wt%) SystemSHTT-CO2SHTA-CO2SHTTF-CO2SHTAF-CO2
Pressure (MPa) Viscosity (mPa·s)
32 °C, 5 wt%7.51.3120.93.3342.495
101.5121.1623.682.676
12.51.691.283.7852.96
151.8031.563.93.27
17.51.9051.774.1023.34
Table 5. The original viscosity data of supercritical CO2 thickened by different types of thickeners at different temperatures under 10 MPa with a thickener content of 5 wt%.
Table 5. The original viscosity data of supercritical CO2 thickened by different types of thickeners at different temperatures under 10 MPa with a thickener content of 5 wt%.
Pressure (MPa)/Content (wt%) SystemSHTT-CO2SHTA-CO2SHTTF-CO2SHTAF-CO2
Temperature (°C) Viscosity (mPa·s)
10 MPa, 5 wt%321.5121.1623.682.676
361.360.9853.512.33
401.110.813.092.17
440.9340.5232.771.79
480.80.42.491.31
Table 6. Under different experimental conditions (content, cosolvent, temperature, pressure), the viscosity data of this study are compared with those of similar thickeners in the existing literature [31].
Table 6. Under different experimental conditions (content, cosolvent, temperature, pressure), the viscosity data of this study are compared with those of similar thickeners in the existing literature [31].
ThickenerConcentration
(wt%)
Cosolvent
(wt%)
Experimental
Condition
Viscosity (mPa·s)Ref.
HBD-15no32 °C/8 MPa3.4[17]
HBD-25no32 °C/8 MPa4.3
SHTTF5no32 °C/10 MPa3.68This study
HS-15no32 °C/10 MPa0.8[19]
HS-25no32 °C/10 MPa2.7
HS-35no32 °C/10 MPa3.0
PDMS (350 mPa·s)510 wt% toluene42 °C/20 MPa1.5[27]
SiO2/HBD-25 wt% HBD-2 + 1 wt% SiO2no32 °C/10 MPa5.82[16]
EEPDMS39 wt% toluene30 °C/8 MPa1.3[28,29]
Ester-branched polydimethylsiloxane2.57.5 wt%
n-hexane
30 °C/12 MPa1.65[30]
Table 7. CED and δ of CO2 and thickened CO2 at 40 °C and 10 MPa.
Table 7. CED and δ of CO2 and thickened CO2 at 40 °C and 10 MPa.
SystemEvan/(J/m3)Eelect/(J/m3)Eother/(J/m3)CED/(J/m3)δ/(J/m3)1/2Δδ/(J/m3)1/2
CO27.958 × 1075.629 × 1073.033 × 1061.389 × 10811.7770
SHTT-CO21.365 × 1089.910 × 1075.454 × 1062.411 × 10815.5123.735
SHTA-CO21.369 × 1089.938 × 1075.413 × 1062.417 × 10815.5073.730
SHTTF-CO27.202 × 1074.727 × 1072.803 × 1061.221 × 10811.0360.741
SHTAF-CO26.082 × 1073.986 × 1072.305 × 1061.030 × 10810.1191.658
Table 8. Intermolecular interaction energy between thickener and CO2 (4 repeating units of the thickener + 1000 CO2 molecules).
Table 8. Intermolecular interaction energy between thickener and CO2 (4 repeating units of the thickener + 1000 CO2 molecules).
System E P o l y m e r C O 2 /(kcal/mol) E P o l y m e r /(kcal/mol) E C O 2 /(kcal/mol)Eint/(kcal/mol)
SHTT-CO2−2449.962372−1406.236231−846.918705−196.807436
SHTA-CO2−2205.395479−1181.816535−847.726976−175.851968
SHTTF-CO2−1635.494087−390.389668−1028.260001−216.844418
SHTAF-CO2−1473.056476−292.508835−977.561245−202.986396
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

Chen, P.; Xiong, Y.; Du, D.; Jiang, R.; Li, J. Synthesis of Linear Modified Siloxane-Based Thickeners and Study of Their Phase Behavior and Thickening Mechanism in Supercritical Carbon Dioxide. Polymers 2025, 17, 2640. https://doi.org/10.3390/polym17192640

AMA Style

Chen P, Xiong Y, Du D, Jiang R, Li J. Synthesis of Linear Modified Siloxane-Based Thickeners and Study of Their Phase Behavior and Thickening Mechanism in Supercritical Carbon Dioxide. Polymers. 2025; 17(19):2640. https://doi.org/10.3390/polym17192640

Chicago/Turabian Style

Chen, Pengfei, Ying Xiong, Daijun Du, Rui Jiang, and Jintao Li. 2025. "Synthesis of Linear Modified Siloxane-Based Thickeners and Study of Their Phase Behavior and Thickening Mechanism in Supercritical Carbon Dioxide" Polymers 17, no. 19: 2640. https://doi.org/10.3390/polym17192640

APA Style

Chen, P., Xiong, Y., Du, D., Jiang, R., & Li, J. (2025). Synthesis of Linear Modified Siloxane-Based Thickeners and Study of Their Phase Behavior and Thickening Mechanism in Supercritical Carbon Dioxide. Polymers, 17(19), 2640. https://doi.org/10.3390/polym17192640

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop