Next Article in Journal
Effects of FPV Drone Frame Materials on Thermal Conditions of Motors Under Extreme Payloads: Experimental and Numerical Analysis
Previous Article in Journal
Cooperative Failure Modes of Overlying Strata and Stressed Distribution Mechanism in Shallow Coal Seam Mining
Previous Article in Special Issue
Prediction of Three Pressures and Wellbore Stability Evaluation Based on Seismic Inversion for Well Huqian-1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Performance and Characteristics of Low-Molecular-Weight Cross-Linked Grafting Terpolymers as Thickening Agents in Reservoir Fracturing Processes

1
College of Safety Science and Engineering, Xi’an University of Science and Technology, Xi’an 710054, China
2
Southwest Branch of China National Coal Group Co., Ltd., Chongqing 400023, China
3
Chongqing Energy Investment Group Technology Co., Ltd., Chongqing 400061, China
*
Author to whom correspondence should be addressed.
Processes 2025, 13(10), 3032; https://doi.org/10.3390/pr13103032
Submission received: 5 August 2025 / Revised: 21 August 2025 / Accepted: 2 September 2025 / Published: 23 September 2025

Abstract

A novel fracture fluid based on a grafting polymer, PAM-co-PAMS-g-PEG (PAM-AMS-AEG), cross-linked by an organic Zr reagent was successfully produced via free-radical polymerization and an in situ cross-linking reaction with a high conversion rate of 96%, resulting in a low molecular weight of 250 kg·mol−1. The effect of fluid constitution on the rheological behavior demonstrates that the P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 (PASG/[Zr]) aqueous solution has the best comprehensive performance. The PASG/[Zr] solution with a low critical associating concentration (CAC) of 0.15 wt% showed faster and steadier disassociation–reassociation processes. The synergy of ionic hydrogen bonds between sulfonic and amine groups and Zr4+-coordination results in steady interactions and fast reconstitution of association, leading to remarkable temperature resistance from 30 to 120 °C and a fast response during thixotropic processes. The PASG/[Zr] solution reduces the damage under high pressure based on the rheological characteristics and compatibility with sand, leading to a low filtration loss of the artificial cores. The PASG/[Zr] solution exhibits a good sand-carrying ability based on the rheological and interfacial performance, resulting in slow settlement and fast suspension. The filtration performance of the PASG/[Zr] fracturing fluid showed that it is not sensitive to the shearing rate, core permeability, or pressure. The comprehensive performance of the PASG/[Zr] fracture fluid is better than that of traditional guar fluid, suggesting that it can be used under various conditions for stratum protection and shale gas extraction.

1. Introduction

Over the past two decades, the rapid expansion of unconventional hydrocarbon development has intensified the demand for advanced stimulation technologies that ensure efficient shale gas recovery while preserving the integrity of sensitive, clay-rich formations [1,2,3]. Without these processes, shale gas reservoirs with low permeability and weak beddings are extremely vulnerable to the invasion of aqueous fluids. The negatively charged basal surfaces of clay minerals can interact with invading fluids [3], leading to the propagation of micro-fractures due to the stress exceeding the tensile strength of the shale matrix. Traditional mitigation strategies include increasing the salinity, adding KCl, or converting to oil-based mud; however, these strategies either pollute the environment or fail under the elevated temperatures (>150 °C) encountered in deep shale clays [4,5].
Polyacrylamide (PAM)-based fracturing fluids have emerged as a new generation of intelligent viscous fluids, serving both as rheology modifiers and as interfacial barriers between the fracturing fluid and shale surfaces [6,7,8,9,10,11]. However, conventional PAM systems suffer from limited thermal stability and poor viscoelasticity under harsh reservoir conditions [12,13,14,15,16]. In order to improve the low thermal stability and viscoelasticity of PAM solutions [12,13,14,15,16], 2-acrylamido-2-methyl-1-propanesulfonic acid (AMS) can be copolymerized with AM monomers [17,18,19,20], resulting in high-density ionic hydrogen bonds [21,22] between the sulfonates and amides in mutually exclusive sulfonic side groups along the unfolding and expanding chain [23]. AMS-containing PAM copolymers retain a high viscosity in brines, which is critical when reusing flowback water as the base fluid in shale operations [24]. Their active ionic hydrogen bonds (iHBs) increase interchain interactions [25,26,27,28]. Grafting polymers with unique topological structures can produce special rheological characteristics, and can be used as pseudoplastic or thixotropic agents [29,30,31,32,33,34,35,36,37]. The dynamic network formed by cross-linked PAM-g-poly(ethylene oxide) (PAM-g-PEG) delays fracture closure, allowing for the deeper penetration of the proppant into complex fracture networks, leading to high-gel preserves under high-temperature and -stress conditions [29,30,31] and enhancement in their thermal stability and salt tolerance [32,33,34]. Furthermore, their Zr4+-coordination can form dynamic network cross-linkers with partial PAMs [38], resulting in active but steady supramolecular interactions in a wide temperature range [39]. Zr4+ coordination in ultra-high-temperature fracturing fluids has been successfully field-tested [40]. Therefore, to the best of our knowledge, a fracture fluid consisting of grafting P(AM-AMS) and with Zr4+-coordination that combines all the advantages mentioned above has never been reported.
Despite these advances, no prior study has reported a fracturing fluid that integrates grafting P(AM-AMS) with Zr4+ coordination to simultaneously leverage the benefits of iHBs, graft architecture, and metal–ligand cross-linking. In this work, we developed a novel fracturing fluid based on low-molecular-weight grafting terpolymers (PAM-AMS-AEG), an organic Zr4+ cross-linker, a surfactant (TBAC), and a gel breaker ((NH4)2S2O8). The adjustable rheological behavior of the fracture fluid is due to the multiple polymer networks being mixed with a sol-associator, which improves its shearing resistance and thixotropy. The high compatibility of the fracture fluid and artificial core leads to efficient sand-carrying and a low filtration loss, which are not sensitive to the environmental conditions, and the fluid shows better performance than a traditional fluid (guar gum).

2. Design of Experiments

Materials: Acrylamide (AM, 99%, Sinopharm Chemical Reagent Co., Ltd, Beijing, China), acrylamido-2-methyl-1-propanesulfonic acid (AMS, 99%, Sinopharm Chemical Reagent Co., Ltd, Beijing, China), and acryloyl polyethylene glycol (AEG, 99%, Mn = 1000 g·mol−1, Shandong Asia Chemical Co., Ltd, Dongying, China) were obtained and purified before polymerization to remove the polymerization inhibitor. K2S2O8 (96%, Chengdu Chron Chemicals Co., Ltd, Chengdu, China), Na2SO3 (98%, Chengdu Chron Chemicals Co., Ltd, Chengdu, China), (NH4)2S2O8 (96%, Chengdu Chron Chemicals Co,. Ltd, Chengdu, China), NaOH (96%, Chengdu Chron Chemicals Co., Ltd, Chengdu, China), tetrabutylammonium chloride (TBAC, 98%, Chengdu Chron Chemicals Co., Ltd, Chengdu, China), ZrOCl2·8H2O (98%, Chengdu Chron Chemicals Co., Ltd, Chengdu, China), isopropyl alcohol (99%, Chengdu Chron Chemicals Co., Ltd, Chengdu, China), L-lactic acid (LLA, 99%, Chengdu Chron Chemicals Co., Ltd, Chengdu, China), and guar gum (98%, Mw = 2000 kg·mol−1, Dongying Xinde Chemicals Co., Ltd, Dongying, China) were also used.
Polymerization of P(AM-AMS-AEG). The free-radical copolymerization of AM, AMS, and AEG was carried out in aqueous solution. The typical process is described as follows: 2.27 g of AM, 1.33 g of AMS, and 4.48 g of AEG were dissolved in 5 mL of deionized water in a flask. The NaOH solution (20 wt%) was added to this flask to adjust the pH = 7, and then 0.09 g of TBAC was added under stirring. The flask was vacuumed and filled with N2 3 times. Amounts of 0.09 g of K2S2O8 and 0.04 g of Na2SO3 were dissolved in 5 mL of deionized water. This initiating solution was injected into the AM/AMS/AEG/TBAC solution to initiate the free-radical copolymerization of AM, AMS, and AEG under stirring, and the polymerization was conducted at 80 °C for 6 h. The conversion of resulting polymers was calculated using the ratio of the original mass of monomers to the resulting mass of polymers.
Preparation of Organic Zr4+-based Cross-linker ([Zr]). The inorganic Zr4+-based cross-linker was prepared by dissolving ZrOCl2·8H2O in deionized water. An amount of 10 mL of ZrOCl2 aqueous solution (30 wt%) in a flask was mixed with 2 eq of isopropyl alcohol under N2 atmosphere. An amount of 2 eq of LLA was added fast into the solution and reacted for 1 h at 20 °C. This solution was neutralized by NaOH solution (20 wt%) at 80 °C for 4 h and vacuumed for 30 min to remove volatiles. The resulting solution was [Zr] solution.
Cross-linking reaction of P(AM-AMS-AEG). The polymerizing solution was injected above the ZrOCl2 aqueous solution (30 wt%) with various dosages to prepare P(AM-AMS-AEG) cross-linked by the inorganic Zr4+-based cross-linker. The polymerizing solution was injected above [Zr] with various dosages to prepare P(AM-AMS-AEG) cross-linked by the organic Zr4+-based cross-linker. The cross-linking reaction was conducted at various temperatures for various durations, and the apparent viscosity of the resulting solution was recorded.
Measurements. Fourier Transform Infrared Spectroscopy (FTIR). Chemical structure analysis of the resulting polymers was performed via Nicolet 6700 FTIR spectroscopy (Thermo Co., MA, USA). All samples were ground on a KBr wafer and dried at 105 °C for 24 h before measurement. IR spectra were recorded at 20 °C in the wavenumber range of 4000~400 cm−1 for 32 scans at a resolution of 4 cm−1.
Proton Nuclear Magnetic Resonance (1H NMR). 1H NMR analysis of the resulting polymers (D2O solution) was carried out. 1H NMR spectra were recorded via a Bruker AV III 400 spectrometer (Bruker Co., MA, USA, 400 MHz) with tetramethylsilane (TMS) as the internal reference.
Gel Permeation Chromatography (GPC). The number-average molecular weight (Mn) and polydispersity index (PDI, Mw/Mn) of terpolymers were determined by an Alliance e2695 chromatographic system (Waters Co., MA, USA) at 30 °C. Terpolymers were dissolved in deionized water (10 μg·mL−1) and deionized water was used as the mobile phase at a flow rate of 1.0 mL·min−1. The column was calibrated against the standard PAM samples.
Viscosity. The apparent viscosity (ηapp) of solution was measured via a CAP 2000+H viscometer (BROOKFIELD Co., MA, USA) at various temperatures and shearing rates. To evaluate the relationship between lnτ and ln γ ˙ , the cross-linking polymer solution (0.27 wt%) was poured into an Ubbelohde viscometer (d = 1 mm) in a DZ-80 constant-temperature bath and the apparent viscosity (ηapp) was recorded at various temperatures.
Interfacial Tension. The interfacial tension of terpolymer solutions was determined by using a JYW-200A interfacial tension meter (Suzhou Qile Electronic Technology Co., LTD).
Rheological Test. The rheological tests of polymer solutions were performed via an MCR 302 rheometer (Anton Paar Co., Graz, Austria). The oscillation stress sweep was performed on the samples with a conical plate (1 Hz, 20 °C). The oscillation frequency sweep was performed on the samples with a conical plate (1000 s−1, 20 °C). The thixotropic loop was recorded in a shearing–recovery cycle in the shearing rate range of 0~450 s−1.
DFT Analysis. Both the geometry optimization and the frequency calculations were performed with the Gaussian 09 program package and run at the D3-corrected levels of the B3LYP functional with an ultrafine grid [41,42].
Low-Field Time-Domain Proton Nuclear Magnetic Resonance (LF-NMR). The transversal relaxation time spectra (T2 spectra) of core samples were recorded on a Magnet2000 spectrometer (NIUMAG Co., Suzhou, China) at 20 °C with the CPMG sequence and inverted by the SIRT algorithm [43,44,45,46].
Static Settling Experiment. An amount of 100 mL of fracturing fluid was poured into a graduated cylinder. Proppant particles with various sizes were placed on the upper surface of the liquid. The settling height, H, of the proppant particles at various observing times, t, was measured and recorded. The settling velocity, vsettle, was determined by the ratio of H to the corresponding t.
Dynamic Settling Experiment. The device used for the dynamic sand-carrying test for fracturing fluid is shown in Figure S9. The fracturing fluid in the storage tank was smoothly injected into the pipeline through a screw pump and processed by the sand addition device to form the mixture with a fixed sand ratio. The mixture through the buffer pipe section entered the horizontal test sections with various diameters. The data acquisition system was connected to display the recording data on the computer. The settling rate of the proppant and the sand-laying pattern were observed with a high-speed camera.
Filtration Experiment. A dynamic filtration system for fracturing fluid was designed in the laboratory, which can simulate the formation environment and the flow and shearing process of fracturing fluid in shales. The schematic diagram of the experimental device is shown in Figures S10 and S11. The artificial cores with a diameter of 25.4 mm and length of 50 mm were soaked in standard brine (2.0% KCl + 5.5% NaCl + 0.45% MgCl2 + 0.55% CaCl2) to saturation. Subsequently, the cores were put into the experimental device, and the fracture fluid was injected into the device. The filtration loss was recorded at various times under various conditions. Finally, the cores were taken out until the environmental medium was replaced with standard brine. The core permeability of artificial cores with a diameter of 25.4 mm and length of 50 mm was determined, and the damage of the core was determined by the ratio of dry core permeability to filtrated core permeability.

3. Results and Discussions

3.1. Synthesis and Chemical Structure of Fracture Fluid Based on Low-Mw Cross-Linking Terpolymer

3.1.1. Polymerization of Terpolymer

The fracture fluid based on the low-molecular-weight cross-linking terpolymers was successfully produced via free-radical polymerization in aqueous solution and subsequent coordination with Zr-based reagent. Considering the viscous nature of the aqueous fracture fluid, acrylamide (AM) was selected as the primary monomer of the terpolymer, due to the good sand-carrying performance. In order to improve the rheological performance of the aqueous fracture fluid, the general functional molecule, 2-acrylamido-2-methylpropanesulfonic acid (AMS), was selected as the second monomer, since the ionic hydrogen bonds can in situ generate between sulfonic groups and amide groups. Moreover, monoacryl poly(ethylene glycol) (AEG) was selected as the third monomer to improve the amphipathy of the fracture fluid to improve compatibility with the formation fluid. In addition, the Na2SO3-K2S2O8 initiating system (I) was selected to realize highly efficient initiation in aqueous solution.
The terpolymer was synthesized through free-radical copolymerization of AM, AMS, and AEG in aqueous solution using a Na2SO3–K2S2O8 initiator system at elevated temperature, as illustrated in Scheme 1. To investigate the influence of polymerization temperature (Tp) and time (tp) on the process, the apparent viscosity (ηapp) and monomer conversion (conv.) were monitored under various conditions, with the molar ratio of AM:AMS:AEG:I fixed at 10:2:1.4:0.18. As shown in Figure 1a, both ηapp and conv. increased with prolonged reaction time at 80 °C: ηapp rose from 0 to 552 mPa·s, while conv. increased from 0% to 92%. The conversion versus time profile followed an exponential trend, indicating that the copolymerization follows apparent first-order kinetics. The effects of temperature were examined by measuring ηapp and conv. after 6 h at temperatures ranging from 30 to 90 °C. As Tp increased from 30 to 80 °C, ηapp increased from 0 to 552 mPa·s, but further increasing the temperature to 90 °C resulted in a decline to 384 mPa·s. Similarly, conversion reached a maximum of 92% at 80 °C. These results indicate that the optimal polymerization time and temperature are 6 h and 80 °C, respectively.
The monomer concentration is a critical factor in the preparation of fracture fluids, as the rheological properties of the solution are highly dependent on it. Given that the rheological behavior is primarily influenced by the concentration of the first monomer, the apparent viscosity (ηapp) was examined by increasing the total monomer content (fm) while maintaining a constant molar ratio of AM:AMS:AEG:I at 10:2:1.4:0.18. The polymerization was carried out at 80 °C for 6 h. As shown in Figure S1, ηapp increased from 498 to 569 mPa·s as fm rose from 20% to 30%, but decreased to 506 mPa·s when fm further increased to 40%. Additionally, the initiator concentration (fi) significantly affects the polymerization process, as it determines the average molecular weight of the polymer and consequently alters the rheological properties of the resulting fracture fluid. To explore the correlation between fi and ηapp, measurements were conducted with fi varying from 0.06% to 0.24%, while keeping the monomer ratio unchanged (AM:AMS:AEG:I = 10:2:1.4:0.18) and maintaining the reaction at 80 °C for 6 h. As depicted in Figure S2, ηapp increased from 223 to 569 mPa·s with fi rising from 0.06% to 0.18%, and then declined to 424 mPa·s when fi was further increased to 0.24%.
Subsequently, the influence of terpolymer composition on rheological properties was evaluated by measuring the apparent viscosity (ηapp) of the polymerizing solution after 6 h at 80 °C. The concentration of each monomer (fam, fams, and fae9) was varied individually across three experimental groups, while maintaining a baseline molar ratio of AM:AMS:AEG:I at 10:2:1.4:0.18. The effects of fam, fams, and fae9 on ηapp are presented in Figure 1b. The results reveal a dual dependence of ηapp on each monomer concentration. Specifically, ηapp reaches a maximum at fam = 30%, fams = 3%, and fae9 = 3.6%, respectively. The non-monotonic behavior observed with varying fam or fams is attributed to ionic hydrogen bonding between -SO3 and -NH2 groups, which is sensitive to the molar ratio of sulfonic to amide groups (Scheme 1). On the other hand, the dual trend in ηapp with fae9 is ascribed to changes in topological conformation resulting from optimal steric repulsion among interpenetrating side chains [37].
Based on the above analysis, the theoretically optimal feed molar ratio of AM:AMS:AEG was determined to be 10:1:1.2, with an optimal AM concentration of 30 mol%. To identify a practical feed ratio that ensures both high monomer conversion and favorable rheological properties, free-radical copolymerization was carried out under five different monomer and surfactant compositions. Each polymerization was conducted at 80 °C for 6 h with an initiator concentration (fi) of 0.18 mol%, using TBAC as a surfactant. The tested molar ratios AM:AMS:AEG:TBAC were as follows: 9:1:2.7:0.2, 9:1:1.7:0.2, 9:2:2.4:0.1, 10:2:0.7:0.2, and 10:2:1.4:0.1. The monomer conversion (conv.) and apparent viscosity (ηapp) of the resulting terpolymers are presented in Figure S3a,b. Among these, the system with the molar ratio of 10:2:1.4:0.1 exhibited the highest comprehensive performance, achieving a conversion of 96% and an apparent viscosity of 582 mPa·s. For simplicity, this terpolymer is denoted as P(AM10-AMS2-AEG1.4)/TBAC0.1.
The representative 1H NMR spectrum of P(AM10-AMS2-AEG1.4) in D2O is given in Figure 1c to clarify the chemical structure of the resulting P(AM-AMS-AEG) polymers. The signal assignments of the 1H NMR spectrum has been reported in the literature [18]. In detail, signals a and b at 1.54 and 2.10 ppm are assigned to H atoms on polymer chains (-CHa2-CRHb-), respectively, in which the integral ratio of a to b (Ia/Ib) equals 2. Signals c and d at 1.07 and 3.54 ppm are assigned to H atoms on the side groups of AMS (-NH-C(CHc3)2-CHd2-SO3H), respectively, in which the integral ratio of a to b (Ia/Ib) equals 3. Signal f at 3.61 ppm is assigned to H atoms on PEG side chains (-O-CHf2-CHf2-). The chemical constitution of P(AM10-AMS2-AEG1.4) can be calculated according to Equations (1) and (2):
F A M S = I d I a × 100 % = 1.84 13.92 × 100 % = 13.22   m o l %
F A E G = I f ÷ 23 ÷ 2 I a × 100 % = 70.9 ÷ 44 13.92 = 11.09   m o l %
where FAMS and FAEG indicate the molar content of AMS and AEG units on the terpolymer chains.
GPC characterization was performed on the five terpolymers mentioned above to quantify the number-average molecular weight (Mn) and molecular weight distribution (MWD) of terpolymers. All the GPC traces shown in Figure 1d present unimodal molecular weight distributions in the MWD range of 1.7 to 2.54, and the terpolymers with higher conversions were obviously shifted toward the higher-Mn region from 152 to 250 kg·mol−1. In brief, the chemical constitution, Mn, and polydispersity (PDI) of these terpolymers are listed in Table 1.
P(AM-AMS-AEG) can be easily synthesized in aqueous solution with the Na2SO3-K2S2O8 initiating system at 80 °C for 6 h. The conv. and Mn of P(AM10-AMS2-AEG1.4) can reach 96% and 250 kg·mol−1, respectively, in which the Mw of P(AM10-AMS2-AEG1.4) is far less than that of guar gum with a typical Mw of 2000 kg·mol−1.

3.1.2. Cross-Linking Process of Terpolymer Solution

The terpolymer solution containing surfactant cannot be directly used as a fracture fluid, due to its still unsatisfactory rheological performance. To further enhance the viscosity of the terpolymer solution, a chemical cross-linking process was introduced in subsequent investigations. Commercial cross-linkers for fracture fluids typically rely on coordination between Zr4+ ions and polar functional groups—such as carboxylate, amine, and amide—and are available in both organic and inorganic zirconium-based formulations. To evaluate the compatibility of these cross-linkers with the representative terpolymer solution P(AM10-AMS2-AEG1.4)/TBAC0.1, comparative tests were conducted (Figure S4). It was observed that the solution turned into a white emulsion immediately after the addition of just 0.3 mol% of the inorganic zirconium reagent. In contrast, the solution remained transparent even after the introduction of up to 1 mol% of the organic zirconium reagent.
Considering the apparent compatibility of the organic Zr-reagent (hereafter denoted as [Zr]), the effect of cross-linking temperature (Txl) or the dosage of [Zr] on the rheological characteristics of this solution is further investigated. To prevent the irreversible gelation during the cross-linking process, the P(AM10-AMS2-AEG1.4)/TBAC0.1 solution was diluted 10 times to a concentration of 0.55 wt% and mixed with [Zr] at various cross-linking temperatures (Txl). With increments in Txl, the cross-linking time (txl) decreases from 600 to 100 s, and the ηapp of the dilute solution increases from 60 to 130 mPa·s (Figure 2a). The plot of txl vs. Txl is in accordance with Boltzmann growth, and the plot of ηapp vs. Txl is in accordance with exponential growth. As the txl or ηapp of the resulting solutions barely changes with Txl ranging from 80 to ~90 °C, the practical Txl is set to 80 °C for continuous synthesizing operation. On the other hand, txl decreases from 1200 to 300 s with the molar ratio of [Zr] to AM (f[Zr]) increasing from 0.1% to 0.45%, and the ηapp of the dilute solution increases from 320 to 650 mPa·s in the process after cooling down (Figure 2a). The plot of txl vs. f[Zr] is in accordance with linear growth, and the plot of ηapp vs. f[Zr] is in accordance with exponential growth. The products with various f[Zr] exhibiting various rheological characteristics can be utilized under different operating conditions, which is described in Section 3.2.2.
The chemical structure of the cross-linked terpolymer, P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1, was analyzed by FTIR characterization, as shown in Figure 2b. For comparison, the FTIR spectra of P and [Zr] are also given in Figure 2b. The absorbance peaks at 1048, 1132, and 1152 cm−1 are assigned to vibration of S-O (νS-O) and symmetric and asymmetric vibration of O-S-O (νs,O-S-O and νs,O-S-O), indicating the existence of AMS units in the terpolymer. The absorbance peaks at 1122 and 954 cm−1 are assigned to vibration of C-O (νS-O) and C-O-C (νC-O-C), indicating the existence of AEG units in the terpolymer. After cross-linking with [Zr], the absorbance peaks assigned to the amide group coordinated with Zr4+ can be observed at 1558 and 3352 cm−1. In addition, the absorbance assigned to ionic hydrogen bonds between -NH2 and -SO3H can be detected at 3207 cm−1, indicating that the multiple networks with covalent cross-linking sites ([Zr] groups) and noncovalent cross-linking sites (ionic hydrogen bonds) can in situ generate in aqueous solution.

3.1.3. Formula of Fracture Fluid

The infiltration of polymer-based fracture fluids into formation cracks can cause significant formation damage due to their rheological properties. To mitigate this risk and enable controllable viscosity reduction, the inclusion of a gel breaker is essential in fracture fluid formulations. In this study, ammonium persulfate ((NH4)2S2O8) was selected as the gel breaker. Its effect on the rheological and surface characteristics was evaluated by adding varying molar ratios of (NH4)2S2O8 to AM (denoted as f_breaker) into a 0.55 wt% aqueous solution of P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1.
As shown in Figure 2c, the apparent viscosity (ηapp) decreases sharply from 100 to 10 mPa·s as fbreaker increases from 0.01% to 0.04%, and then declines gradually to 5 mPa·s when fbreaker reaches 0.06%. Similarly, the surface tension (γs) drops markedly from 40 to 20 mN·m−1 over the same initial range of fbreaker (0.01% to 0.04%), and stabilizes between 10 and 20 mN·m−1 with the further increase to 0.06%. These results indicate that the effective adjustment range for fbreaker lies between 0.01% and 0.04%, with the optimal dosage dependent on specific operational requirements.
In addition to viscosity control, surfactants play a crucial role in fracture fluids by enhancing surface activity, which is vital for effective proppant transport. To investigate the influence of TBAC concentration on interfacial behavior, the surface tension (γs) and sand–fluid interfacial tension (γi) of the P(AM10-AMS2-AEG1.4)/[Zr]0.35 solution were measured across a range of TBAC-to-AM molar ratios (fs) from 0.2% to 1.0% (Figure S5a,b). Both γs and γi decrease consistently with increasing fs: γs declines from 34.65 to 27.13 mN·m−1, and γi decreases from 7.30 to 1.39 mN·m−1. These results suggest excellent compatibility between TBAC and sand when fs ≥ 0.8%.
Furthermore, the surfactant content also affects the apparent viscosity of the fracture fluid. As fs increases, ηapp rises slightly from 610 to 685 mPa·s, which can be attributed to enhanced phase interactions induced by the surfactant.
The P(AM-AMS-AEG)/[Zr]/TBAC/[(NH4)2S2O8] fracture fluid can be in situ synthesized via free-radical polymerization for 6 h and subsequent cross-linking reaction with [Zr] at 80 °C. The effect of the feeding ratio of the terpolymer to the cross-linking reagent, gel breaker, and surfactant on the rheological characteristics was investigated, in which the P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 aqueous solution reveals the best comprehensive performance.

3.2. Rheological Behavior of Fracture Fluid Under Various Conditions

3.2.1. Associating Behavior and Segmental Motion in P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 Aqueous Solution

The polymer solution has a general transition process from dissociating status to associating status with increasing concentration, as shown in Figure 3a. The associating polymer in aqueous solution reveals remarkable non-Newtonian characteristics, whereas the dissociating sample can be regarded as a Newtonian fluid. In order to clarify the critical associating concentration (CAC) of samples, the increase in ηapp with increments in cterpolymer was analyzed first (Figure 3b). ηapp enhances from 120 to 150 mPa·s with a linear slope of 375 in the cterpolymer range of 0.08~0.16 wt%, and then dramatically enhances to 420 mPa·s with a linear slope of 1688 in the cterpolymer range of 0.16~0.32 wt%. The cterpolymer of 0.16 wt% at the inflection point can be recognized as the CAC.
To analyze in-depth the CAC and corresponding rheological behavior, the variations in shearing stress (τ) with shearing rate ( γ ˙ ) were recorded for terpolymer solutions with various concentrations. The plots of lnτ vs. ln γ ˙ are given in Figure S6 and the relationships between the consistency coefficient (K) or flow index (n) and cterpolymer are shown in Figure 3c, according to Equation (3).
l n τ = n l n γ ˙ + l n K
Figure 3c shows that K increases from 0.35 to 0.95 and n decreases from 0.44 to 0.39 with increasing cterpolymer. The CAC can be determined by the inflection point at 0.15 wt%, which is basically in accordance with the value according to the ηapp-cterpolymer plot.
Furthermore, the special pseudoplastic characteristic of the branched terpolymer has been proved due to the topological structure, which can adjust the rheological behavior of fracture fluid. In order to study the effect of fAEG on the sol–gel transition process, the solation time (tsol) of terpolymer solution with various fAEG values of P(AM10-AMS2-AEG) under the γ ˙ of 1000 s−1 at 20 °C was recorded, as shown in Figure S7. The tsol of terpolymer solution increases from 15 to 60 min, in which the relationship between tsol and fAEG is in accordance with exponential growth. The increase in tsol has a dramatic transition point at an fAEG of 0.8%, indicating that the topological structure has a remarkable effect when fAEG > 0.8%.
According to the above analysis, cterpolymer is selected as 0.27 wt% for a stable associating structure in a static state in the fracture fluid. The dynamic stability of the fracture fluid was studied by the variation in storage (G′) and loss modulus (G″). The modulus comparison of P(AM10-AMS2-AEG1.4) and guar gum aqueous solutions (0.27 wt%) with increasing shearing stress is shown in Figure 3d. It is shown in Figure 3d that the edge of the linear viscoelastic zone (LVZ) of guar gum is at a stress of 0.1 Pa, whereas that of P is at a stress of 0.17 Pa. Compared with guar gum solution, the higher G′ and G″ of P(AM10-AMS2-AEG1.4) solution have no section point in the testing range, indicating that the P(AM10-AMS2-AEG1.4) solution is more robust than guar gum solution. Ultimately, the G′ and G″ can be further improved by cross-linking, leading to a steady associator in P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 (hereinafter denoted as PASG/[Zr]) solution. The oscillation frequency results of P(AM10-AMS2-AEG1.4) in the absence or presence of [Zr] and guar gum are given in Figure 3e. The steep decreases in G′ and G″ during the shearing process of guar gum with decreasing frequency are noticed, in which a section point is found at a frequency of 0.2 Hz. In contrast, the G′ of PASG/[Zr] solution is higher than the corresponding G″ from beginning to end, in which the modulus values are still higher than the corresponding values of guar gum. Moreover, the G′ and G″ of PASG/[Zr] solution with a gentle decrease are higher than those of P(AM10-AMS2-AEG1.4) solution due to the cross-linking sites. Meanwhile, the loss factor (tanδ) of PASG/[Zr] solution is lower than those of P(AM10-AMS2-AEG1.4) and guar gum solutions, since the multiple cross-linking networks can inhibit the segmental motion of polymers.
The PASG/[Zr] solution with the CAC of 0.15 wt% at 20 °C exhibits a fast disassociation–reassociation process, leading to a steadier associator in PASG/[Zr] solution compared to the guar gum solution.

3.2.2. Effect of Temperature on the Rheological Performance of Fracture Fluid

The associating affinity of PASG/[Zr] solution is dominated by the environmental temperature and oscillating strength. Fracture fluid with strong affinity exhibits excellent performances under extreme conditions. In order to evaluate the affinity of the PASG/[Zr] solution, the relationships between ηapp of PASG/[Zr] solution (0.27 wt%) and γ ˙ at various temperatures are recorded in Figure 4a. According to the linear relationships, the effect of temperature on K plotted in Figure 4a is in accordance with the Arrhenius Function shown in Equation (4).
η a p p = A e E a R T
where A, R, and Ea indicate the pre-exponential factor, universal gas constant, and disassociating active energy during the damage process. Ea is 304.96 kJ·mol−1 according to Equation (4).
The affinity of the associating solution is also dependent on the dosage of [Zr], in which a higher f[Zr] can lead to a higher resistance to high temperature. The effect of f[Zr] on the temperature resistance of PASG/[Zr] solution is given in Figure 4c,d to describe the importance of f[Zr]. The plots of ηapp and oscillating temperature (Tos) can be fitted by the Boltzmann Function according to Equation (5).
η a p p = η a p p , m a x 1 + e T o s T o s , d d T o s
where ηapp,max indicates the maximum ηapp in the process. Tos,d and dTos indicate the disassociating temperature and the disassociating rate at the disassociating temperature in the process. Tos,d according to Equation (5) increases from 38 to 101 °C with the increase in f[Zr] from 0.1% to 0.35%, indicating that the temperature resistance can be remarkably improved by increments in f[Zr] (Figure 4c–d).
On the basis of the above analysis, the fracture fluid with various f[Zr] values can be utilized under various conditions. The disassociation of P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution with f[Zr] values of 0.25%, 0.35%, and 0.45% at 60, 90, and 120 °C, respectively, is shown in Figure S8. The associators in the P(AM10-AMS2-AEG1.4)/[Zr]0.25/TBAC0.1 solution are totally broken after shearing for 8 h at 60 °C, whereas the associators in P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 and P(AM10-AMS2-AEG1.4)/[Zr]0.45/TBAC0.1 solutions are broken after shearing for 6 h at 90 °C and shearing for 4 h at 120 °C. This indicates that the f[Zr] of fracture fluid can be adjusted according to practical application.
The remarkable temperature resistance of PASG/[Zr] solution is due to the high disassociating active energy (Ea) of 304.96 kJ·mol−1. Meanwhile, the higher f[Zr] results in a higher affinity of the associating solution and higher rheological performance at high temperature.

3.2.3. Thixotropy and Self-Recovery Ability of Fracture Fluid

The thixotropic and self-recovery properties of the associative networks are critical for practical applications, as they rely on the dynamic reformation of cross-linked structures under shear. In the PASG/[Zr] system, abundant ionic hydrogen bonds (iHBs) between sulfonic and amine groups enable a rapid response to shear, whereas the Zr4+ coordination responds more slowly. To clarify the role of iHBs in balancing association and dissociation, density functional theory (DFT) analysis was performed (Figure 4b). The average bond length of N···H or O···H in these iHBs is 1.46 Å, significantly shorter than that of conventional hydrogen bonds. Moreover, the total bond energy reaches 92.6 kJ·mol−1, exceeding typical hydrogen bond strengths. These results indicate that the iHBs are both strong and highly dynamic, facilitating rapid breakdown and reformation under shear. Consequently, the presence of active iHBs supports the construction of stable yet responsive associative networks, ensuring efficient self-recovery and robust thixotropic behavior.
To study the recovery process of associators in PASG/[Zr] solution (0.27 wt%), the recovery rate of ηapp (η/η0) of PASG/[Zr] solution was recorded in Figure 4c,d under strong shearing ( γ ˙ = 1000 s−1) and weak shearing ( γ ˙ = 500 s−1) for increasing duration. The damage of strong shearing is always stronger than that of weak shearing, leading to low η/η0 in each of the three groups. However, the increment in concentration results in an increase in η/η0 from 83% to 92% after 100 min even under strong shearing. Considering that f[Zr] is settled as 0.35% in solution, the increases in η/η0 are mainly attributed to the greater amount of iHBs in concentrated solution. In addition, Figure S9 shows that the interfacial tension can be recovered under shearing conditions, which is also not sensitive to temperature.
The thermal relaxing characteristic endows the polymer network with the ability of rapidly balancing disassociation and reassociation. Identical to the above analysis, the contribution of [Zr] to the recovery ability can also be quantified by investigating the effect of f[Zr] on the shearing balance of P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution given in Figure 4e. The plots of ηapp and tos can also be fitted by the Boltzmann Function according to Equation (5).
η a p p = η a p p , m a x 1 + e t o s t o s , d d t o s
where tos,d and dtos indicate the relaxing time and the disassociating rate at the relaxing time in the process. tos,d according to Equation (5) increases from 11 to 130 min with the increase in f[Zr] from 0.1% to 0.35%. This indicates that the shearing resistance can be remarkably improved by increments in f[Zr], whereas the recovery ability is inhibited in the process. Therefore, the robustness of the polymer associator is decided by f[Zr], but the recovery and thixotropy ability are decided by iHBs. The multiple networks have synergetic effects, resulting in different applications under various conditions.
In addition, the thixotropy of PASG/[Zr] solution (0.27 wt%) can be evaluated by the cyclic thixotropic test (Figure 4f). The maximum stress of the PASG/[Zr] solution (0.27 wt%) of 22.96 Pa is higher than that of guar gum solution (0.27 wt%) of 18.18 in the γ ˙ range of 100~500 s−1. Meanwhile, the area of the thixotropic ring (hysteresis) of PASG/[Zr] solution (0.27 wt%) is 14.08 kPa·s−1 in a γ ˙ range of 100~500 s−1. For comparison, the value of guar gum solution (0.27 wt%) is 10.89 Pa·s−1, which is smaller than that of the PASG/[Zr] solution. This confirms that the thixotropic ability of PASG/[Zr] solution fracture fluid is superior to traditional guar gum fracture fluid.
The PASG/[Zr] solution, with its stable network structure, exhibits excellent disassociation–reassociation behavior under shear, owing to its branched topological architecture, intrinsic hydrogen bonds (iHBs), and Zr4+ coordination. The Zr4+ coordination provides strong associative strength, while the iHBs contribute to rapid recovery and thixotropic properties. By adjusting the formulation, the operational parameters of the fracturing fluid can be finely tuned, making it suitable for application under extreme conditions.
The abundant iHBs between sulfonic and amine groups in the PASG/[Zr] system respond rapidly during thixotropic processes, promoting stable interactions and fast reconstruction of multiple networks through the synergistic effect of iHBs and Zr4+ coordination. Furthermore, increasing the Zr4+ content enhances the network reconstruction activity and improves the frequency resistance of the P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution. The outstanding thixotropic performance of the PASG/[Zr] solution—characterized by rapid recovery and low hysteresis—is crucial for thickening agents to prevent circulation loss and extend thickening time.

3.3. Fracturing Fluid as Thickening Agent Protecting Cores

3.3.1. Sand-Carrying Ability of PASG/[Zr]-Based Fracturing Fluid

To protect the stratum from the corrosion by natural fluid, a kind of non-Newtonian fluid with high compatibility and good impact resistance should be applied as the thickening reagent for shale gas extraction. To reduce the damage caused by precipitated sands, the sand-carrying ability of the above fracture fluid is investigated. The increases in settling volume (Vs) of sand from the surface at various settling durations (ts) in the static mixture of PASG/[Zr] solution (0.27 wt%) and sand with various particle sizes at 20 °C are plotted in Figure 5a. For comparison, the increases in Vs in guar gum are also given in Figure 5a under corresponding conditions, with larger particle sizes and faster settling rates. The Vs of sand with constant particle size in PASG/[Zr] solution is always lower than that in guar gum solution. The Vs-ts plots can be fitted by the Boltzmann Function according to Equation (6).
V s = V s , m a x 1 + e t s t s , s e m i d t s
where Vs,max indicates the maximum Vs in the process. ts,semi and dts indicate the semi-settling time and the settling rate at the semi-settling time in the process. The ts,semi of sand in the PASG/[Zr] solution decreases from 28.24 to 24.90 h with the increase in average particle size from 30 to 70 mesh, which is always higher than each ts,semi of the sand in guar gum solution (Figure 5b).
Meanwhile, the effect of temperature is also noticed, due to the possible disassociation and sand aggregation. The sand-carrying ability of the above fracture fluid at various temperatures is analyzed. The increase in Vs with increasing ts in the static mixture of PASG/[Zr] solution (0.27 wt%) and sand (30 mesh) at various temperatures is plotted in Figure 5a. For comparison, the increasing Vs in guar gum is also given in Figure 5a under corresponding conditions, with higher temperature and faster settling rate. Identical to the above analysis, the Vs of sand in PASG/[Zr] solution is always lower than that in guar gum solution at the same temperature. The Vs-ts plots can also be fitted by the Boltzmann Function according to Equation (6). The ts,semi of sand in PASG/[Zr] solution decreases from 19.54 to 8.74 h with the increase in temperature from 30 to 90 °C, which is always higher than each ts,semi of sand in the guar gum solution (Figure 5b).
The sand-carrying ability of flowing PASG/[Zr] solution can be further studied by carrying experiments in a horizontal pipe (dpipe = 2.5 cm, Scheme 2). The pressure gradient (∇×P) and flow velocity (v) of the mixture with solute and sand concentrations (Csolute, Csand) of 0.27 and 30 wt%, respectively, were recorded with increasing experimental duration at 20 °C, as shown in Figure 5c. Initially, the sand suspended in the fluid can smoothly move through the pipe in 2 min with a slight change in ∇×P and decrease in v. ∇×P sharply increases from 0.1 to 2.9 kPa·s−1 and v decreases from 0.24 to 0.02 m·s−1 in 2~13 min with the sand gradually settling on the bottom of the pipe, due to the reduction in actual internal area. Subsequently, the sand begins to be recarried in the fluid because of the high ∇×P, leading to an enhancement of the fluid. When v reaches 0.26 m·s−1 at 24 min, the sand is totally suspended in the fluid without any settlement on the bottom. The v of fluid beginning to settle is denoted as the settling velocity (vc,settle), and the v of sand totally suspended in fluid is denoted as the suspending velocity (vc,suspend). The vc,settle and vc,suspend of the above experiment are 0.24 and 0.26 m·s−1.
In order to clarify the dynamic sand-carrying ability of PASG/[Zr] fracturing fluid during the flowing process at 20 °C, the relationships between the diameter of the pipe (dpipe) or Csand and vc,settle or vc,suspend of PASG/[Zr] fluid are given in Figure 5d. For comparison, the corresponding relationships of guar gum fluid are also given in Figure 5d. vc,suspend is always higher than vc,settle as the driving force of sand suspension is always higher than that of sand settlement. All of the velocity values reveal increases with an increase in dpipe or Csand, according to the Bernoulli Principle. Significantly, the vc,settle or vc,suspend of guar gum fluid is always higher than that of PASG/[Zr] fracturing fluid, whether the variable is dpipe or Csand, indicating that the sand-carrying ability of PASG/[Zr] fracturing fluid is much better than that of guar gum fluid during the flowing process.
Consequently, the PASG/[Zr] fracturing fluid exhibits a good sand-carrying ability due to its great rheological performance and interfacial activity, leading to low Vs during the static sand-carrying test and low vc,suspend during the dynamic sand-carrying test.

3.3.2. Filtration System Based on PASG/[Zr] Fracturing Fluid

The filtration performance is a critical factor in evaluating the applicability of fracturing fluids, as it reflects the degree of core damage caused by fluid leakage. To quantify the filtration loss of rock with the fracture fluid through the core, the filtration loss volume (Vf) of the artificial core was measured at 20 °C during the flowing process with various γ ˙ and a constant filtration differential pressure (Δpf) of 3.5 MPa by using the filtration experimental device shown in Figure S10. The plots of Vf vs. t0.5 of PASG/[Zr] and guar gum solutions with various γ ˙ are given in Figure 6a. The filtration coefficient (cf) and core damage rate can be calculated according to Equation (8).
c f = 0.005 × m A
where m indicates the slope of the linear stage of the Vf vs. t0.5 plot. A indicates the cross-sectional area of the flowing pipe. cf reveals a low increase from 2 × 10−4 to 4.2 × 10−4 m·min−0.5 by using PASG/[Zr] fracturing fluid, which is lower than the increase from 8 × 10−4 to 16 × 10−4 m·min−0.5 by using guar gum fracturing fluid. On the other hand, the increase cf from 2.8 × 10−4 to 30.6 × 10−4 m·min−0.5 is gentler than the increase in cf from 7.6 × 10−4 to 59.4 × 10−4 m·min−0.5 with an increment in core permeability from 5 to 96 mD. This indicates that the PASG/[Zr] fracturing fluid exhibits better filtration performance than guar gum fracturing fluid due to the good compatibility and rheological performance of PASG/[Zr] fracturing fluid. In addition, the core permeability has a more prominent effect on cf than γ ˙ . Δpf is another vital practical parameter in the field of stratum protection, which determines the filtration and damaging process of cores (Figure S12). cf slightly increases from 2.2 × 10−4 to 2.3 × 10−4 m·min−0.5 with an increment in Δpf from 3.5 to 11.0 MPa by using PASG/[Zr] fracturing fluid, and Rd also slightly increases from 15.4% to 21.1% in the process. For comparison, the increase in cf from 8.2 × 10−4 to 8.6 × 10−4 m·min−0.5 and increase in Rd from 30.5% to 37.7% in the filtration system with guar gum fracturing fluid can lead to severe damage to cores. It also indicates that the PASG/[Zr] fracture fluid can also serve well in drilling operations and as cement slurry because of its low filtration and active thixotropy, which also exhibits a high displacement efficiency of 68.74%, calculated according to references [6,7,8,9].
In addition, the fracture fluid with various f[Zr] can exhibit exceptional filtration performance at various temperatures. To illustrate the filtration performance of fracture fluid at various temperatures, the relationships between the change in ηappηapp) of P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution (0.27%) with an f[Zr] of 0.25%, 0.35%, and 0.45% at 60, 90, and 120 °C, respectively, and t0.5 are shown in Figure S11. The Δηapp in P(AM10-AMS2-AEG1.4)/[Zr]0.25/TBAC0.1 solution increases from 8.5 to 17.8 mPa·s at 60 °C from 9 to 36 min, that of P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 increases from 10.2 to 19.7 mPa·s at 90 °C from 9 to 36 min, and that of P(AM10-AMS2-AEG1.4)/[Zr]0.45/TBAC0.1 solutions increases from 10.9 to 20.3 mPa·s at 120 °C from 9 to 36 min. A high f[Zr] leads to high-temperature resistance during the filtration process, confirming that the f[Zr] of fracture fluid mediates the filtration performance for practical application.
LF-NMR analysis was performed on artificial cores treated with different fluids to evaluate formation damage under various conditions. Two groups of identical artificial cores were first saturated with standard brine (2.0% KCl + 5.5% NaCl + 0.45% MgCl2 + 0.55% CaCl2) to measure initial water permeability. After 100 min of exposure to the fracturing fluids, the cores were subjected to LF-NMR analysis (Figure 6c). The transverse relaxation time (T2) reflects the mobility of hydrogen atoms in hydrogen-bearing components within the filtration system: a longer T2 indicates higher mobility and stronger associative or absorptive interactions [43,44,45,46]. In the T2 spectra of cores saturated with standard brine, three signal groups were observed at 2.42 ms, 13.92 ms, and 177.7 ms, corresponding to small pores, large pores, and free water, respectively, based on previous studies [47,48,49,50]. After treatment with the PASG/[Zr] fracturing fluid, the T2 values were detected at 2.96 ms, 13.99 ms, and 177.8 ms for small pores, large pores, and free water, respectively (Figure 6d). The integral area ratios of these three signals showed negligible changes. In contrast, after exposure to guar gum fracturing fluid, the T2 values shifted to 2.96 ms, 21.23 ms, and 239.1 ms (Figure 6d), indicating an increase in pore size and a reduction in core–fluid affinity. The integral ratio of the small-pore signal increased from 15.1% to 18.9%, while those of large pores and free water decreased from 21.4% and 63.5% to 20.6% and 60.5%, respectively. These results suggest that the guar gum fracturing fluid causes structural loosening within the core and impairs fluid–core interactions, ultimately leading to formation damage.
The PASG/[Zr] solution with admirable viscosity, rheological characteristics, and affinity between the core and fluid reduces the damage from high pressure and leads to low filtration loss of the artificial cores during the filtration process. The filtration performance of PASG/[Zr] fracture fluid is better than that of traditional fluid (guar gum), which can be used as a novel thickening agent for stratum protection and shale gas extraction.

4. Conclusions

A novel fracturing fluid based on the graft copolymer P(AM-AMS-AEG), cross-linked by an organic zirconium reagent, was successfully developed via free-radical polymerization followed by in situ cross-linking. The P(AM-AMS-AEG) polymer, with a molecular weight (Mw) lower than that of typical guar gum (2000 kg·mol−1), was readily synthesized in aqueous solution using a Na2SO3-K2S2O8 initiating system at 80 °C for 6 h. Under these conditions, the conversion rate and number-average molecular weight (Mn) of P(AM10-AMS2-AEG1.4) reached 96% and 250 kg·mol−1, respectively. The P(AM-AMS-AEG)/[Zr]/TBAC/[(NH4)2S2O8] fracturing fluid was subsequently prepared through in situ cross-linking by adding zirconium reagent at 80 °C for 1200 to 300 s, with the zirconium mass fraction (f[Zr]) varying from 0.1% to 0.45%. The effects of the feed ratios of f[Zr], the breaker (fb), and salt (fs) on rheological properties were systematically investigated. Among the formulations, the P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 aqueous solution demonstrated the most comprehensive performance. At concentrations exceeding 0.15 wt% and 20 °C, the PASG/[Zr] solution exhibited faster and more stable disassociation–reassociation behavior compared to guar gum. Its high activation energy (Ea) of 304.96 kJ·mol−1 contributes to exceptional temperature resistance. The abundant intrinsic hydrogen bonds (iHBs) between sulfonic and amine groups enable a rapid response during thixotropic processes, promoting stable interactions and rapid reconstruction of multiple networks through synergistic effects with Zr4+ coordination. Increasing f[Zr] enhanced the associative affinity and high-temperature rheological performance. Notably, the cross-linked structures in the P(AM10-AMS2-AEG1.4)/[Zr]0.45/TBAC0.1 solution remained unbroken even after 4 h of shearing at 120 °C. Higher f[Zr] also improved the network reconstitution activity and frequency resistance, resulting in a higher hysteresis area of 14.08 kPa·s−1 for the PASG/[Zr] solution. Owing to its pseudoplasticity, thixotropy, and good sand compatibility, the PASG/[Zr] fluid effectively minimizes formation damage under high pressure and shows low filtration loss in artificial cores. It also exhibits excellent sand-carrying capacity, attributed to its superior rheological properties and interfacial activity. In static sand-suspension tests, the settling velocity (Vs) increased only slowly between 30 and 90 °C, while in dynamic tests conducted in a horizontal pipe (d = 2.5 cm), the critical settling velocity (vc,suspend) was as low as 0.26 m·s−1. The filtration coefficient (cf) of the PASG/[Zr] fracturing fluid ranged from 2.2 × 10−4 to 2.5 × 10−4 m·min−0.5, showing low sensitivity to shear rate ( γ ˙ ), core permeability, and differential pressure (Δpf). This indicates significantly improved filtration performance compared to conventional guar gum-based fluids. With its combination of robust rheological properties, minimal formation damage, and effective suspension characteristics, the PASG/[Zr] fracturing fluid demonstrates strong potential as a novel thickening agent for applications in reservoir protection and shale gas extraction.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr13103032/s1, Figure S1: Relationship between ηapp and fM. (Conditions: AM:AMS:AEG:I = 10:2:1.4:0.18, Tp = 80 °C, tp = 6 h); Figure S2: Relationship between ηapp and fI. (Conditions: AM:AMS:AEG = 10:2:1.4, fM = 30%, Tp = 80 °C, tp = 6 h); Figure S3: The (a) conv. and (b) ηapp of 5 groups polymers with various AM:AMS:AEG:TBAC including 9:1:2.7:0.2, 9:1:1.7:0.2, 9:2:2.4:0.1, 10:2:0.7:0.2, 10:2:1.4:0.1. (Conditions: fM = 30%, Tp = 80 °C, tp = 6 h); Figure S4: Relationship between txl or ηapp and f[Zr]. (Conditions: AM:AMS:AEG:I = 10:2:1.4:0.18, fM = 30%, Tp = 80 °C, tp = 6 h, Txl = 80 °C); Figure S5: Relationship between γs, γi or ηapp and f[Zr]. (Conditions: AM:AMS:AEG:I = 10:2:1.4:0.18, fM = 30%, f[Zr] = 0.35%); Figure S6: The lnτ-ln γ ˙ plots of PASG/[Zr] solution at 20 °C; Figure S7: Relationship between tsol and fAEG of PASG/[Zr] solution (0.27 wt%) at 20 °C; Figure S8: Reduction of ηapp of P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution (0.27 wt%) with various f[Zr] of 0.25%, 0.35% and 0.45% at 60, 90 and 120 °C, respectively; Figure S9: The recovering process of γi of PASG/[Zr] solution with under strong (1000 s−1) or weak shearing (500 s−1) at various temperature; Figure S10: Schematical illustration of the filtrating sand-carrying experimental device; Figure S11: Schematical illustration of the filter loss meter in the filtrating sand-carrying experimental device; Figure S12: Effect of Δpf on cf and Rd in the fracturing experiment of PASG/[Zr] solution (0.27 wt%) and artificial core at 20 °C; Figure S13: Incremental of Δηapp of P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution (0.27 wt%) with various f[Zr] of 0.25%, 0.35% and 0.45% at 60, 90 and 120 °C, respectively; Table S1: Physical parameters of P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1 (PASG/[Zr]) system.

Author Contributions

K.W., Conceptualization, Investigation, Methodology, Writing—Original Draft Preparation, Funding Acquisition. C.G., Investigation, Methodology, Writing—Original Draft Preparation. Q.G., Formal Analysis, Writing—Original Draft Preparation. G.L., Data Curation, Writing—Original Draft Preparation. C.Z., Data Curation, Writing—Review and Editing. T.J., Investigation, Writing—Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

Chongqing’s Talent Program—Innovative Leading Talent Project (Project Number: CQYC20220304204); Key Project of the National Natural Science Foundation of China (Joint Fund for Regional Innovation and Development) (Grant No. U24A2090).

Data Availability Statement

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

Conflicts of Interest

Author Chenye Guo and Gen Li were employed by the company Southwest Branch of China National Coal Group Co., Ltd., Chongqing 400023, China. Author Qisen Gong, Cuilan Zhang and Teng Jiang were employed by the company Chongqing Energy Investment Group Technology Co., Ltd., Chongqing 400061, China. 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.

References

  1. Liang, H.; Qi, Z.; Wang, S.; Huang, X.; Yan, W.; Yuan, Y.; Li, Z. Adsorption Models for Shale Gas: A Mini-Review. Energy Fuels 2022, 36, 12946–12960. [Google Scholar] [CrossRef]
  2. Striolo, A.; Cole, D.R. Understanding Shale Gas: Recent Progress and Remaining Challenges. Energy Fuels 2017, 31, 10300–10310. [Google Scholar] [CrossRef]
  3. Sun, C.; Nie, H.; Dang, W.; Chen, Q.; Zhang, G.; Li, W.; Lu, Z. Shale Gas Exploration and Development in China: Current Status, Geological Challenges, and Future Directions. Energy Fuels 2021, 35, 6359–6379. [Google Scholar] [CrossRef]
  4. You, L.; Zhang, N.; Kang, Y.; Xu, J.; Cheng, Q.; Zhou, Y. Zero Flowback Rate of Hydraulic Fracturing Fluid in Shale Gas Reservoirs: Concept, Feasibility, and Significance. Energy Fuels 2021, 35, 5671–5682. [Google Scholar] [CrossRef]
  5. Da, Q.; Yao, C.; Zhang, X.; Li, L.; Lei, G. Reservoir Damage Induced by Water-Based Fracturing Fluids in Tight Reservoirs: A Review of Formation Mechanisms and Treatment Methods. Energy Fuels 2024, 38, 18093–18115. [Google Scholar] [CrossRef]
  6. Yang, J.; Dong, T.; Yi, J.; Jiang, G. Development of Multiple Crosslinked Polymers and Its Application in Synthetic-Based Drilling Fluids. Gels 2024, 10, 120. [Google Scholar] [CrossRef]
  7. Huang, J.; Chen, J.; Lv, L.; Yu, Z.; Yao, W.; Cheng, H.; Niu, W.; Wang, J.; Zhang, J.; Qi, H. Design and Verification of a Wearable Micro-Capacitance Test System for POC Biosensing. IEEE Trans. Instrum. Meas. 2025, 74, 1–11. [Google Scholar] [CrossRef]
  8. Li, J.; Wen, M.; Yang, J.; Liu, Y.; Jiang, Z.; Chen, J. Synthesis and analysis of magnetic nanoparticles within foam matrix for foam drainage gas production. Geoenergy Sci. Eng. 2024, 238, 212887. [Google Scholar] [CrossRef]
  9. He, S.; Ding, L.; Xiong, Z.; Spicer, R.; Farnsworth, A.; Valdes, P.; Wang, C.; Cai, F.; Wang, H.; Sun, Y.; et al. A distinctive Eocene Asian monsoon and modern biodiversity resulted from the rise of eastern Tibet. Sci. Bull. 2022, 21, 2245–2258. [Google Scholar] [CrossRef]
  10. Al-Hajri, S.; Negash, B.M.; Rahman, M.M.; Haroun, M.; Al-Shami, T.M. Perspective Review of Polymers as Additives in Water-Based Fracturing Fluids. ACS Omega 2022, 7, 7431–7443. [Google Scholar] [CrossRef] [PubMed]
  11. Davoodi, S.; Al-Shargabi, M.; Wood, D.A.; Rukavishnikov, V.S.; Minaev, K.M. Synthetic polymers: A review of applications in drilling fluids. Pet. Sci. 2024, 21, 475–518. [Google Scholar] [CrossRef]
  12. Yang, S.; Yu, W.; Zhao, M.; Ding, F.; Zhang, Y. A Review of Weak Gel Fracturing Fluids for Deep Shale Gas Reservoirs. Gels 2024, 10, 345. [Google Scholar] [CrossRef]
  13. Ahmad, H.M.; Iqbal, T.; Kamal, M.S.; Al-Harthi, M.A. Influence of Hydrophobically Modified Polymer and Titania Nanoparticles on Shale Hydration and Swelling Properties. Energy Fuels 2020, 34, 16456–16468. [Google Scholar] [CrossRef]
  14. Movahedi, H.; Schiefler, A.A.; Dwivedi, S.; Sørensen, H.O.; Poulsen, H.-F.; Bovet, N. Novel Temperature-Responsive Organic Polymer Gel for Efficient Abandonment of Aged Porous Chalk Formations Characterized by X-ray Computed Tomography. Energy Fuels 2024, 38, 19478–19493. [Google Scholar] [CrossRef]
  15. Protsak, I.S.; Morozov, Y.M. Fundamentals and advances in stimuli-responsive hydrogels and their applications: A review. Gels 2025, 11, 30. [Google Scholar] [CrossRef] [PubMed]
  16. Zhu, D.; Bai, B.; Hou, J. Polymer gel systems for water management in high-temperature petroleum reservoirs: A chemical review. Energy Fuels 2017, 17, 13063–13087. [Google Scholar] [CrossRef]
  17. de Leon, A.C.C.; da Silva, Í.G.M.; Pangilinan, K.D.; Chen, Q.; Caldona, E.B.; Advincula, R.C. High performance polymers for oil and gas applications. React. Funct. Polym. 2021, 162, e104878. [Google Scholar] [CrossRef]
  18. Liu, X.; Liu, K.; Gou, S.; Liang, L.; Luo, C.; Guo, Q. Water-Soluble Acrylamide Sulfonate Copolymer for Inhibiting Shale Hydration. Ind. Eng. Chem. Res. 2014, 53, 2903–2910. [Google Scholar] [CrossRef]
  19. Bagaria, H.G.; Yoon, K.Y.; Neilson, B.M.; Cheng, V.; Lee, J.H.; Worthen, A.J.; Xue, Z.; Huh, C.; Bryant, S.L.; Bielawski, C.W.; et al. Stabilization of Iron Oxide Nanoparticles in High Sodium and Calcium Brine at High Temperatures with Adsorbed Sulfonated Copolymers. Langmuir 2013, 29, 3195–3206. [Google Scholar] [CrossRef]
  20. Wang, J.; Ran, D.; Lu, H.; Zhang, J.; Wu, Y. Realizing the Efficient Dissolution of Drag Reducer upon pH-Induced Demulsification of Inverse Polymer Emulsion. Langmuir 2024, 40, 15869–15876. [Google Scholar] [CrossRef]
  21. Yang, D.; Liu, C.; Piao, H.; Quan, P.; Fang, L. Enhanced Drug Loading in the Drug-in-Adhesive Transdermal Patch Utilizing a Drug-Ionic Liquid Strategy: Insight into the Role of Ionic Hydrogen Bonding. Mol. Pharm. 2021, 18, 1157–1166. [Google Scholar] [CrossRef]
  22. Yonamine, Y.; Yoshimatsu, K.; Lee, S.-H.; Hoshino, Y.; Okahata, Y.; Shea, K.J. Polymer Nanoparticle−Protein Interface. Evaluation of the Contribution of Positively Charged Functional Groups to Protein Affinity. ACS Appl. Mater. Interfaces 2013, 5, 374–379. [Google Scholar] [CrossRef]
  23. Nguyen, C.A.; Argun, A.A.; Hammond, P.T.; Lu, X.; Lee, P.S. Layer-by-Layer Assembled Solid Polymer Electrolyte for Electrochromic Devices. Chem. Mater. 2011, 23, 2142–2149. [Google Scholar] [CrossRef]
  24. Lei, S.; Sun, J.; Lv, K.; Zhang, Q.; Yang, J. Types and Performances of Polymer Gels for Oil-Gas Drilling and Production: A Review. Gels 2022, 8, 386. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, J.; Zhou, X.; Hu, Q.; Zhou, K.; Zhang, Y.; Dong, S.; Zhao, G.; Zhang, S. Concentration-induced spontaneous polymerization of protic ionic liquids for efficient in situ adhesion. Nat. Comm. 2024, 15, e4265. [Google Scholar] [CrossRef] [PubMed]
  26. Abdel-Azeim, S. Phase Behavior and Interfacial Properties of Salt-Tolerant Polymers: Insights from Molecular Dynamics Simulations. ACS Appl. Polym. Mater. 2021, 3, 6488–6501. [Google Scholar] [CrossRef]
  27. Si, S.; Liu, Y.; Dai, C.; Zou, C.; Zhuo, L.; Li, X.; Yang, N.; Ma, D. Salt Resistance Mechanism and Oil Displacement Performance Evaluation of a New Zwitterionic Polymer. Energy Fuels 2025, 39, 13443–13453. [Google Scholar] [CrossRef]
  28. Wang, S.; Sun, Z.; Zhu, Q.; Cao, X.; Feng, Y.; Yin, H. In Situ Generated Hydrogels Exhibiting Simultaneous High-Temperature and High-Salinity Resistance for Deep Hydrocarbon Reservoir Exploitation. Ind. Eng. Chem. Res. 2024, 63, 18263–18278. [Google Scholar] [CrossRef]
  29. Chen, Y.; Li, J.; Song, Q.; Bu, C.; Ye, K.; He, F.; Lv, Z.; Chen, X. Research on Flow Patterns of Two-Phase Fracturing Fluid in Hydraulic Fracture Considering the Fluid Leak-off. ACS Omega 2024, 9, 2432–2442. [Google Scholar] [CrossRef]
  30. Huang, C.; Mu, L.; Gong, X. Development and Characterization of Environmentally Responsive Thickening Agents for Fracturing Fluids in Shale Gas Reservoir Stimulation. Processes 2025, 13, 1253. [Google Scholar] [CrossRef]
  31. Yu, Y.; Zhang, H.; Xu, P.; Zhang, X.; Wang, H.; Hu, M.; Guo, J. Effect of Ultra-High Temperature Degradation on the Physical Properties and Chemical Structure of an AMPS-Based Copolymer Oil-Well Cement Additive PADIM in Aqueous Solution. Polymers 2025, 17, 591. [Google Scholar] [CrossRef] [PubMed]
  32. Srivastava, A.; Mandal, P.; Kumar, R. Solid state thermal degradation behaviour of graft copolymers of carboxymethyl cellulose with vinyl monomers. Int. J. Biol. Macromol. 2016, 87, 357–365. [Google Scholar] [CrossRef]
  33. Liu, X.; Wang, A.; Wang, C.; Qu, J.; Wen, Y.; Chen, B.; Wang, Z.; Wu, B.; Yuan, Z.; Wei, B. Preparation and Performance of Salt Tolerance and Thermal Stability Cellulose Nanofibril Hydrogels and Their Application in Drilling Engineering. Pap. Biomater. 2019, 4, 10–19. [Google Scholar] [CrossRef]
  34. Wang, S.; Xiong, P.; Liu, X.; Meng, F.; Ma, Q.; Song, C.; Zhang, Q.; Che, Y. A Hydroxypropyl Methyl Cellulose-Based Graft Copolymer with Excellent Thermothickening and Anti-salt Ability for Enhanced Oil Recovery. Energy Fuels 2022, 36, 2488–2502. [Google Scholar] [CrossRef]
  35. Balding, P.; Li, M.; Wu, Q.; Volkovinsky, R.; Russo, P. Cellulose Nanocrystal–Polyelectrolyte Hybrids for Bentonite Water-Based Drilling Fluids. ACS Appl. Bio Mater. 2020, 3, 3015–3027. [Google Scholar] [CrossRef]
  36. Hao, H.; Yao, E.; Pan, L.; Chen, R.; Wang, Y.; Xiao, H. Exploring heterogeneous drivers and barriers in MaaS bundle subscriptions based on the willingness to shift to MaaS in one-trip scenarios. Transp. Res. Part A Policy Pract. 2025, 199, 104525. [Google Scholar] [CrossRef]
  37. Bahamdan, A.; Daly, W.H. Poly(oxyalkylene) grafts to guar gum with applications in hydraulic fracturing fluids. Polym. Adv. Technol. 2006, 17, 679–681. [Google Scholar] [CrossRef]
  38. Lu, W.; Wang, J.; Wang, T.; Zhang, K.; Jiang, X.; Zhao, H. Visual style prompt learning using diffusion models for blind face restoration. Pattern Recognit. 2025, 161, 111312. [Google Scholar] [CrossRef]
  39. Shi, S.L.; Sun, J.S.; Mu, S.B.; Lv, K.H.; Liu, J.P.; Bai, Y.R.; Wang, J.T.; Huang, X.B.; Jin, J.F.; Li, J. Effect of Chemical Composition of Metal–Organic Crosslinker on the Properties of Fracturing Fluid in High-Temperature Reservoir. Molecules 2024, 29, 2798. [Google Scholar] [CrossRef]
  40. Xu, Z.Z.; Zhao, M.W.; Sun, N.; Meng, X.J.; Yang, Z.T.; Xie, Y.X.; Ding, F.; Dong, Y.B.; Gao, M.W.; Wu, Y.N.; et al. Delayed Crosslinking Gel Fracturing Fluid with Dually Crosslinked Polymer Network for Ultra-Deep Reservoir: Performance and Delayed Crosslinking Mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2025, 708, e135967. [Google Scholar] [CrossRef]
  41. Zhou, Y.; Xia, H.; Yu, D.; Cheng, J.; Li, J. Outlier detection method based on high-density iteration. Inf. Sci. 2024, 662, 120286. [Google Scholar] [CrossRef]
  42. Wang, Y.-C.; Zha, H.; Cheng, P.-Y.; Zhang, S.; Liu, X.-W.; Wu, Y.-X. Vanadium(V) Complexes Containing Unsymmetrical N-Heterocyclic Carbene Ligands: Highly Efficient Synthesis and Catalytic Behavior towards Ethylene/Propylene Copolymerization. Chin. J. Polym. Sci. 2024, 42, 32–41. [Google Scholar] [CrossRef]
  43. Gao, Y.; Li, A.; Chen, J.; Cui, Z.; Wu, Y. Quaternized Sodium Alginate-g-Ethyl-Oxazoline Copolymer Brushes and Their Supramolecular Networks via Hydrogen Bonding. ACS Biomater. Sci. Eng. 2022, 8, 3424–3437. [Google Scholar] [CrossRef]
  44. Li, A.; Li, J.-L.; Zhang, J.-M.; Ma, J.-Y.; Wu, Y.-X. In-situ Synthesis of Chemically Stable Hybrid Networks of Poly(thioctic acid) with Fe3+ via Controlled/living Cationic Ring-opening Polymerization. Macormol. Rapid. Commun. 2025, 4, e2401115. [Google Scholar] [CrossRef]
  45. Qi, H.; Hu, Z.; Yang, Z.; Zhang, J.; Wu, J.J.; Cheng, C.; Wang, C.; Zheng, L. Capacitive aptasensor coupled with microfluidic enrichment for real-time detection of trace SARS-CoV-2 nucleocapsid protein. Anal. Chem. 2022, 6, 2812–2819. [Google Scholar] [CrossRef]
  46. Li, A.; Zhang, J.; Yi, Q.; Wang, Y.; Wu, Y. Optical Performance of Hybrid Co-Networks of Poly(thioctic Acid) Coordinated with Fe3+: Optical Focusing and Photothermal Conversion. Adv. Opt. Mater. 2025, 13, e01148. [Google Scholar] [CrossRef]
  47. Fang, Y.; Yue, T.; Zhang, S.L.; Liu, J.; Zhang, L. Molecular Dynamics Simulations of Self-Healing Topological Copolymers with a Comblike Structure. Macromolecules 2021, 54, 1095–1105. [Google Scholar] [CrossRef]
  48. Sun, L.; Shi, W.; Tian, X.; Li, J.; Zhao, B.; Wang, S.; Tan, J. A plane stress measurement method for CFRP material based on array LCR waves. NDT E Int. 2025, 151, 103318. [Google Scholar] [CrossRef]
  49. Zhang, M.; Li, X. Understanding the relationship between coopetition and startups’ resilience: The role of entrepreneurial ecosystem and dynamic exchange capability. J. Bus. Ind. Mark. 2025, 2, 527–542. [Google Scholar] [CrossRef]
  50. Liu, Y.; Zhou, L.; Wan, X.; Tang, Y.; Liu, Q.; Li, W.; Liao, J. Synthesis and Characterization of a Temperature-Sensitive Microcapsule Gelling Agent for High-Temperature Acid Release. ACS Omega 2024, 9, 20849–20858. [Google Scholar] [CrossRef]
Scheme 1. The structure of multiple networks based on P(AM-AMS-AEG) grafting terpolymer, synthesized by copolymerization of AM, AMS, and AEG and subsequent in situ cross-linking reaction with [Zr].
Scheme 1. The structure of multiple networks based on P(AM-AMS-AEG) grafting terpolymer, synthesized by copolymerization of AM, AMS, and AEG and subsequent in situ cross-linking reaction with [Zr].
Processes 13 03032 sch001
Figure 1. (a) Plots of ηapp vs. fAM, fAMS, or fAEG. Conditions: 80 °C, 6 h. (b) The relationship between ηapp or conv. and tp or Tp. The baseline of AM:AMS:AEG:I is set as 10:2:1.4:0.18. (c) The representative 1H NMR spectrum of P(AM10-AMS2-AEG1.4) in D2O. (d) GPC traces of 5 groups of terpolymers with the column calibrated against the standard PAM samples.
Figure 1. (a) Plots of ηapp vs. fAM, fAMS, or fAEG. Conditions: 80 °C, 6 h. (b) The relationship between ηapp or conv. and tp or Tp. The baseline of AM:AMS:AEG:I is set as 10:2:1.4:0.18. (c) The representative 1H NMR spectrum of P(AM10-AMS2-AEG1.4) in D2O. (d) GPC traces of 5 groups of terpolymers with the column calibrated against the standard PAM samples.
Processes 13 03032 g001
Figure 2. (a) Relationships between ηapp or txl and f[Zr] or Txl. (b) FTIR spectra of (1) [Zr], (2) P(AM10-AMS2-AEG1.4)/TBAC0.1, and (3) P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1. (c) The plots of γs or ηapp vs. fbreaker. (d) The effect of fs on the surficial activity and ηapp.
Figure 2. (a) Relationships between ηapp or txl and f[Zr] or Txl. (b) FTIR spectra of (1) [Zr], (2) P(AM10-AMS2-AEG1.4)/TBAC0.1, and (3) P(AM10-AMS2-AEG1.4)/[Zr]0.35/TBAC0.1. (c) The plots of γs or ηapp vs. fbreaker. (d) The effect of fs on the surficial activity and ηapp.
Processes 13 03032 g002
Figure 3. (a) Schematical illustration of the disassociation–reassociation process of PASG/[Zr] solution during oscillation. (The legends are referred from Scheme 1). (b) The relationship between ηapp and cterpolymer of PASG/[Zr] solution at 20 °C. (c) The effect of cterpolymer on the non-Newtonian characteristics of PASG/[Zr] solution at 20 °C. (d) The shearing damage process of P(AM10-AMS2-AEG1.4), PASG/[Zr], and guar gum solutions with a concentration of 0.27 wt% at 20 °C. (e) The oscillation frequency sweep ( γ ˙ = 1000 s−1) of P(AM10-AMS2-AEG1.4), PASG/[Zr], and guar gum solutions with a concentration of 0.27 wt% at 20 °C.
Figure 3. (a) Schematical illustration of the disassociation–reassociation process of PASG/[Zr] solution during oscillation. (The legends are referred from Scheme 1). (b) The relationship between ηapp and cterpolymer of PASG/[Zr] solution at 20 °C. (c) The effect of cterpolymer on the non-Newtonian characteristics of PASG/[Zr] solution at 20 °C. (d) The shearing damage process of P(AM10-AMS2-AEG1.4), PASG/[Zr], and guar gum solutions with a concentration of 0.27 wt% at 20 °C. (e) The oscillation frequency sweep ( γ ˙ = 1000 s−1) of P(AM10-AMS2-AEG1.4), PASG/[Zr], and guar gum solutions with a concentration of 0.27 wt% at 20 °C.
Processes 13 03032 g003
Figure 4. (a) Plots of ηapp of PASG/[Zr] solution (0.27 wt%) and γ ˙ at various temperatures and the corresponding K-T−1 plot fitted by the Arrhenius Function. (b) DFT analysis of iHBs between an AM unit and an AMS unit. (c) The relationship between ηapp and tos,d or Tos of P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution with various f[Zr] values. (d) The effect of f[Zr] on tos,d or Tos,d according to (c). (e) The recovery process of PASG/[Zr] solution with various concentrations under strong (1000 s−1) or weak shearing (500 s−1). (f) Thixotropic rings of PASG/[Zr] and guar gum (0.27 wt%) solutions.
Figure 4. (a) Plots of ηapp of PASG/[Zr] solution (0.27 wt%) and γ ˙ at various temperatures and the corresponding K-T−1 plot fitted by the Arrhenius Function. (b) DFT analysis of iHBs between an AM unit and an AMS unit. (c) The relationship between ηapp and tos,d or Tos of P(AM10-AMS2-AEG1.4)/[Zr]/TBAC0.1 solution with various f[Zr] values. (d) The effect of f[Zr] on tos,d or Tos,d according to (c). (e) The recovery process of PASG/[Zr] solution with various concentrations under strong (1000 s−1) or weak shearing (500 s−1). (f) Thixotropic rings of PASG/[Zr] and guar gum (0.27 wt%) solutions.
Processes 13 03032 g004
Figure 5. (a) Effects of sand size and settling temperature on the settling process with PASG/[Zr] and guar gum solutions (Csolute = 0.27 wt%, Csand = 30 wt%) as the proppant, and the corresponding relationships between sand size or settling temperature and ts,semi. (b) The relationship between PASG/[Zr] and guar gum. (c) The horizontal tube-carrying test of PASG/[Zr] solution (Csolute = 0.27 wt%, Csand = 30 wt%, dpipe = 2.5 cm) and (d) the effects of dpipe or Csand on vc,settle and vc,suspend.
Figure 5. (a) Effects of sand size and settling temperature on the settling process with PASG/[Zr] and guar gum solutions (Csolute = 0.27 wt%, Csand = 30 wt%) as the proppant, and the corresponding relationships between sand size or settling temperature and ts,semi. (b) The relationship between PASG/[Zr] and guar gum. (c) The horizontal tube-carrying test of PASG/[Zr] solution (Csolute = 0.27 wt%, Csand = 30 wt%, dpipe = 2.5 cm) and (d) the effects of dpipe or Csand on vc,settle and vc,suspend.
Processes 13 03032 g005
Scheme 2. The sand-carrying experimental device. The assignments of mark numbers are listed as follows: 1. Fracturing fluid storage tank; 2. blender; 3. screw pump; 4. valve; 5. pressure gauge; 6. thermometer; 7. flowmeter; 8. check valve; 9. add sand to the sand tank; 10. gas cylinder; 11. frequency converter; 12. differential pressure sensor; 13. visualized pipe sections; 14. camera; 15. computer; 16. waste liquid tank; 17. bypass.
Scheme 2. The sand-carrying experimental device. The assignments of mark numbers are listed as follows: 1. Fracturing fluid storage tank; 2. blender; 3. screw pump; 4. valve; 5. pressure gauge; 6. thermometer; 7. flowmeter; 8. check valve; 9. add sand to the sand tank; 10. gas cylinder; 11. frequency converter; 12. differential pressure sensor; 13. visualized pipe sections; 14. camera; 15. computer; 16. waste liquid tank; 17. bypass.
Processes 13 03032 sch002
Figure 6. (a) Relationships between the filtration volume of fracturing fluid and filtration time at various γ ˙ values. (b) The effect of γ ˙ and core permeability on cf in the fracturing system of PASG/[Zr] solution (0.27 wt%) and the artificial core at 20 °C after 100 min. (c) T2 analysis for core damage with various fluids after 100 min and (d) the T2 signal and integral ratio of small pores, large pores, and free water according to (c).
Figure 6. (a) Relationships between the filtration volume of fracturing fluid and filtration time at various γ ˙ values. (b) The effect of γ ˙ and core permeability on cf in the fracturing system of PASG/[Zr] solution (0.27 wt%) and the artificial core at 20 °C after 100 min. (c) T2 analysis for core damage with various fluids after 100 min and (d) the T2 signal and integral ratio of small pores, large pores, and free water according to (c).
Processes 13 03032 g006aProcesses 13 03032 g006b
Table 1. Brief illustration of the structural characterization of P(AM-AMS-AEG), including FAMS, FAEG, Mn, and PDI, according to 1H NMR and GPC analysis.
Table 1. Brief illustration of the structural characterization of P(AM-AMS-AEG), including FAMS, FAEG, Mn, and PDI, according to 1H NMR and GPC analysis.
SampleFAMS (%)FAEG (%)Mn (kg·mol−1)PDI
P(AM10-AMS2-AEG1.4)13.2211.09250.21.71
P(AM10-AMS2-AEG0.7)15.035.84205.31.88
P(AM9-AMS2-AEG2.4)22.4217.96202.11.92
P(AM9-AMS1-AEG1.7)21.2713.65200.72.07
P(AM9-AMS1-AEG2.7)20.8020.31152.62.54
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, K.; Guo, C.; Gong, Q.; Li, G.; Zhang, C.; Jiang, T. Performance and Characteristics of Low-Molecular-Weight Cross-Linked Grafting Terpolymers as Thickening Agents in Reservoir Fracturing Processes. Processes 2025, 13, 3032. https://doi.org/10.3390/pr13103032

AMA Style

Wang K, Guo C, Gong Q, Li G, Zhang C, Jiang T. Performance and Characteristics of Low-Molecular-Weight Cross-Linked Grafting Terpolymers as Thickening Agents in Reservoir Fracturing Processes. Processes. 2025; 13(10):3032. https://doi.org/10.3390/pr13103032

Chicago/Turabian Style

Wang, Kai, Chenye Guo, Qisen Gong, Gen Li, Cuilan Zhang, and Teng Jiang. 2025. "Performance and Characteristics of Low-Molecular-Weight Cross-Linked Grafting Terpolymers as Thickening Agents in Reservoir Fracturing Processes" Processes 13, no. 10: 3032. https://doi.org/10.3390/pr13103032

APA Style

Wang, K., Guo, C., Gong, Q., Li, G., Zhang, C., & Jiang, T. (2025). Performance and Characteristics of Low-Molecular-Weight Cross-Linked Grafting Terpolymers as Thickening Agents in Reservoir Fracturing Processes. Processes, 13(10), 3032. https://doi.org/10.3390/pr13103032

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