Next Article in Journal
Histological Evaluation of Chemo Mechanical Caries Removal with a Babaco-Based Formulation Gel
Previous Article in Journal
Agarose Hydrogels for Bone Tissue Engineering, from Injectables to Bioprinting
Previous Article in Special Issue
Study on a Strong Polymer Gel by the Addition of Micron Graphite Oxide Powder and Its Plugging of Fracture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Urea Delays High-Temperature Crosslinking of Polyacrylamide for In Situ Preparation of an Organic/Inorganic Composite Gel

by
Li Liang
1,2,
Junlong Li
1,2,
Dongxiang Li
1,2,*,
Jie Xu
1,2,
Bin Zheng
1,2 and
Jikuan Zhao
1,2,*
1
Key Laboratory of Optic-Electro Sensing and Analytical Chemistry of Life Science, Ministry of Education, Qingdao University of Science and Technology, Qingdao 266042, China
2
Lab of Colloids and Functional Nanostructures, College of Chemistry and Molecular Engineering, Qingdao University of Science and Technology, Qingdao 266042, China
*
Authors to whom correspondence should be addressed.
Gels 2025, 11(4), 256; https://doi.org/10.3390/gels11040256
Submission received: 18 February 2025 / Revised: 20 March 2025 / Accepted: 26 March 2025 / Published: 31 March 2025
(This article belongs to the Special Issue Advanced Gels for Oil Recovery (2nd Edition))

Abstract

:
To address the rapid crosslinking reaction and short stability duration of polyacrylamide gel under high salinity and temperature conditions, this paper proposes utilizing urea to delay the nucleophilic substitution crosslinking reaction among polyacrylamide, hydroquinone, and formaldehyde. Additionally, urea regulates the precipitation of calcium and magnesium ions, enabling the in situ preparation of an organic/inorganic composite gel consisting of crosslinked polyacrylamide and carbonate particles. With calcium and magnesium ion concentrations at 6817 mg/L and total salinity at 15 × 104 mg/L, the gelation time can be controlled to range from 6.6 to 14.1 days at 95 °C and from 2.9 to 6.5 days at 120 °C. The resulting composite gel can remain stable for up to 155 days at 95 °C and 135 days at 120 °C. The delayed gelation facilitates longer-distance diffusion of the gelling agent into the formation, while the enhancements in gel strength and stability provide a solid foundation for improving the effectiveness of profile control and water shut-off in oilfields. The urea-controlling method is novel and effective in extending the high-temperature cross-linking reaction time of polyacrylamide. By converting calcium and magnesium ions into inorganic particles, it enables the in situ preparation of organic/inorganic composite gels, enhancing their strength and stability.

Graphical Abstract

1. Introduction

The gelation time, stability duration, and mechanical strength of polymer gels are pivotal parameters for their successful application in oilfield profile control and water shut-off treatments [1,2,3,4,5,6,7]. These parameters directly influence the duration and effectiveness of on-site production operations. However, high temperatures can accelerate gelation, while high salinity conditions can drastically shorten the stability duration of polymer gels, presenting ongoing challenges for researchers [8,9,10,11]. Simply adjusting the concentrations of polymers and crosslinking agents has proven insufficient for effectively delaying gelation at elevated temperatures [12,13,14,15], with limited literature addressing the modulation of crosslinking reactions based on fundamental reaction mechanisms. To overcome these limitations and enhance gel performance, organic/inorganic composite gels (CG) have attracted significant attention [2,4,6,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. The integration of organic and inorganic components within composite gels leverages various interactions including electrostatic forces, hydrogen bonding, and coordination bonds to markedly improve the strength and stability of the resulting materials [16,17,18,19]. In addition, inorganic particles can also serve as multifunctional crosslinkers to form physical crosslinking points with polymer molecules. Organic/inorganic composite gels exhibit improved hydrophilicity, thermal stability, salt resistance, and higher elastic modulus, and have become an important research direction for enhanced oil recovery [20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. A wide array of inorganic particles, including silica [16,19,20,21,22], titanium dioxide [23], alumina [24], zirconia [25], fly ash [26], Fe3O4 [27], montmorillonite [28,29,30], bentonite [31], laponite [17,32], magnesium aluminum layered double hydroxide [33], zirconium hydroxide [18], graphene oxide [23,34], micron graphite oxide powder [35], and carbon nanotubes [23], have been explored for this purpose. However, incorporating these inorganic particles often requires regulating particle size [19,20], layer thickness [28,30,33], and surface properties [29,31] to ensure compatibility. Some systems struggle with high-salinity formation waters, necessitating the use of low-concentration brines or freshwater for gel preparation [17,19,34], which limits practical applications. The in situ synthesis of organic/inorganic composite gels from high-salinity polymer solutions remains underexplored. It is well-known that divalent metal ions such as Ca2+ and Mg2+ in formation brine can catalyze polymer hydrolysis, leading to over-crosslinking and syneresis at high temperature [18,36,37]. There is a need for scientific approaches to utilize mineral resources within formations for the effective in situ preparation of composite gels.
Significant research has explored the crosslinking reaction mechanisms involving polyacrylamide (PAM), hydrolyzed polyacrylamide (HPAM), binary and ternary copolymers containing acrylamide [1,4,5,6,7,12,13,14,15,38,39,40,41,42,43]. Organic crosslinkers that either contain or can produce formaldehyde are frequently utilized, including formaldehyde itself, hexamethylenetetramine (HMTA), phenol/catechol/resorcinol/hydroquinone (HQ)-formaldehyde or HMTA, and melamine-formaldehyde, etc., [12,13,14,15,38,39,40,42,43,44]. The reaction mechanisms of these crosslinking systems typically fall into two categories [42]. The first involves hydroxymethylation reactions between formaldehyde and PAM or phenolic substances, leading to the introduction of hydroxymethyl groups on the amide nitrogen and ortho/para positions of phenolic hydroxyls, followed by condensation dehydration to form a three-dimensional network [1,4,43]. The second category encompasses hydroxymethylation reactions between formaldehyde and phenolic or melamine molecules, introducing hydroxymethyl groups that subsequently react with amide groups in the polymer through condensation, ultimately forming a robust three-dimensional gel structure [12,13,14,40].
HMTA, which gradually decomposes into formaldehyde and ammonia above 80 °C [14], has emerged as a more environmentally friendly alternative to formaldehyde. Studies [10,13,44] have examined the gelation performance of PAM/HMTA/HQ and HPAM/HMTA/methyl p-hydroxybenzoate crosslinking systems across varying pH levels. The reaction system can form a stable gel when the pH is maintained between 7.5 and 9.5. However, gelation does not occur when the pH exceeds 9.5 [10]. By lowering the pH of the reaction system from 7.9 to 4.5 using an acid solution, the gelation time is significantly reduced from 10.6 h to 3.2 h, and the gel strength increases from 0.053 MPa to 0.062 MPa [13]. It has been proposed that an acidic environment promotes the decomposition of HMTA into formaldehyde, which further accelerates the crosslinking reaction, leading to a faster reaction rate and increased crosslink density. In contrast, an alkaline environment inhibits the decomposition of HMTA and the subsequent crosslinking reactions [5,13]. Notably, in gel systems involving HMTA, the decomposition of HMTA generates ammonia gas, which reacts with water to dissociate and release hydroxide ions, thereby increasing the system’s alkalinity. Consequently, when analyzing the impact of pH on gelation performance, the primary focus has been on how acidity and alkalinity influence the decomposition of HMTA, which is considered the main factor affecting the gelation performance of the reaction system. However, the direct effect of pH on the crosslinking reaction itself has often been overlooked.
Bryant et al. [15] investigated the gelation performance of acrylamide and the acrylamide-based AM-AMPS binary copolymer within a phenol-formaldehyde crosslinking system, with a particular focus on the influence of pH. In this system, formaldehyde acts directly as a reactant, and its quantity or concentration remains independent of the solution’s pH. The experiments demonstrated that gelation occurs within a pH range of 1.0 to 8.5, with the crosslinking reaction proceeding rapidly at pH 1, while the rate of crosslinking is the slowest under neutral conditions. Feng et al. [39] utilized 13C-NMR spectroscopy to explore the reaction process between PAM and formaldehyde. The study revealed that at a high pH (9), formaldehyde reacts with PAM to produce N-hydroxymethylated polyacrylamide without leading to crosslinking. Conversely, at low pH (1), formaldehyde can react with PAM to form methylene-bis-acrylamide crosslinked structures. These findings indicate that pH plays a significant regulatory role in the crosslinking reaction between formaldehyde or phenolic formaldehyde crosslinking agents and acrylamide polymers.
Marandi et al. [38] explored the crosslinking reaction of formaldehyde with PAM catalyzed by hydrochloric acid at pH 5. Under acidic conditions, the protonation of the hydroxymethyl group facilitates the nucleophilic attack of amide nitrogen atoms on the carbon atoms of the hydroxymethyl groups, resulting in the formation of a gel structure with methylene diacrylamide serving as the crosslinking point. The crosslinking reaction proceeds via a nucleophilic reaction mechanism. In another study, melamine/formaldehyde (MF) resins rich in hydroxymethyl groups were used to crosslink HPAM [40], with optimal pH levels for stable, high-strength gels ranging from 5 to 9. Outside this range, gel viscosity decreases and strength weakens due to polymer chain coiling and hydrolysis. It is suggested that the reaction between the amide groups and the hydroxymethyl groups in the MF resin follows a nucleophilic substitution reaction mechanism [40]. Acidic conditions promote the protonation of hydroxymethyl groups, thereby enhancing the crosslinking between amide groups and hydroxymethyl groups. Conversely, an increase in the pH of the reaction system inhibits the protonation of hydroxymethyl groups, which impedes the nucleophilic attack of the nitrogen atom in the amide group on the carbon of the hydroxymethyl group, ultimately preventing the establishment of a dense crosslinked network structure within the system.
Although a pH range of 5–9 is favorable for the formation of high-strength gels, acidic conditions can cause corrosion of the internal walls of the wellbore. To mitigate this issue, ammonium chloride was utilized as a catalyst to regulate the crosslinking reaction between HPAM and MF at pH levels of 8–9 [40]. At a temperature of 80 °C, this approach successfully reduced the gelation time from 7 days to a mere 8 h while simultaneously enhancing gel strength. It is suggested that the ammonium ions generated from the ionization of ammonium chloride donate protons to the hydroxymethyl groups, thereby accelerating the crosslinking reaction within the described system [40]. This study not only validates the nucleophilic substitution mechanism of the reaction between PAM and organic crosslinkers but also provides valuable insights for regulating gelation time based on this mechanism. References [45,46] have documented the use of ammonium salts to accelerate polymer gelation, particularly to facilitate the crosslinking reaction between polyacrylamide and phenolic formaldehyde compounds under low-temperature conditions. These findings will contribute to the rapid sealing of formations and the protection of oil and gas reservoirs during production activities.
Based on the nucleophilic substitution crosslinking reaction mechanism between PAM and phenolic formaldehyde compounds [38,40], this research employs urea additives to regulate both the gelation time and strength of the composite gel. Urea decomposes at elevated temperatures, producing ammonia and carbon dioxide, which further generate hydroxide and carbonate anions in the solution. These anions inhibit the protonation of the methylol group, thereby extending the gelation time of the gelant. In high-salinity solutions, divalent calcium and magnesium ions interact with carboxylate ions produced from the hydrolysis of amide groups through electrostatic or coordination bonds. Additionally, these calcium and magnesium ions can also react with carbonate or hydroxide ions generated from the decomposition of urea to form inorganic particles, such as carbonates or basic carbonates. These particles subsequently bind closely with polymer molecules, resulting in the formation of high-strength composite gels. This study utilizes divalent metal ions in formation brine as raw materials to in situ synthesize organic/inorganic composite gels. This approach not only mitigates the destabilizing impact of inorganic salts on polymer gels but also enhances the strength and stability of the composite gel through the incorporation of generated inorganic particles, thereby achieving integrated utilization of subsurface mineral resources.

2. Results and Discussion

In this study, urea was employed to regulate the crosslinking reaction between polyacrylamide and hydroquinone/formaldehyde crosslinkers, as well as the precipitation reactions of calcium and magnesium ions. The amount of urea was carefully controlled to maintain a specific molar ratio between urea and the divalent metal ions in the solution, denoted as N = nurea/(nCa2+ + nMg2+), where nurea, nCa2+ and nMg2+ are the amounts of substance of urea, calcium ions, and magnesium ions, respectively. The gelling solution or gelant transforms into an organic/inorganic composite gel (CG) with defined viscoelastic properties after high-temperature aging in the oven. For performance comparisons, freshwater gel (FG) and brine gel (BG) samples without urea were also prepared. In this paper, the samples of freshwater gel, brine gel, and composite gel are designated as FG-t, BG-Tx-t, and CG-Tx-UN-t, respectively, where t signifies the aging temperature of the sample in degrees Celsius (°C), Tx indicates that the total dissolved solids or salinity in the gelant equals to x × 104 mg/L, and UN represents the molar ratio N between urea and the sum of calcium and magnesium ions in the reaction system.

2.1. FT-IR and XRD

The FT-IR spectra of the PAM polymer and FG sample are presented in Figure 1a. The absorption peaks observed at 3443, 1641, and 1104 cm−1 correspond to the stretching vibrations of υN-H, υC=O, and υC-N, respectively [9,18,40,47,48]. Furthermore, the peaks at 2920, 2851, 1453, and 1390 cm−1 represent the asymmetric stretching, symmetric stretching, asymmetric bending, and symmetric bending absorption peaks of the C–H bonds in the methylene group [9,18,47,48]. These findings confirm the presence of the primary functional groups in polyacrylamide. In the FT-IR spectrum of PAM FG sample, two new absorption peaks at 1570 and 1231 cm−1 appear, which are attributed to the stretching vibrations of carbon–carbon conjugated double bonds and the C–O bond stretching vibrations in phenolic compounds, respectively [47,49]. These FT-IR data suggest that a gel structure is formed through a crosslinking reaction between hydroquinone, formaldehyde, and PAM via covalent bonding.
X-ray diffraction (XRD) is an effective technique for analyzing the structure of inorganic materials. In this study, XRD tests were conducted on inorganic particles formed by the reaction of Ca2+ and Mg2+ ions with hydrolysis products of urea, with the results illustrated in Figure 1b. The characteristic peaks observed at 2θ values of 26.1°, 27.4°, 33.4°, 36.1°, 38.7°, and 46.0° correspond to the (111), (021), (012), (200), (022), and (221) crystal plane diffractions of CaCO3 (JCPDS#75-2230) [50]. A prominent peak at 2θ of 30.1° is attributed to the (104) crystal plane diffraction of (Mg0.129Ca0.871)CO3 (JCPDS#86-2336) [51]. The peaks at 2θ values of 12.0°, 15.2°, and 23.7° can be indexed as the (020), (100), and (040) crystal plane diffractions of Mg7(CO3)5(OH)4·24H2O (JCPDS#47-188) [52]. The XRD results indicate that Ca2⁺ and Mg2⁺ ions in brine can react with OH and CO32− generated from urea hydrolysis to form various inorganic particles, including calcium carbonate, magnesium calcium carbonate, and basic magnesium carbonate. The diffraction peak intensities of calcium carbonate and magnesium calcium carbonate are relatively strong, while the intensity of the basic magnesium carbonate peak is lower, suggesting that the crystal structures of the former two inorganic compounds are more well defined, whereas basic magnesium carbonate exhibits significantly greater amorphicity [52]. The equations involving urea decomposition and its products further interacting with calcium and magnesium ions are detailed in Formulas (1) to (7) [53,54].
Urea decomposition reaction equations:
CO(NH2)2 + H2O = CO2↑ + 2NH3
NH 3 + H 2 O NH 3 · H 2 O NH 4 + + OH
CO 2 + H 2 O = H 2 CO 3 H + + HCO 3
HCO 3 + OH CO 3 2 + H 2 O
Precipitation reaction equations for Ca2⁺ and Mg2⁺ ions:
CO32− + Ca2+→CaCO3
CO32− + 0.129Mg2+ + 0.871Ca2+→ (Mg0.129 Ca0.871)CO3
5CO32− + 7Mg2+ + 4OH + 24H2O→ Mg7(CO3)5(OH)4·24 H2O
The amide groups in PAM undergo hydrolysis at elevated temperatures, leading to the formation of carboxyl groups, which subsequently dissociate into carboxylate ions that interact with cations in the solution through electrostatic or coordination forces [1,8,36,55]. The electrostatic interactions between Ca2⁺ and Mg2⁺ ions and carboxylate ions in the solution are significantly stronger than those with monovalent cations. Additionally, divalent cations can bind to carboxylate ions from different molecular chains, resulting in crosslinking that reduces the hydrophilicity of the polymer, and may even lead to excessive crosslinking and syneresis [1,36,55]. This phenomenon is a major contributor to the reduced stability of the polymer under high mineralization conditions. In PAM solutions with a high concentration of Ca2+ and Mg2+ ions, the availability of carboxylate ions is relatively limited, meaning that only a fraction of the divalent ions can interact with carboxylate ions through electrostatic or coordination interactions. At elevated temperatures, urea decomposes to produce NH3 and CO2. As the concentration of NH3 increases within the reaction system, it further reacts with water to yield NH4+ and OH, resulting in an increase in the system’s alkalinity. Concurrently, CO2 can generate carbonate ions (CO32−) under alkaline conditions. The cations Ca2+ and Mg2+ that associate with carboxylate ions can precipitate with OH and CO32− on the surface of the polymer, leading to the formation of inorganic particles such as calcium carbonate, magnesium carbonate, and basic magnesium carbonate. This process resembles biomineralization [56]. These inorganic particles interact with the carboxylate ions of the polymer through electrostatic or coordination forces, resulting in the formation of an organic/inorganic composite gel. Scheme 1 illustrates the deposition of calcium and magnesium inorganic particles on the surface of polyacrylamide.

2.2. SEM

The experimental results reveal that both the BG sample (BG-T15-95, Figure 2a) and CG sample (CG-T15-U1.00-95, Figure 2d) exhibit a three-dimensional network morphology. During the freeze-drying process, a substantial amount of inorganic salts precipitates continuously on the surface of the crosslinked polymer network in both gels. Figure 2b,e display the SEM images of the BG and CG samples after being washed with water, respectively. While the gel samples maintain their three-dimensional network structure, the surface of the polymer gel becomes smoother after washing, indicating the removal of water-soluble inorganic salts. Upon further magnification of the electron microscope images, it was observed that no significant particles were attached to the surface of the polymer in the BG sample (Figure 2c). In contrast, numerous solid particles remained adhered to the surface of the polymer in the CG sample (Figure 2f), with an average particle size of approximately 80–100 nm. This morphological characteristic is similar to that of the PAM/PEI/CAF composite gels [26]. It is inferred that the solid particles attached to the polymer surface in the composite gel correspond to inorganic particles of calcium carbonate, magnesium carbonate, or basic magnesium carbonate, as indicated by the XRD test results. These insoluble substances, formed from calcium and magnesium ions in conjunction with the decomposition products of urea, cannot be removed through water washing. Moreover, electrostatic or coordination interactions between divalent cations and the carboxylate anions of the polymer facilitate the stable adhesion of the particles to the polymer surface. Consequently, the crosslinked polymer forms an organic/inorganic composite gel with the inorganic particles.
PAM is a linear polymer that undergoes a crosslinking reaction involving its amide groups when exposed to hydroquinone formaldehyde crosslinking agents, resulting in the development of a three-dimensional gel network. When a substantial amount of sodium chloride, calcium chloride, and magnesium chloride is dissolved in water, a three-dimensional brine gel is formed. According to the salinity composition of the brine gel, a certain amount of urea is introduced into the reaction system. Under high-temperature conditions, urea decomposes and facilitates the precipitation reaction of divalent cations, leading to the formation of inorganic particles such as calcium carbonate, magnesium carbonate, and basic magnesium carbonate. These inorganic components interact with the organically crosslinked polymer, resulting in the creation of a three-dimensional composite gel. The precipitation reaction consumes a significant amount of divalent metal ions, thereby reducing their concentrations in the liquid phase and effectively mitigating their adverse effects on the stability of the polymer gel. Concurrently, the inorganic particles are uniformly distributed within the crosslinked polymer matrix, significantly enhancing the elastic modulus, strength, and stability of the composite gel [16,17,26,28].

2.3. Gelation Performance

The gelation process and stability of gel samples were recorded based on the gel strength codes proposed by Sydansk et al. [8,57,58,59]. The classification of gel strength levels is detailed in Table 1. The gel strength developments of FG, BG, and CG samples throughout the gelation process at temperatures of 95 °C and 120 °C, are presented in Table 2 and Table 3. Concentrations of PAM (CPAM), HQ (CHQ), and HCHO (CHCHO) for the gel samples are also given. The FG can achieve a maximum strength of I at 95 °C without syneresis. At 120 °C, FG can also reach the strength of I level, but begins to dehydrate after being heated for 126 days. BG samples similarly attain their maximum strength I at 95 °C. However, brine gels exhibit a highest strength of H and these samples are more susceptible to dehydration at 120 °C. CG samples demonstrate the capability to reach a peak strength of I at both 95 °C and 120 °C. Moreover, the CG samples show superior stability under higher urea proportion conditions (N = 1.0 − 2.0).
The time required for the gel sample to attain a strength code of D was recorded as the gelation time (tD) [58,59]. The stability time (tS) is defined as the duration needed for the gel to dehydrate by 5 mL from a 20 mL gel sample. This work conducts a comparative analysis of the effects of various factors, including urea concentration (Curea), degree of mineralization, and temperature, on both tD and tS.
tD and tS data of FG and BG samples are shown in Table 4. The experimental results indicate that the FG samples are characterized by a longer stability time, exceeding 176 days and 126 days at 95 °C and 120 °C, respectively. For the BG samples, as the salinity increases from 5 × 104 mg/L to 20 × 104 mg/L, both the gelation time and stability time are significantly shortened. For example, at 95 °C, the tD of BG decreases from 12.4 days to 7.9 days, while the tS drops from 42.0 days to 24.5 days. Notably, BG samples exhibit significantly shorter stability times compared to FG samples. When comparing high-temperature samples with low-temperature samples, both gelation time and stability time are shorter for the former. These findings suggest that high temperature and high salinity are the primary factors contributing to the reduced stability of polymer gels [1,8,36,37,55]. Under high-temperature conditions, the thermal motion of polymer and crosslinking agent molecules intensifies, providing higher energy that enables more molecules to participate in the crosslinking reaction, thereby shortening the gelation time [10,16,26]. Under high mineralization conditions, divalent calcium and magnesium ions facilitate polymer hydrolysis, converting amide groups into carboxyl groups, which can further dissociate into carboxylate ions and hydrogen ions. Divalent cations link with carboxylate ions of different polymer chains through ionic interactions [55], while the dissociated hydrogen ions promote nucleophilic substitution crosslinking reactions between amide groups and hydroquinone formaldehyde crosslinking agents. These reactions result in over-crosslinking and syneresis within the system, significantly diminishing the stability of the gel samples [36,37,55,60].
At 95 °C, we investigated the effects of CPAM, CHQ, and CHCHO on tD and tS of the composite gel, which had a salinity of 15 × 104 mg/L and a fixed Curea of 0.68 wt% (N = 0.5). The experimental data are summarized in Table 5, where CPAM, CHQ and CHCHO are maintained ranging from 0.5 to 2.0 wt%, 0.06 to 0.31 wt%, and 0.07 to 0.33 wt%, respectively. Under conditions of low polymer concentration (CPAM = 0.5 wt%) and low crosslinker concentrations (CHQ = 0.06 wt%, CHCHO = 0.07 wt%), the stability times of the composite gels are found to be 37.8 days and 69.2 days, respectively. However, the structures of these gel samples remain incomplete and susceptible to dehydration, as their maximum gel strengths correspond to the codes of H and F, rather than the code of I. In contrast, when the PAM concentration is excessively high (1.5–2.0 wt%), the gelation time is significantly reduced to 0.7–1.2 days, resulting in over-crosslinking, syneresis, and a loss of stability. Consequently, the optimized concentrations for CPAM, CHQ and CHCHO at 95 °C are determined to be 1 wt%, 0.20 wt%, and 0.21 wt%, respectively. At 120 °C, the preferred CPAM remains at 1 wt%, while both CHQ and CHCHO are adjusted to 0.16 wt% to delay the crosslinking reaction.
To further investigate the effects of salinity and Curea on the gelation performance of the composite gel, we examined urea-containing crosslinking systems with N values of 0.5 and 1.0 across salinities ranging from 5 to 20 × 104 mg/L. The gelation time and stability performance of the composite gels are presented in Table 6. Due to the proportional amount of urea and divalent cations in the reaction solution, Curea increases along with salinity, leading to a gradual extension of the gelation time for the composite gel. When N is set at 0.5, an increase in salinity results in an extended gelation time from 2.1 days to 5.2 days, with stability times ranging from 72.1 to 85.2 days, significantly surpassing those of the brine gel. When the N value is increased to 1.0, as salinity rises from 5 to 20 × 104 mg/L, the gelation time extends from 2.2 days to 7.0 days, while the stability time of the composite gel further increases to over 155.0 days. These findings underscore the critical role of urea in delaying the crosslinking reaction between polyacrylamide and the hydroquinone formaldehyde crosslinker at elevated temperatures, thereby enhancing the stability of the composite gel.
This paper systematically investigates the effects of Curea on the gelation and stability times of a polymer solution with a salinity of 15 × 104 mg/L and a PAM concentration of 1 wt% at temperatures of 95 °C and 120 °C, as illustrated in Figure 3. An increase in urea concentration results in gelation times ranging from 2.5 to 14.1 days at 95 °C and from 1.4 to 6.5 days at 120 °C (Figure 3a), demonstrating that urea effectively regulates the gelation time tD of the composite gel. Although the tD at 120 °C is shorter than that at 95 °C, higher urea concentrations can extend the tD by 3.6 to 4.6 times, leading to significantly longer gelation times for CG samples compared to FG and BG specimens.
In both laboratory experiments and field production, a common approach to prolong gelation time is to adjust the concentrations of crosslinking agents and polymers. For instance, references [12,13,14] indicate that adjustable gelation time ranges are 21–41 h at 100.8 °C, 6.3–15.5 h at 110 °C, and 7–14 h at 150 °C. Although these regulatory measures demonstrate some effectiveness, the overall findings highlight that short gelation times and limited controllable ranges at elevated temperatures remain significant technical challenges. A comprehensive understanding of the crosslinking reaction mechanism, coupled with precise regulation of the concentrations of reaction components, is crucial for effectively extending the high-temperature crosslinking reaction time of the polymer.
In this paper, urea is employed as a base source. At high temperatures, it decomposes into ammonia and carbon dioxide. This process leads to the formation of weak electrolytes, such as hydrated ammonia and carbonic acid, which partially dissociate into NH4+, OH, H+, HCO3 and CO32− ions. The anions interact with NH4+ and H+ ions through electrostatic forces, effectively suppressing the protonation of hydroxymethyl groups and thereby prolonging the crosslinking reaction time at elevated temperatures [40]. The ionization equilibrium of the weak electrolyte shifts forward only when Ca2+ and Mg2+ ions in the solution consume a sufficient amount of OH and CO32− ions, allowing NH4+ and H+ ions to facilitate the crosslinking reaction. Based on the nucleophilic reaction mechanism, the gelation time of the PAM solution can be extended, thereby creating favorable conditions for deep profile control and water plugging in oilfields.
Figure 3b illustrates the correlation between the stability time tS of the composite gel and the urea concentration Curea. At 95 °C, it is clear that as Curea increases from 0.14 wt% (N = 0.1) to 0.41 wt% (N = 0.3), the stability time of the composite gel extends from 36.0 days to 55.7 days. Further increasing Curea to 0.68 wt% (N =0.5) and 1.02 wt% (N = 0.75) results in a significant increase in stability time, reaching 78 days and 122 days, respectively. Notably, when Curea is elevated to the range of 1.35–2.71 wt% (N = 1.0–2.0), the stability time can extend up to 155 days. At 120 °C, the stability time of the composite gel is slightly reduced compared to that at 95 °C. Specifically, as Curea increases from 0.14 wt% (N = 0.1) to 1.02 wt% (N = 0.75), the stability time gradually rises from 9.75 days to 31.6 days. This trend suggests that high-temperature conditions may lead to over-crosslinking and syneresis, which can reduce stability time. However, further increasing Curea significantly prolongs the stability time of the composite gel at 120 °C. For example, when Curea increases from 1.35 wt% (N = 1.0) to 2.71 wt% (N = 2.0), the stability time of the composite gel samples can reach 135 days.
Urea facilitates the precipitation of Ca2+ and Mg2+ ions, leading to the formation of carbonate particles. This process not only effectively reduces the concentration of divalent cations that undermine gel stability but also enhances the stability of the composite gel by incorporating inorganic particles. Under the influence of urea, the stability time of the composite gel significantly increases, allowing it to maintain strength at I or H levels for an extended period under high-temperature and high-salinity conditions (Table 2 and Table 3). This stability is crucial for improving production efficiency in oilfields.

2.4. Dynamic Rheological Property

It is well established that during the aging process of the gelant, urea gradually decomposes, and the alkaline environment effectively inhibits hydroxymethyl protonation, thereby extending the gelation time at elevated temperatures. Concurrently, the OH and CO32− anions present in the reaction system can precipitate with Ca2+ and Mg2+ ions in highly mineralized solutions, leading to the formation of inorganic particles that subsequently create organic/inorganic composite gels with the crosslinked polymers. To analyze the impact of the composite structure on gel strength, this study evaluated the dynamic rheological properties of the gel samples. The elastic modulus and viscous modulus of FG, BG, and CG samples were measured, and the effects of urea concentration and salinity on gel performance were assessed, as illustrated in Figure 4.
Figure 4a illustrates the elastic modulus and the viscous modulus curves of FG, BG, and CG samples. The composite gels were prepared under varying urea concentration conditions, with a salinity of 15 × 104 mg/L. All samples were aged at 95 °C for 44.7 days. Firstly, the overall experimental results indicate that the elastic modulus G′ of all gel samples is significantly greater than the viscous modulus G”, demonstrating that the gel networks formed through crosslinking reactions possess excellent deformation characteristics and quasi-solid properties [21,31,61], which aid in plugging large pores in the formation and controlling the water absorption profile. Secondly, under experimental conditions with oscillation frequencies ranging from 0.01 to 10 Hz, the elastic modulus G′ of the freshwater gel FG-95 is the lowest (45.3–55.6 Pa), followed by the brine gel BG-T15-95 (51.3–73.7 Pa), while the composite gel exhibits the highest elastic modulus (Figure 4a). It can be analyzed that, although the polymer molecular chains in the FG system are the most extended, the strength of the pure organic gel is the lowest. In contrast, the BG containing inorganic salts demonstrates a higher elastic modulus than the FG under the same conditions, attributed to calcium and magnesium ions promoting the hydrolysis of PAM and crosslinking with the formed carboxylate groups [1,8,36], thereby enhancing the gel strength. When urea is introduced into the reaction solution, the inorganic particles in the prepared composite gel interact with the polymers through electrostatic or coordination forces, and this composite structure imparts a higher elastic modulus to the product [16,17]. It is evident that both the crosslinked structure and the composite structure directly influence the enhancement of gel strength. Thirdly, the elastic modulus of the composite gel increases with rising urea concentration. In Figure 4a, there are six CG samples including CG-T15-U0.30-95, CG-T15-U0.50-95, CG-T15-U0.75-95, CG-T15-U1.00-95, CG-T15-U1.25-95, and CG-T15-U1.50-95. The salinity and polymer concentration were fixed at 15 × 104 mg/L and 1 wt%, respectively. As the N value increases from 0.30 to 1.50, the elastic modulus ranges for the six CG samples are 64.5–75.6 Pa, 60.0–82.4 Pa, 75.2–90.9 Pa, 94.9–131.2 Pa, 103.5–144.0 Pa, and 103.8–150.3 Pa, respectively. These results indicate that the elastic modulus of CG-T15-U0.30-95 is slightly higher than that of BG-T15-95 (51.3–73.7 Pa). As the N value further increases to 0.50 and 0.75, the strengths of these two composite gel samples rise, showing an increase of 10–15 Pa in elastic modulus compared to that of CG-T15-U0.30-95. When the N value reaches 1.00 in sample CG-T15-U1.00-95, there is a significant increase in gel strength, with an increase of about 20–40 Pa. When the N value is further increased to 1.25–1.50, the elastic modulus G′ of the composite gel continues to show an upward trend, but the increase is relatively small. Under low urea concentration conditions (N = 0.3 − 1.0), as the urea concentration increases, the reaction between calcium and magnesium ions in the composite gel and the hydrolysis products of urea can produce a larger number of inorganic particles, which helps to increase the strength of the composite gel [17,21,26]. Under high urea concentration conditions (N = 1.25 − 1.50), further increasing the urea content creates an alkaline environment due to urea decomposition, which is unfavorable for the PAM crosslinking reaction. As shown in Figure 3a, high urea concentration conditions can significantly prolong the gelation time of the composite gel. Therefore, at N = 1.25 − 1.50, the increase in the elastic modulus of the composite gel diminishes.
Figure 4b presents the elastic modulus G′-oscillation frequency curves of three urea-containing composite gels (CG-T15-U1.00-120, CG-T15-U1.25-120, and CG-T15-U1.50-120) aged 7.3 days at 120 °C, with N values of 1.00, 1.25, and 1.50, respectively. Under the experimental conditions, the elastic modulus ranges of these three composite gels are 40.2–54.8 Pa, 41.1–56.6 Pa, and 41.7–57.8 Pa, respectively. As the N value increases from 1.00 to 1.50, the elastic modulus of the composite gels increases slightly. This observation is similar to the phenomenon noted in Figure 4a. The influence of urea on the growth of inorganic particles and its effect on the PAM crosslinking reaction jointly determine the elastic modulus of the composite gel. In Figure 4b, the elastic modulus of the CG sample CG-T15-U1.00-120 (120 °C, aged for 7.3 days, CPAM = 1.0 wt%, CHQ = CHCHO = 0.16 wt%) is lower than that of the CG sample CG-T15-U1.00-95 depicted in Figure 4a (95 °C, aged for 44.7 days, CPAM = 1.0 wt%, CHQ = 0.20 wt%, CHCHO = 0.21 wt%). When the composition of the gel agent is the same, high temperature promotes the polymer crosslinking reaction and the formation of inorganic particles, resulting in a composite gel with a faster gelation rate and higher gel strength. To delay the crosslinking reaction of the composite gel under high temperature conditions of 120 °C, we selected a lower concentration of crosslinking agent to prolong the gelation time and control the gel strength. Additionally, since the aging time of the composite gel at 120 °C is relatively short in this study, the elastic modulus of the composite gel samples in Figure 4b is relatively small. In fact, delaying the gelation rate of the gel system under high-temperature conditions and controlling the increase in gel strength is more advantageous for the gelant to penetrate deeply into the formation. Furthermore, as indicated in Table 3, the CG samples under high-temperature conditions demonstrate long-term stability, with gel strength reaching level I.
Figure 4c depicts the elastic modulus of composite gel samples, i.e., CG-T5-U1.00-95, CG-T10-U1.00-95, CG-T15-U1.00-95, and CG-T20-U1.00-95. As salinity and urea concentration increase, the elastic modulus of the composite gel increases correspondingly. For salinities of 5, 10, 15, and 20 × 104 mg/L, the elastic modulus ranges for the CG samples are 68.0–95.1 Pa, 77.0–114.2 Pa, 94.9–131.2 Pa, and 119.1–156.4 Pa, respectively. These results confirm the role of inorganic particles formed in brine in enhancing the elastic modulus of the composite gel. Increased urea concentrations and salinities result in a higher quantity of inorganic particles in the system, thus strengthening the composite gel [17,21,26]. This study employs a precipitation reaction between products of urea decomposition and calcium and magnesium ions to treat inorganic salts, converting destabilizing metal ions into beneficial inorganic particles. Simultaneously, these generated inorganic particles are used to construct the composite gel in situ with organic crosslinked polymers, thereby improving its elastic modulus. This process exemplifies an effective integration of inorganic salt treatment and application, converting adverse factors that destabilize the gel into favorable conditions that improve its strength and stability. This research has introduced a novel and efficient method for the in situ preparation of organic/inorganic composite gels from concentrated brine.

2.5. Thermal Stability

Figure 5 presents the thermogravimetric (TG), differential scanning calorimetry (DSC), and differential thermogravimetric (DTG) curves for FG, BG, and CG samples. The thermal analysis is divided into low-, medium-, and high-temperature ranges based on the weight loss observed in the DTG curves, as illustrated in Figure 5. The primary changes within each temperature range are summarized as follows. In the low-temperature range, adsorbed water is desorbed from the gel samples [16,17,62]. Within the medium-temperature range, imidization reactions occur through pendant amide groups [16,62,63,64], marking the onset of significant structural changes in the polymer. In the high-temperature range, processes such as polymer backbone cleavage, gel crosslinking bond breakage, and inorganic decomposition are evident [9,62,63,64]. The thermal decomposition temperature serves as a critical threshold at which the gel samples start to lose thermal stability.
In Figure 5a, the FG-95 sample shows a weight loss of approximately 4% in the temperature range of 30–110 °C, with the weight loss concentrated around 74 °C. This process corresponds to the removal of adsorbed water from the polymer gel, which is consistent with the results reported in reference [65]. For the BG-T15-95 sample (Figure 5b), dehydration spans from 30 to 143 °C, featuring notable weight losses at around 57, 85, and 116 °C. The weight loss at 116 °C corresponds to the removal of adsorbed water from magnesium salts in the brine gel [66], indicating a higher dehydration temperature relative to FG, with a total weight loss of 7.7%. In the case of the CG-T15-U1.00-95 sample (Figure 5c), dehydration occurs between 30 and 134 °C, with the maximum weight loss rate peaking around 95 °C, leading to a weight loss of approximately 2.8%. This lower weight loss suggests a reduced hydrophilicity of the inorganic components in the composite gel. The DSC curves for all gel samples during the dehydration process exhibit endothermic behavior, reflecting the energy absorbed during these transformations.
The imidization reactions for FG, BG, and CG samples occur within moderate temperature ranges of 110–446 °C, 143–738 °C, and 134–761 °C, respectively. The DSC curves for these processes exhibit exothermic behavior. Side-chain decomposition in FG, BG, and CG predominantly takes place at 357 °C, 423 °C and 427 °C. The notably higher thermal decomposition temperatures of BG and CG compared to FG suggest an improvement in thermal stability attributed to the incorporation of inorganic materials. The XRD test results indicate that magnesium ions in the composite gel form basic magnesium carbonate with crystallization water in the presence of urea [52]. In Figure 5c, the composite gel sample exhibits a noticeable weight loss around 214 °C, which corresponds to a typical endothermic peak. This process is associated with the thermal decomposition of magnesium carbonate losing its crystallization water, consistent with the results reported in reference [67].
In the high-temperature range from 446 to 900 °C for FG, 738–900 °C for BG, and 761–900 °C for CG, the DSC curves show endothermic events corresponding to the decomposition of the polymer main chain, rupture of crosslinking bonds within the gel, and decomposition of inorganic components. While FG exhibits substantial decomposition around 637 °C, nearly reaching completion by 900 °C (Figure 5a), BG and CG undergo significant decomposition only beyond their respective onset temperatures of 738 °C and 761 °C (Figure 5b,c). Consequently, due to the presence of inorganic constituents, BG and CG retain residual masses of 59.6% and 57.7%, respectively, at 900 °C.

3. Conclusions

This study innovatively employs urea as a precipitating agent to synthesize organic/inorganic composite gels, with low molecular weight PAM as the primary agent, HQ and formaldehyde serving as crosslinking agents. The decomposition of urea not only inhibits nucleophilic substitution reactions, thereby extending the gelation time, but also promotes the in situ transformation of destabilizing calcium and magnesium ions into stable inorganic particles such as calcium carbonate, magnesium carbonate, and basic magnesium carbonate. Consequently, the interplay between the organic and inorganic components results in a robust three-dimensional composite structure that significantly enhances the mechanical strength and stability of the gel.
The gelation time for these composite materials can be precisely controlled within a range of 6.6 to 14.1 days at 95 °C and 2.9 to 6.5 days at 120 °C, with stability durations reaching 155 days and 135 days, respectively. This work pioneers an efficient method for the in situ preparation of organic/inorganic composite gels, exploiting urea’s unique ability to delay high-temperature crosslinking reactions of PAM while simultaneously addressing the challenge posed by high salinity inorganic salts present in formation waters. The resulting heat-resistant and salt-tolerant composite gels represent a significant advancement in material science, offering valuable insights for profile control and water shut-off operations in high-temperature, high-salinity reservoirs.
The method reported in this study for preparing organic/inorganic composite gels demonstrates good adaptability at high temperatures (95–120 °C) and high mineralization levels (5–20 × 104 mg/L). This adaptability can meet the varying production scale needs for enhancing oil recovery in oil fields. The gelation time, gel strength, and stability of the composite gels can be easily controlled. The polymer, cross-linking agents, and urea used in this method are widely sourced and cost-effective, which can help improve the economic efficiency of oil fields and make it suitable for industrial application.

4. Materials and Methods

4.1. Materials

Acrylamide (AM, AR, 99.0%, batch number E2107072) was supplied by Shanghai Aladdin Biochemical Technology Co., Ltd (Shanghai, China). The other reagents employed in the experimental process, including potassium persulfate (K2S2O8, batch number 20230116), urea (batch number 20221104), hydroquinone (HQ, batch number 20220905), formaldehyde (37 wt%, batch number 20230322), sodium chloride, calcium chloride, and hexahydrate magnesium chloride, were all of analytical grade and purchased from Sinopharm Chemical Reagent Co., Ltd (Shanghai, China). Deionized water was used to prepare the saline solution and to synthesize polyacrylamide and gel samples.

4.2. Polyacrylamide Synthesis

Polyacrylamide was synthesized in the laboratory, and the preparation process was similar to our previous report [47], with some modifications. In a typical procedure, 30 g of AM monomer was dissolved in 160 mL of deionized water and transferred into a 500 mL three-neck round-bottom flask at 70 °C. Nitrogen gas was purged through the solution for 30 min at a flow rate of 50 mL/min to remove dissolved oxygen. Subsequently, 60 mg of K2S2O8 was dissolved in 10 mL of deionized water and added dropwise to the flask to initiate polymerization. The solution was continuously stirred under nitrogen and heated at 70 °C for 4.5 h. Then, 400 mL of deionized water was gradually added to dilute the polymer solution, resulting in a uniform polymer stock liquor with a concentration of 5 wt%. The viscosity-average molecular weight of the synthesized PAM was measured to be 1.40 × 106 g/mol using an Ubbelohde viscometer (produced by Sichuan Shubo (Group) Co., Ltd., Chongzhou, China). The degree of hydrolysis was determined to be 1.3% through titration.

4.3. Gel Sample Preparation

Sodium chloride, calcium chloride, and magnesium chloride were used to prepare a brine stock solution with a total dissolved solids of 30 × 104 mg/L, utilizing deionized water. After dilution, brine solutions with different mineralization levels were produced to simulate formation water. The composition of the brine stock solution is outlined in Table 7.
In preparing the gel reaction solution, the polymer, brine stock solutions, and deionized water were combined in specified proportions to ensure that the sample’s salinity ranged from 5 to 20 × 104 mg/L [12,48], while the polyacrylamide concentration was maintained between 0.5 and 2.0 wt%. Crosslinking agents, including formaldehyde (0.15–0.40 wt%), HQ (0.16–0.42 wt%), and thiourea stabilizer (0.05 wt%), were subsequently incorporated. Urea was then added and mixed thoroughly, resulting in a gel reaction solution, commonly referred to as gelant. After preparing the solution sample, 20 mL of the gelant was accurately transferred into an ampoule using a disposable syringe. Since thiourea was added to the sample as a stabilizer, primarily to remove oxygen in the solution, an inert gas atmosphere was not used during the sample encapsulation process. The ampoule was sealed using an alcohol blast burner, and the sample was then placed in an oven for constant temperature aging. Regular observations and tests were conducted.

4.4. Structure Characterization and Morphology Observation

The polymer samples and freshwater gel samples were freeze-dried separately and then crushed. They were subsequently mixed with KBr and ground at a mass ratio of 1:100 (polymer to KBr) before being pressed into pellets. The resulting samples were analyzed using a Bruker Vertex 70 Fourier Transform Infrared Spectrometer (manufactured by Bruker Corporation, Billerica, Massachusetts, United States) over a wavenumber range of 400 to 4000 cm−1.
To analyze the structure of the inorganic components in the composite gel, a brine solution with a mineralization degree of 15 × 104 mg/L was used. Urea was added in an amount equivalent to that of the composite gel with N = 1, excluding polymers and crosslinking agents. The sample was heated at 95 °C for 30 days to ensure adequate precipitation. Following washing with water, drying, and grinding, X-ray diffraction (XRD) analysis was conducted using a Japan Rigaku X-ray diffractometer (D/MAX-2500/PC) (manufactured by Rigaku Corporation, Tokyo, Japan) equipped with a Cu Kα radiation source (40 kV, 150 mA, λ = 1.54051 Å) to examine the crystalline structure of the inorganic components. The analysis was performed over a diffraction angle range of 2θ = 5-70 °, with a scanning speed of 10 °/min.
The morphology of the samples was examined using a field emission scanning electron microscope (JSM-6700F) from JEOL Ltd (Tokyo, Japan), facilitating a comparative analysis of the morphological characteristics of BG and CG. Due to the significant presence of inorganic salts in both BG and CG samples, inorganic salt particles precipitated from the liquid phase and adhered to the surface of the polymer during the direct observation of the gel samples. Consequently, this study also included washing the BG and CG samples with water to remove soluble inorganic salts. Afterward, the samples were freeze-dried and coated with gold for surface treatment, allowing for morphological observations to compare and analyze the characteristics of the BG and CG samples before and after washing.

4.5. Gelation Performance Observation

The evaluation of gel strength, gelation time, and stability time was conducted following the gel strength code method [8,57,58,59]. The procedure involved sealing 20 mL of gelant in an ampoule and placing it in a temperature-controlled oven for heating. The gel strength was observed and the corresponding code was recorded at certain time.

4.6. Rheological Testing

Dynamic rheological performance tests of FG, BG, and CG were conducted using a parallel plate rotational rheometer (ARES-G2) from TA Instruments, USA (New Castle, DE, USA). The primary focus was on the measurement of the elastic modulus (G′) and viscous modulus (G″), as well as the influence of factors such as urea concentration and salinity on the strength of the composite gel. During the experiment, the oscillation frequency was initially set at 1 Hz (6.28 rad/s) for strain scanning to identify the linear viscoelastic region of the sample, where the complex modulus remains constant with shear stress. Following this, frequency scanning was carried out within the established linear viscoelastic region, with oscillation frequencies ranging from 0.01 to 10 Hz, to determine G′ and G″, while maintaining the experimental temperature at 25 °C.

4.7. Thermal Stability Testing

Thermogravimetric and differential scanning calorimetry (TG-DSC) analyses of the FG, BG, and CG samples were performed using a PerkinElmer STA 8000 thermal analyzer (manufactured by PerkinElmer, Inc., Waltham, MA, USA) to assess the thermal stability of the dried gel samples. During the experiment, the samples were heated in a nitrogen atmosphere at a rate of 10 °C/min, with the temperature range set between 30 and 900 °C.

Author Contributions

Conceptualization, D.L. and J.Z.; methodology, J.Z.; software, L.L.; validation, L.L. and J.L.; formal analysis, L.L.; investigation, L.L. and J.L.; resources, D.L.; data curation, L.L. and J.Z.; writing—original draft preparation, L.L.; writing—review and editing, J.Z.; visualization, J.X.; supervision, B.Z.; project administration, J.Z.; funding acquisition, D.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 21975139), the Natural Science Foundation of Shandong Province (No. ZR2017MB042), China.

Institutional Review Board Statement

Not applicable.

Informed Consent 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 authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Moradi-Araghi, A. A review of thermally stable gels for fluid diversion in petroleum production. J. Pet. Sci. Eng. 2000, 26, 1–10. [Google Scholar] [CrossRef]
  2. Naficy, S.; Brown, H.R.; Razal, J.M.; Spinks, G.M.; Whitten, P.G. Progress toward robust polymer hydrogels. Aust. J. Chem. 2011, 64, 1007–1025. [Google Scholar] [CrossRef]
  3. Bai, B.; Zhou, J.; Yin, M. A comprehensive review of polyacrylamide polymer gels for conformance control. Pet. Explor. Develop. 2015, 42, 525–532. [Google Scholar] [CrossRef]
  4. Zhu, D.; Bai, B.; Hou, J. Polymer gel systems for water management in high-temperature petroleum reservoirs: A chemical review. Energy Fuels 2017, 31, 13063–13087. [Google Scholar] [CrossRef]
  5. Amir, Z.; Said, I.M.; Jan, B.M. In situ organically cross-linked polymer gel for high-temperature reservoir conformance control: A review. Polym. Adv. Technol. 2018, 30, 13–39. [Google Scholar] [CrossRef]
  6. Kang, W.; Kang, X.; Lashari, Z.A.; Li, Z.; Zhou, B.; Yang, H.; Sarsenbekuly, B.; Aidarova, S. Progress of polymer gels for conformance control in oilfield. Adv. Colloid Interface Sci. 2021, 289, 102363. [Google Scholar] [CrossRef]
  7. Guzmán-Lucero, D.; Martínez-Palou, R.; Palomeque-Santiago, J.F.; Vega-Paz, A.; Guzmán-Pantoja, J.; López-Falcón, D.A.; Guevara-Rodríguez, F.d.J.; García-Muñoz, N.A.; Castillo-Acosta, S.; Likhanova, N.V. Water control with gels based on synthetic polymers under extreme conditions in oil wells. Chem. Eng. Technol. 2022, 45, 998–1016. [Google Scholar] [CrossRef]
  8. Zhang, G.; Chen, L.; Ge, J.; Jiang, P.; Zhu, X. Experimental research of syneresis mechanism of HPAM/Cr3+ gel. Colloids Surf. A Physicochem. Eng. Asp. 2015, 483, 96–103. [Google Scholar] [CrossRef]
  9. Xiong, C.; Wei, F.; Li, W.; Liu, P.; Wu, Y.; Dai, M.; Chen, J. Mechanism of polyacrylamide hydrogel instability on high-temperature conditions. ACS Omega 2018, 3, 10716–10724. [Google Scholar] [CrossRef]
  10. Sengupta, B.; Sharma, V.P.; Udayabhanu, G. Gelation studies of an organically cross-linked polyacrylamide water shut-off gel system at different temperatures and pH. J. Pet. Sci. Eng. 2012, 81, 145–150. [Google Scholar] [CrossRef]
  11. Xu, Z.; Zhao, M.; Sun, N.; Meng, X.; Yang, Z.; Xie, Y.; Ding, F.; Dong, Y.; Gao, M.; Wu, Y.; 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, 135967. [Google Scholar] [CrossRef]
  12. Fang, J.; Zhang, X.; He, L.; Zhao, G.; Dai, C. Experimental research of hydroquinone (HQ)/hexamethylene tetramine (HMTA) gel for water plugging treatments in high-temperature and high-salinity reservoirs. J. Appl. Polym. Sci. 2016, 134, 43359. [Google Scholar] [CrossRef]
  13. Liu, Y.; Dai, C.; Wang, K.; Zhao, M.; Zhao, G.; Yang, S.; Yan, Z.; You, Q. New insights into the hydroquinone (HQ)-hexamethylenetetramine (HMTA) gel system for water shut-off treatment in high temperature reservoirs. J. Ind. Eng. Chem. 2016, 35, 20–28. [Google Scholar] [CrossRef]
  14. Zhu, D.; Hou, J.; Wei, Q.; Wu, X.; Bai, B. Terpolymer gel system formed by resorcinol-hexamethylenetetramine for water management in extremely high-temperature reservoirs. Energy Fuels 2017, 31, 1519–1528. [Google Scholar] [CrossRef]
  15. Bryant, S.L.; Bartosek, M.; Lockhart, T.P. Laboratory evaluation of phenol-formaldehyde/polymer gelants for high-temperature applications. J. Pet. Sci. Eng. 1997, 17, 197–209. [Google Scholar] [CrossRef]
  16. Li, Q.; Yu, X.; Wang, L.; Qu, S.; Wu, W.; Ji, R.; Luo, Y.; Yang, H. Nano-silica hybrid polyacrylamide/polyethylenimine gel for enhanced oil recovery at harsh conditions. Colloids Surf. A Physicochem. Eng. Asp. 2022, 633, 127898. [Google Scholar] [CrossRef]
  17. Tongwa, P.; Nygaard, R.; Bai, B. Evaluation of a nanocomposite hydrogel for water shut-off in enhanced oil recovery applications: Design, synthesis, and characterization. J. Appl. Polym. Sci. 2012, 128, 787–794. [Google Scholar] [CrossRef]
  18. Michael, F.M.; Fathima, A.; AlYemni, E.; Jin, H.; Almohsin, A.; Alsharaeh, E.H. Enhanced polyacrylamide polymer gels using zirconium hydroxide nanoparticles for water shutoff at high temperatures: Thermal and rheological investigations. Ind. Eng. Chem. Res. 2018, 57, 16347–16357. [Google Scholar] [CrossRef]
  19. Chen, L.; Wang, J.; Yu, L.; Zhang, Q.; Fu, M.; Zhao, Z.; Zuo, J. Experimental investigation on the nanosilica-reinforcing polyacrylamide/polyethylenimine hydrogel for water shutoff treatment. Energy Fuels 2018, 32, 6650–6656. [Google Scholar] [CrossRef]
  20. Azimi Dijvejin, Z.; Ghaffarkhah, A.; Sadeghnejad, S.; Vafaie Sefti, M. Effect of silica nanoparticle size on the mechanical strength and wellbore plugging performance of SPAM/chromium (III) acetate nanocomposite gels. Polym. J. 2019, 51, 693–707. [Google Scholar] [CrossRef]
  21. Liu, Y.; Dai, C.; Wang, K.; Zou, C.; Gao, M.; Fang, Y.; Zhao, M.; Wu, Y.; You, Q. Study on a novel cross-linked polymer gel strengthened with silica nanoparticles. Energy Fuels 2017, 31, 9152–9161. [Google Scholar] [CrossRef]
  22. Telin, A.; Safarov, F.; Yakubov, R.; Gusarova, E.; Pavlik, A.; Lenchenkova, L.; Dokichev, V. Thermal degradation study of hydrogel nanocomposites based on polyacrylamide and nanosilica used for conformance control and water shutoff. Gels 2024, 10, 846. [Google Scholar] [CrossRef]
  23. Almeida, A.I.A.d.R.; Carvalho, L.d.O.; Lopes, R.C.F.G.; Sena, L.E.B.; do Amparo, S.Z.S.; de Oliveira, C.P.M.; Calado, H.D.R.; Silva, G.G.; de Vasconcelos, C.K.B.; Viana, M.M. Enhanced polyacrylamide polymer hydrogels using nanomaterials for water shutoff: Morphology, thermal and rheological investigations at high temperatures and salinity. J. Mol. Liq. 2024, 405, 125041. [Google Scholar] [CrossRef]
  24. Pérez-Robles, S.; Cortés, F.B.; Franco, C.A. Effect of the nanoparticles in the stability of hydrolyzed polyacrylamide/resorcinol/formaldehyde gel systems for water shut-off/conformance control applications. J. Appl. Polym. Sci. 2019, 136, 47568. [Google Scholar] [CrossRef]
  25. Otsuka, T.; Chujo, Y. Poly(methyl methacrylate) (PMMA)-based hybrid materials with reactive zirconium oxide nanocrystals. Polym. J. 2010, 42, 58–65. [Google Scholar] [CrossRef]
  26. Adewunmi, A.A.; Ismail, S.; Sultan, A.S. Study on strength and gelation time of polyacrylamide/polyethyleneimine composite gels reinforced with coal fly ash for water shut-off treatment. J. Appl. Polym. Sci. 2014, 132, 41392. [Google Scholar] [CrossRef]
  27. Wang, K.; Wang, S.; Wang, X.; Wang, X.; Zheng, L.; Wen, J.; Yang, H.; Zhang, H. Modulation of syneresis rate and gel strength of PAM-PEI gels by nanosheets and their mechanisms. Colloids Surf. A Physicochem. Eng. Asp. 2025, 704, 135525. [Google Scholar] [CrossRef]
  28. Gao, G.; Du, G.; Sun, Y.; Fu, J. Self-healable, tough, and ultrastretchable nanocomposite hydrogels based on reversible polyacrylamide/montmorillonite adsorption. ACS Appl. Mater. Interfaces 2015, 7, 5029–5037. [Google Scholar] [CrossRef]
  29. Helvacıoğlu, E.; Aydın, V.; Nugay, T.; Nugay, N.; Uluocak, B.G.; Şen, S. High strength poly(acrylamide)-clay hydrogels. J. Polym. Res. 2011, 18, 2341–2350. [Google Scholar] [CrossRef]
  30. Zolfaghari, R.; Katbab, A.A.; Nabavizadeh, J.; Tabasi, R.Y.; Nejad, M.H. Preparation and characterization of nanocomposite hydrogels based on polyacrylamide for enhanced oil recovery applications. J. Appl. Polym. Sci. 2006, 100, 2096–2103. [Google Scholar] [CrossRef]
  31. Pereira, K.A.B.; Aguiar, K.; Oliveira, P.F.; Vicente, B.M.; Pedroni, L.G.; Mansur, C.R.E. Synthesis of hydrogel nanocomposites based on partially hydrolyzed polyacrylamide, polyethyleneimine, and modified clay. ACS Omega 2020, 5, 4759–4769. [Google Scholar] [CrossRef] [PubMed]
  32. Bai, Y.; Shang, X.; Wang, Z.; Zhao, X. Experimental study of low molecular weight polymer/nanoparticle dispersed gel for water plugging in fractures. Colloids Surf. A Physicochem. Eng. Asp. 2018, 551, 95–107. [Google Scholar] [CrossRef]
  33. Fu, P.; Xu, K.; Song, H.; Chen, G.; Yang, J.; Niu, Y. Preparation, stability and rheology of polyacrylamide/pristine layered double hydroxide nanocomposites. J. Mater. Chem. 2010, 20, 3869–3876. [Google Scholar] [CrossRef]
  34. Zhang, H.-P.; Cao, J.-J.; Jiang, W.-B.; Yang, Y.-Q.; Zhu, B.-Y.; Liu, X.-Y.; Wu, Y.; Sun, X.; Essouma, A.F.B.E.; Liu, J.; et al. Synthesis and mechanical properties of polyacrylamide gel doped with graphene oxide. Energies 2022, 15, 5714. [Google Scholar] [CrossRef]
  35. Shi, B.; Zhang, G.; Zhang, L.; Wang, C.; Li, Z.; Chen, F. Study on a strong polymer gel by the addition of micron graphite oxide powder and its plugging of fracture. Gels 2024, 10, 304. [Google Scholar] [CrossRef]
  36. Moradi-Araghi, A.; Doe, P.H. Hydrolysis and precipitation of polyacrylamides in hard brines at elevated temperatures. SPE Reserv. Eng. 1987, 2, 189–198. [Google Scholar] [CrossRef]
  37. Ma, Q.; Shuler, P.J.; Aften, C.W.; Tang, Y. Theoretical studies of hydrolysis and stability of polyacrylamide polymers. Polym. Degrad. Stab. 2015, 121, 69–77. [Google Scholar] [CrossRef]
  38. Marandi, G.B.; Esfandiari, K.; Biranvand, F.; Babapour, M.; Sadeh, S.; Mahdavinia, G.R. pH sensitivity and swelling behavior of partially hydrolyzed formaldehyde-crosslinked poly(acrylamide) superabsorbent hydrogels. J. Appl. Polym. Sci. 2008, 109, 1083–1092. [Google Scholar] [CrossRef]
  39. Fong, D.W.; Kowalski, D.J. An investigation of the crosslinking of polyacrylamide with formaldehyde using 13C nuclear magnetic resonance spectroscopy. J. Polym. Sci., Part A Polym. Chem. 1993, 31, 1625–1627. [Google Scholar] [CrossRef]
  40. Cheng, L.; Qin, Y.; Su, Y.; Pan, Y.; Wang, Y.; Liao, R.; Li, Z. Development of a high-strength and adhesive polyacrylamide gel for well plugging. ACS Omega 2022, 7, 6151–6159. [Google Scholar] [CrossRef]
  41. Ma, Z.; Zhao, M.; Yang, Z.; Wang, X.; Dai, C. Development and gelation mechanism of ultra-high-temperature-resistant polymer gel. Gels 2023, 9, 726. [Google Scholar] [CrossRef] [PubMed]
  42. Zhu, D.; Hou, J.; Meng, X.; Zheng, Z.; Wei, Q.; Chen, Y.; Bai, B. Effect of different phenolic compounds on performance of organically cross-linked terpolymer gel systems at extremely high temperatures. Energy Fuels 2017, 31, 8120–8130. [Google Scholar] [CrossRef]
  43. Reddy, B.R.; Eoff, L.; Dalrymple, E.D.; Black, K.; Brown, D.; Rietjens, M. A natural polymer-based crosslinker system for conformance gel systems. SPE J. 2003, 8, 99–106. [Google Scholar] [CrossRef]
  44. Li, Z.-Y.; Li, X.-G.; Du, K.; Liu, H.-K. Development of a new high-temperature and high-strength polymer gel for plugging fractured reservoirs. Upstream Oil Gas Technol. 2020, 5, 100014. [Google Scholar] [CrossRef]
  45. Fox, K.B.; Evans, A., Jr. Acceleration of Gelation of Water Soluble Polymers. United States patent US 5447986, 5 September 1995. [Google Scholar]
  46. Ren, Q.; Jia, H.; Yu, D.; Pu, W.-F.; Wang, L.-L.; Li, B.; Yang, J.-J.; Ni, J.-H.; Chen, L. New insights into phenol-formaldehyde-based gel systems with ammonium salt for low-temperature reservoirs. J. Appl. Polym. Sci. 2014, 131, 40657. [Google Scholar] [CrossRef]
  47. Li, Y.; Wei, Q.; Wang, R.; Zhao, J.; Quan, Z.; Zhan, T.; Li, D.; Xu, J.; Teng, H.; Hou, W. 3D hierarchical porous nitrogen-doped carbon/Ni@NiO nanocomposites self-templated by cross-linked polyacrylamide gel for high performance supercapacitor electrode. J. Colloid Interface Sci. 2020, 570, 286–299. [Google Scholar] [CrossRef]
  48. Lu, G.; Zhao, J.; Li, S.; Chen, Y.; Li, C.; Wang, Y.; Li, D. Incorporation of partially hydrolyzed polyacrylamide with zwitterionic units and poly(ethylene glycol) units toward enhanced tolerances to high salinity and high temperature. Front. Mater. 2021, 8, 788746. [Google Scholar] [CrossRef]
  49. Yang, H.; Yan, R.; Chen, H.; Lee, D.H.; Zheng, C. Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel 2007, 86, 1781–1788. [Google Scholar] [CrossRef]
  50. Wang, C. Control the polymorphism and morphology of calcium carbonate precipitation from a calcium acetate and urea solution. Mater. Lett. 2008, 62, 2377–2380. [Google Scholar] [CrossRef]
  51. De Silva, P.; Bucea, L.; Sirivivatnanon, V. Chemical, microstructural and strength development of calcium and magnesium carbonate binders. Cem. Concr. Res. 2009, 39, 460–465. [Google Scholar] [CrossRef]
  52. Douglas, T.E.L.; Łapa, A.; Samal, S.K.; Declercq, H.A.; Schaubroeck, D.; Mendes, A.C.; Van der Voort, P.; Dokupil, A.; Plis, A.; De Schamphelaere, K.; et al. Enzymatic, urease-mediated mineralization of gellan gum hydrogel with calcium carbonate, magnesium-enriched calcium carbonate and magnesium carbonate for bone regeneration applications. J. Tissue Eng. Regener. Med. 2017, 11, 3556–3566. [Google Scholar] [CrossRef] [PubMed]
  53. Krajewska, B. Urease-aided calcium carbonate mineralization for engineering applications: A review. J. Adv. Res. 2018, 13, 59–67. [Google Scholar] [CrossRef] [PubMed]
  54. Wu, Y.; Ajo-Franklin, J.B.; Spycher, N.; Hubbard, S.S.; Zhang, G.; Williams, K.H.; Taylor, J.; Fujita, Y.; Smith, R. Geophysical monitoring and reactive transport modeling of ureolytically-driven calcium carbonate precipitation. Geochem. Trans. 2011, 12, 7. [Google Scholar] [CrossRef] [PubMed]
  55. Digiacomo, P.M.; Schramm, C.M. Mechanism of polyacrylamide gel syneresis determined by C-13 NMR. Paper SPE 11787. In Proceedings of the International Symposium on Oilfield and Geothermal Chemistry, Denver, CO, USA, 1–3 June 1983; Society of Petroleum Engineers: Houston, TX, USA, 1983. [Google Scholar] [CrossRef]
  56. Rauner, N.; Meuris, M.; Dech, S.; Godde, J.; Tiller, J.C. Urease-induced calcification of segmented polymer hydrogels-A step towards artificial biomineralization. Acta. Biomater. 2014, 10, 3942–3951. [Google Scholar] [CrossRef]
  57. Hakiki, F.; Arifurrahman, F. Cross-linked and responsive polymer: Gelation model and review. J. Ind. Eng. Chem. 2023, 119, 532–549. [Google Scholar] [CrossRef]
  58. Xu, P.; Shang, Z.; Yao, M.; Li, X. Mechanistic insight into improving strength and stability of hydrogels via nano-silica. J. Mol. Liq. 2022, 357, 119094. [Google Scholar] [CrossRef]
  59. Zhu, D.; Hou, J.; Chen, Y.; Wei, Q.; Zhao, S.; Bai, B. Evaluation of terpolymer-gel systems crosslinked by polyethylenimine for conformance improvement in high-temperature reservoirs. SPE J. 2019, 24, 1726–1740. [Google Scholar] [CrossRef]
  60. Albonico, P.; Lockhart, T.P. Divalent ion-resistant polymer gels for high-temperature applications: Syneresis Inhibiting Additives. Paper SPE 25220. In Proceedings of the SPE International Symposium on Oilfield Chemistry, New Orleans, LA, USA, 2–5 March 1993; Society of Petroleum Engineers: Houston, TX, USA, 1993. [Google Scholar] [CrossRef]
  61. Gu, C.; Lv, Y.; Fan, X.; Zhao, C.; Dai, C.; Zhao, G. Study on rheology and microstructure of phenolic resin crosslinked nonionic polyacrylamide (NPAM) gel for profile control and water shutoff treatments. J. Pet. Sci. Eng. 2018, 169, 546–552. [Google Scholar] [CrossRef]
  62. Caulfield, M.J.; Qiao, G.G.; Solomon, D.H. Some aspects of the properties and degradation of polyacrylamides. Chem. Rev. 2002, 102, 3067–3084. [Google Scholar] [CrossRef]
  63. Burrows, H.D.; Ellis, H.A.; Utah, S.I. Adsorbed metal ions as stabilizers for the thermal degradation of polyacrylamide. Polymer 1981, 22, 1740–1744. [Google Scholar] [CrossRef]
  64. Yang, M.-H. The Two-stages thermal degradation of polyacrylamide. Polym. Test 1998, 17, 191–198. [Google Scholar] [CrossRef]
  65. Ahmadyari-Sharamin, M.; Hassanzadeh-Tabrizi, S.A. Polyacrylamide gel synthesis, characterization, and optical properties of Co1-xNixCr2O4 spinel nanopigment. J. Sol-Gel Sci. Technol. 2021, 99, 560–566. [Google Scholar] [CrossRef]
  66. Lee, K.E.; Khan, I.; Morad, N.; Teng, T.T.; Poh, B.T. Thermal behavior and morphological properties of novel magnesium salt–polyacrylamide composite polymers. Polym. Compos. 2011, 32, 1515–1522. [Google Scholar] [CrossRef]
  67. Bhattacharjya, D.; Selvamani, T.; Mukhopadhyay, I. Thermal decomposition of hydromagnesite effect of morphology on the kinetic parameters. J. Therm. Anal. Calorim. 2012, 107, 439–445. [Google Scholar] [CrossRef]
Figure 1. FR-IR spectra of PAM and FG (a), XRD curves of inorganic samples (b).
Figure 1. FR-IR spectra of PAM and FG (a), XRD curves of inorganic samples (b).
Gels 11 00256 g001
Scheme 1. Schematic diagram of inorganic particles deposited on the surface of polyacrylamide.
Scheme 1. Schematic diagram of inorganic particles deposited on the surface of polyacrylamide.
Gels 11 00256 sch001
Figure 2. SEM images of BG-T15-95 (ac) and CG-T15-U1.00-95 (d,e). (a,d) (without washing); (b,e) (after washing); (c,f) (magnified images after washing).
Figure 2. SEM images of BG-T15-95 (ac) and CG-T15-U1.00-95 (d,e). (a,d) (without washing); (b,e) (after washing); (c,f) (magnified images after washing).
Gels 11 00256 g002
Figure 3. CG gelation time (a) and stability time (b) curves vs. Curea at 95 °C and 120 °C. Total dissolved solids and CPAM are 15 × 104 mg/L and 1%, respectively.
Figure 3. CG gelation time (a) and stability time (b) curves vs. Curea at 95 °C and 120 °C. Total dissolved solids and CPAM are 15 × 104 mg/L and 1%, respectively.
Gels 11 00256 g003
Figure 4. Elastic modulus and viscous modulus of gel samples aged 44.7 days at 95 °C (a), elastic modulus of gel samples aged 7.3 days at 120 °C (b), effects of salinity on the elastic modulus of CG samples aged 44.7 days at 95 °C (c).
Figure 4. Elastic modulus and viscous modulus of gel samples aged 44.7 days at 95 °C (a), elastic modulus of gel samples aged 7.3 days at 120 °C (b), effects of salinity on the elastic modulus of CG samples aged 44.7 days at 95 °C (c).
Gels 11 00256 g004
Figure 5. TG-DSC-DTG curves of FG (a), BG (b) and CG (c). (CPAM = 1.0 wt%, CHQ = 0.20 wt%, CHCHO = 0.21 wt%. For BG and CG preparation, salinity is fixed as 15 × 104 mg/L, and N is selected as 1.00 to obtain CG.
Figure 5. TG-DSC-DTG curves of FG (a), BG (b) and CG (c). (CPAM = 1.0 wt%, CHQ = 0.20 wt%, CHCHO = 0.21 wt%. For BG and CG preparation, salinity is fixed as 15 × 104 mg/L, and N is selected as 1.00 to obtain CG.
Gels 11 00256 g005
Table 1. Gel strength code and description.
Table 1. Gel strength code and description.
Gel Strength CodeDescription
ANo detectable gel formed: The viscosity of the gelant is as same as that of polymer solution.
BHighly flowable gel: The viscosity of the gelant is slightly higher than that of the initial polymer solution.
CFlowable gel: The majority of the gelant can flow upon inversion.
DMedium flowable gel: 90–95% of the gel can flow upon inversion.
ESlightly flowable gel: less than 85% of the gel can flow upon inversion.
FHighly deformable gel: The gel is highly deformed but does not flow upon inversion.
GModerately deformable gel: The gel does not flow and the gel deforms to the middle of the bottle upon inversion.
HSlightly deformable gel: The gel does not flow and only the surface of the gel undergoes slight deformation upon inversion.
IRigid gel: The surface of the gel does not deform upon inversion.
Table 2. Gel sample strength code at 95 °C across different time intervals (CPAM = 1.0 wt%, CHQ = 0.20 wt%, CHCHO = 0.21 wt%).
Table 2. Gel sample strength code at 95 °C across different time intervals (CPAM = 1.0 wt%, CHQ = 0.20 wt%, CHCHO = 0.21 wt%).
Samplestime/d
0.52.54.65.46.67.98.49.79.811.212.414.124.535.036.040.842.078.0122155
FG- 95ABBBBBBDDFGHIIIIIIII
BG-T5-95AABBBBBBBCDEFHII3 *I5///
BG-T10-95AABBBBBBBDDEII0.5I1I5////
BG-T15-95ABBBBCDEEFFFHH5//////
BG-T20-95ABBBCDEFFHHH0.5H5///////
CG-T15-U0.10-95ADHIIIIIIIII0.1F0.5F4F5/////
CG-T15-U0.50-95ABDEHHHIIIIIIIIIII5//
CG-T15-U0.75-95ABCDGGHIIIIIIIIIIII5/
CG-T15-U1.00-95ABBCDFGIIIIIIIIIIIII
CG-T15-U1.50-95ABBBBBBCDEFFIIIIIIII
CG-T15-U2.00-95ABBBBBBBBBCDIIIIIIII
* The superscript number on the right of the gel strength symbol represents the dehydration volume (in mL) of a 20 mL gel sample. Stop observation when the dehydration volume reaches 5 mL.
Table 3. Gel sample strength code at 120 °C across different time intervals (CPAM = 1.0 wt%, CHQ = CHCHO = 0.16 wt%).
Table 3. Gel sample strength code at 120 °C across different time intervals (CPAM = 1.0 wt%, CHQ = CHCHO = 0.16 wt%).
Samplestime/d
0.51.41.81.92.32.72.93.04.55.66.57.59.714.315.820.731.6126135
FG- 120ABBBCDDDHHHIIIIIII0.2I0.2
BG-T5-120ABBBBCCDGH0.1 *H0.2H0.5H2H4H5////
BG-T10-120ABBBBCDGGH0.5H0.5H1H3H5/////
BG-T15-120ABBBDFFFF0.5F0.5F1F3F5//////
BG-T20-120ABCDFGGGHH0.5H1H5///////
CG-T15-U0.10-120ADHHF0.5F0.5F0.5F0.5F0.5F1F2F2.5F5//////
CG-T15-U0.50-120ABDDFGGHIIIII0.5I0.5I1I5///
CG-T15-U0.75-120ABCDEFFFGGGHI0.2I0.5H1H2H5//
CG-T15-U1.00-120ABBBBCDDFGHHHHIIII2I2.5
CG-T15-U1.50-120ABBBBBBBCDGHHHIIII1I1.5
CG-T15-U2.00-120ABBBBBBBBCDFGHIIII0.5I0.8
* The superscript number on the right of the gel strength symbol represents the dehydration volume (in mL) of a 20 mL gel sample. Stop observation when the dehydration volume reaches 5 mL.
Table 4. Gelation time (tD) and stability time (tS) of FG, BG samples.
Table 4. Gelation time (tD) and stability time (tS) of FG, BG samples.
SamplestD/dtS/d
FG-959.7>176
BG-T5-9512.442.0
BG-T10-9511.240.8
BG-T15-958.435.0
BG-T20-957.924.5
FG-1202.7>126
BG-T5-1203.015.8
BG-T10-1202.914.3
BG-T15-1202.39.7
BG-T20-1201.97.5
At 95 °C, CPAM = 1.0 wt%, CHQ = 0.20 wt%, CHCHO = 0.21 wt%; At 120 °C, CPAM = 1.0 wt%, CHQ = CHCHO = 0.16 wt%.
Table 5. Effects of polymer concentration and crosslinker concentration on gelation time and stability time of composite gel CG-T15-U0.5-95.
Table 5. Effects of polymer concentration and crosslinker concentration on gelation time and stability time of composite gel CG-T15-U0.5-95.
CPAM/%CHQ/wt%CHCHO/wt%tD/dtS/d
1.00.060.0712.669.2
1.00.120.147.771.0
1.00.200.214.678.0
1.00.250.271.956.1
1.00.310.331.643.9
0.50.200.219.137.8
1.50.200.211.257.0
2.00.200.210.757.0
Table 6. Effects of salinity and urea concentration on gelation time and stability time of composite gels.
Table 6. Effects of salinity and urea concentration on gelation time and stability time of composite gels.
SamplesCurea/wt%tD/dtS/d
CG-T5- U0.50-950.232.185.2
CG-T10- U0.5-950.452.876.7
CG-T15- U0.5-950.684.678.0
CG-T20- U0.5-950.905.272.1
CG-T5- U1.0-950.452.2167.4
CG-T10- U1.0-950.903.6161.2
CG-T15- U1.0-951.356.6155.0
CG-T20- U1.0-951.817.0170.0
Table 7. Composition of brine stock solution.
Table 7. Composition of brine stock solution.
Ion Concentration/mg·L−1Total Dissolved Solids/mg·L−1
Na+Ca2+Mg2+Cl
99,998.76816.06818.6186,366.73.0 × 105
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

Liang, L.; Li, J.; Li, D.; Xu, J.; Zheng, B.; Zhao, J. Urea Delays High-Temperature Crosslinking of Polyacrylamide for In Situ Preparation of an Organic/Inorganic Composite Gel. Gels 2025, 11, 256. https://doi.org/10.3390/gels11040256

AMA Style

Liang L, Li J, Li D, Xu J, Zheng B, Zhao J. Urea Delays High-Temperature Crosslinking of Polyacrylamide for In Situ Preparation of an Organic/Inorganic Composite Gel. Gels. 2025; 11(4):256. https://doi.org/10.3390/gels11040256

Chicago/Turabian Style

Liang, Li, Junlong Li, Dongxiang Li, Jie Xu, Bin Zheng, and Jikuan Zhao. 2025. "Urea Delays High-Temperature Crosslinking of Polyacrylamide for In Situ Preparation of an Organic/Inorganic Composite Gel" Gels 11, no. 4: 256. https://doi.org/10.3390/gels11040256

APA Style

Liang, L., Li, J., Li, D., Xu, J., Zheng, B., & Zhao, J. (2025). Urea Delays High-Temperature Crosslinking of Polyacrylamide for In Situ Preparation of an Organic/Inorganic Composite Gel. Gels, 11(4), 256. https://doi.org/10.3390/gels11040256

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