Next Article in Journal
Experimental Study of Dye Degradation in a Single-Jet Cavitation System
Previous Article in Journal
Molecularly Imprinted Polymers for Pollutant Capture and Degradation: A Snapshot Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analysis of Carbon Dioxide Mineralization in Carbonates from Tampico-Misantla Basin, Mexico: Effect of Organic Matter Content

by
Roxana López-Dinorín
1,
Ana María Mendoza-Martínez
1,*,
Diana Palma-Ramírez
2,*,
Héctor Dorantes-Rosales
3,
Ricardo García-Alamilla
1,
Issis Claudette Romero-Ibarra
4 and
David Salvador García-Zaleta
5
1
Tecnológico Nacional de México—Instituto Tecnológico de Ciudad Madero, Centro de Investigación en Petroquímica, Prol. Bahía de Aldhair y Av. de las Bahías, Parque de la Pequeña y Mediana Industria, Altamira 89600, Tamaulipas, Mexico
2
Unidad Profesional Interdisciplinaria de Ingeniería Campus Hidalgo (UPIIH), Department of Polymers, Instituto Politécnico Nacional (IPN), San Agustín Tlaxiaca 42162, Hidalgo, Mexico
3
Department of Metallurgical and Materials Engineering, Escuela Superior de Ingeniería Química e Industrias Extractivas (ESIQIE), Instituto Politécnico Nacional (IPN), Mexico City 07738, Mexico
4
Instituto Politécnico Nacional—UPIITA, Mexico City 02580, Mexico
5
Department of División Académica Multidisciplinaria de Jalpa de Méndez, Universidad Juárez Autónoma de Tabasco (UJAT), Villahermosa 86690, Tabasco, Mexico
*
Authors to whom correspondence should be addressed.
Processes 2025, 13(4), 1087; https://doi.org/10.3390/pr13041087
Submission received: 1 March 2025 / Revised: 31 March 2025 / Accepted: 1 April 2025 / Published: 4 April 2025
(This article belongs to the Section Environmental and Green Processes)

Abstract

:
The pursuit of effective climate change mitigation strategies is driving research into geological carbon dioxide (CO2) storage. The present work explores the interaction of CO2 with carbonate rocks from the El Abra formation in the Tampico-Misantla basin, focusing on the comparative influence of organic matter (OM) content on mineralization processes, hypothesizing that variations in OM content significantly modulate the mineralization process affecting both the rate and type of carbonate formation. Expanding on a previous study, CO2 is studied and injected under high-pressure (1350-2350 PSI) and high-temperature (60–110 °C) conditions into two contrasting samples: one with high OM content and another with low OM content. Structural, morphological, and physical adsorption changes were evaluated through Fourier transform infrared (FT-IR) spectroscopy, X-ray diffraction (XRD), scanning electron microscopy (SEM), and Brunauer–Emmett–Teller (BET) analyses. The findings indicate that the mineralogy of El Abra promotes secondary carbonate precipitation, with rock–fluid interactions significantly enhanced by brine presence. Samples with high OM exhibited a dramatic reduction in average particle size from 13 μm to 2 μm, along with the formation of metastable phases, such as vaterite—evidenced by XRD peak shifting and modifications in the FT-IR spectrum of carbonate bands. Meanwhile, low-OM samples showed an increase in particle size from 1.6 μm to between 3.26 and 4.12 μm, indicating predominant recrystallization. BET analysis confirmed a significant porosity enhancement in high-OM samples (up to 2.918 m2/g). Therefore, OM content plays a critical role in modulating both the rate and type of mineralization, potentially enhancing physical storage capacity in low-OM samples. These integrated findings demonstrate that OM critically governs calcite dissolution, secondary carbonate formation, and microstructural evolution, providing key insights for optimizing CO2 storage in complex carbonate reservoirs.

1. Introduction

The mineralization of carbon dioxide (CO2) in geological formations has been established as a key strategy for capturing and permanently storing carbon, standing out for its ability to convert CO2 into stable carbonates through rock–water–fluid interactions [1]. In the context of climate change, the search for effective and sustainable solutions to reduce CO2 emissions has led to a growing interest in geological storage technologies [2]. This process is crucial for mitigating climate change, particularly in areas where lithological and mineralogical characteristics facilitate efficient chemical reactions [3]. Globally, different strategies for carbon capture have been studied. However, challenges remain, associated with the process efficiency, long-term safety, and the variability of available geological formations. This phenomenon is relevant in diverse geological environments, such as carbonate rocks, where factors including mineralogical composition, porosity, and thermodynamic conditions play a crucial role in achieving practical carbon storage [4].
The Tampico-Misantla Basin in Mexico is one of the most significant hydrocarbon-producing regions in the country, characterized by a complex lithological framework including the El Abra Formation, a carbonate-based formation recognized for its high potential for CO2 storage. Located in eastern Mexico along the Gulf Coast, between the states of Tamaulipas and Veracruz, the Tampico-Misantla Basin’s geographical context is shown in Figure 1. Recent studies have identified two principal aquifers, the Cretaceous Tamaulipas and Jurassic Cahuasas formations, as suitable candidates for CO2 storage, with an estimated adequate storage capacity of 972.3 to 1346.0 million tons of CO2 [5,6].
The CNH has extensively studied the El Abra Formation due to its significant hydrocarbon reserves. This formation developed within the passive margin of the Gulf of Mexico (as documented in the Atlas Geológico Tampico-Misantla). The passive margin was formed during the Jurassic as a result of the separation of the North and South American, and African continental masses, creating a tectonically stable environment. This environment facilitated the accumulation of large volumes of carbonate sediments, characteristic of the El Abra Formation [8]. The tectonic and stratigraphic framework described by the Atlas highlights carbonate platform depositional environments, with a predominance of reef and high-energy facies, which has favored the development of porosity and permeability [5]. Moreover, early oil production in the Tampico-Misantla Basin initially targeted the Medium Cretaceous El Abra Formation, underscoring its historical and economic significance [6].
Despite its potential, studies on CO2 mineralization in these formations have been limited, raising the need to evaluate their feasibility and efficiency in the Mexican geological context. Recent research in the Gulf of Mexico suggests that carbonate formations are more likely to favor the precipitation of secondary carbonates, such as aragonite and dolomite, under conditions of CO2 injection [9]. Carbonate rocks, such as limestone and dolomite, are among the most abundant sedimentary rocks on Earth and play a crucial role in geological CO2 storage due to their high reactivity with CO2. The mineralization of CO2 in these rocks involves conversion into stable carbonates, such as calcite and dolomite, through interactions between the rock, water, and fluids [2]. This process is particularly effective in carbonate formations, where the presence of calcium and magnesium ions facilitates the precipitation of secondary carbonates [3,10].
According to the CNH (2016), the El Abra Formation is located at variable depths ranging from 500 to 1800 m in terrestrial zones and 1500 to 2800 m in shallow-water areas (Figure 2a). These depth variations are attributed to subsidence processes associated with the passive margin, which also promoted hydrocarbon migration and emplacement (as documented in the Atlas Geológico Tampico-Misantla) [8]. Additionally, the architecture and diagenetic history of these rocks significantly influence their storage potential, as illustrated in the stratigraphic column (Figure 2b).
For example, studies in the Lesser Himalayas have demonstrated the importance of fault and fracture networks in controlling hydrothermal dolomitization in Jurassic carbonates [11]. Similarly, research related to the Trans Indus Ranges highlights the complex diagenetic evolution of Jurassic carbonates and its impact on reservoir properties [12]. These studies emphasize the importance of understanding the diagenetic history and structural characteristics of carbonate formations when evaluating their suitability for CO2 storage. Mineralization in these rocks is particularly relevant, since it contributes to the long-term stability of stored CO2 by forming stable carbonates [13]. Despite these findings, a debate persists regarding the comparative efficiency of formations in different geological settings [12], underscoring the need for a more detailed analysis.
Previous studies have shown that mineralization in carbonate rocks can improve storage security as carbonates form and seal the pores of the rocks, strengthening their containment capacity [14]. The role of organic matter (OM) in these rocks is critical as it can influence the porosity, permeability, and reactivity of the rock matrix. OM-rich carbonates tend to exhibit high rates of secondary carbonate precipitation, but the resulting minerals may be less stable over time. In contrast, low-OM carbonates may favor physical trapping mechanisms, where CO2 is stored in pore spaces without significant chemical reactions [10]. Microscale lithological heterogeneity, such as fractures and secondary porosity, is a critical factor determining the success of mineralization, as observed in reactive transport models applied in the Gulf of Mexico [1].
The El Abra Formation exhibits a depositional architecture characteristic of carbonate platforms, with reef and high-energy facies predominating. These features facilitated the development of porosity and permeability, which are critical for CO2 storage. The formation has also undergone diagenetic processes, including cementation, dissolution, and dolomitization, which have modified its microstructure and storage properties, as highlighted in the Atlas Geológico de la Cuenca Tampico-Misantla. The Tampico-Misantla basin, including the Abra Formation, has a complex tectonic and stratigraphic history influenced by the formation of extensive carbonate platforms, as characterized by its tectonic and stratigraphic history, documented by the National Hydrocarbons Commission (CNH).
Additionally, the rocks of the El Abra Formation have undergone a range of diagenetic processes, including cementation, dissolution, and dolomitization, which have significantly altered their microstructure and storage properties. These processes have played a crucial role in modifying porosity and permeability, influencing the formation’s capacity for CO2 storage. For instance, dolomitization and secondary porosity development have enhanced the formation’s ability to host and retain CO2. The Atlas emphasizes that the structural and diagenetic evolution of these carbonate rocks is a critical factor in assessing their suitability for long-term carbon storage solutions [5,6,8,15].
The use of advanced analytical techniques, such as Fourier transform infrared (FTIR) spectroscopy, X-ray diffraction (XRD), scanning electron microscopy (SEM), and Brunauer–Emmett–Teller (BET) analysis, has proven to be essential for characterizing structural, morphological, and physical adsorption changes in rocks after CO2 injection [16,17,18,19,20]. For example, calcification and leaching with ammonium salts in industrial asphalt waste have proven to be effective strategies for producing calcium carbonates, such as vaterite, a metastable polymorph with industrial applications [21]. Similarly, the controlled synthesis of pure vaterite using CO2 storage materials (CO2SM) highlights the importance of optimizing parameters such as pH and Ca2⁺ concentration to maximize carbonate precipitation [22]. These advances enable a deeper understanding of induced mineralization; however, questions remain about the efficiency of these processes and how different factors influence their impact on each geological formation.
Internationally, projects such as CarbFix in Iceland and Wallula in the USA have demonstrated that the injection of CO2 into geological formations generates accelerated mineralization, yielding promising results within a short period, typically less than two years [23]. However, the applicability of those models in regions with different geological characteristics, such as the Tampico-Misantla basin, remains an open area of research. These studies reinforce the viability of mineralization in carbonate rocks under controlled conditions [24].
Additionally, technologies are being developed for the use of industrial waste in the mineralization of CO2, highlighting the use of carbide slag, an alkaline waste generated in industrial processes, as a viable alternative for carbon capture. This approach is particularly relevant since it allows for the simultaneous use of CO2 emissions and solid waste from the same industry due to their contents of calcium and magnesium ions, which favor the mineralization of stable phases such as calcite. Recent studies have shown that pretreatment strategies, such as slag calcification and the use of ultrasound and magnetic stirring, can significantly improve CO2 mineralization efficiency, achieving capture rates of up to 637 kg of CO2 per ton of carbide slag, with a purity of the mineralized product of more than 94% [25].
On the other hand, the implementation of CO2 storage projects faces both technical and operational challenges, which have driven the development of emerging technologies, such as the use of artificial intelligence, for monitoring and optimizing these processes. Machine learning has been used at various stages of geological CO2 storage, including evaluating injection sites, predicting storage capacity, and monitoring potential CO2 leaks. This technological integration has allowed us to improve the security and reliability of the projects, facilitating their large-scale implementation [26]. However, the effectiveness of these approaches still depends on the availability of high-resolution data and the adaptation of algorithms to specific geological conditions.
Therefore, this research aims to evaluate the influence of OM during CO2 mineralization in the El Abra formation of the Tampico-Misantla Basin, exploring the differentiated impact on the mineralization process. This work aims to answer the following question: How does the content of organic matter influence CO2 mineralization in carbonate rocks under simulated reservoir conditions?
In Mexico, the General Law on Climate Change establishes the basis for implementing political strategies to address climate change, ensuring the right to a healthy environment and reducing greenhouse gas emissions. The National Climate Change Strategy, aligned with international commitments such as the Paris Agreement, defines ambitious emission reduction goals, including a 22% reduction in greenhouse gases by 2030 [27,28]. These commitments have driven the need to develop new CO2 capture and storage technologies, as well as to continue scientific research in this field. Within this context, evaluating mineralization in Mexican geological formations is crucial to determine its viability as an effective mitigation solution.
A previous study reported the evaluation of the Zuloaga, Agua Nueva, and Abra formations in the Tampico-Misantla basin. OM was analyzed through FTIR and UV-Vis methods, which showed evidence of a high content of OM. Selected rocks from the Abra formation were assessed by comparing the effects of pressure (980, 1250, and 1500 PSI), temperature (60 °C), time (2.5 and 24 h), and the presence or absence of brine (B or WB) on the structural, morphological, and physical adsorption properties after CO2 injection. In the Tampico-Misantla Basin, the presence of organic matter in carbonate formations, such as the El Abra and Agua Nueva formations, has enhanced the precipitation of secondary carbonates. However, the stability of these minerals may vary depending on the OM content [16,29]. Additionally, the lithological heterogeneity of the basin, including fractures and secondary porosity, plays a critical role in determining the success of CO2 mineralization [30]. This work is an extension that aims to advance scientific knowledge of the mineralization of CO2 in the Abra formation for evaluating its potential application in the capture and storage of carbon, focusing on higher temperatures and pressures than 60 °C and 1500 PSI in the presence (B) and absence (WB) of brine. A comparison with a low-OM rock is also analyzed to determine if it generates significant differences in mineralization, with a more pronounced development of secondary carbonates and a predominant physical storage mechanism. FTIR spectroscopy, XRD, SEM, and BET methods are used to provide valuable information on the long-term potential and optimize geological storage strategies in Mexico, laying the foundation for future research and offering scientific evidence for developing public policies focused on climate change mitigation.
Furthermore, recent studies on the Late Cretaceous–Paleocene stratigraphic and structural evolution of central Mexico have demonstrated that foreland basin deposition and regional tectonic processes have significantly influenced sediment provenance and diagenetic alterations in adjacent carbonate platforms, including the El Abra Formation [31]. This broader tectonic context reinforces the importance of considering regional structural evolution when evaluating the CO2 storage potential of the El Abra Formation.

2. Materials and Methods

2.1. Selection Sample

The National Lithotheque of the Hydrocarbon Industry provided the samples from the CNH (Hidalgo, Mexico). Fragments of representative cores of the El Abra geological formation, located in the Tampico-Misantla basin, were selected based on a prior assessment of their mineralogical composition and organic matter content, as detailed in our previous work [30]. Details of the selected samples, including their depth, OM content, age, and hydrocarbon-generating potential, are summarized in Table 1. Different rock types were analyzed, including El Abra, found to be predominantly composed of calcite, with varying levels of organic matter. The El Abra formation was chosen due to its geological importance, extensive study in the petroleum sector, and marine origin, as recognized by the CNH as one of the most significant reservoirs in the Tampico-Misantla basin. Its mineralogical composition and structural composition properties make it highly representative of carbonate reservoirs suitable for CO2 mineralization studies.

2.2. Sample Preparation

The rocks were dried at 40 °C for 24 h. Subsequently, some samples were saturated with a 25,000 ppm saline solution (NaCl) under vacuum conditions to remove air from the pores, while others were saturated with distilled water. This was performed to create a low-pressure environment during saturation and ensure uniform brine distribution in core samples. The reservoir was simulated using a saline solution, a representative condition for reservoirs in carbonate formation. This was ensured through monitoring weight changes. The confining pressure values, ranging from 16 to 46 MPa, were selected to represent the confining pressure at the site where the samples were collected, starting from the shallowest depth suitable for storage, i.e., 800 m, and extending to a deeper depth where the confining stress reaches 46 MPa. This method is consistent with other similar reports [32,33,34,35].

2.3. CO2 Injection in High Pressure and Temperature Reactor

CO2 injection experiments were conducted using a Parr model 4547 reactor, coupled to a Parr 4848 controller (Parr Instrument Company, Davenport, FL, USA), which is capable of reaching temperatures of up to 500 °C and pressures of up to 5000 PSI. The experimental procedure is shown in Figure 3.
The gas used was industrial-grade, with a purity of 99.9%. The samples were placed inside the reactor, and CO2 was injected at pressures between 1350 and 2350 PSI and temperatures from 60 to 110 °C for 24 h. These conditions were selected to replicate the subsurface environment of the Tampico-Misantla basin, where the rock samples were collected. According to records from the National Lithotheque of the Hydrocarbon Industry of the CNH [36], the samples were extracted from depths ranging between 1000 and 1500 m. Based on this depth range, reservoir temperatures and pressures were estimated using an average geothermal gradient of 25–30 °C/km and a hydrostatic pressure gradient of approximately 10 MPa/km (~1450 psi/km). These estimations yielded a temperature between 60 and 110 °C and a pressure range of 980–1500 PSI (6.8–10.3 MPa), which are consistent with subsurface conditions in carbonate formations at those depths. The specific experimental conditions for each sample are summarized in Table 2 [37]. The conditions in other studies ranged from 8 to 30 MPa (1160–4350 PSI) [4,23,29]. The range falls within the spectrum, particularly on the lower end, which is consistent with the depth-dependent pressure and temperature conditions of carbonate formations in this region. For ease of reading, specific codes were assigned to the samples, where the subscripts t, P, and T represent time, pressure, and temperature, respectively. Experiments were performed in triplicate.

2.4. Characterization Before and After CO2 Injection

To evaluate the changes in mineralogical composition, morphology, and surface properties before and after CO2 injection, a combination of advanced analytical techniques was applied. The methodologies employed include:
  • Fourier-transform infrared spectroscopy (FT-IR) using attenuated total reflectance (ATR). Infrared spectra were acquired in a PerkinElmer Spectrum 100 ONE spectrophotometer (PerkinElmer, Waltham, MA, USA)equipped with an ATR accessory of diamond crystal. This technique facilitated the identification of functional groups and structural modifications in the samples before and after CO2 injection. Measurements were conducted within the spectral range of 4000–400 cm−1, with a resolution of 4 cm−1 and 32 scans per spectrum;
  • SEM with energy-dispersive X-ray spectroscopy (EDS). A JEOL JSM 7401-F field emission microscope (JEOL Ltd., Akishima, Tokyo, Japan) coupled with an EDAX X1 system was used to obtain high-resolution images of the rock surface. To ensure conductivity, the samples were coated with a thin carbon layer before analysis. Imaging was performed at an accelerating voltage of 15 kV, with magnifications ranging from 10× to 1000×, allowing for a detailed examination of microstructural changes;
  • XRD analysis: The crystalline phases present in the samples were identified using a Bruker D2 PHASER diffractometer (Bruker Corporation, Billerica, MA, USA) equipped with a Cu-Kα radiation source (λ = 0.154 nm). Diffraction patterns were recorded over the 2θ range of 10–90°, with a step size of 0.02° and a counting time of 1 s per step, ensuring precise mineralogical characterization;
  • Surface area and porosity analysis: The specific surface area of the samples was determined using the Brunauer–Emmett–Teller (BET) method on a Quantachrome Nova 2200e analyzer (Quantachrome Instruments, Boynton Beach, FL, USA). Before measurement, nitrogen physisorption was conducted at 77 K to generate adsorption isotherms. The resulting data were processed to calculate the specific surface area and assess the porous structure of the rock samples.

3. Results and Discussion

This section presents the results obtained after the injection of CO2 in the Abra formation samples to compare the structural, morphological, and physical adsorption changes induced by the rock–fluid interaction. Samples exposed to controlled pressure and temperature conditions for 24 h were analyzed, considering the effect of brine on the mineralization process.

3.1. Fourier Transform Infrared (FT-IR) Spectroscopy to Monitor Functional Groups

FTIR spectroscopy is a suitable non-destructive technique for evaluating the molecular structure of either organic or inorganic materials. In the case of geological sciences, it enables the investigation of the structural properties of rocks before and after they are subjected to various processes, such as CO2 injection at the micro level [38]. It provides an understanding of functional groups and structures in petroleum source rocks, shales, and coals by analyzing the position of a peak [39]. This technique was employed to evaluate the functional groups in dependence on the CO2 injection parameters of the Abra cores, focusing on correlating changes in spectra to modifications in the crystalline lattice and mineral precipitation detected by XRD and SEM. The spectra of the subjected samples are shown in Figure 4. Figure 4a illustrates samples with a high organic matter content. The Abra initially displays the following signals: OH stretching (3540–3270 cm−1), stretching of aliphatic CH (2990–2810 cm−1), combined v 1 + v 4 and v 1 + v 3 modes of carbonates (2555–1780 cm−1), asymmetric stretching v 3 of free carbonate ion bands in calcite (1420 cm−1), out-of-plane bending ( v 2 ) of calcite (870 cm−1), a low intense band for aragonite (855 cm−1), and v 4 of calcite and aragonite (712 cm−1) [40,41,42].
The FTIR spectra reveal notable changes after CO2 injection. The signals between 4000 and 3050 cm−1 corresponding to OH vibrations of adsorbed water [43] in Abrat=24 P=1500 T=90 WB and Abrat=24 P=1600 T=95 WB samples disappear after CO2 injection. Simultaneously, the intensity of bands associated with aliphatic CH stretching (initially at 2985–2866 cm−1) and C=O stretching from carbonates (~2990 cm−1) decreases markedly, indicating the loss or transformation of organic components [44,45]. A remarkable change in the asymmetric stretching of free carbonate ion bands at 1415 cm−1 was observed after the process, and it became a very broad and intense signal; this could be attributed to structural modifications under pressure that promote the reorganization of the molecules in the lattice. This feature will be later corroborated by XRD analysis. Similarly, the out-of-plane bending mode of carbonate ion at 874 cm−1 is still present, but is slightly low in intensity. On the other hand, the band corresponding to in-plane bending at 715 cm−1 for free carbonate shifts toward a high wavenumber (727 cm−1). These spectral shifts indicate that pressure affects the bond lengths and angles within the carbonate structure, thereby modifying the lattice parameters and facilitating the formation of secondary phases.
The structural analysis by FT-IR indicates a relationship with pressure, as the CO3 bands were slightly more intense when using lower temperatures (90 °C) and pressures (1500 PSI) in Abrat=24 P=1500 T=90 WB than in Abrat=24 P=1600 T=95 WB. In addition, the presence of brine does not appear to significantly alter the functional groups post-injection, implying that the chemical interactions governing organic matter transformation are primarily driven by CO2 and pressure, rather than the aqueous medium. It is essential to note that the samples, after injection, exhibited evidence of oil release. FTIR analysis was performed on the extract, and the resulting spectrum is shown in Figure 4b, alongside a spectrum of an extract derived from the Abra core that was processed through the Soxhlet method. The signal at 3472 cm−1 in the band corresponds to O-H functional groups, indicating the presence of hydroxyl groups in the samples. This is followed by the signal at 2972 cm−1, which corresponds to the bending of methyl groups (CH3). The signal at 2866 cm−1 is assigned to the bending of the methylene (CH2) groups, followed by the 1718 cm−1 band corresponding to the stretching of the C=C. The signal at 1473 cm−1 is assigned to the stretching of the C-O bond due to the presence of carbonyl or ether groups, followed by the signals at 1200 cm−1, 748 cm−1, and 522 cm−1, which are assigned to the stretching of the C-H bond, confirming the presence of alkyl groups. These assignments offer valuable insights into the chemical composition of rock samples, enabling the characterization of their origin and properties.
For samples treated under different conditions, such as Abrat=24 P=1700 T=85 B, the FTIR spectrum revealed shifts in the CH3 and CH2 peaks (now appearing at 2920 cm−1/1381 cm−1 and 2852 cm−1, respectively) and C=C stretching (at 1741 cm−1). Most of the characteristic bands decreased in intensity, except for the 2970 cm−1 band, which not only shifts to 2920 cm−1, but also becomes significantly more intense. For Abrat=24 P=1600 T=95 B subjected to high temperature and low pressure, the shifting was found at 1709 cm−1, 1356 cm−1, and 1222 cm−1 for C=C, CH3, and CH, respectively, and are reduced in intensity. The shifts in the signals in the bands associated with functional groups, such as C=O, C=C, and -OH, were due to the formation or transformation of organic compounds by interacting with CO2 [46]. These changes in the functional group bands are indicative of the formation or transformation of organic compounds as they interact with CO2, a process that is consistent with the structural modifications detected in the crystalline lattice by XRD and with the precipitation phenomena observed in SEM.
Finally, Figure 4c displays the FTIR spectra of a sample with a low organic matter (OM) content from the Abra cores. In these samples, the spectral changes were minimal, suggesting that the interaction between CO2 and organic compounds was less pronounced. This minimal modification aligns with the reduced influence on the crystalline structure and the more minor changes in particle size observed after injection compared with high-OM samples. This observation opens the possibility of increasing the reaction time in future studies to enhance carbonation in low-OM samples.

3.2. Structural Analysis Through X-Ray Diffraction Method

As previously reported, the Abra core with high OM content exhibits the typical crystallographic planes of calcite, as indicated in PDF #96-900-9669, with a trigonal structure. The XRD patterns are shown in Figure 5a. After CO2 injection, a noticeable reduction in calcite peak intensities was observed, accompanied by the emergence of new peaks attributed to metastable phases, such as vaterite. A shift was observed in the (102) plane signal (Δ = 1.57°), which increased by 0.967°, 1.104°, and 1.191° as the pressure decreased from 1700, 1600, and 1500 PSI, respectively. The shifting increased with an increase in temperature. The shift in XRD peaks suggests the occurrence of dissolution and reprecipitation processes in the crystalline structure of the samples. According to previous studies, the dissolution of calcite in the presence of CO2 and water generates a redistribution of ions in solution, which induces stress in the crystalline lattice and promotes the formation of new mineral phases (Ca2⁺ + CO32− → CaCO3). In particular, the precipitation of vaterite and aragonite under conditions of Ca2⁺ and CO32− supersaturation has been reported in similar systems, which aligns with the changes observed in this study. These changes suggest that the presence of organic matter promotes partial calcite dissolution and the subsequent reprecipitation of secondary carbonates. Moreover, the observed peak shifts toward higher 2θ values suggest increased lattice stress and recrystallization under high-pressure and high-temperature conditions. This interpretation is consistent with the FTIR findings, which showed alterations in the carbonate bands and a reduction in the intensity of organic functional groups [47].
The primary and most intense crystallographic plane exhibited the same tendency, with increases of approximately 1.43°, 1.58°, and 1.65° as the pressure decreased. All signals present the same tendency. This behavior has been detected in ambient CO2 capture and mineral carbonation [48,49]. Another observation is that the presence of brine does not affect the crystallographic structure. Therefore, pressure and temperature are parameters that can modify the crystalline structure of the Abra cores. The pressure effect during CO2 injection induces the dissolution and precipitation of new mineral phases, such as vaterite, aragonite, dolomite, and quartz, as observed in 5a. On the other hand, the XRD pattern of the Abra core with low OM content (5b) displayed a shifting of all crystallographic planes compared with that of high OM content, which was c.a. 0.389°. The presence of OM tends to change the lattice [50]. After CO2 injection, the shifting of the signals was very smooth, since there was only a modification of 0.105° and 0.2105°, which shifted to low 2θ values. This observation is interesting, since it displays the contrary effect when organic matter is present. Stresses are responsible for the shifts in crystallographic planes, either towards higher or lower values. An increase in spacing corresponds to lower 2θ values, while a decrease in spacing results in higher 2θ values [51]. The shifting toward low values has also been reported to be due to the increase of pore diameter [52]. Therefore, this work demonstrates that the presence of a low OM increases the lattice parameters, whereas a high OM content decreases them.

3.3. Morphology Evaluation Through Scanning Electron Microscopy

SEM is a valuable technique for observing the surface morphology of geological samples before and after they are subjected to CO2 injection, providing information on mineral dissolution, precipitation, and textural changes that help to understand fluid-rock interactions and their implications in mineralization processes [53]. The structural changes observed in the SEM micrographs provide a visual confirmation of the structural changes detected by XRD, and they can be explained by the dissolution and precipitation mechanisms previously reported in the literature. Specifically, the dissolution of calcite in the presence of CO2 results in the release of Ca2⁺ and HCO3, promoting the nucleation of new carbonate minerals. In samples with high organic matter content (AbraHigh-OM), the interaction with CO2 appears to be influenced by organic compounds, which may affect the kinetics of mineralization and the stability of the precipitates. On the other hand, the precipitation of secondary carbonates, such as vaterite and aragonite, was more pronounced in samples with low organic matter content (AbraLow-OM), suggesting that mineralization occurs more efficiently in the absence of organic matter. These findings align with previous studies that have observed a higher degree of mineralization in systems with lower organic content, favoring the stable capture of CO2 in a solid carbonate form. In particular, the interaction between the rock and the brine plays a fundamental role in these processes since the ionic composition of the fluid can affect the dissolution kinetics and the precipitation of new secondary carbonates [54].
Figure 6 shows the analysis of the SEM micrographs. The Abra formation with a high organic matter (OM) content (AbraHigh-OM) exhibited a typical calcite crystal feature characterized by variable morphology, ranging from euhedral to subhedral shapes. It showed a dense texture with well-defined crystals and a compact surface, which suggested lower initial porosity. Significant differences were observed in particle size depending on the organic matter content. Analysis with ImageJ (version 1.53t) revealed that AbraHigh-OM displayed an average particle size of 13 μm (Figure 7a). The particle size distribution may be related to the interaction of organic matter with the mineral surface, favoring the precipitation of new crystals. Following CO2 injection, SEM micrographs revealed the formation of new mineral precipitates with varied morphologies, indicated by a yellow color, including nano-aggregates that correspond to the vaterite phase identified by XRD. It was observed that the presence of brine influences the nucleation of these mineral phases, modifying the morphology of the precipitates and promoting the stability of metastable forms, such as vaterite [54]. The micrographs corresponding to the injection of CO2 in samples with high organic matter content showed significant alterations. In Figure 6c, corresponding to sample Abrat=24 P=1500 T=90 WB, crystals with a mostly irregular morphology were observed, compared with the calcite crystals before injecting CO2, where no signs of precipitation of new minerals were detected. This confirms that rock–fluid–brine interaction is a key factor for the precipitation of new mineral phases, since the ionic composition of the fluid can modulate the solubility of calcium carbonate and the nucleation of crystals [54].
In comparison, the samples with low organic matter content (AbraLow-OM, Figure 6i–l) presented an average particle size of 1.6 μm (Figure 7b). These observations align with those in the existing literature, which indicates that the presence of organic compounds promotes the nucleation and growth of carbonate crystals [21]. On the other hand, Figure 6i shows tiny calcite crystals, but less compact compared with those in samples with high OM content. In Figure 6k, euhedral minerals with shapes similar to those of the hexagonal crystal system, highlighted in yellow in Figure 6l, are distinguished. The size distribution (Figure 6j) revealed smaller and heterogeneous particles, as well as minerals arranged in a spheroidal shape, which is characteristic of vaterite [14].
In contrast to the samples with low organic matter (OM) and high pressures, the SEM images revealed that the calcite crystals became more extensive and more well-defined after CO2 treatment, suggesting a dominant recrystallization process, rather than dissolution. This observation aligns with the relatively minor changes detected in the XRD patterns for these samples (6m-p); the crystalline forms of the prisms are still observed. However, undefined amorphous precipitates can be seen on them. These results are consistent with those of previous studies in which it has been reported that the ionic composition of the brine affects the stability of the precipitates, favoring the formation of amorphous structures and metastable phases under certain conditions of pressure and temperature [54] and with results of research in which it has been reported that vaterite forms under controlled conditions of pressure and temperature, generating spherical particles with diameters similar to those observed in this study. The precipitation of these secondary carbonates confirms the mineralogical alteration observed in the FTIR and XRD analyses.
Additionally, the treatments applied to the samples significantly affected the particle size distribution, as shown in Table 3, which summarizes the variations observed under different experimental conditions. The particle size distribution further supports the findings from SEM and XRD. The initial AbraHigh-OM sample had an average particle size of 13 μm. After CO2 injection, the particle sizes in the high-OM samples decreased dramatically to approximately 2.11 μm (Abrat=24 P=1500 T=90 WB), 1.88 μm (Abrat=24 P=1700 T=85 B), and 2.12 μm (Abrat=24 P=1600 T=95 B). This significant reduction suggests that the combined effects of CO2-induced dissolution and mechanical fragmentation, enhanced by organic matter, lead to a marked decrease in grain size.
On the other hand, the AbraLow-OM samples exhibited an increase in particle size following treatment. The average particle size increased from 1.6 μm in the untreated state to 3.26 μm (Abrat=24 P=1350 T=60 WB) and 4.12 μm (Abrat=24 P=2350 T=110 WB) post-injection, indicating that recrystallization and grain growth processes are more dominant when OM is limited [53,54].
These results indicate that particle size is not only influenced by the organic matter content, but also by the experimental conditions. The samples treated at high pressures and temperatures exhibit a reduction in particle size, which may be attributed to increased dissolution and recrystallization processes, as supported by XRD patterns. Notably, the Abrat=24 P=2350 T=110 WB sample, which experienced the most extreme treatment conditions, showed the largest particle size (4.12 μm), suggesting a coarsening effect potentially linked to secondary mineral growth (Figure 7a–c).

3.4. Specific Surface Analysis by Brunauer–Emmett–Teller Analysis

The results of the BET analysis revealed significant differences in the specific surface characteristics of the samples from the Abra Formation, both as a function of the organic matter content (high vs. low) and the treatment conditions to which they were subjected. Table 4 summarizes the values obtained for the BET surface area, pore surface area, pore volume, and pore diameter of each sample.
The treatment conditions had a significant impact on the BET surface area. According to the results obtained, the samples with a high organic matter content (AbraHigh-OM) exhibited a considerable increase in surface area after treatment, particularly in sample Abrat=24 P=1600 T=95 B, which reached a value of 2.918 m2/g. This enhancement in porosity was indicative of mineral dissolution followed by the formation of new pore spaces through the reprecipitation of secondary carbonates, a process that was promoted by the presence of organic matter.
On the other hand, temperature and brine had significant effects, mainly in sample Abrat=24 P=1600 T=95 B, which had the largest surface area. In contrast, the sample Abrat=24 P=1700 T=85 B, treated with brine at a lower temperature (85 °C), showed a smaller surface area (0.831 m2/g). This indicates that the combination of high temperature and brine was more effective in increasing the BET surface area in these samples [55].
It was also determined that the organic matter content had a significant effect. In samples with a low organic matter content (AbrLow-OM), the injection of CO2 had a slight effect on the surface area. Sample Abrat=24 P=2350 T=110 WB, treated at high pressure and temperature (without brine), showed a larger surface area (1.893 m2/g) than sample Abrat=24 P=1350 T=60 WB, treated under less intense conditions (without brine). This suggests that the structural changes induced by CO2 were less extensive in the absence of significant organic matter, resulting in a low degree of porosity enhancement.
Nitrogen adsorption isotherms (Figure 8) provide valuable information on the porous structure of the treated samples. The isotherms of samples Abrat=24 P=1500 T=90 WB, Abrat=24 P=1700 T=85 B, and Abrat=24 P=1600 T=95 B showed higher nitrogen adsorption than AbraHigh-OM. This confirms that the porosity and the accessible surface area increased after the injection of CO2 [56]. Sample Abrat=24 P=1600 T=95 B presented the isotherm with the highest adsorption, which correlated with its high BET surface area value (Figure 8a).
Additionally, the low OM isotherms of samples Abrat=24 P=1350 T=60 WB and Abrat=24 P=2350 T=110 WB in Figure 8b also showed an increase in adsorption compared with the original sample (AbraLow-OM). However, the increase was less pronounced than that in the high-organic-matter samples, suggesting that the porous structure of these samples was less susceptible to surface changes related to brine-promoted dissolution processes [54]. These samples exhibited more significant heterogeneity in their pore structure, characterized by more open hysteresis loops in the isotherms, which suggests the presence of pore bottlenecks that limit their adsorption capacity [57,58].

3.5. Analysis of Surface Area and Pore Size Distribution by the BJH Method

The analysis of the pore size distribution using the BJH method (Figure 9) provides additional information on the porous structure of the samples from the Abra formation. This indicates that these rocks were mesoporous, although they also had macropores with diameters greater than 50 nm, complementing the results of the BET analysis [58].

3.5.1. Pore Diameter and Pore Network Characteristics

The AbraHigh-OM sample presented low values of pore volume and surface area (Figure 9a). However, after treatment with CO2, both pore volume and surface area increased, especially the sample Abrat=24 P=1600 T=95 B, which exhibited the highest pore volume and surface area. These results indicated that the mesoporosity of the sample increased due to processes associated with dissolution processes or diagenesis of the rock, where initially the pores were occupied by organic matter and, finally, they were mobilized to the exterior due to the effects of pressure and temperature to which the samples were subjected [59].
In contrast, the low-organic-matter samples in Figure 9b indicated that the original AbraLow-OM sample exhibited a low pore volume and surface area across the entire diameter range. The treated sample Abrat=24 P=2350 T=110 WB showed a slight increase in pore volume compared with the original sample. Still, the surface area remained small, suggesting the presence of larger pores (macropores) in a smaller proportion.

3.5.2. Influence of Organic Matter on Pore Structure

The observed increases in surface area and pore volume, especially in samples with high organic matter content treated with brine at high temperature and pressure, are related to a combination of processes [56]:
  • Dissolution/Precipitation of minerals: The injection of CO2 into the reactor, together with the high-pressure and -temperature conditions, may have induced the dissolution of minerals susceptible to acidification, such as carbonates. Dissolving these minerals creates new pore spaces, increasing the surface area and pore volume. At the same time, the alteration of the physicochemical conditions may have favored the precipitation of new mineral phases, such as silica or clays, which may also contribute to the modification of the pore structure [60]. SEM, XRD, and FTIR analyses provide evidence of these dissolution and precipitation processes;
  • Organic matter maturation and hydrocarbon release: The Abra formation is hydrocarbon-producing and located in the Tampico-Misantla basin. The temperatures and pressures used in the treatments (90–110 °C and 1350–2350 PSI) are within the hydrocarbon generation window for source rocks. The oil expulsion observed in the organic matter-rich samples during the treatment suggests that an OM maturation process is being simulated, where kerogen is converted into oil and gas. The release of these hydrocarbons from the rock matrix could generate new pore spaces, improving permeability and the accessible surface area [61]. This process is particularly relevant in hydrocarbon source rocks, such as the Abra formation, where porosity and permeability are fundamental [62].

3.5.3. Effect on CO2 Storage Potential and EOR Implications

CO2 injection tends to facilitate the mobilization and removal of some of the organic matter present in the pores, especially in the form of oil. This process would also contribute to pore cleaning and to increasing the accessible surface area. The results of this study suggest that high-pressure and high-temperature CO2 treatments can improve the porosity and permeability of the formation, which could have positive implications for hydrocarbon production through enhanced oil recovery (EOR) techniques [61]. Furthermore, the increase in surface area and pore volume could increase the CO2 storage capacity in the formation, contributing to climate change mitigation.

4. Conclusions

The mineralization of CO2 in the Abra carbonate formation has been successfully demonstrated in this study under controlled laboratory conditions, providing essential data for the assessment of geological carbon storage strategies in Mexico. This research highlights the beneficial role of brine in enhancing rock–fluid interactions and promoting the precipitation of secondary carbonates, which contributes to the long-term stability of stored CO2. Temperature and pressure conditions were found to influence the extent of mineralization, suggesting that the optimization of these parameters could further improve CO2 storage efficiency.
Crucially, this study revealed that the content of organic matter plays a significant role in modulating the mineralization process, with low-OM samples potentially favoring physical trapping mechanisms, offering a differentiated approach to CO2 storage. Analytical techniques, including FT-IR, XRD, SEM, and BET, proved effective in characterizing the structural and morphological changes induced by CO2 injection.
The results of this study contribute valuable data for assessing the feasibility of geological carbon storage in the Tampico-Misantla basin and support the development of informed climate change mitigation strategies in Mexico. Future research should focus on optimizing CO2 injection parameters, investigating the long-term stability of mineralized carbonates under in situ conditions, and expanding the scope to other formations within the basin.

5. Recommendations for Future Research

It is recommended to analyze the influence of organic matter on CO2 mineralization efficiency in greater detail within carbonate formations. Specifically, systematic studies should be conducted to evaluate the effects of varying OM levels on the dissolution of calcite, the formation of secondary phases, and the resultant changes in particle size and porosity. In particular, it would be relevant to evaluate the long-term stability of the precipitated carbonates, as well as the effect of brine composition on the nucleation and growth of these mineral phases. Moreover, large-scale experimental studies are needed to validate laboratory findings and to provide key information for optimizing CO2 capture strategies in real-world geological formations. Additionally, the implementation of thermodynamic and kinetic models to predict the evolution of mineralization processes as a function of variables such as pressure, temperature, OM content, and brine composition is suggested. This would contribute to improving the efficiency of geological CO2 storage.

Author Contributions

Conceptualization, funding acquisition, A.M.M.-M. and D.P.-R.; Conceptualization, methodology, validation, and data curation, R.L.-D.; SEM formal analysis, H.D.-R.; FTIR acquisition, R.G.-A.; BET formal analysis and validation, I.C.R.-I.; XRD formal analysis and validation, D.S.G.-Z.; original draft preparation, writing—review and editing, all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was granted by the TecNM 13439.21 and IPN SIP20251064 projects. The authors gratefully acknowledge the financial support and resources provided.

Data Availability Statement

Data are contained within the article.

Acknowledgments

R.L.-D. is grateful for her CONAHCYT grant. The authors are grateful Litoteca Nacional de la Industria de Hidrocarburos (CNIH), Comisión Nacional de Hidrocarburos (CNH) Hidalgo, Mexico for providing the core samples.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mishra, A.; Boon, M.M.; Benson, S.M.; Watson, M.N.; Haese, R.R. Reconciling predicted and observed carbon mineralization in siliciclastic formations. Chem. Geol. 2023, 619, 121324. [Google Scholar] [CrossRef]
  2. Eyitayo, S.; Arbad, N.; Okere, C.; Gamadi, T.; Watson, M. Advancing geological storage of carbon dioxide (CO2) with emerging technologies for climate change mitigation. Int. J. Environ. Sci. Technol. 2025, 22, 5023–5056. [Google Scholar] [CrossRef]
  3. Rochelle, C.A.; Czernichowski-Lauriol, I.; Milodowski, A.E. The impact of chemical reactions on CO2 storage in geological formations: A brief review. In Geological Storage of Carbon Dioxide; Baines, S.J., Worden, R.H., Eds.; Geological Society of London: London, UK, 2004; Volume 233, pp. 87–106. [Google Scholar]
  4. Kim, K.; Kim, D.; Na, Y.; Song, Y.; Wang, J. A review of carbon mineralization mechanism during geological CO2 storage. Heliyon 2023, 9, e23135. [Google Scholar] [CrossRef] [PubMed]
  5. Medina, E.; Levresse, G.; Carrera-Hernández, J.J.; Cerca, M. A basin scale assessment framework of onshore aquifer-based CO2 suitability storage in Tampico Misantla basin, Mexico. Int. J. Greenh. Gas Control 2023, 125, 103874. [Google Scholar] [CrossRef]
  6. Vega-Ortiz, C.; Beti, D.R.; Setoyama, E.; McLennan, J.D.; Ring, T.A.; Levey, R.; Martínez-Romero, N. Source rock evaluation in the central-western flank of the Tampico Misantla Basin, Mexico. J. South Am. Earth Sci. 2020, 100, 102552. [Google Scholar] [CrossRef]
  7. Comision Nacional de Hidrocarburos. Mapa de Hidrocarburos. 2025. Available online: https://mapa.hidrocarburos.gob.mx (accessed on 28 February 2025).
  8. Guzmán, A.E. Tampico-Misantla: A premier super basin in waiting. AAPG Bull. 2022, 106, 495–516. [Google Scholar] [CrossRef]
  9. Wendt, A.; Sheriff, A.; Shih, C.Y.; Vikara, D.; Grant, T. A multi-criteria CCUS screening evaluation of the Gulf of Mexico, USA. Int. J. Greenh. Gas Control 2022, 118, 103688. [Google Scholar] [CrossRef]
  10. Saleh, M.A.; Ryan, M.P.; Trusler, J.P.M.; Krevor, S. The interfacial processes controlling carbon dioxide mineralisation in magnesium and calcium silicates. Fuel 2025, 380, 132969. [Google Scholar] [CrossRef]
  11. Rahim, H.u.; Shah, M.M.; Khalil, R.; Jamil, M. Evaluation of fault and fracture controlled hydrothermal dolomitization in the Jurassic carbonate succession of the lesser Himalayas, NW Pakistan. Carbonates Evaporites 2025, 40, 32. [Google Scholar] [CrossRef]
  12. Jamil, M.; Ullah, I.; Rahim, H.U.; Khan, I.; Abbas, W.; Rehman, M.U.; Rashid, A.; Umar, M.; Ali, A.; Siddiqui, N.A. Tracking Depositional Architecture and Diagenetic Evolution in the Jurassic Carbonates, Trans Indus Ranges, NW Himalayas. Minerals 2024, 14, 1170. [Google Scholar] [CrossRef]
  13. Raza, A.; Glatz, G.; Gholami, R.; Mahmoud, M.; Alafnan, S. Carbon mineralization and geological storage of CO2 in basalt: Mechanisms and technical challenges. Earth-Sci. Rev. 2022, 229, 104036. [Google Scholar] [CrossRef]
  14. Pusparizkita, Y.M.; Schmahl, W.W.; Ambarita, M.; Kholid, H.N.; Sadewa, A.Y.; Ismail, R.; Jamari, J.; Bayuseno, A.P. Mineralizing CO2 and producing polymorphic calcium carbonates from bitumen-rock asphalt manufacturing solid residues. Clean. Eng. Technol. 2023, 12, 100602. [Google Scholar] [CrossRef]
  15. Mexico Gobierno de la Republica. Atlas Geológico Cuenca Tampico—Misantla; National Hydrocarbons Commission: Mexico City, Mexico, 2023; Available online: https://hidrocarburos.gob.mx/media/3091/atlas_geologico_cuenca_tampico-misantla_v3.pdf (accessed on 28 February 2025).
  16. Isah, A.; Mahmoud, M.; Aljawad, M.S.; Arif, M.; Hussaini, S.R.; Amao, A.; Raza, A.; Kamal, M.S. Enforced CO2 mineralization in anhydrite-rich rocks. Energy 2024, 305, 132323. [Google Scholar] [CrossRef]
  17. Sanguinito, S.; Goodman, A.; Tkach, M.; Kutchko, B.; Culp, J.; Natesakhawat, S.; Fazio, J.; Fukai, I.; Crandall, D. Quantifying dry supercritical CO2-induced changes of the Utica Shale. Fuel 2018, 226, 54–64. [Google Scholar] [CrossRef]
  18. Moita, P.; Berrezueta, E.; Abdoulghafour, H.; Beltrame, M.; Pedro, J.; Mirão, J.; Miguel, C.; Galacho, C.; Sitzia, F.; Barrulas, P.; et al. Mineral Carbonation of CO2 in Mafic Plutonic Rocks, II—Laboratory Experiments on Early-Phase Supercritical CO2—Brine—Rock Interactions. Appl. Sci. 2020, 10, 5083. [Google Scholar] [CrossRef]
  19. Morais, M.M.; Lucas-Oliveira, E.; Bonagamba, T.J.; Aum, P.T.P.; Lucas, C.R.d.S.; da Silva, D.N.N.; Fortulan, C.A. Fabrication and petrophysical characterization of artificial carbonate rocks with multiscale porosity sintered in a CO2 atmosphere. Geoenergy Sci. Eng. 2023, 229, 212096. [Google Scholar] [CrossRef]
  20. Jayasekara, D.W.; Ranjith, P.G.; Wanniarachchi, W.A.M.; Rathnaweera, T.D.; Van Gent, D. CO2-brine-caprock interaction: Reactivity experiments on mudstone caprock of South-west Hub geo-sequestration project. J. Pet. Sci. Eng. 2020, 189, 107011. [Google Scholar] [CrossRef]
  21. Ma, L.; Zhao, L.; Li, Y.; Zhang, J. Controllable crystallization of pure vaterite using CO2-storage material and different Ca2+ sources. J. CO2 Util. 2021, 48, 101520. [Google Scholar] [CrossRef]
  22. Durucan, S.; Korre, A.; Parlaktuna, M.; Senturk, E.; Wolf, K.-H.; Chalari, A.; Stork, A.; Nikolov, S.; Kunder, R.; Sigfusson, B.; et al. SUCCEED: A CO2 storage and utilisation project aimed at mitigating against greenhouse gas emissions from geothermal power production. SSRN Electron. J. 2021, 1, 1–7. [Google Scholar] [CrossRef]
  23. Sun, L.; Liu, Y.; Cheng, Z.; Jiang, L.; Lv, P.; Song, Y. Review on Multiscale CO2 Mineralization and Geological Storage: Mechanisms, Characterization, Modeling, Applications and Perspectives. Energy Fuels 2023, 37, 14512–14537. [Google Scholar] [CrossRef]
  24. Meckel, T.A.; Bump, A.P.; Hovorka, S.D.; Trevino, R.H. Carbon capture, utilization, and storage hub development on the Gulf Coast. Greenh. Gases Sci. Technol. 2021, 11, 619–632. [Google Scholar] [CrossRef]
  25. Lu, H.; Wang, W.; Zhang, J.; Wu, M.; Deng, Y.; Li, G.; Tian, M. Accelerated Direct Mineralization of CO2 by Carbide Slag under Ambient Temperature and Pressure Reaction Conditions. Energy Fuels 2024, 38, 14483–14492. [Google Scholar] [CrossRef]
  26. Li, N.; Feng, W.; Yu, J.; Chen, F.; Zhang, Q.; Zhu, S.; Hu, Y.; Li, Y. Recent Advances in Geological Storage: Trapping Mechanisms, Storage Sites, Projects, and Application of Machine Learning. Energy Fuels 2023, 37, 10087–10111. [Google Scholar] [CrossRef]
  27. Jamaluddin, N.; Hsu, Y.-I.; Asoh, T.-A.; Uyama, H. Optically Transparent and Toughened Poly(methyl methacrylate) Composite Films with Acylated Cellulose Nanofibers. ACS Omega 2021, 6, 10752–10758. [Google Scholar] [CrossRef] [PubMed]
  28. Rahman, S.S.; Mahmud, M.B.; Omranpour, H.; Salehi, A.; Monfared, A.R.; Park, C.B. Highly Tough Yet Stiff, Transparent, and Recyclable PMMA Nanocomposites Incorporating TPU Nanofibril Networks with High Thermal Stability and Strong Interfacial Adhesion. ACS Appl. Mater. Interfaces 2024, 16, 42687–42703. [Google Scholar] [CrossRef]
  29. Ye, H.; Liu, Q.; Bao, Q.; Wang, Z.; Xie, Y.; Michelle, T.; Zhao, W.; Xian, C. Review on in-situ CO2 mineralization sequestration: Mechanistic understanding and research frontiers. Int. J. Coal Sci. Technol. 2025, 12, 15. [Google Scholar] [CrossRef]
  30. López-Dinorín, R.; Mendoza-Martínez, A.M.; Palma-Ramírez, D.; Luna-Dominguez, J.H.; Dorantes-Rosales, H. Structural and morphological evaluation of three types of carbonate hydrocarbon-bearing rocks from Tampico-Misantla basin: Structural effects after CO2 injection. MRS Adv. 2024, 9, 1023–1030. [Google Scholar] [CrossRef]
  31. Juárez-Arriaga, E.; Lawton, T.F.; Stockli, D.F.; Solari, L.; Martens, U. Late Cretaceous-Paleocene stratigraphic and structural evolution of the central Mexican fold and thrust belt, from detrital zircon (U-Th)/(He-Pb) ages. J. South Am. Earth Sci. 2019, 95, 102264. [Google Scholar] [CrossRef]
  32. Peter, A.; Jin, X.; Fan, X.; Eshiet, K.I.-I.; Sheng, Y.; Yang, D. Effect of CO2 Phase on Pore Geometry of Saline Reservoir Rock. Rock Mech. Rock Eng. 2022, 55, 1907–1930. [Google Scholar] [CrossRef]
  33. Xue, J.; Gao, H.; Ma, Z.; Shi, H.; Li, X.; Li, T.; Cheng, Z.; Wang, C.; Li, P.; Zhang, N.C. Analyzing the Microscopic Production Characteristics of CO2 Flooding after Water Flooding in Tight Oil Sandstone Reservoirs Utilizing NMR and Microscopic Visualization Apparatus. Atmosphere 2024, 15, 487. [Google Scholar] [CrossRef]
  34. González-Partida, E.; Camprubí, A.; Canet, C.; Sánchez, F. Fisicoquímica de salmueras e hidrocarburos en cuencas petroleras y en depósitos minerales tipo Mississippi Valley y asociados. Parte II: Ejemplos de la Cuenca de Sabinas y la Cuenca del Sureste, México. Boletín Soc. Geológica Mex. 2007, 60, 23–42. [Google Scholar] [CrossRef]
  35. Salem, A.M.; Shedid, S.A. Variation of petrophysical properties due to carbon dioxide (CO2) storage in carbonate reservoirs. J. Pet. Gas Eng. 2013, 4, 91–102. [Google Scholar]
  36. National Hydrocarbons Commission. Reporte de Reservas de Hidrocarburos; National Hydrocarbons Commission: Mexico City, Mexico, 2023; Available online: https://datos.gob.mx/busca/dataset/reporte-de-reservas-de-hidrocarburos (accessed on 28 February 2025).
  37. Kim, K.; Vilarrasa, V.; Makhnenko, R.Y. CO2 Injection Effect on Geomechanical and Flow Properties of Calcite-Rich Reservoirs. Fluids 2018, 3, 66. [Google Scholar] [CrossRef]
  38. Chen, Y.; Zou, C.; Mastalerz, M.; Hu, S.; Gasaway, C.; Tao, X. Applications of Micro-Fourier Transform Infrared Spectroscopy (FTIR) in the Geological Sciences—A Review. Int. J. Mol. Sci. 2015, 16, 30223–30250. [Google Scholar] [CrossRef]
  39. Pejcic, B.; Heath, C.; Pagès, A.; Normore, L. Analysis of carbonaceous materials in shales using mid-infrared spectroscopy. Vib. Spectrosc. 2021, 112, 103186. [Google Scholar] [CrossRef]
  40. Toffolo, M.B.; Regev, L.; Dubernet, S.; Lefrais, Y.; Boaretto, E. FTIR-Based Crystallinity Assessment of Aragonite–Calcite Mixtures in Archaeological Lime Binders Altered by Diagenesis. Minerals 2019, 9, 121. [Google Scholar] [CrossRef]
  41. Xu, B.; Poduska, K.M. Linking crystal structure with temperature-sensitive vibrational modes in calcium carbonate minerals. Phys. Chem. Chem. Phys. 2014, 16, 17634–17639. [Google Scholar] [CrossRef]
  42. Florek, M.; Fornal, E.; Gómez-Romero, P.; Zieba, E.; Paszkowicz, W.; Lekki, J.; Nowak, J.; Kuczumow, A. Complementary microstructural and chemical analyses of Sepia officinalis endoskeleton. Mater. Sci. Eng. C 2009, 29, 1220–1226. [Google Scholar] [CrossRef]
  43. Prasad, P.S.R.; Srinivasa Sarma, D.; Nirmal Charan, S. Mineral trapping and sequestration of carbon-dioxide in deccan basalts: SEM, FTIR and raman spectroscopic studies on secondary carbonates. J. Geol. Soc. India 2012, 80, 546–552. [Google Scholar] [CrossRef]
  44. Ortega-Villamagua, E.; Arcos, M.; Romero, M.; Vasquez, C.; Palma-Cando, A. Precipitación de carbonatos inducida microbiológicamente como potencial estrategia en la restauración de estructuras patrimoniales. Ge-Conservacion 2022, 21, 224–234. [Google Scholar] [CrossRef]
  45. Matalkah, F.; Soroushian, P. Role of CO2 in enhancing geopolymer properties formulated with fluidized bed combustion ash. J. CO2 Util. 2023, 71, 102462. [Google Scholar] [CrossRef]
  46. Krukowski, E.G.; Goodman, A.; Rother, G.; Ilton, E.S.; Guthrie, G.; Bodnar, R.J. FT-IR study of CO2 interaction with Na+ exchanged montmorillonite. Appl. Clay Sci. 2015, 114, 61–68. [Google Scholar] [CrossRef]
  47. Ali, M.; Jha, N.K.; Pal, N.; Keshavarz, A.; Hoteit, H.; Sarmadivaleh, M. Recent advances in carbon dioxide geological storage, experimental procedures, influencing parameters, and future outlook. Earth-Sci. Rev. 2022, 225, 103895. [Google Scholar] [CrossRef]
  48. Chen, M.; Zhang, Q.; Li, Z.; Hu, H.; Wang, C. Insights into the mechanochemical interfacial interaction between calcite and serpentine: Implications for ambient CO2 capture. J. Clean. Prod. 2023, 401, 136715. [Google Scholar] [CrossRef]
  49. Zoveidavianpoor, M.; Jalilavi, M.; Mohsin, R. Investigating mineral carbonation dynamics by variable CO2 injection in Bera sandstone and Tapah limestone under controlled weathering. Int. J. Pet. Eng. 2024, 4, 106–132. [Google Scholar] [CrossRef]
  50. Ali, A.; Chiang, Y.W.; Santos, R.M. X-ray Diffraction Techniques for Mineral Characterization: A Review for Engineers of the Fundamentals, Applications, and Research Directions. Minerals 2022, 12, 205. [Google Scholar] [CrossRef]
  51. Elhamdi, I.; Souissi, H.; Taktak, O.; Elghoul, J.; Kammoun, S.; Dhahri, E.; Costa, B.F.O. Experimental and modeling study of ZnO:Ni nanoparticles for near-infrared light emitting diodes. RSC Adv. 2022, 12, 13074–13086. [Google Scholar] [CrossRef]
  52. Jiménez-Gómez, C.P.; Cecilia, J.A.; Alba-Rubio, A.C.; Cassidy, A.; Moreno-Tost, R.; García-Sancho, C.; Maireles-Torres, P. Tailoring the selectivity of Cu-based catalysts in the furfural hydrogenation reaction: Influence of the morphology of the silica support. Fuel 2022, 319, 123827. [Google Scholar] [CrossRef]
  53. Chen, F.; Xu, J.; Tang, L.; Yu, Q.; Gu, J. Microscopic investigation of chemical and physical alteration behaviors in sandstone minerals induced by CO2-H2O-rock interactions during CO2 saline aquifer storage. Sci. Total Environ. 2025, 958, 177897. [Google Scholar] [CrossRef]
  54. Md Yusof, M.A.; Ibrahim, M.A.; Idress, M.; Idris, A.K.; Saaid, I.M.; Mohd Rodzi, N.; Mohsin, S.; Azhari Awangku Matali, A.A. Effects of CO2/Rock/Formation Brine Parameters on CO2 Injectivity for Sequestration. In Proceedings of the Offshore Technology Conference Asia, Online, 2–6 November 2020. [Google Scholar]
  55. Duan, R.; Xu, Z.; Dong, Y.; Liu, W. Characterization and classification of pore structures in deeply buried carbonate rocks based on mono- and multifractal methods. J. Pet. Sci. Eng. 2021, 203, 108606. [Google Scholar] [CrossRef]
  56. Cheng, Y.; Zeng, M.; Lu, Z.; Du, X.; Yin, H.; Yang, L. Effects of Supercritical CO2 Treatment Temperatures on Mineral Composition, Pore Structure and Functional Groups of Shale: Implications for CO2 Sequestration. Sustainability 2020, 12, 3927. [Google Scholar] [CrossRef]
  57. Fu, S.; Fang, Q.; Li, A.; Li, Z.; Han, J.; Dang, X.; Han, W. Accurate characterization of full pore size distribution of tight sandstones by low-temperature nitrogen gas adsorption and high-pressure mercury intrusion combination method. Energy Sci. Eng. 2021, 9, 80–100. [Google Scholar] [CrossRef]
  58. Li, T.; Huang, Z.; Zhao, J.; Xu, X.; Guo, X. Pore structure characteristics and their influencing factors: A case study from the middle jurassic mixed siliciclastic carbonate rocks, Turpan-Hami basin, Northwest China. J. Pet. Sci. Eng. 2021, 203, 108611. [Google Scholar] [CrossRef]
  59. Kuila, U.; Prasad, M. Specific surface area and pore-size distribution in clays and shales. Geophys. Prospect. 2013, 61, 341–362. [Google Scholar] [CrossRef]
  60. Zhao, H.; Luo, X.; Zhang, H.; Sun, N.; Wei, W.; Sun, Y. Carbon-based adsorbents for post-combustion capture: A review. Greenh. Gases Sci. Technol. 2018, 8, 11–36. [Google Scholar] [CrossRef]
  61. Petrovic, B.; Gorbounov, M.; Masoudi Soltani, S. Influence of surface modification on selective CO2 adsorption: A technical review on mechanisms and methods. Microporous Mesoporous Mater. 2021, 312, 110751. [Google Scholar] [CrossRef]
  62. Rashid, A.; Siddiqui, N.A.; Ahmed, N.; Jamil, M.; El-Ghali, M.A.K.; Ali, S.H.; Zaidi, F.K.; Wahid, A. Field attributes and organic geochemical analysis of shales from early to middle Permian Dohol Formation, Peninsular Malaysia: Implications for hydrocarbon generation potential. J. King Saud Univ.-Sci. 2022, 34, 102287. [Google Scholar] [CrossRef]
Figure 1. Geological map of the Tampico-Misantla Basin location in eastern Mexico [7].
Figure 1. Geological map of the Tampico-Misantla Basin location in eastern Mexico [7].
Processes 13 01087 g001
Figure 2. Schematic representation of the tectonic and stratigraphic framework of the El Abra Formation in the Tampico-Misantla Basin: (a) Basin profile showing depth variations of the El Abra Formation, highlighting subsidence processes associated with the passive margin; (b) Stratigraphic column of the El Abra Formation, indicating predominant facies and marking the study samples analyzed in this work.
Figure 2. Schematic representation of the tectonic and stratigraphic framework of the El Abra Formation in the Tampico-Misantla Basin: (a) Basin profile showing depth variations of the El Abra Formation, highlighting subsidence processes associated with the passive margin; (b) Stratigraphic column of the El Abra Formation, indicating predominant facies and marking the study samples analyzed in this work.
Processes 13 01087 g002
Figure 3. Generalized flowchart of the CO2 mineralization experimental procedure.
Figure 3. Generalized flowchart of the CO2 mineralization experimental procedure.
Processes 13 01087 g003
Figure 4. FTIR spectra for Abra samples (a) with high content of OM (AbraHigh-OM), (b) released-extract, and (c) low content of OM (AbraLow-OM), after CO2 injection.
Figure 4. FTIR spectra for Abra samples (a) with high content of OM (AbraHigh-OM), (b) released-extract, and (c) low content of OM (AbraLow-OM), after CO2 injection.
Processes 13 01087 g004
Figure 5. XRD pattern for Abra samples (a) AbraHigh-OM and (b) AbraLow-OM after CO2 injection.
Figure 5. XRD pattern for Abra samples (a) AbraHigh-OM and (b) AbraLow-OM after CO2 injection.
Processes 13 01087 g005
Figure 6. SEM micrographs of El Abra Formation samples under varying experimental conditions. (a) AbraHigh-OM: dense calcite crystals with euhedral to subhedral morphologies (average size: 13 μm), (b) AbraHigh-OM: euhedral to subhedral calcite crystals with well-defined angular faces, (c,d) Abrat=24 P=1500 T=90 WB: irregular crystal morphologies after CO2 injection, showing secondary carbonate precipitation (yellow highlights). (e,f) Abrat=24 P=1700 T=85 B: varied crystal morphologies influenced by brine (B), promoting metastable phases like vaterite, (g,h) Abrat=24 P=1600 T=95 B: well-defined crystal growth, highlighting brine’s role in stabilizing secondary carbonates, (il) AbraLow-OM: tiny calcite crystals (average size: 1.6 μm) with spheroidal vaterite particles, indicating low OM content and high porosity, (m,n) Abrat=24 P=1350 T=60 WB: prismatic calcite with amorphous precipitates, suggesting less stable phases under lower pressure and temperature, (o,p) Abrat=24 P=2350 T=110 WB: euhedral crystals with amorphous precipitates, consistent with high-pressure conditions favoring metastable phases.
Figure 6. SEM micrographs of El Abra Formation samples under varying experimental conditions. (a) AbraHigh-OM: dense calcite crystals with euhedral to subhedral morphologies (average size: 13 μm), (b) AbraHigh-OM: euhedral to subhedral calcite crystals with well-defined angular faces, (c,d) Abrat=24 P=1500 T=90 WB: irregular crystal morphologies after CO2 injection, showing secondary carbonate precipitation (yellow highlights). (e,f) Abrat=24 P=1700 T=85 B: varied crystal morphologies influenced by brine (B), promoting metastable phases like vaterite, (g,h) Abrat=24 P=1600 T=95 B: well-defined crystal growth, highlighting brine’s role in stabilizing secondary carbonates, (il) AbraLow-OM: tiny calcite crystals (average size: 1.6 μm) with spheroidal vaterite particles, indicating low OM content and high porosity, (m,n) Abrat=24 P=1350 T=60 WB: prismatic calcite with amorphous precipitates, suggesting less stable phases under lower pressure and temperature, (o,p) Abrat=24 P=2350 T=110 WB: euhedral crystals with amorphous precipitates, consistent with high-pressure conditions favoring metastable phases.
Processes 13 01087 g006
Figure 7. Particle size distribution of Abra Formation samples. (a) AbraHigh-OM sample before CO2 injection, showing an average particle size of 13.0 μm. (b) AbraHigh-OM samples after of CO2 injection under different conditions: Abrat=24 P=1500 T=90 WB (2.11 μm), Abrat=24 P=1700 T=85 B (1.88 μm), and Abrat=24 P=1600 T=95 B (2.12 μm), demonstrating particle size reduction due to dissolution and recrystallization. (c) AbraLow-OM sample before CO2 injection, with an average particle size of 1.6 μm, and (d) AbraLow-OM samples after treatment: Abrat=24 P=1350 T=60 WB (3.26 μm) and Abrat=24 P=2350 T=110 WB (4.12 μm), where increased temperature and pressure led to particle growth, likely due to secondary mineral precipitation.
Figure 7. Particle size distribution of Abra Formation samples. (a) AbraHigh-OM sample before CO2 injection, showing an average particle size of 13.0 μm. (b) AbraHigh-OM samples after of CO2 injection under different conditions: Abrat=24 P=1500 T=90 WB (2.11 μm), Abrat=24 P=1700 T=85 B (1.88 μm), and Abrat=24 P=1600 T=95 B (2.12 μm), demonstrating particle size reduction due to dissolution and recrystallization. (c) AbraLow-OM sample before CO2 injection, with an average particle size of 1.6 μm, and (d) AbraLow-OM samples after treatment: Abrat=24 P=1350 T=60 WB (3.26 μm) and Abrat=24 P=2350 T=110 WB (4.12 μm), where increased temperature and pressure led to particle growth, likely due to secondary mineral precipitation.
Processes 13 01087 g007
Figure 8. Nitrogen adsorption isotherms of samples from the Abra formation with (a) high organic matter content and (b) low organic matter content.
Figure 8. Nitrogen adsorption isotherms of samples from the Abra formation with (a) high organic matter content and (b) low organic matter content.
Processes 13 01087 g008
Figure 9. Pore size distribution analysis of Abra formation samples using the BJH method, illustrating the prevalence of mesopores and the existence of macropores in samples with (a) high organic matter content and (b) low organic matter content. PV = pore volume and SA = surface area in plots.
Figure 9. Pore size distribution analysis of Abra formation samples using the BJH method, illustrating the prevalence of mesopores and the existence of macropores in samples with (a) high organic matter content and (b) low organic matter content. PV = pore volume and SA = surface area in plots.
Processes 13 01087 g009
Table 1. Details of the samples based on the CNH National Lithotheque Registry for the El Abra Formation in the Tampico-Misantla Basin.
Table 1. Details of the samples based on the CNH National Lithotheque Registry for the El Abra Formation in the Tampico-Misantla Basin.
SampleDepth
(m)
OM ContentHydrocarbon Generating PotentialAdditional Details
AbraHigh-OM~1010HighHigh hydrocarbon generating potential (CNH)Representative high-OM sample; evaluated for CO2 storage potential and diagenetic alteration.
AbraLow-OM~1450LowLow hydrocarbon generating potential (CNH)Representative low-OM sample; compared against high-OM sample to assess differences in secondary carbonate precipitation.
Table 2. Experimental conditions of CO2 injection in samples of Abra formation.
Table 2. Experimental conditions of CO2 injection in samples of Abra formation.
Sample Time
(h)
Pressure
(PSI)
Temperature
(°C)
B or WB
Abrat=24 P=1500 T=90 WB24150090WB
Abrat=24 P=1700 T=85 B24170085B
Abrat=24 P=1600 T=95 B24160095B
Abrat=24 P=1350 T=60 WB24135060WB
Abrat=24 P=2350 T=110 WB242350110WB
Table 3. Average particle size of Abra Formation samples after CO2 injection.
Table 3. Average particle size of Abra Formation samples after CO2 injection.
SampleParticle Size (μm)
AbraHigh-OM13
Abrat=24 P=1500 T=90 WB2.11
Abrat=24 P=1700 T=85 B1.88
Abrat=24 P=1600 T=95 B2.12
AbraLow-OM1.6
Abrat=24 P=1350 T=60 WB3.26
Abrat=24 P=2350 T=110 WB4.12
Table 4. BET data from the Abra formation.
Table 4. BET data from the Abra formation.
Sample BET Surface Area
(m2 g−1)
Pore Surface Area
(m2 g−1)
Pore Diameter
(nm)
AbraHigh-OM0.6110.4363.587
Abrat=24 P=1500 T=90 WB0.9770.4445.5564
Abrat=24 P=1700 T=85 B0.8310.3914.917
Abrat=24 P=1600 T=95 B2.9184.1254.884
AbraLow-OM0.7440.4424.327
Abrat=24 P=1350 T=60 WB0.5390.2614.896
Abrat=24 P=2350 T=110 WB1.8930.254.847
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

López-Dinorín, R.; Mendoza-Martínez, A.M.; Palma-Ramírez, D.; Dorantes-Rosales, H.; García-Alamilla, R.; Romero-Ibarra, I.C.; García-Zaleta, D.S. Analysis of Carbon Dioxide Mineralization in Carbonates from Tampico-Misantla Basin, Mexico: Effect of Organic Matter Content. Processes 2025, 13, 1087. https://doi.org/10.3390/pr13041087

AMA Style

López-Dinorín R, Mendoza-Martínez AM, Palma-Ramírez D, Dorantes-Rosales H, García-Alamilla R, Romero-Ibarra IC, García-Zaleta DS. Analysis of Carbon Dioxide Mineralization in Carbonates from Tampico-Misantla Basin, Mexico: Effect of Organic Matter Content. Processes. 2025; 13(4):1087. https://doi.org/10.3390/pr13041087

Chicago/Turabian Style

López-Dinorín, Roxana, Ana María Mendoza-Martínez, Diana Palma-Ramírez, Héctor Dorantes-Rosales, Ricardo García-Alamilla, Issis Claudette Romero-Ibarra, and David Salvador García-Zaleta. 2025. "Analysis of Carbon Dioxide Mineralization in Carbonates from Tampico-Misantla Basin, Mexico: Effect of Organic Matter Content" Processes 13, no. 4: 1087. https://doi.org/10.3390/pr13041087

APA Style

López-Dinorín, R., Mendoza-Martínez, A. M., Palma-Ramírez, D., Dorantes-Rosales, H., García-Alamilla, R., Romero-Ibarra, I. C., & García-Zaleta, D. S. (2025). Analysis of Carbon Dioxide Mineralization in Carbonates from Tampico-Misantla Basin, Mexico: Effect of Organic Matter Content. Processes, 13(4), 1087. https://doi.org/10.3390/pr13041087

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