Next Article in Journal
Wind Turbine Blade Damage Evaluation under Multiple Operating Conditions and Based on 10-Min SCADA Data
Previous Article in Journal
Fast Power Coefficient vs. Tip–Speed Ratio Curves for Small Wind Turbines with Single-Variable Measurements following a Single Test Run
Previous Article in Special Issue
Research and Application of Carbon Capture, Utilization, and Storage–Enhanced Oil Recovery Reservoir Screening Criteria and Method for Continental Reservoirs in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbon Capture and Storage in Depleted Oil and Gas Reservoirs: The Viewpoint of Wellbore Injectivity

by
Reyhaneh Ghorbani Heidarabad
and
Kyuchul Shin
*
School of Chemical Engineering and Applied Chemistry, Kyungpook National University, 80 Daehak-ro, Buk-gu, Daegu 41566, Republic of Korea
*
Author to whom correspondence should be addressed.
Energies 2024, 17(5), 1201; https://doi.org/10.3390/en17051201
Submission received: 31 January 2024 / Revised: 16 February 2024 / Accepted: 29 February 2024 / Published: 2 March 2024

Abstract

:
Recently, there has been a growing interest in utilizing depleted gas and oil reservoirs for carbon capture and storage. This interest arises from the fact that numerous reservoirs have either been depleted or necessitate enhanced oil and gas recovery (EOR/EGR). The sequestration of CO2 in subsurface repositories emerges as a highly effective approach for achieving carbon neutrality. This process serves a dual purpose by facilitating EOR/EGR, thereby aiding in the retrieval of residual oil and gas, and concurrently ensuring the secure and permanent storage of CO2 without the risk of leakage. Injectivity is defined as the fluid’s ability to be introduced into the reservoir without causing rock fracturing. This research aimed to fill the gap in carbon capture and storage (CCS) literature by examining the limited consideration of injectivity, specifically in depleted underground reservoirs. It reviewed critical factors that impact the injectivity of CO2 and also some field case data in such reservoirs.

1. Introduction

Global warming is one of the main concerns of human beings currently and the global average temperature and sea level will reach 3.5 °C and 95 cm, respectively, by 2100 if the anthropogenic greenhouse gas (GHG) emissions continue to increase [1,2,3,4,5,6,7]. CO2 accounts for ~64–76% of the total global GHG emissions and is one of the pollutants that endanger public health and welfare [6,7,8,9,10,11].
Several strategies have been introduced to reduce the carbon footprint, including shifting the energy mix to less carbon-intensive sources, reducing energy consumption, replacing fossil fuels with fuels that have shorter carbon chains, improving energy efficiency, and long-term carbon capture and sequestration in underground structures [5,6,12,13]. Among these, carbon capture and storage (CCS) has garnered attention because it decelerates the rate of the increase in atmospheric CO2 concentrations [5,7,12,14,15,16,17,18,19,20,21,22,23]. CCS can contribute to ~12% of the global decarbonization target by 2050 and stabilize atmospheric GHG concentrations to ~450 ppm CO2 eq by 2100 [5,7,12,14,15,16,17,18,19,20,21,22,23].
Deep saline aquifers, depleted hydrocarbon reservoirs, and hydrocarbon reservoirs of enhanced oil recovery (EOR) and enhanced gas recovery (EGR) projects, unmineable coal seams of enhanced coal bed methane (ECBM) projects, and salt domes/mined caverns are primary mediums of CO2 sequestration [3,6,7,8,12,13,18,22,24,25,26,27,28,29,30,31,32,33,34]. CO2 is trapped in these storage mediums via (i) structural trapping, where an impermeable caprock stops the CO2 plume; (ii) capillary trapping, where CO2 remains in pore spaces as residual CO2 gas saturation; (iii) solubility trapping, where CO2 dissolution produces dense CO2-saturated brine; and (iv) mineral trapping caused by the reaction of minerals with CO2-saturated brine [8,27,35,36,37]. Table 1 shows storage media based on their capacity, cost, integrity, and technical feasibility.
The majority of studies on carbon capture, and storage (CCS) focused on assessing the storage potential of deep saline aquifers; however, oil and gas reservoirs, despite their relatively smaller storage potential, are ideal for CCS owing to their high capacity, containment, reservoir structure, and surface facilities that can be adapted for CO2 storage operations [3,4,27,38]. However, it should be noted that oil is considered as a hydrophobic fluid and has no harmful effect on pipeline walls, but CO2 is a moisture content that in contact with water causes sweet corrosion in pipelines [39,40,41,42,43,44]. The facilities already existing are designed for hydrocarbons so it makes the CO2 corrosion a significant problem in oil and gas production and transportation facilities that the cost of remediation can be higher than replacing the facilities [41,42,45,46,47].
Oil and gas reservoirs have also trapped hydrocarbons under caprock sealing for millions of years at high pressures, ensuring rock integrity for long-term CO2 geosequestration with less environmental impact [48,49,50]. CO2 sequestration in these reservoirs can provide an economic incentive, which is additional revenue from oil and gas recovered from CO2 injection with EOR/EGR technology [7,51,52,53,54,55]. Technologies for CCS can increase the oil recovery through EOR/EGR up to 40%, whereas a recovery factor of 1–18% can be obtained after CO2 injection in gas reservoirs [6,7,28,48,56]. A study using a Silurian core drilled from Door District in northeast Wisconsin, United States revealed an oil recovery of 34% during CO2 flooding [57]. Studies conducted on gas reservoirs and CO2–EGR field pilots have revealed that common depleted gas reservoirs have CO2 storage capacities of 390–750 gigatons [51,58].
The feasibility of CCS operation mainly depends on adequate storage capacity and threshold well injectivity, which ensure that the desired amount of CO2 is injected at acceptable rates through a minimum number of wells [3,5,29,59,60]. In other words, even with enormous storage capacities and high-quality overlying seals, CCS operation might not be financially viable without obtaining a minimum level of CO2 injectivity into a formation [3,5,29,59,60]. Injectivity refers to the fluid’s ability to be injected into a geological formation, i.e., the rate at which it can be injected into a storage medium without causing caprock fractures [7,12]. An adequate CO2 injectivity is a prerequisite for CCS projects and considerably influences their economics.
CO2 injectivity is strongly affected by the interactions between injected CO2 and rock minerals or fluids in storage sites [5,7,9,12,28]. Although CO2 storage mechanisms in oil and gas reservoirs have been extensively studied, the implementation of CCS at a field scale is met with some limitations [12,27]. Moreover, although the storage and factors affecting oil/gas recovery have been explored, factors influencing wellbore injectivity in depleted hydrocarbon reservoirs remain understudied [12,27]. Herein, wellbore injectivity was investigated along with the main factors affecting the pressure build-up in CCS projects in active/depleted oil and gas reservoirs.

2. Factors Affecting the Injectivity of CO2

Well injectivity issues are detrimental to CCS projects because large volumes of CO2 must be stored for long periods in geological time scales [61]. CO2 injectivity is mainly influenced by innate reservoir properties, residual gas/water saturation, residual oil/condensate saturation, injected fluid properties, mineral dissolution, salt precipitation, asphaltene precipitation, thermodynamic phase behavior of CO2 in the wellbore, clay swelling, injection rate, and wettability alteration [7,9,12,27,60,62,63,64,65]. These factors and their influences on the wellbore injectivity of CO2 geosequestration are discussed in subsequent sections.

2.1. Innate Reservoir Properties

CO2 injection affects petrophysical characteristics, necessitating in-depth investigation [49]. Therefore, the innate reservoir properties that influence CO2 injectivity were investigated herein: permeability, porosity, rock strength, composition of fluids, and heterogeneity level of the storage medium [12,18,66,67].

2.1.1. Permeability

Permeability is the ability of a porous medium to facilitate fluid flow and is measured in m2 (in the metric system) or Darcy (D) or milli-Darcy (mD in the oilfield system) [68]. The permeability of a formation is closely linked to injectivity [7]. However, Effective wellbore permeability is also a critical parameter for estimating wellbore leakage potential, which significantly influences the CO2 leak rate [69]. Given the close tie between well permeability and CO2 leakage, quantitatively assessing well permeability uncertainty is crucial for evaluating CO2 leakage risk [27,69]. Geochemical reactions as well as temperature can cause changes in permeability at both microscopic and mesoscale [9,70].
For effective CO2 storage, low permeability is essential, but >100 mD is needed for good injectivity near the well [67]. However, the permeability of a medium should be low for permanent CO2 storage assurance [67]. CO2 flow in tight, low-permeability rocks is controlled by reservoir heterogeneity and permeability, which demand significant capillary pressure for CO2 to penetrate the pores [48,71]. Positive total skin factor and wellbore permeability decline can result from partial penetration and formation damage [72].
Relative permeability significantly controls the injectivity, and pressure build-up in the reservoir [61,64,73]. CO2 migration through caprocks involves a two-phase flow with capillary effects; however, measuring relative permeability curves is challenging due to low caprock permeability [62].

2.1.2. Porosity

Petroleum reservoirs are porous rocks containing hydrocarbons and connate water [68]. Porosity quantifies the pore volume relative to the total volume [68]. Naturally fractured reservoirs (NFRs) have matrices and fracture zones [68]. Matrices have higher porosity, whereas natural fractures (NF) have higher permeability [68]. The alteration of rock permeability and porosity is a case-specific phenomenon and is influenced by the composition of injection rate of fluid, rock mineral type, pore geometry, and thermodynamics [61].

2.1.3. Pressure

During CCS processes, a significant pressure difference between discharge and target pressure is created, which causes high storage density in depleted reservoirs [38]. Fluid injection into reservoirs is a complicated process, which is a thermoelastic, poroelastic, and chemoelastic coupled problem, and is accompanied by the state change of in situ stresses in reservoirs [19,27,62,74]. Considering the effect of injection on the storage site, changes in pore pressure have a direct impact on rocks’ poroelastic properties [27].
High reservoir pressure in CO2 injection can lead to mechanical stress and deformation, impacting both the reservoir and caprock sealing, potentially reducing injectivity and emphasizing the importance of preventing reservoir pressure from surpassing caprock fracture pressure for maintaining CO2 containment and evaluating seal integrity to prevent fracture initiation [22,67]. The amount of oil produced rises and the time it takes for gas to break through reduces with increasing CO2 injection pressure [55].
According to a study on how reservoir depletion affects stress, as pore pressure drops during reservoir depletion, effective horizontal in situ stress rises by 50 to 80% [74]. CO2 enhances oil recovery in low-permeability reservoirs due to low viscosity and miscibility [49]. The injectivity is also controlled by the integrity of well bores, the failure of which causes rapid escape of injected CO2. In turn, the integrity of well bores also depends on the injecting pressure and the upper limit for injecting pressure.

2.1.4. Well Configuration

The reliability analysis of storage sites emphasizes seal capacity, geometry, and integrity as crucial factors [66,67]. The sealing effectiveness of faults, influenced by pore throat size and mineral properties like water-wettability, particularly benefits from minerals such as mica, muscovite, and phlogopite, enhancing the ability to contain CO2 plumes [67]. By increasing the pressure during the injection, the normal stress on a fault surface decreases and can lead to mechanical breakdown (reactivation) [27]. Moreover, CO2 leakage is a significant challenge in CCS, occurring through various pathways like casing-cement interfaces, cement-rock contacts, and degraded or fractured materials. Wellbore integrity is critical for preventing these leaks [75].

2.1.5. Heterogeneity Level of the Storage Medium

It is unknown how the storage reservoir’s injectivity is affected by compartmentalization and geological heterogeneities [9,66]. Geological heterogeneities are often classified into two main categories: the presence of alternating layers with varied mechanical properties, pore pressure, and/or lithology and permeability, and the presence of faults and compartmentalization within the specified storage reservoir [9,66]. Reservoir compartmentalization is influenced by fault structures and deposition history, impacting permeability between deposition units [9]. Existing faults and fractures before CO2 injection may also enhance and/or postpone fluid migration rates [76]. This is the reason why it is included as a vital factor in fully understanding the locations, geometries, and permeabilities of the reservoirs [76].
The multiphase flow of CO2-brine can be expressed in terms of permeability variation influenced by heterogeneity [77]. Storage capacity depends on heterogeneities and porosity; however, injectivity relies on petrophysical properties such as permeability [77]. Heterogeneities can cause unexpected outcomes in simulating injection processes and CO2 plume behavior in storage reservoirs, especially in depleted oil or gas fields [9]. Remediation options include acid injection to create high-permeability pathways and surfactant formulations to alter wettability and counteract CO2 trapping [9].

2.2. Capillary Trapping

Capillary trapping is a physical phenomenon in which CO2 is trapped as residual gas saturation (SgrCO2) in pore spaces due to capillary force [27]. Capillary trapping, a mechanism impacting injectivity, leads to residual CO2 saturation in rock pores, influencing storage capacity [78]. The minimum saturation level at which gas can start to flow is called residual gas saturation [79]. In other words, injected CO2 trapped in rock pores surrounded by water forms residual CO2 saturation during capillary trapping [78]. When there is no mobility threshold above the saturation level, this parameter is equal to the residual gas saturation [79]. Residual gas in depleted reservoirs can increase or decrease storage capacity, and decrease brine mobility, density, and viscosity of gas mixtures when dissolved in supercritical CO2 [67]. Based on a flooding experiment on four core samples (one composite and three Berea samples drilled from the Waarre C Formation in the CO2CRC’s Naylor Filed) it was concluded that early on in the CO2 injection process, residual natural gas in the depleted reservoir lowered the CO2 injectivity [80].

2.3. Residual Oil/Condensate Saturation

After primary and secondary oil recovery, residual oil saturation in reservoirs often ranges from 50 to 60% of the original oil-in-place (OOIP) [56,81]. Higher oil saturation linearly decreases storage capacity, but at 40% oil saturation, storage capacity does not vary very much [73].

2.4. Fluid Properties

Carbon dioxide transitions between gas, liquid, and solid states based on temperature and pressure variations [7]. Low levels of residual gas, water, and condensate in the reservoir are required for efficient CO2 injectivity [48]. Injecting supercritical or liquid CO2 in low-pressure reservoirs may cause evaporation in the tubing or wellbore [4,48]. Some properties are presented in Table 2 and the following section to understand more about the effect of fluid properties on wellbore injectivity.

2.4.1. Viscosity and Density of Injected CO2

Density and viscosity are crucial properties in the storage process, receiving significant attention for their impact on storage capacity and enhanced oil recovery rate [82]. The low-density gas phase reduces hydrostatic pressure, impacting flow stability and potentially causing cavitation during phase changes [7]. In the sites with a depth higher than 800 m, the pressures and temperatures reach above the critical points of 7.38 MPa and 31.1 °C, respectively, and CO2 can be injected as a supercritical fluid [12,22]. Therefore, sequestering CO2 in shallow reservoirs (<2600 ft. (800 m)) is discouraged because of nonsupercritical conditions, although such complexities can be avoided with an understanding of changing multiphase behavior [7]. Theoretically, CO2 temperature at the injector well bottom correlates positively with CO2 injectivity [18]. This is due to density and viscosity decrease with rising bottom hole temperature, enhancing CO2 mobility and injectivity [18]. The density of the supercritical CO2 is more like liquid, but the viscosity is like gas [71]. Furthermore, residual CH4 in reservoirs alters supercritical CO2 density and viscosity in pore space [48]. In a study conducted by Nicot, et al. [83] on the impact of viscosity on the geologic storage capacity in shallow depth, results showed that a decrease in viscosity of the CO2 mixtures leads to approximately the same proportion loss in the storage capacity [83].
Above 31.1 °C and critical pressure, CO2 is supercritical with a mass density of ~0.3−0.8 g/cm3 (less dense than coexisting brine), which is a crucial behavior for storage considerations [7,84]. CO2’s higher density promotes the stability of displacement fronts, and its supercritical state enhances efficient subsurface storage [85]. CO2 density influences hydrocarbon extraction; heavier hydrocarbons are extracted at higher densities [71]. In addition, the efficiency of CO2 storage increases at higher CO2 densities, enhancing safety by reducing the buoyancy force [12]. CO2 density at reservoir condition (the temperature between 293 K and the pressure between 25 bar and 700 bar) can be estimated by [67]:
ρ = α + βT + γT2 + θT3
α = (A1 + B1P + C1P2 + D1P3)
β = (A2 + B2P + C2P2 + D2P3)
γ = (A3 + B3P + C3P2 + D3P3)
θ = (A4 + B4P + C4P2 + D4P3)
In the mentioned equations, α, β, and γ are the temperature coefficients. P is the pressure in bar scale and T is the temperature in Kelvin. ρ is the density with the scale of kg/m3, and θ is the contact angle representing the medium wettability [67].

2.4.2. Injected CO2 Purity

CO2 can be captured from fossil fuel power plants but it comprises a variation of impurities such as N2, NOx, Ar, O2, and SO2 in different concentrations [86]. Injecting this CO2 may affect the amount of storage in a geological medium [27]. According to Wang, et al. [87], who performed a study on the effect of H2S and SO2 on CO2 injectivity, coinjecting impurities with CO2 for storage is cost-effective, but impurities negatively affect transport, injection, and storage [87]. Acid impurities such as SOx and NOx react with rocks, impacting injectivity and storage integrity; hazardous impurities pose environmental risks in the case of a CO2 leakage [87].
The coinjection of these impurities may also impact well injectivity and wellbore integrity, thereby reducing porosity, cap rock integrity, CO2, and water containment altering formation salinity near the well and mineral composition in the reservoir [9,83,86]. It also affects static capacity by altering the density and viscosity of the CO2 mixture [83]. Lower density impacts CO2 capacity due to impurity space and generally lower impurity density [83]. Impurity type influences thermal front location, delineating the radial zone with significant induced reservoir cooling [88]. They have a more pronounced impact on plume shape at shallow depths [67]. Separating impurities before injection is crucial for maintaining storage capacity, and removing CO2 moisture is necessary to prevent corrosion and hydration-related costs [67].
Impurity removal is expensive; therefore, their coinjection will considerably reduce the cost of CO2 capture [86]. The cost of a CCS project depends on CO2 separation, with the impurity level affecting storage capacity [59]. The higher the level of impurity, the lower the storage capacity for CO2 and the lower the CO2 injected [59]. In addition, injecting a pure CO2 stream that is free of impurities and water, prevents corrosion and formation damage caused by insoluble iron precipitates, thereby preserving injectivity in porous media [89].

2.4.3. Injectant Temperature

Injectant temperature directly influences total horizontal stress, with lower temperatures reducing near-wellbore stress significantly [19]. Repetitive thermal loading can cause the failure of well barrier material, particularly in CO2 injection wells that experience temperature variations from injected fluid and rock [21]. The temperature fluctuations can range from 15 °C to 25 °C and 6 °C to 7 °C for onshore and offshore transport, respectively [21]. Thermal stress-induced wellbore damage is influenced by factors such as injection and formation temperature, formation stress state, and the thermal/mechanical properties of well barrier materials [21].
Table 2. The impacts of fluid properties on wellbores injectivity.
Table 2. The impacts of fluid properties on wellbores injectivity.
ReferenceStudy Remarks
Jin, Pekot, Smith, Salako, Peterson, Bosshart, Hamling, Mibeck, Hurley and Beddoe [37]CO2 saturated Mead-Strawn stock-tank oil at 135° F showed that the density of oil increases when CO2 is dissolved in the oil [90].
The gas storage rate in the Bell Creek oil field is linked to the injection rate, decreasing as the injection stabilizes.
Kazemzadeh, et al. [91]The minimum miscible pressure (MMP) is the ideal pressure for cost-effective injection in oil recovery.
Barrufet, Bacquet and Falcone [59]The duration of a project on a gas condensate fluid from the Cusiana field located in the northeast of Bogota, Colombia, in the Lianos basin is determined by injectivity, injection rates, and the number of wells; injection rates do not affect the eventual storage capacity.
Tawiah, Duer, Bryant, Larter, O’Brien and Dong [18]Injection rates in reservoir rocks near the wellbore are influenced by injection pressures, fluid saturation, and fluid mobility.

2.5. Mineral Dissolution/Precipitation

Mineral dissolution/precipitation in CO2 (supercritical) and water–rock interactions enhance CO2 trapping and alter mineral surface wettability, which is crucial for residual trapping [16]. CO2 injection induces mineral dissolution and precipitation determined via the compositions of the original formation water and rock samples [49,92]. Increasing temperature and decreasing fluid pressure led to reduced carbonate solubility and CO2 degassing [93]. An increased concentration of Ca or Mg ions from the dissolution of rock can lead to the rapid mineralization of CO2 [36]. CO2 dissolution in brine for storage triggers mineral reactions, which transform reservoir mineralogy and influence petrophysical properties such as porosity and permeability [94,95]. Mineral dissolution (chemical mineral dissolution of pore-filling cements such as carbonate and anhydrite [94]) improves formation permeability [7,93,94,96], porosity [93,94,96] and the proportions of pore-exposed grain-rimming clay coatings [94]. Dissolution near the injection site may increase CO2 storage capacity within the medium, increasing the amount of localized CO2 storage [97], altering the transport of CO2 [16], and inducing surface cracking, which can increase the reactive surface area [70]. Uniform mineral dissolution was observed during the simulation of pure quartz sandstone, with slow surface reaction during CO2 injection into sandstone [98]. The porosity and permeability of the reservoir are increased when rock minerals dissolve, whereas they are decreased when carbonate or sulfate compounds precipitate [49]. Sokama-Neuyam and Ursin’s [99] study presents evidence that mineral dissolution negatively impacts CO2 injectivity, reducing the efficiency of CO2 injection. Injectivity impairment experienced a reduction of 9% when brine salinity was halved. The experimented sample belonged to Kocurek Industries, Caldwell, TX, USA, and the impairment of injectivity was measured via the pressure drop measurements [99].
Hydrated well cement, composed primarily of C-S-H and Portandite, is a reactive component in the near-well zone [95]. However, dissolved minerals that aggregate into fine particles in rock formations can enhance chemical reactions, causing pore throat blockage, reducing permeability, and impacting CO2 mobility near plume boundaries [7,16,94,95]. Moreover, the permeability and porosity increase because of calcite [7], anhydrite [7,94], and carbonate dissolution; it also provides improved fluid pathways [94]. Pore throat sealing minimally affects total rock porosity but significantly deteriorates its permeability [94].
The main reactions that occur during the dissolution of CO2 in water are as follows [98]:
CO2 + H2O H2CO3
H2CO3  HCO3 + H+
HCO3  CO32− + H+
CO2 dissolves into formation water and generates H+, H C O 3 , and C O 3 2 ions, which then react with specific ions in the formation water and rocks [49]. The concentrations of C O 3 2 and H C O 3 , are increased by a larger volume fraction of injected water and an increased Ca2+ content in the formation water [49].

2.6. Salt Precipitation

Salt precipitation due to water evaporation has been recognized both experimentally and numerically and proven to lead to the abandonment of wells [61,95]. Formation water that is rich in ions reacts with injected CO2, vaporizing and precipitating salts in reservoirs [7]. Salt precipitation also occurs in gas condensate reservoirs [61] and is a longstanding issue in the gas and petroleum industry [100]. CO2 partially dissolves in brine and vaporizes saline water [63]. Evaporated water increases brine salt content, potentially leading to halite deposition if the solubility limit exceeds (~26.5% by weight) [63]. If the salt concentration exceeds the solubility limit under reservoir conditions, minerals precipitate [63,101]. The precipitated salt fills the porous space and clogs the pore network of the formation, modifying the flow channels [7,63,101,102]. It can reduce porosity and permeability, change capillary forces, and cause injectivity loss around the wellbore [5,7,24,27,61,63,95,100,101,102,103] completely blocking the injection [101] and reducing oil/gas productivity [63]. The decrease in permeability reduces the overall porosity and affects the pore space geometry as well as the precipitate distribution within the pore space [101]. The impact of precipitation on permeability varies based on reservoir chemistry and pore structure [103].
Concentrated halite precipitation results from sufficient brine mobility caused by a capillary pressure gradient, which impacts injectivity [101]. Simulations demonstrated that even when dry CO2 is introduced into a depleted gas reservoir that contains medium-salinity brine, halite can still precipitate [101]. In addition, the combined effect of salt precipitation and fine migration can decrease the permeability three-fold compared to salt precipitation alone [102].
Salt precipitation is high around the injection well because the fluxes, concentrations, and saturation gradients of injected fluids are highest [5]. Even minimal salt deposition near the injection zone can cause significant CO2 injectivity impairment [5]. However, this effect varies based on initial liquid saturation [103]. With sufficiently mobile brine, the continuously recharged precipitation front that is driven by capillary pressure considerably reduces the formation permeability [103].
For reducing salt precipitation, the most common remediating injectivity method is to inject chemicals [9]. In addition, salt precipitation in production wells can be reduced by diluting produced water with low-salinity water downhole and in production systems [5].

2.6.1. Effects of CO2 Flow Rate

A critical scCO2 injection flow rate influences particle migration in porous media, affecting salt precipitation and permeability [102]. Evaporation rate, directly linked to injection rate, influences brine concentration; increased gas volumes enhance halite precipitation [63]. High CO2 injection rates may detach formation fines, causing pore clogging and reduced injectivity [5]. Optimal brine salinity and injection rate mitigate salt precipitation efficiently [5]. Because of the availability of NaCl, solid saturation rises with initial brine saturation, except at very low rates (0.1 kg/s) [101]. Moreover, high injection rates create a higher pressure gradient, suppressing capillary backflow and reducing the possibility of intensive salt accumulation [100]. Injectivity significantly decreases to a low of 40% point at a CO2 injection rate of >5 mL/min possibly due to the uneven mineral distribution of a Berea sandstone core with a diameter of 3.81 cm and length of 7.62 cm ± 0.05 [102].

2.6.2. Effects of Brine Salinity

Brine salinity intricately affects CO2 injectivity and reservoir properties [63]. Initial brine salinity is crucial for determining the residual salinity and salt precipitation; it also influences porosity and reduces permeability [63]. Higher brine salinity increases both salt precipitation and permeability reduction [100]. Injectivity of the Berea sandstone cores improves with reduced brine salinity but declines below 21.102 g/L [102]. Significantly reducing brine salinity by almost half results in only a marginal 9% injectivity decline, highlighting nonlinear mineral precipitation and injectivity dependency [5]. Low brine salinity, such as LSW, induces colloidal particle detachment, causing pore bridging and reduced CO2 injectivity [5]. Extremely low salinity may cause simulation deviations, causing chemical interactions and impairing injectivity [5]. The positive correlation observed between brine salinity and injectivity impairment emphasizes the importance of brine salinity in assessing the effectiveness of CO2 injection [102].

2.6.3. Effects of Pore Size

Initial permeability considerably influences subsequent permeability losses during CO2 injection; lower initial permeability causes more significant reductions in permeability [63]. Higher permeability rocks have larger, less susceptible pore throats; thus, the permeability decreases only slightly [63]. Moreover, the pore channel size plays a crucial role in the dissolution of precipitates [5]. The impact of precipitates on injectivity is expected to be more pronounced in low-permeability rocks due to the potential plugging of narrow pore channels, even with the same precipitation rate [5].

2.6.4. Effects of Particle Size

While the correlation between particle size and pore throat size is a crucial factor, there is a theoretical understanding that the concentration of the suspension may significantly influence plugging mechanisms [102]. This is attributed to the shortened distance between suspended particles as the concentration increases, thereby intensifying the multiparticle blocking of invaded pores [102]. Injected dry CO2 induces persistent water evaporation and halite precipitation around the well [101]. Despite potential halite effects, a high CO2 injection rate can mitigate injectivity issues depending on reservoir factors such as the formation’s characteristics, initial reservoir thermodynamic conditions, initial brine saturation, and salinity [101].

2.6.5. Effects of Water Saturation

Irreducible water saturation (Swi) determines the maximum relative permeability of CO2; higher Swi results in lower maximum permeability [63]. A decrease in water saturation triggers brine capillary backflow, sustaining evaporation and precipitation [63].

2.6.6. Effects of Temperature

Elevated temperatures decrease water saturation and increase halite and salt precipitation in high-temperature gas reservoirs [63]. Rising temperature increases CO2 phase water solubility, causing rapid saturation and salt precipitation [100]. Temperature only slightly impacts salt precipitation compared with factors such as injection rate and capillary pressure [100]. Numerical simulations were conducted to assess the dry-out process during CO2 injection; however, the impact of halite precipitation on field operations could not be accurately predicted due to undefined relations between changes in porosity and permeability [103].

2.7. Asphaltene Precipitation

Crude oil is a complex mixture containing different hydrocarbons, primarily saturates, aromatics, resins, and asphaltenes (SARA) [56,104,105]. The most polar components are asphaltenes because they contain heteroatoms like oxygen, sulfur, or nitrogen, while saturated and aromatic compounds are nonpolar [55,106]. The stability of asphaltene depends on aromatic-to-saturate and resin-to-asphaltene ratios in crude oil [105]. Resins are an excellent bridging agent for all components of crude oil because they contain both polar and nonpolar sites [55,105,107].
Asphaltenes, the heaviest and most polar crude oil fractions, have a molecular weight of 100 to 10,000 g/mol [55,56,81,104,105,106,107,108,109,110,111]. They are composed mainly of carbon and hydrogen, with condensed aromatic rings surrounded by insoluble aliphatic chains [55,56,81,104,105,106,107,108,109,110,111]. While insoluble in light hydrocarbon solvents, asphaltenes are soluble in aromatics such as toluene and xylene [55,56,81,104,105,106,107,108,109,110,111]. They are generally incompatible with light petroleum fractions leading to undesirable effects in many stages of the petroleum industry such as pipeline routes, oil and gas production, and EOR/EGR [56]. The nature of asphaltenes is not clearly understood due to their complex nature and because various parameters affect their precipitation [56,104]. Thus, a universal model to predict and simulate their precipitation has not been developed yet [56,104,109]; their exact chemical structure also remains unknown [109].
Crude oil components are in equilibrium under reservoir conditions, but precipitation can occur due to pressure or temperature changes during oil field operations, such as high flow rates or gas introduction [55,104,107]. Asphaltene precipitation can be problematic and has caused major issues with flow assurance for the oil industry in recent decades [106,111]. For instance, when CO2 is injected into the oil reservoir formation, the equilibrium state of the asphaltene–oil colloidal system in porous media is altered by variations in thermodynamic conditions, such as pressure (e.g., during primary depletion of highly under-saturated reservoirs [109]), temperature, asphaltene concentration, CO2 content, oil composition, flow condition, and chemical additives; this results in a liquid-like solid precipitation with high viscosity [52,54,81,106,107,109]. Asphaltene precipitation can occur from reservoirs containing even very small amounts of asphaltene [107]. This will be far more severe when the reservoir pressure drops below the bubble point if the reservoir pressure is initially above the bubble point [55,104]. The molar volume of oil increases as the pressure decreases, subsequently decreasing the asphaltene solubility [105]. The maximum amount of asphaltene is precipitated at the bubble point [104,105] both with/without CO2 injection [104] because of high amounts of dissolved gas by volume in oil [104,105]. In addition, precipitation is more at higher CO2 injection rates [54].
The interaction coefficient between CO2 and asphaltene is considerably higher than other components (natural gas and nitrogen [110]) with asphaltene [56,110]. Elevated CO2 concentration increases the bubble point, precipitation, and interaction coefficient with asphaltene due to shared polarity [56,104].
Asphaltenes precipitated during CO2 flooding, can either remain suspended or deposit onto surfaces, particularly high-specific-area clay minerals [28,52,54,55,56,81,104,105,107,109,110]. This deposition can lead to reservoir and wellbore pore plugging, reduced porosity, permeability, and CO2 injectivity, and altered wettability from water-wet to oil-wet [28,52,54,55,56,81,104,105,107,109,110]. Formation damage and a negative impact on production efficiency may result from pressure drop increase and changes in pore surface wettability [28,52,54,55,56,81,104,105,107,109,110]. Precipitation also threatens the capacity of surface facilities by plugging tubular and flow lines and clogging production separators, such as damage to pumps [56,104,106]. Despite all these disadvantages, asphaltene precipitation yields oil with less asphaltene content [81,110], making it lighter and less viscous than crude oil [56].
The degree of asphaltene precipitation during CO2 injection is influenced by factors such as injection method, pressure, and miscibility [55]. It also depends on the asphaltene content in crude oil, correlated with injected CO2, reservoir conditions, pressure, and temperature [28,112].

2.7.1. Effects of Permeability

Asphaltene deposition reduces permeability [112] and improves oil displacement by water due to increased oil relative permeability [108]. In particular, permeability affects the stability of asphaltenes and has a significant impact on deposition [111]. Deposited asphaltenes are more stable and migrate significantly less in more permeable reservoirs [111]. With increasing rock permeability, the impact of asphaltene deposition on the decreasing core permeability and oil recovery reduces [113]. Oil reservoirs with lower permeability exhibit more severe formation damages caused by simultaneous sulfur and asphaltene deposition than those with higher permeability [112,113]. Permeability reduction fluctuates between 40% and 90%, influenced by fluid composition, porous medium pore structure, and proximity to the core inlet or outlet [111]. Maximum asphaltene precipitation was reported at the core inlet, with a linear relation between permeability-reduction factor and asphaltene amount [107].

2.7.2. Effects of Pore Size Distribution

The characteristics of crude oil, the pore-plugging mechanism, and the deposition process are all impacted by the size distribution of precipitated asphaltene [56]. The analysis of pore size distribution can be utilized to determine how much asphaltene deposition reduces permeability [112]. Pore size distribution must be determined to identify fluid transport properties of porous media [108]. With increasing pore sizes (pore diameter > 8 μm [111]), the weight percent of asphaltene decreased and the oil recovery rate increased [55,111]. Small pores (<8.0 µm [111] or 9.0 µm [112] in radius) showed higher sensitivity to more asphaltene deposition, indicating their significant contribution to reducing permeability [111,112]. Asphaltene precipitation in unconventional reservoirs may be more severe than in conventional reservoirs due to considerable differences in pore sizes [55].

2.7.3. Effects of Temperature

Temperature is recognized as a relevant parameter for asphaltene stability [106]. When the temperature is raised during the CO2-EOR process, asphaltene from crude oil can become unstable since CO2 is usually in a supercritical state [54]. In multiphase flows, low heat transfer increases asphaltene deposition depending on the concentration and velocity of reactants [28]. The temperature of the reservoir considerably influences the kinetics of asphaltene precipitation, controlling the precipitation onset time and the size of deposits [54].
Oil recovery increases with increasing temperature; however, CO2 breakthrough time decreases due to low oil viscosity [55]. Higher temperature also leads to increased asphaltene weight percent in bypassed oil due to resin instability [55]. Moreover, the asphaltene precipitation onset point increases with CO2 injection at higher temperatures, which also increases the solubility parameter difference value [114]. In contrast, lower temperatures with liquid CO2 decrease asphaltene stability due to its nonpolar behavior [106]. At higher temperatures, when CO2 becomes supercritical, polar–polar interactions are enhanced and asphaltene precipitation is reduced due to steric repulsion and increased side-chain motion [106].

2.7.4. Effects of CO2 Concentration

The amount of asphaltene precipitation depends on the concentration of injected CO2 gas and rapidly increases when the CO2 concentration exceeds its critical value [107]. The formation of a gas phase can cause the asphaltene phase volume to reduce at very high CO2 concentrations [109]. Maximum amounts of asphaltene precipitates were obtained at saturation pressure, which gradually increased with an increase in the mole of CO2 gas [107]. Higher CO2 concentration increases asphaltene precipitation in the single-phase region and bubble point [115]. At higher concentrations of injected gas and reservoir pressures above and below the bubble point, asphaltene precipitation increases [104].

2.7.5. Effects of Porosity

The porosity and asphaltene accumulation along the core are correlated [111]. Porosity decreases in response to any increase in asphaltene deposits in a particular core region, and vice versa [111]. When the injection pressure was roughly at the MMP, there was an increase in oil recovery from the smaller pores [112]. CO2 penetrated the small pores during miscible flooding, improving the oil recovery [112]. This highlights how crucial the MMP is for improving oil recovery from small pores [112].

2.7.6. Effects of Pressure

Asphaltene precipitation increases with injection pressure [55,112,114]. By increasing the injection pressure, more CO2 molar percentage is required to achieve asphaltene precipitation [114]. For samples of crude oil, there was a small increase in the mean asphaltene particle size as the pressure was raised [116]. In contrast, Lei, Pingping, Ying, Jigen, Shi and Aifang [52] reported that at a constant injection pressure, the asphaltene precipitate amount first increases and decreases as the injected CO2 amount increases; the asphaltene precipitate amount reaches its maximum when the gas phase occurs in the CO2–oil system [52].

2.7.7. Effects of Viscosity

The gas phase, the liquid phase rich in hydrocarbons, and the liquid phase rich in asphaltene have different viscosities depending on the phase composition, temperature, and pressure [109]. The dissolution of CO2 in crude oil and reduction in its asphaltene content considerably reduce oil viscosity, thereby decreasing the weight percent of asphaltene [109]. This is one of the main reasons for lower amounts of oil production at higher viscosity [55]. The weight percentage of asphaltene showed a positive correlation with the rise in viscosity for both the oil extracted and the oil that bypassed the system [55].

2.7.8. Effects of Flow Rate

By increasing the CO2 flow rate, pressure decreases considerably, indicating asphaltene deposition and permeability reduction [28,113,117]. Thus, by reducing the flow rate, the formation damage can be considerably reduced when producing crude oils with medium and high contents of sulfur and asphaltene [113].

2.8. Fine Mobilization

Mobile particles with an equivalent diameter of <40 μm are known as fines [9]. They are assumed to be initially located on the surface of quartz grains [118]. The nonswelling and swelling clays that detach from the pore-grain interface can migrate [9,119]. Fines mobilize through chemical or mechanical interaction with pore fluid, including clay swelling and fluid flow-induced mobilization [9]. During CO2 injection, fine particles are lifted into the reservoir and possibly plug the pore channels depending on the petrophysical characteristics of the rock, particle sizes, hydrodynamic conditions, solid concentrations, reservoir properties, and ionic strength and/or pH of the carrying fluid [7,9,60,89,94,95,120,121]. Born repulsion, electrical double-layer repulsion, and London–van der Waals attraction contribute to the potential energy that characterizes fine detachment [122]. The short-range potential known as the Born repulsive potential is created by the electron clouds’ overlap [122]. Fines can originate in the injected fluid due to its contamination by contact with casing cement, drilling fluid filter cake residue on the wellbore wall, or from the formation itself [9,60,94]. The result of experiments on the Berea core sample with permeability (60–100 mD) and porosity (19–20%) showed that fines mobilization harms injectivity more than drying or high salt concentration [60].
The critical salt concentration (CSC) is defined as a specific salt concentration at which the fine particles may be released [122]. It can impair formation permeability (which some laboratory observations suggest may be permanent [119]) by blocking/bridging the pore throats [7,60,94,118,120,121] and productivity [7], but negligible porosity change [121]. Significant injectivity reduction could be caused by minor particles in the pore fluid or wellbore fluid in the immediate injection area [9,60,94].
The modeling of fine migration and effects of mineral dissolution can be expressed by [123]:
k k 0 = 1 [ 1 + β s σ s β ( φ φ i ) ]
where β and βs are the dissolution coefficient and the formation damage coefficient, respectively, and σs is the volumetric concentration of retained fines [123]. Their values for the same brine salinity can be achieved by Wang, Bedrikovetsky, Yin, Othman, Zeinijahromi and Le-Hussain [123]. k and k0 are permeability and initial permeability, respectively. φ is defined as porosity and φi is the initial porosity [123].
Well impairment can lead to many challenges, including the possibility of an uncontrolled rise in cap rock pore pressure above the permissible fracturing pressure, the start and spread of fractures, and preferred flow pathways for CO2 leaks [120]. Injectable wells may need to be abandoned in severe cases [120]. Clogging caused by fine release, migration, and capture is commonly considered an irreversible process [120]. However, some remedies can be suggested such as high injection velocities for deposition away from the wellbore due to radial flow because of the radially divergent nature of the injection flow [9]. In addition, greater permeability impairment is caused by core plugs with higher clay contents, most likely as a result of larger clay particles and smaller pore throats that make blocking and bridging easier [121]. Injecting at higher permeability intervals may further reduce the possibility of injectivity loss [121]. Clay fines can only detach when the pore space’s salinity falls below a particular critical salinity, which can be prevented by raising the injected water’s salinity above the CSC [119].

2.8.1. Effects of Permeability

In typically highly permeable sandstones, fine migration has been studied as the main cause of permeability decrease [9]. The injectivity will rapidly decrease due to a decrease in permeability resulting from increasing pore pressure [120]. Until the permeability is restored through additional physical or chemical treatments, the fluid can no longer be injected into the damaged formation at the required high flow rates [120]. A decrease in permeability implies that higher pressure will be required to inject the fluid into the formation, causing changes in injectivity [7]. The low-permeability reservoir rocks showed a more noticeable decrease in permeability [121]. Therefore, low-permeability reservoirs show lower injectivity loss [7,121]. Even within a short timeframe, changes in reservoir permeability affect injectivity, productivity, and CO2 flow dynamics, impacting saturation distribution [120].

2.8.2. Particle Concentration

Particle concentration considerably affects CO2 injectivity [7]. The concentration of fine particles in the fluid has a direct impact on the deposition rate [28]. The interaction between the particle and the pore throat increases as the particle concentration rises [7]. CO2 injectivity impairment increased with an increase in the concentration of fine particles [7,60].

2.8.3. Injection Rate

Injectivity loss increased with increasing CO2 injection rate [7]. Turbulence from a higher flow rate makes the particle stack tighter and enables their even redistribution in the pore network [7]. In contrast, Sokama-Neuyam, Ginting, Timilsina and Ursin [60] concluded that the CO2 injection rate did considerably influence the reduction in CO2 injectivity induced by fine entrapment [60]. However, at high injection velocities, the strong fluxes may dislodge the particles bridging the pore channels, opening some of the clogged channels [60]. The fluid’s density, the square of its velocity, viscosity, and compressibility, as well as the size and form of the particles, all affect how much force is needed to raise them [60]. Hydrodynamic force acting on the particles is increased by supercritical CO2’s gas-like viscosity and liquid-like density [60]. Typical CO2 injection rates under storage conditions are approximately 1 Mt/pa [60].

2.8.4. Particle Size

Particle size and pore constriction, or more pertinently, the size of the fine particle to pore constriction, is crucial for determining the entrapment or piping mechanism occurring in the pore throat [7,60]. The larger the size of the particle, the higher the injectivity loss [7].

2.9. Clay Swelling

Many targeted geological storage sites are sealed by shale or mudstone rich in clay minerals such as sandstone gas reservoirs, which are heterogeneous and mainly composed of clay minerals and silica with small amounts of carbonates [35,58,124,125]. Clays are interlayer aluminosilicate minerals that have different structures [58]. They are crucial for the geological storage of CO2 [126]. CO2 is stored for a long term by overlying caprocks that act as low-permeability barriers to upward fluid flow [17,124,125,126,127].
Clay minerals are generally <1 mm in diameter and are therefore known as micro to nanocrystalline materials [128]. Clay minerals’ layered structure and atom arrangement give them a platy morphology [128]. Illite, chlorite, smectite, and kaolinite are the main types of clay found in typical sandstone rocks [58]. A property that sets swelling clay minerals apart from other clays and micas is their easy interchangeability with the surrounding environment [125]. Of the different common rock-forming minerals, calcite is the most reactive because of its high solubility and kinetics [24]. These rocks are frequently composed of smectites, or expanding clays like montmorillonites, which are frequently found in faults that laterally seal possible storage reservoirs at depths and temperatures of up to 3.5 km and 100 °C, respectively [17,124]. At greater depths, smectites generally begin to transform into nonswelling clay minerals [124]. Smectites, a mixed layer of illite and smectite, and a mixed layer of chlorite and smectite are most sensitive to water and are hydrophilic [129]. These minerals, when present in the pore network, can drastically reduce the intrinsic permeability [129]. When the clay is exposed to solutions containing cations, the present cations may be exchanged with other cations [129]. The highest amount of total cations in clays that can be exchanged with a solution of a particular pH is known as the cation exchange capacity of clays [17,129]. The different capabilities of CO2 adsorption and selectivity of mixed clays are caused by a variety of elements in clay structures, including cation exchange capacity (which is low in nonswelling forms such as kaolinite), charge on the clay surface, and interlayer distance [58]. Smectite clays are often used as samples due to their high swelling capacity [129]. The quantity and thermodynamic properties of water in the system have a significant impact on the intercalation and retention of CO2 in smectite interlayers [130]. Illite and kaolinite, not known for interlayer expansion under any experimental conditions, can adsorb significant amounts of CO2 [17]. If the surface area of the clays contacting the solution is large, the activities of divalent cations may be reduced either via sorption onto clay surfaces or via cation exchange [128]. Clays also increase the adsorption capacity of sandstone rocks because of their high surface area [58].
Clay swelling is a major cause of damage formation in hydrocarbon reservoirs and can substantially reduce nanoparticle mobility in porous media [127]. It can cause substantial changes in the reservoir structure, blocking the swelling pores, and causing dynamical behavior of the intercalated fluid molecules concerning the bulk fluid phase [17,125,130]. Another way to clog pore throats and reduce injection rates is by the mobilization of detrital or diagenetic clays [67]. The term “swelling clay” is derived from the ability of clay particles to increase their molar volume, thereby shrinking or contracting the interlayer pores based on the migration of polar molecules, such as water or organic molecules, into and out of the interlayers [9,124,125,127,130]. This migration depends on the clay mineral type and its hydration state [9,124,125,127,130]. The impact of clay swelling is most severe when incompatible fluids (e.g., oil) come in contact with swelling clays, leading to reduced formation permeability [127]. The same phenomenon is observed in the presence of CO2 in brine solution or in contact with resident brine [9,131]. CO2 can laterally displace the fluid and brine from the rock to assist in the intercalation of other chemical species in the clay, particularly during the early stages of injection [35,125,130].
The long-term efficacy of impermeable cap rocks in sealing the reservoir against the leakage of injected CO2 needs further evaluation [125]. To ensure a safe CO2 injection into reservoir formations, caprock failure and subsidence must be prevented, and the overlaying caprock must operate effectively [132].

2.9.1. Effects of Pressure

Pressure and temperature changes do not impact the clay nanoscale structure, suggesting stable seal quality during CO2 injection [84]. At lower pressures, the CO2 adsorption selectivity is enhanced, whereas at higher pressures, the adsorption amount increases [53]. Nonswelling clays can exhibit swelling under high pressure and temperature with CO2 and water coexistence [133]. Supercritical CO2 can alter capillary entry pressure in swelling clays, affecting injectivity, and may reduce long-term CO2 storage capacity due to mineral dissolution and precipitation [9,131].

2.9.2. Effects on Strain

Because of the strain that the intercalated molecules may produce, the pore can swell or shrink [133]. Strain changes induced by CO2 injection are considerably larger than those from pore-pressure changes alone [132,133]. CO2 injection initially increases the strain rapidly and then stabilizes [35]. Strain changes vary spatially due to sedimentary structure, with local low-porosity layers acting as barriers [35]. Swelling strain rapidly increases with CO2 pressure and decreases at higher pressures [17]. Linear strain is the length change relative to the initial length, and volumetric strain is the volume change relative to the initial volume [134]. Swelling strain can be predicted when micropore volume is known [133].
Material strain varies with the consolidation state [132]. Consolidated conditions exhibit higher strains than overconsolidated conditions due to microstructural differences [132]. Larger strain changes are observed due to CO2 adsorption, particularly in kaolinite-rich rock [35]. Moreover, swelling strain can be substantial, leading to differential matrix swelling during brine displacement [35]. CO2-induced swelling stress in Na-SWy-1 montmorillonite decreases as the effective stress and burial depth increase [124].

2.9.3. Effects of Temperature

At higher temperatures, CO2 adsorption on sandstone rocks decreases [58]. Bandera sandstone showed the least reduction due to favorable CO2 adsorption conditions on carbonate minerals [58]. By increasing the treatment temperature, the aromatic and aliphatic hydrocarbons in clay-rich shales are reduced [131]. Hot CO2 injection doubles natural gas production and improves CO2 sequestering in depleted gas reservoirs [58].

2.10. Hydrate Formation

The expansion of CO2 is linked with the Joule-Thomson phenomenon, potentially causing the formation of dry ice or hydrates, consequently diminishing the injectivity of CO2 [4]. Hydrates are formed due to the interaction of water with CO2 and gaseous hydrocarbons such as methane during injection into depleted reservoirs [89]. Such formations depend on specific conditions such as injection rate, pressure, and lower temperatures [89].
Figure 1 compares the impact of injectivity loss with cases involved with salt deposition, mineral dissolution, fine migration, hydrates, and also without injectivity disruption [89].
In Figure 1 case 1 was modeled using the following reaction in CMG simulators [89]:
Halite Cl + Na+
The value for case 2 was obtained from modeling the below reactions based on the reservoir rock mineralogy of a sample composed of 1.1% Calcite and 0.4% Dolomite [89].
Calcite + H+  Ca2+ + HCO3
Dolomite + H+  Ca2+ + Mg2+ + 2HCO3
Case 3 was calculated from reaction (11) and in the case of hydrate, it was modeled in CMG-STARS assuming that the permeability of the block was zero from the initial hydrate formation time [89]. Injectivity decreased to 73% from the base case in specific conditions [89].

3. CO2 Injectivity in Field Cases

This field case study aimed to improve a general understanding of CO2 injectivity into depleted hydrocarbon reservoirs by analyzing practical examples. The results are expected to provide concrete insights and useful recommendations to those who lead the responsibility of influencing environmental and energy management. However, the rapidly evolving confidential matters in CCS political settings may have resulted in uncertainties, leading to the potential for outdated information.

3.1. Niagaran Pinnacle Reef Oil Field

In the depleted pinnacle reef fields in Michigan, USA, CCS operations are performed by the Midwest Regional Carbon Sequestration Partnership (MRCSP) [135,136,137,138]. In the Michigan Basin, there are several hundred pinnacle reef structures, including this field [136]. It is a pinnacle reef that is late Silurian in age and has undergone significant primary and secondary recovery phases [136]. The shallow shelf carbonate depositional system that covered the Lower Peninsula of Michigan, northern Indiana, northeastern Illinois, eastern Wisconsin, northwest Ohio, and the Bruce Peninsula of Ontario gave rise to the reefs along the Niagaran Pinnacle Reef Trend oil fields [135].
MRCSP studies an extensive reef trend with over 880 closely spaced, highly compartmentalized, and laterally discontinuous reefs (Figure 2) [138]. The northern reef trend is divided into gas, oil, and water zones [138].

3.2. Netherlands Fields

The Netherlands leads the European Union in natural gas production with over 190 exploited gas fields. Fields in mature or depleted phases in the Netherlands are used for CO2 storage [139]. Induced seismicity affects 15% with ML w3.6. CO2 storage in depleted gas fields is favored due to proven seal quality and there is no record of seal integrity failure due to fault reactivation in seismically active Netherlands [139]. The most suitable sites to store CO2 are depleted gas fields (79%), followed by aquifers (19%) [139].
The number of wells, injection rates, duration of the project, and reservoir thickness and depth all affect drilling costs [23]. The costs increase linearly with the length of drilling up to a 3 km depth [23]. After 3 km, costs increase exponentially with depth [23]. The injectivity index assesses the ability of wells to inject fluids into a porous, permeable formation in petroleum reservoir engineering [140]. The injectivity index is the ratio of injection rate to pressure difference [140]. Because data are not available for every site, average injection rates based on the stratigraphic unit were utilized instead of site-specific injection rates (Table 3) [23].

3.3. Malaysia

Malaysia explores CCS to cut greenhouse gas emissions and environmental impact [141]. Malaysia’s largest gas field was drilled with about 20 wells and they are largely depleted [142]. The structure of the field is defined by a mild anticline, exhibiting a typical north–south orientation [142]. The average porosity of the reservoir exceeds 26%, and its permeability is favorable at an average of 1000 mD for CO2 storage [142].
Fractures may initiate during injection if pressure surpasses the caprock fracture threshold [33]. To address the challenges with injectivity, the optimal injection rate must be determined before field-scale injection [50]. CO2 injection within safe limits prevents formation damage from fine migration in heterogeneous formations around the wellbore [50]. By maintaining injection below the critical limit, considering rock permeability, wellbore formation damage can be prevented [50].

3.4. Goldeneye

Goldeneye, a depleted gas field and platform in the northern North Sea, UK, was operated by Shell (2004–2011) about 100 km northeast of Moray Firth, Scotland [143,144]. This site is ideal for CO2 storage because of factors such as young facilities, a dedicated pipeline, excellent Darcy sandstone quality, and good containment [145]. This storage site is located at 58°0′10.8″ N, 0°22′48″ W, and has a 120 m water depth [144]. The seafloor, besides pipelines and offshore platforms, is mostly flat with occasional pockmarks and abandoned wells [144]. The current infrastructure facilitates cost-effective appraisal and expansion into connected saline aquifer systems near or overlying the Goldeneye field [145].
Depleted field storage projects may stop injection below preproduction pressures to ensure containment and address uncertainties in capacity [146]. In cases such as Goldeneye, the “pressure-sink” effect acts as a risk-mitigating factor, causing water to flow back into the pressure anomaly [146]. Demonstrating long-term stability for site closure in the EU, particularly with potential pressure increases, can be challenging [146].

3.5. Cranfield

The depleted deep clastic Cranfield field in southwestern Mississippi, USA, presents potential for CCS operations [34,64,146,147]. From its discovery in 1943 until 1966, oil was produced there [34,64,146,147]. In 2008, the EOR CO2 injection project began [34]. At a depth of 3300 m, the Detailed Area Study (DAS) pilot project aims to reach the Cretaceous Lower Tuscaloosa Formation [34]. The CO2 injection pilot, which commenced on 1 December 2009, began at ~175 kg/min; the value was increased to ~300 kg/min after 20 days and ~500 kg/min after 156 days [34]. Over 31/4 years, approximately 3 million metric tons of CO2 have been injected into the Cranfield field [148]. The CO2 injection zone in the Lower Tuscaloosa Formation forms a four-way anticline with a diameter of 6.4 km (4 mi), shaped by an inactive salt dome [147].
Injected CO2 flow rates are directly proportional to its relative permeability, which is a crucial parameter [34]. Low gas relative permeabilities lead to slower CO2 plume spreading and increased pressure build-up [34]. Moreover, salt precipitation from water evaporation may impact injectivity, thereby increasing the CO2 flow and reducing the porosity [149]. Pressure build-up is yet to be quantitatively determined [149].

4. Conclusions

The assessment of injectivity is vital for the secure sequestration of CO2 in targeted depleted reservoirs. This study addressed various factors affecting injectivity and presents field case data. It was concluded that hydrophysical, chemical, and geo-mechanical processes significantly influence injectivity. The reviewed papers in the literature highlight that asphaltene precipitation, mineral dissolution, and fine mobilization can disturb injectivity by causing core plugging. Additionally, properties like purity, density, and viscosity of injected CO2 play a crucial role in ensuring effective injection and storage.
A comprehensive understanding of diverse trapping mechanisms and the adept modeling of CCS in depleted reservoirs is mandatory for achieving successful CO2 injection. Therefore, it is advisable to pursue further research in this domain.

Author Contributions

R.G.H.: Conceptualization, Writing—original draft. K.S.: Writing—review & editing, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by NRF Korea [grant numbers 2021R1F1A1047221].

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the support by NRF Korea (2021R1F1A1047221) funded by the Ministry of Science and ICT of Korea.

Conflicts of Interest

The authors declare that they have no known competing financial interests.

References

  1. Iddphonce, R.; Wang, J.; Zhao, L. Review of CO2 injection techniques for enhanced shale gas recovery: Prospect and challenges. J. Nat. Gas Sci. Eng. 2020, 77, 103240. [Google Scholar] [CrossRef]
  2. Yan, H.; Zhang, J.; Rahman, S.S.; Zhou, N.; Suo, Y. Predicting permeability changes with injecting CO2 in coal seams during CO2 geological sequestration: A comparative study among six SVM-based hybrid models. Sci. Total Environ. 2020, 705, 135941. [Google Scholar] [CrossRef]
  3. Park, Y.-C.; Kim, S.; Lee, J.H.; Shinn, Y.J. Effect of reducing irreducible water saturation in a near-well region on CO2 injectivity and storage capacity. Int. J. Greenh. Gas Control 2019, 86, 134–145. [Google Scholar] [CrossRef]
  4. Hoteit, H.; Fahs, M.; Soltanian, M.R. Assessment of CO2 injectivity during sequestration in depleted gas reservoirs. Geosciences 2019, 9, 199. [Google Scholar] [CrossRef]
  5. Sokama-Neuyam, Y.A.; Ursin, J.R. Experimental and theoretical investigations of CO2 injectivity. AGH Drill. Oil Gas 2016, 33, 245–258. [Google Scholar] [CrossRef]
  6. Abba, M.K. Enhanced gas recovery by CO2 injection: Influence of monovalent and divalent brines and their concentrations on CO2 dispersion in porous media. J. Nat. Gas Sci. Eng. 2020, 84, 103643. [Google Scholar] [CrossRef]
  7. Ginting, P. Effect of Colloidal Transport on CO2 Injectivity; University of Stavanger: Stavanger, Norway, 2016. [Google Scholar]
  8. Wang, X.; Alvarado, V.; Swoboda-Colberg, N.; Kaszuba, J.P. Reactivity of dolomite in water-saturated supercritical carbon dioxide: Significance for carbon capture and storage and for enhanced oil and gas recovery. Energy Convers. Manag. 2013, 65, 564–573. [Google Scholar] [CrossRef]
  9. Torsæter, M.; Cerasi, P. Geological and geomechanical factors impacting loss of near-well permeability during CO2 injection. Int. J. Greenh. Gas Control 2018, 76, 193–199. [Google Scholar] [CrossRef]
  10. Peysson, Y.; André, L.; Azaroual, M. Well injectivity during CO2 storage operations in deep saline aquifers—Part 1: Experimental investigation of drying effects, salt precipitation and capillary forces. Int. J. Greenh. Gas Control 2014, 22, 291–300. [Google Scholar] [CrossRef]
  11. Abba, M.K.; Al-Otaibi, A.; Abbas, A.J.; Nasr, G.G.; Burby, M. Influence of permeability and injection orientation variations on dispersion coefficient during enhanced gas recovery by CO2 injection. Energies 2019, 12, 2328. [Google Scholar] [CrossRef]
  12. Raza, A.; Gholami, R.; Rezaee, R.; Rasouli, V.; Rabiei, M. Significant aspects of carbon capture and storage—A review. Petroleum 2019, 5, 335–340. [Google Scholar] [CrossRef]
  13. Al-Hasami, A.; Ren, S.; Tohidi, B. CO2 Injection for Enhanced Gas Recovery and Geo-Storage: Reservoir Simulation and Economics. In Proceedings of the SPE Europec/EAGE Annual Conference, Madrid, Spain, 13–16 June 2005. [Google Scholar]
  14. Kasahara, J.; Tsuruga, K. An Innovative Method for the 4D Monitor of Storage in CCS(Carbon Dioxide Capture and Storage) and Oil and Gas Reservoirs and Aqufers. In Proceedings of the 2nd Joint BCSR-JCCP Environmental Symposium, Bahrain, 8–10 February 2010. [Google Scholar]
  15. Valle, L.; Rodríguez, R.; Grima, C.; Martínez, C. Effects of supercritical CO2 injection on sandstone wettability and capillary trapping. Int. J. Greenh. Gas Control 2018, 78, 341–348. [Google Scholar] [CrossRef]
  16. Min, Y.; Kim, D.; Jun, Y.-S. Effects of Na+ and K+ exchange in interlayers on biotite dissolution under high-temperature and high-CO2-pressure conditions. Environ. Sci. Technol. 2018, 52, 13638–13646. [Google Scholar] [CrossRef]
  17. De Jong, S.; Spiers, C.; Busch, A. Development of swelling strain in smectite clays through exposure to carbon dioxide. Int. J. Greenh. Gas Control 2014, 24, 149–161. [Google Scholar] [CrossRef]
  18. Tawiah, P.; Duer, J.; Bryant, S.L.; Larter, S.; O’Brien, S.; Dong, M. CO2 injectivity behaviour under non-isothermal conditions—Field observations and assessments from the Quest CCS operation. Int. J. Greenh. Gas Control 2020, 92, 102843. [Google Scholar] [CrossRef]
  19. Jung, H.; Espinoza, D.N.; Hosseini, S.A. Wellbore injectivity response to step-rate CO2 injection: Coupled thermo-poro-elastic analysis in a vertically heterogeneous formation. Int. J. Greenh. Gas Control 2020, 102, 103156. [Google Scholar] [CrossRef]
  20. Ren, B.; Ren, S.; Zhang, L.; Chen, G.; Zhang, H. Monitoring on CO2 migration in a tight oil reservoir during CCS-EOR in Jilin Oilfield China. Energy 2016, 98, 108–121. [Google Scholar] [CrossRef]
  21. Roy, P.; Morris, J.P.; Walsh, S.D.; Iyer, J.; Carroll, S. Effect of thermal stress on wellbore integrity during CO2 injection. Int. J. Greenh. Gas Control 2018, 77, 14–26. [Google Scholar] [CrossRef]
  22. Zhao, Y.; Yu, Q. CO2 breakthrough pressure and permeability for unsaturated low-permeability sandstone of the Ordos Basin. J. Hydrol. 2017, 550, 331–342. [Google Scholar] [CrossRef]
  23. Ramírez, A.; Hagedoorn, S.; Kramers, L.; Wildenborg, T.; Hendriks, C. Screening CO2 storage options in the Netherlands. Int. J. Greenh. Gas Control 2010, 4, 367–380. [Google Scholar] [CrossRef]
  24. Rohmer, J.; Pluymakers, A.; Renard, F. Mechano-chemical interactions in sedimentary rocks in the context of CO2 storage: Weak acid, weak effects? Earth Sci. Rev. 2016, 157, 86–110. [Google Scholar] [CrossRef]
  25. Mackay, E.J. 3—Modelling the Injectivity, Migration and Trapping of CO2 in Carbon Capture and Storage (CCS). In Geological Storage of Carbon Dioxide (CO2); Gluyas, J., Mathias, S., Eds.; Woodhead Publishing: Sawston, UK, 2013; pp. 45–70.e. [Google Scholar]
  26. Al-Khdheeawi, E.A.; Vialle, S.; Barifcani, A.; Sarmadivaleh, M.; Iglauer, S. Influence of injection well configuration and rock wettability on CO2 plume behaviour and CO2 trapping capacity in heterogeneous reservoirs. J. Nat. Gas Sci. Eng. 2017, 43, 190–206. [Google Scholar] [CrossRef]
  27. Raza, A.; Rezaee, R.; Gholami, R.; Rasouli, V.; Bing, C.H.; Nagarajan, R.; Hamid, M.A. Injectivity and quantification of capillary trapping for CO2 storage: A review of influencing parameters. J. Nat. Gas Sci. Eng. 2015, 26, 510–517. [Google Scholar] [CrossRef]
  28. Khurshid, I.; Choe, J. Analysis of asphaltene deposition, carbonate precipitation, and their cementation in depleted reservoirs during CO2 injection. Greenh. Gases Sci. Technol. 2015, 5, 657–667. [Google Scholar] [CrossRef]
  29. Kalra, S.; Wu, X. CO2 Injection for Enhanced Gas Recovery. In Proceedings of the SPE Western North American and Rocky Mountain Joint Meeting, Denver, CO, USA, 17–18 April 2014. [Google Scholar]
  30. Gao, S.; Wang, Y.; Jia, L.; Wang, H.; Yuan, J.; Wang, X. CO2–H2O–coal interaction of CO2 storage in coal beds. Int. J. Min. Sci. Technol. 2013, 23, 525–529. [Google Scholar] [CrossRef]
  31. Santibanez-Borda, E.; Govindan, R.; Elahi, N.; Korre, A.; Durucan, S. Maximising the dynamic CO2 storage capacity through the optimisation of CO2 injection and brine production rates. Int. J. Greenh. Gas Control 2019, 80, 76–95. [Google Scholar] [CrossRef]
  32. Borda, E.S.; Govindan, R.; Elahi, N.; Korre, A.; Durucan, S. The Development of a Dynamic CO2 Injection Strategy for the Depleted Forties and Nelson Oilfields Using Regression-based Multi-objective Programming. Energy Procedia 2017, 114, 3335–3342. [Google Scholar] [CrossRef]
  33. Raza, A.; Gholami, R.; Rezaee, R.; Bing, C.H.; Nagarajan, R.; Hamid, M.A. Well selection in depleted oil and gas fields for a safe CO2 storage practice: A case study from Malaysia. Petroleum 2017, 3, 167–177. [Google Scholar] [CrossRef]
  34. Soltanian, M.R.; Amooie, M.A.; Cole, D.R.; Graham, D.E.; Hosseini, S.A.; Hovorka, S.; Pfiffner, S.M.; Phelps, T.J.; Moortgat, J. Simulating the Cranfield geological carbon sequestration project with high-resolution static models and an accurate equation of state. Int. J. Greenh. Gas Control 2016, 54, 282–296. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Xue, Z.; Park, H.; Shi, J.Q.; Kiyama, T.; Lei, X.; Sun, Y.; Liang, Y. Tracking CO2 plumes in clay-rich rock by distributed fiber optic strain sensing (DFOSS): A laboratory demonstration. Water Resour. Res. 2019, 55, 856–867. [Google Scholar] [CrossRef]
  36. Jeon, P.R.; Lee, C.-H. Effect of surfactants on CO2 solubility and reaction in CO2-brine-clay mineral systems during CO2-enhanced fossil fuel recovery. Chem. Eng. J. 2020, 382, 123014. [Google Scholar] [CrossRef]
  37. Jin, L.; Pekot, L.J.; Smith, S.A.; Salako, O.; Peterson, K.J.; Bosshart, N.W.; Hamling, J.A.; Mibeck, B.A.; Hurley, J.P.; Beddoe, C.J. Effects of gas relative permeability hysteresis and solubility on associated CO2 storage performance. Int. J. Greenh. Gas Control 2018, 75, 140–150. [Google Scholar] [CrossRef]
  38. Valbuena, E.; Barrufet, M. A generalized partial molar volume algorithm provides fast estimates of CO2 storage capacity in depleted oil and gas reservoirs. Fluid Phase Equilibria 2013, 359, 45–53. [Google Scholar] [CrossRef]
  39. Chauhan, D.S.; Quraishi, M.; Qurashi, A. Recent trends in environmentally sustainable Sweet corrosion inhibitors. J. Mol. Liq. 2021, 326, 115117. [Google Scholar] [CrossRef]
  40. Usman, B.J.; Ali, S.A. Carbon dioxide corrosion inhibitors: A review. Arab. J. Sci. Eng. 2018, 43, 1–22. [Google Scholar] [CrossRef]
  41. Perez, T.E. Corrosion in the oil and gas industry: An increasing challenge for materials. Jom 2013, 65, 1033–1042. [Google Scholar] [CrossRef]
  42. Nejad, A.M. A review of contributing parameters in corrosion of oil and gas wells. Anti Corros. Methods Mater. 2018, 65, 73–78. [Google Scholar] [CrossRef]
  43. Florez, J.J.A.; Ferrari, J.V. Fluid flow effects on CO2 corrosion: A review of applications of rotating cage methodology. Anti-Corros. Methods Mater. 2019, 66, 507–519. [Google Scholar] [CrossRef]
  44. Yang, H.-M. Role of organic and eco-friendly inhibitors on the corrosion mitigation of steel in acidic environments—A state-of-art review. Molecules 2021, 26, 3473. [Google Scholar] [CrossRef]
  45. Liu, Q.; Mao, L.; Zhou, S. Effects of chloride content on CO2 corrosion of carbon steel in simulated oil and gas well environments. Corros. Sci. 2014, 84, 165–171. [Google Scholar] [CrossRef]
  46. Askari, M.; Aliofkhazraei, M.; Jafari, R.; Hamghalam, P.; Hajizadeh, A. Downhole corrosion inhibitors for oil and gas production—A review. Appl. Surf. Sci. Adv. 2021, 6, 100128. [Google Scholar] [CrossRef]
  47. Chamkalani, A.; Nareh’ei, M.A.; Chamkalani, R.; Zargari, M.H.; Dehestani-Ardakani, M.R.; Farzam, M. Soft computing method for prediction of CO2 corrosion in flow lines based on neural network approach. Chem. Eng. Commun. 2013, 200, 731–747. [Google Scholar] [CrossRef]
  48. Hamza, A.; Hussein, I.A.; Al-Marri, M.J.; Mahmoud, M.; Shawabkeh, R.; Aparicio, S. CO2 enhanced gas recovery and sequestration in depleted gas reservoirs: A review. J. Pet. Sci. Eng. 2021, 196, 107685. [Google Scholar] [CrossRef]
  49. Cui, G.; Yang, L.; Fang, J.; Qiu, Z.; Wang, Y.; Ren, S. Geochemical reactions and their influence on petrophysical properties of ultra-low permeability oil reservoirs during water and CO2 flooding. J. Pet. Sci. Eng. 2021, 203, 108672. [Google Scholar] [CrossRef]
  50. Mat Razali, N.Z.; Mustapha, K.A.; Kashim, M.Z.; Misnan, M.S.; Md Shah, S.S.; Abu Bakar, Z.A. Critical rate analysis for CO2 injection in depleted gas field, Sarawak Basin, offshore East Malaysia. Carbon Manag. 2022, 13, 294–309. [Google Scholar] [CrossRef]
  51. Mahmoud, M.; Hussein, I.; Carchini, G.; Shawabkeh, R.; Eliebid, M.; Al-Marri, M.J. Effect of rock mineralogy on Hot-CO2 injection for enhanced gas recovery. J. Nat. Gas Sci. Eng. 2019, 72, 103030. [Google Scholar] [CrossRef]
  52. Lei, H.; Pingping, S.; Ying, J.; Jigen, Y.; Shi, L.; Aifang, B. Prediction of asphaltene precipitation during CO2 injection. Pet. Explor. Dev. 2010, 37, 349–353. [Google Scholar] [CrossRef]
  53. Zhou, W.; Wang, H.; Yan, Y.; Liu, X. Adsorption mechanism of CO2/CH4 in kaolinite clay: Insight from molecular simulation. Energy Fuels 2019, 33, 6542–6551. [Google Scholar] [CrossRef]
  54. Li, X.; Chi, P.; Guo, X.; Sun, Q. CO2-induced asphaltene deposition and wettability alteration on a pore interior surface. Fuel 2019, 254, 115595. [Google Scholar] [CrossRef]
  55. Fakher, S.; Imqam, A. Asphaltene precipitation and deposition during CO2 injection in nano shale pore structure and its impact on oil recovery. Fuel 2019, 237, 1029–1039. [Google Scholar] [CrossRef]
  56. Zanganeh, P.; Dashti, H.; Ayatollahi, S. Visual investigation and modeling of asphaltene precipitation and deposition during CO2 miscible injection into oil reservoirs. Fuel 2015, 160, 132–139. [Google Scholar] [CrossRef]
  57. Baban, A.; Hosseini, M.; Keshavarz, A.; Ali, M.; Hoteit, H.; Amin, R.; Iglauer, S. Robust NMR Examination of the Three-Phase Flow Dynamics of Carbon Geosequestration Combined with Enhanced Oil Recovery in Carbonate Formations. Energy Fuels 2024, 38, 2167–2176. [Google Scholar] [CrossRef]
  58. Hamza, A.; Hussein, I.A.; Al-Marri, M.J.; Mahmoud, M.; Shawabkeh, R. Impact of clays on CO2 adsorption and enhanced gas recovery in sandstone reservoirs. Int. J. Greenh. Gas Control 2021, 106, 103286. [Google Scholar] [CrossRef]
  59. Barrufet, M.A.; Bacquet, A.; Falcone, G. Analysis of the storage capacity for CO2 sequestration of a depleted gas condensate reservoir and a saline aquifer. J. Can. Pet. Technol. 2010, 49, 23–31. [Google Scholar] [CrossRef]
  60. Sokama-Neuyam, Y.A.; Ginting, P.U.R.; Timilsina, B.; Ursin, J.R. The impact of fines mobilization on CO2 injectivity: An experimental study. Int. J. Greenh. Gas Control 2017, 65, 195–202. [Google Scholar] [CrossRef]
  61. Jeddizahed, J.; Rostami, B. Experimental investigation of injectivity alteration due to salt precipitation during CO2 sequestration in saline aquifers. Adv. Water Resour. 2016, 96, 23–33. [Google Scholar] [CrossRef]
  62. Li, Z.; Dong, M.; Li, S.; Huang, S. CO2 sequestration in depleted oil and gas reservoirs—Caprock characterization and storage capacity. Energy Convers. Manag. 2006, 47, 1372–1382. [Google Scholar] [CrossRef]
  63. Tang, Y.; Yang, R.; Kang, X. Modeling the effect of water vaporization and salt precipitation on reservoir properties due to carbon dioxide sequestration in a depleted gas reservoir. Petroleum 2018, 4, 385–397. [Google Scholar] [CrossRef]
  64. Hosseini, S.A.; Lashgari, H.; Choi, J.W.; Nicot, J.-P.; Lu, J.; Hovorka, S.D. Static and dynamic reservoir modeling for geological CO2 sequestration at Cranfield, Mississippi, USA. Int. J. Greenh. Gas Control 2013, 18, 449–462. [Google Scholar] [CrossRef]
  65. Goudarzi, A.; Meckel, T.A.; Hosseini, S.A.; Treviño, R.H. Statistical analysis of historic hydrocarbon production data from Gulf of Mexico oil and gas fields and application to dynamic capacity assessment in CO2 storage. Int. J. Greenh. Gas Control 2019, 80, 96–102. [Google Scholar] [CrossRef]
  66. Ambrose, W.; Lakshminarasimhan, S.; Holtz, M.; Núñez-López, V.; Hovorka, S.D.; Duncan, I. Geologic factors controlling CO2 storage capacity and permanence: Case studies based on experience with heterogeneity in oil and gas reservoirs applied to CO2 storage. Environ. Geol. 2008, 54, 1619. [Google Scholar] [CrossRef]
  67. Raza, A.; Rezaee, R.; Gholami, R.; Bing, C.H.; Nagarajan, R.; Hamid, M.A. A screening criterion for selection of suitable CO2 storage sites. J. Nat. Gas Sci. Eng. 2016, 28, 317–327. [Google Scholar] [CrossRef]
  68. Shirbazo, A.; Taghavinejad, A.; Bagheri, S. CO2 Capture and Storage Performance Simulation in Depleted Shale Gas Reservoirs as Sustainable Carbon Resources. J. Constr. Mater. 2021. [Google Scholar] [CrossRef]
  69. Gan, M.; Nguyen, M.C.; Zhang, L.; Wei, N.; Li, J.; Lei, H.; Wang, Y.; Li, X.; Stauffer, P.H. Impact of reservoir parameters and wellbore permeability uncertainties on CO2 and brine leakage potential at the Shenhua CO2 Storage Site, China. Int. J. Greenh. Gas Control 2021, 111, 103443. [Google Scholar] [CrossRef]
  70. Akono, A.T.; Druhan, J.L.; Dávila, G.; Tsotsis, T.; Jessen, K.; Fuchs, S.; Crandall, D.; Shi, Z.; Dalton, L.; Tkach, M.K. A review of geochemical-mechanical impacts in geological carbon storage reservoirs. Greenh. Gases Sci. Technol. 2019, 9, 474–504. [Google Scholar] [CrossRef]
  71. Yu, W.; Lashgari, H.R.; Wu, K.; Sepehrnoori, K. CO2 injection for enhanced oil recovery in Bakken tight oil reservoirs. Fuel 2015, 159, 354–363. [Google Scholar] [CrossRef]
  72. Kelley, M.; Abbaszadeh, M.; Mishra, S.; Mawalkar, S.; Place, M.; Gupta, N.; Pardini, R. Reservoir characterization from pressure monitoring during CO2 injection into a depleted pinnacle reef-MRCSP commercial-scale CCS demonstration project. Energy Procedia 2014, 63, 4937–4964. [Google Scholar] [CrossRef]
  73. Islam, M.R.; Chakma, A. Storage and utilization of CO2 in petroleum reservoirs—A simulation study. Energy Convers. Manag. 1993, 34, 1205–1212. [Google Scholar] [CrossRef]
  74. Streit, J.E.; Hillis, R.R. Building Geomechanical Models for the Safe Underground Storage of Carbon Dioxide in Porous Rock. In Proceedings of the Greenhouse Gas Control Technologies—6th International Conference, Kyoto, Japan, 1–4 October 2002; pp. 495–500. [Google Scholar]
  75. Recasens, M.; Garcia, S.; Mackay, E.; Delgado, J.; Maroto-Valer, M.M. Experimental study of wellbore integrity for CO2 geological storage. Energy Procedia 2017, 114, 5249–5255. [Google Scholar] [CrossRef]
  76. Kaldi, J.; Daniel, R.; Tenthorey, E.; Michael, K.; Schacht, U.; Nicol, A.; Underschultz, J.; Backe, G. Containment of CO2 in CCS: Role of Caprocks and Faults. Energy Procedia 2013, 37, 5403–5410. [Google Scholar] [CrossRef]
  77. Raza, A.; Gholami, R.; Rezaee, R.; Rasouli, V.; Bhatti, A.A.; Bing, C.H. Suitability of depleted gas reservoirs for geological CO2 storage: A simulation study. Greenh. Gases Sci. Technol. 2018, 8, 876–897. [Google Scholar] [CrossRef]
  78. Karine, S.; Gérald, H.; Joël, B.; Vincent, T. SCA2003-14: Residual Gas Saturation of Sample Originally at Residual Water Saturation in Heterogeneous Sandstone Reservoirs. Available online: https://www.jgmaas.com/SCA/2003/SCA2003-14.pdf (accessed on 21 September 2003).
  79. Raza, A.; Gholami, R.; Rezaee, R.; Bing, C.H.; Nagarajan, R.; Hamid, M.A. CO2 storage in depleted gas reservoirs: A study on the effect of residual gas saturation. Petroleum 2018, 4, 95–107. [Google Scholar] [CrossRef]
  80. Saeedi, A.; Rezaee, R. Effect of residual natural gas saturation on multiphase flow behaviour during CO2 geo-sequestration in depleted natural gas reservoirs. J. Pet. Sci. Eng. 2012, 82, 17–26. [Google Scholar] [CrossRef]
  81. Cao, M.; Gu, Y. Oil recovery mechanisms and asphaltene precipitation phenomenon in immiscible and miscible CO2 flooding processes. Fuel 2013, 109, 157–166. [Google Scholar] [CrossRef]
  82. Tan, Y.; Nookuea, W.; Li, H.; Thorin, E.; Yan, J. Property impacts on Carbon Capture and Storage (CCS) processes: A review. Energy Convers. Manag. 2016, 118, 204–222. [Google Scholar] [CrossRef]
  83. Nicot, J.-P.; Solano, S.; Lu, J.; Mickler, P.; Romanak, K.; Yang, C.; Zhang, X. Potential Subsurface Impacts of CO2 Stream Impurities on Geologic Carbon Storage. Energy Procedia 2013, 37, 4552–4559. [Google Scholar] [CrossRef]
  84. Rother, G.; Ilton, E.S.; Wallacher, D.; Hauβ, T.; Schaef, H.T.; Qafoku, O.; Rosso, K.M.; Felmy, A.R.; Krukowski, E.G.; Stack, A.G. CO2 sorption to subsingle hydration layer montmorillonite clay studied by excess sorption and neutron diffraction measurements. Environ. Sci. Technol. 2013, 47, 205–211. [Google Scholar] [CrossRef] [PubMed]
  85. Oldenburg, C.M.; Benson, S.M. CO2 Injection for Enhanced Gas Production and Carbon Sequestration. In Proceedings of the SPE International Petroleum Conference and Exhibition in Mexico, Villahermosa, Mexico, 10–12 February 2002. [Google Scholar]
  86. Vu, H.P.; Black, J.R.; Haese, R.R. The geochemical effects of O2 and SO2 as CO2 impurities on fluid-rock reactions in a CO2 storage reservoir. Int. J. Greenh. Gas Control 2018, 68, 86–98. [Google Scholar] [CrossRef]
  87. Wang, J.; Ryan, D.; Anthony, E.J.; Wildgust, N.; Aiken, T. Effects of impurities on CO2 transport, injection and storage. Energy Procedia 2011, 4, 3071–3078. [Google Scholar] [CrossRef]
  88. Ziabakhsh-Ganji, Z.; Kooi, H. Sensitivity of Joule-Thomson cooling to impure CO2 injection in depleted gas reservoirs. Appl. Energy 2014, 113, 434–451. [Google Scholar] [CrossRef]
  89. Machado, M.V.B.; Delshad, M.; Sepehrnoori, K. Injectivity assessment for CCS field-scale projects with considerations of salt deposition, mineral dissolution, fines migration, hydrate formation, and non-Darcy flow. Fuel 2023, 353, 129148. [Google Scholar] [CrossRef]
  90. Holm, L.; Josendal, V. Mechanisms of oil displacement by carbon dioxide. J. Pet. Technol. 1974, 26, 1427–1438. [Google Scholar] [CrossRef]
  91. Kazemzadeh, Y.; Parsaei, R.; Riazi, M. Experimental study of asphaltene precipitation prediction during gas injection to oil reservoirs by interfacial tension measurement. Colloids Surf. A Physicochem. Eng. Asp. 2015, 466, 138–146. [Google Scholar] [CrossRef]
  92. Gaus, I.; Audigane, P.; André, L.; Lions, J.; Jacquemet, N.; Durst, P.; Czernichowski-Lauriol, I.; Azaroual, M. Geochemical and solute transport modelling for CO2 storage, what to expect from it? Int. J. Greenh. Gas Control 2008, 2, 605–625. [Google Scholar] [CrossRef]
  93. Bacci, G.; Korre, A.; Durucan, S. An experimental and numerical investigation into the impact of dissolution/precipitation mechanisms on CO2 injectivity in the wellbore and far field regions. Int. J. Greenh. Gas Control 2011, 5, 579–588. [Google Scholar] [CrossRef]
  94. Pudlo, D.; Henkel, S.; Reitenbach, V.; Albrecht, D.; Enzmann, F.; Heister, K.; Pronk, G.; Ganzer, L.; Gaupp, R. The chemical dissolution and physical migration of minerals induced during CO2 laboratory experiments: Their relevance for reservoir quality. Environ. Earth Sci. 2015, 73, 7029–7042. [Google Scholar] [CrossRef]
  95. Belgodere, C.; Sterpenich, J.; Pironon, J.; Randi, A.; Birat, J.-P. Experimental Study of CO2 Injection in the Triassic Sandstones of Lorraine (Eastern France)–Investigation of Injection Well Injectivity Impairment by Mineral Precipitations. In Proceedings of the Le Studium Conference. Geochemical Reactivity in CO2 Geological Storage Sites, Orleans, France, 25–26 February 2014; pp. 25–26. [Google Scholar]
  96. Pudlo, D.; Henkel, S.; Enzmann, F.; Heister, K.; Werner, L.; Ganzer, L.; Reitenbach, V.; Albrecht, D.; Gaupp, R. The Relevance of Mineral Mobilization and Dissolution on the Reservoir Quality of Sandstones in CO2 Storage Sites. Energy Procedia 2014, 59, 390–396. [Google Scholar] [CrossRef]
  97. Wellman, T.P.; Grigg, R.B.; McPherson, B.J.; Svec, R.K.; Lichtner, P.C. Evaluation of CO2-Brine-Reservoir Rock Interaction with Laboratory Flow Tests and Reactive Transport Modeling. In Proceedings of the International Symposium on Oilfield Chemistry, Houston, TX, USA, 5–7 February 2003. [Google Scholar]
  98. Al-Yaseri, A.; Zhang, Y.; Ghasemiziarani, M.; Sarmadivaleh, M.; Lebedev, M.; Roshan, H.; Iglauer, S. Permeability Evolution in Sandstone Due to CO2 Injection. Energy Fuels 2017, 31, 12390–12398. [Google Scholar] [CrossRef]
  99. Sokama-Neuyam, Y.A.; Ursin, J.R. CO2 Well Injectivity: Effect of Viscous Forces on Precipitated Minerals. In Proceedings of the International Petroleum Technology Conference, Doha, Qatar, 6–9 December 2015. [Google Scholar]
  100. Miri, R.; Hellevang, H. Salt precipitation during CO2 storage—A review. Int. J. Greenh. Gas Control 2016, 51, 136–147. [Google Scholar] [CrossRef]
  101. Giorgis, T.; Carpita, M.; Battistelli, A. 2D modeling of salt precipitation during the injection of dry CO2 in a depleted gas reservoir. Energy Convers. Manag. 2007, 48, 1816–1826. [Google Scholar] [CrossRef]
  102. Yusof, M.A.M.; Mohamed, M.A.; Akhir, N.A.M.; Ibrahim, M.A.; Mardhatillah, M.K. Combined Impact of Salt Precipitation and Fines Migration on CO2 Injectivity Impairment. Int. J. Greenh. Gas Control 2021, 110, 103422. [Google Scholar] [CrossRef]
  103. Bacci, G.; Korre, A.; Durucan, S. Experimental investigation into salt precipitation during CO2 injection in saline aquifers. Energy Procedia 2011, 4, 4450–4456. [Google Scholar] [CrossRef]
  104. Ashoori, S.; Balavi, A. An investigation of asphaltene precipitation during natural production and the CO2 injection process. Pet. Sci. Technol. 2014, 32, 1283–1290. [Google Scholar] [CrossRef]
  105. Hemmati-Sarapardeh, A.; Ahmadi, M.; Ameli, F.; Dabir, B.; Mohammadi, A.H.; Husein, M.M. Modeling asphaltene precipitation during natural depletion of reservoirs and evaluating screening criteria for stability of crude oils. J. Pet. Sci. Eng. 2019, 181, 106127. [Google Scholar] [CrossRef]
  106. Cruz, A.A.; Amaral, M.; Santos, D.; Palma, A.; Franceschi, E.; Borges, G.R.; Coutinho, J.A.P.; Palácio, J.; Dariva, C. CO2 influence on asphaltene precipitation. J. Supercrit. Fluids 2019, 143, 24–31. [Google Scholar] [CrossRef]
  107. Kalantari-Dahaghi, A.; Gholami, V.; Moghadasi, J.; Abdi, R. Formation damage through asphaltene precipitation resulting from CO2 gas injection in Iranian carbonate reservoirs. SPE Prod. Oper. 2008, 23, 210–214. [Google Scholar] [CrossRef]
  108. Shedid, S.A. Influences of asphaltene precipitation on capillary pressure and pore size distribution of carbonate reservoirs. Pet. Sci. Technol. 2001, 19, 503–519. [Google Scholar] [CrossRef]
  109. Nasrabadi, H.; Moortgat, J.; Firoozabadi, A. New three-phase multicomponent compositional model for asphaltene precipitation during CO2 injection using CPA-EOS. Energy Fuels 2016, 30, 3306–3319. [Google Scholar] [CrossRef]
  110. Mohammed, S.; Gadikota, G. The influence of CO2 on the structure of confined asphaltenes in calcite nanopores. Fuel 2019, 236, 769–777. [Google Scholar] [CrossRef]
  111. Papadimitriou, N.; Romanos, G.; Charalambopoulou, G.C.; Kainourgiakis, M.; Katsaros, F.; Stubos, A. Experimental investigation of asphaltene deposition mechanism during oil flow in core samples. J. Pet. Sci. Eng. 2007, 57, 281–293. [Google Scholar] [CrossRef]
  112. Wang, C.; Li, T.; Gao, H.; Zhao, J.; Gao, Y. Quantitative study on the blockage degree of pores due to asphaltene precipitation in low-permeability reservoirs with NMR technique. J. Pet. Sci. Eng. 2018, 163, 703–711. [Google Scholar] [CrossRef]
  113. Shedid, S.A.; Zekri, A.Y. Formation damage caused by simultaneous sulfur and asphaltene deposition. SPE Prod. Oper. 2006, 21, 58–64. [Google Scholar] [CrossRef]
  114. Saeedi Dehaghani, A.; Shadman, M.; Ahmadi, S.; Assaf, M. Modeling the onset point of asphaltene precipitation using the solubility parameter in CO2 injection. Energy Sources Part A Recovery Util. Environ. Eff. 2017, 45, 1–7. [Google Scholar]
  115. Srivastava, R.; Huang, S.; Dong, M. Asphaltene deposition during CO2 flooding. SPE Prod. Facil. 1999, 14, 235–245. [Google Scholar] [CrossRef]
  116. Nielsen, B.B.; Svrcek, W.Y.; Mehrotra, A.K. Effects of Temperature and Pressure on Asphaltene Particle Size Distributions in Crude Oils Diluted with n-Pentane. Ind. Eng. Chem. Res. 1994, 33, 1324–1330. [Google Scholar] [CrossRef]
  117. Jafari Behbahani, T.; Ghotbi, C.; Taghikhani, V.; Shahrabadi, A. Investigation on Asphaltene Deposition Mechanisms during CO2 Flooding Processes in Porous Media: A Novel Experimental Study and a Modified Model Based on Multilayer Theory for Asphaltene Adsorption. Energy Fuels 2012, 26, 5080–5091. [Google Scholar] [CrossRef]
  118. Rosenbrand, E.; Kjøller, C.; Riis, J.F.; Kets, F.; Fabricius, I.L. Different effects of temperature and salinity on permeability reduction by fines migration in Berea sandstone. Geothermics 2015, 53, 225–235. [Google Scholar] [CrossRef]
  119. Cihan, A.; Petrusak, R.; Bhuvankar, P.; Alumbaugh, D.; Trautz, R.; Birkholzer, J.T. Permeability Decline by Clay Fines Migration around a Low-Salinity Fluid Injection Well. Groundwater 2022, 60, 87–98. [Google Scholar] [CrossRef] [PubMed]
  120. Sbai, M.A.; Azaroual, M. Numerical modeling of formation damage by two-phase particulate transport processes during CO2 injection in deep heterogeneous porous media. Adv. Water Resour. 2011, 34, 62–82. [Google Scholar] [CrossRef]
  121. Xie, Q.; Saeedi, A.; Delle Piane, C.; Esteban, L.; Brady, P.V. Fines migration during CO2 injection: Experimental results interpreted using surface forces. Int. J. Greenh. Gas Control 2017, 65, 32–39. [Google Scholar] [CrossRef]
  122. Takahashi, S.; Kovscek, A.R. Wettability estimation of low-permeability, siliceous shale using surface forces. J. Pet. Sci. Eng. 2010, 75, 33–43. [Google Scholar] [CrossRef]
  123. Wang, Y.; Bedrikovetsky, P.; Yin, H.; Othman, F.; Zeinijahromi, A.; Le-Hussain, F. Analytical model for fines migration due to mineral dissolution during CO2 injection. J. Nat. Gas Sci. Eng. 2022, 100, 104472. [Google Scholar] [CrossRef]
  124. Zhang, M.; de Jong, S.; Spiers, C.J.; Busch, A.; Wentinck, H.M. Swelling stress development in confined smectite clays through exposure to CO2. Int. J. Greenh. Gas Control 2018, 74, 49–61. [Google Scholar] [CrossRef]
  125. Giesting, P.; Guggenheim, S.; Van Groos, A.F.K.; Busch, A. Interaction of carbon dioxide with Na-exchanged montmorillonite at pressures to 640 bars: Implications for CO2 sequestration. Int. J. Greenh. Gas Control 2012, 8, 73–81. [Google Scholar] [CrossRef]
  126. Narayanan Nair, A.K.; Cui, R.; Sun, S. Overview of the Adsorption and Transport Properties of Water, Ions, Carbon Dioxide, and Methane in Swelling Clays. ACS Earth Space Chem. 2021, 5, 2599–2611. [Google Scholar] [CrossRef]
  127. Pham, H.; Nguyen, Q.P. Effect of silica nanoparticles on clay swelling and aqueous stability of nanoparticle dispersions. J. Nanoparticle Res. 2014, 16, 1–11. [Google Scholar] [CrossRef]
  128. Bibi, I.; Icenhower, J.; Niazi, N.K.; Naz, T.; Shahid, M.; Bashir, S. Clay minerals: Structure, chemistry, and significance in contaminated environments and geological CO2 sequestration. Environ. Mater. Waste 2016, 543–567. [Google Scholar] [CrossRef]
  129. Anderson, R.; Ratcliffe, I.; Greenwell, H.; Williams, P.; Cliffe, S.; Coveney, P. Clay swelling—A challenge in the oilfield. Earth-Sci. Rev. 2010, 98, 201–216. [Google Scholar] [CrossRef]
  130. Loganathan, N.; Bowers, G.M.; Yazaydin, A.O.; Schaef, H.T.; Loring, J.S.; Kalinichev, A.G.; Kirkpatrick, R.J. Clay swelling in dry supercritical carbon dioxide: Effects of interlayer cations on the structure, dynamics, and energetics of CO2 intercalation probed by XRD, NMR, and GCMD simulations. J. Phys. Chem. C 2018, 122, 4391–4402. [Google Scholar] [CrossRef]
  131. Fatah, A.; Mahmud, H.B.; Bennour, Z.; Hossain, M.; Gholami, R. Effect of supercritical CO2 treatment on physical properties and functional groups of shales. Fuel 2021, 303, 121310. [Google Scholar] [CrossRef]
  132. Favero, V.; Laloui, L. Impact of CO2 injection on the hydro-mechanical behaviour of a clay-rich caprock. Int. J. Greenh. Gas Control 2018, 71, 133–141. [Google Scholar] [CrossRef]
  133. Pang, J.; Liang, Y.; Masuda, Y.; Matsuoka, T.; Zhang, Y.; Xue, Z. Swelling phenomena of the nonswelling Clay induced by CO2 and Water cooperative adsorption in janus-surface micropores. Environ. Sci. Technol. 2020, 54, 5767–5773. [Google Scholar] [CrossRef] [PubMed]
  134. Mukherjee, M.; Misra, S. A review of experimental research on Enhanced Coal Bed Methane (ECBM) recovery via CO2 sequestration. Earth Sci. Rev. 2018, 179, 392–410. [Google Scholar] [CrossRef]
  135. Miller, J.; Sullivan, C.; Larsen, G.; Kelley, M.; Rike, W.; Gerst, J.; Gupta, N.; Paul, D.; Pardini, R.; Modroo, A. Alternative conceptual geologic models for CO2 injection in a Niagaran pinnacle reef oil field, Northern Michigan, USA. Energy Procedia 2014, 63, 3685–3701. [Google Scholar] [CrossRef]
  136. Ganesh, P.R.; Mishra, S.; Mawalkar, S.; Gupta, N.; Pardini, R. Assessment of CO2 injectivity and storage capacity in a depleted pinnacle reef oil field in northern Michigan. Energy Procedia 2014, 63, 2969–2976. [Google Scholar] [CrossRef]
  137. Barnes, D.; Harrison, B.; Grammer, G.M.; Asmus, J. CO2/EOR and geological carbon storage resource potential in the Niagaran Pinnacle Reef Trend, lower Michigan, USA. Energy Procedia 2013, 37, 6786–6799. [Google Scholar] [CrossRef]
  138. Mishra, S.; Haagsma, A.; Valluri, M.; Gupta, N. Assessment of CO2-enhanced oil recovery and associated geologic storage potential in the Michigan Northern Pinnacle Reef Trend. Greenh. Gases Sci. Technol. 2020, 10, 32–49. [Google Scholar] [CrossRef]
  139. Orlic, B. Geomechanical effects of CO2 storage in depleted gas reservoirs in the Netherlands: Inferences from feasibility studies and comparison with aquifer storage. J. Rock Mech. Geotech. Eng. 2016, 8, 846–859. [Google Scholar] [CrossRef]
  140. Gupta, N.; Kelley, M.; Place, M.; Conner, A.; Mawalkar, S.; Mishra, S.; Sminchak, J. Integrated Monitoring Volume: A Summary of Monitoring Studies Conducted in Niagaran Carbonate Pinnacle Reefs During Enhanced Oil Recovery with CO2; Battelle Memorial Inst.: Columbus, OH, USA, 2020. [Google Scholar]
  141. Sazali, Y.; Sazali, W.; Ibrahim, J.; Dindi, M.; Graham, G.; Gödeke, S. Investigation of high temperature, high pressure, scaling and dissolution effects for carbon capture and storage at a high CO2 content carbonate gas field offshore Malaysia. J. Pet. Sci. Eng. 2019, 174, 599–606. [Google Scholar] [CrossRef]
  142. Brownsort, P.; Scott, V.; Sim, G.; Haszeldine, S. Carbon Dioxide Transport Plans for Carbon Capture and Storage in the North Sea Region. Available online: https://era.ed.ac.uk/bitstream/handle/1842/16481/wp-2015-02.pdf?sequence=1&isAllowed=y (accessed on 20 July 2015).
  143. Dale, A.W.; Sommer, S.; Lichtschlag, A.; Koopmans, D.; Haeckel, M.; Kossel, E.; Deusner, C.; Linke, P.; Scholten, J.; Wallmann, K. Defining a biogeochemical baseline for sediments at Carbon Capture and Storage (CCS) sites: An example from the North Sea (Goldeneye). Int. J. Greenh. Gas Control 2021, 106, 103265. [Google Scholar] [CrossRef]
  144. Esposito, M.; Martinez-Cabanas, M.; Connelly, D.P.; Jasinski, D.; Linke, P.; Schmidt, M.; Achterberg, E.P. Water column baseline assessment for offshore Carbon Dioxide Capture and Storage (CCS) sites: Analysis of field data from the Goldeneye storage complex area. Int. J. Greenh. Gas Control 2021, 109, 103344. [Google Scholar] [CrossRef]
  145. Tucker, O.; Holley, M.; Metcalfe, R.; Hurst, S. Containment risk management for CO2 storage in a depleted gas field, UK North Sea. Energy Procedia 2013, 37, 4804–4817. [Google Scholar] [CrossRef]
  146. Hannis, S.; Lu, J.; Chadwick, A.; Hovorka, S.; Kirk, K.; Romanak, K.; Pearce, J. CO2 storage in depleted or depleting oil and gas fields: What can we learn from existing projects? Energy Procedia 2017, 114, 5680–5690. [Google Scholar] [CrossRef]
  147. Núñez-López, V.; Gil-Egui, R.; Hosseini, S.A. Environmental and operational performance of CO2-EOR as a CCUS technology: A Cranfield example with dynamic LCA considerations. Energies 2019, 12, 448. [Google Scholar] [CrossRef]
  148. Choi, J.-W.; Nicot, J.-P.; Hosseini, S.A.; Clift, S.J.; Hovorka, S.D. CO2 recycling accounting and EOR operation scheduling to assist in storage capacity assessment at a US gulf coast depleted reservoir. Int. J. Greenh. Gas Control 2013, 18, 474–484. [Google Scholar] [CrossRef]
  149. Kim, S.; Hosseini, S.A. Above-zone pressure monitoring and geomechanical analyses for a field-scale CO2 injection project in Cranfield, MS. Greenh. Gases Sci. Technol. 2014, 4, 81–98. [Google Scholar] [CrossRef]
Figure 1. Comparison between the impact of injectivity reduction in hydrates with some other effects on injectivity [Reproduced with permission [89] Copyright 2023 Elsevier].
Figure 1. Comparison between the impact of injectivity reduction in hydrates with some other effects on injectivity [Reproduced with permission [89] Copyright 2023 Elsevier].
Energies 17 01201 g001
Figure 2. Northern pinnacle reef trend modeled; spatial extent is highlighted in green [Reproduced [138] Copyright 2020 Elsevier].
Figure 2. Northern pinnacle reef trend modeled; spatial extent is highlighted in green [Reproduced [138] Copyright 2020 Elsevier].
Energies 17 01201 g002
Table 1. Evaluating geologically suitable storage reservoirs [Reproduced with permission [27] Copyright 2015 Elsevier].
Table 1. Evaluating geologically suitable storage reservoirs [Reproduced with permission [27] Copyright 2015 Elsevier].
Storage OptionRelative CapacityRelative CostStorage IntegrityTechnical Feasibility
Active Oil Well (EOR)SmallVery LowGoodHigh
Coal BedsUnknownLowUnknownUnknown
Depleted oil/gas wellsModerateLowGoodHigh
Deep AquifersLarge UnknownUnknownUnknown
Mined caverns/salt domesLarge Very HighGoodHigh
Table 3. The average CO2 injection rate of hydrocarbon fields per stratigraphic unit in Mt/y per well [Reproduced with permission [23] Copyright 2010 Elsevier].
Table 3. The average CO2 injection rate of hydrocarbon fields per stratigraphic unit in Mt/y per well [Reproduced with permission [23] Copyright 2010 Elsevier].
FormationHydrocarbon Field
Lower Cretaceous GroupVlieland Sandstone Fm1
Lower Germanic Trias GroupLower Buntsandstein Fm0.4
Niedersachsen GroupFriese front Fm (sandstone members)0.4
Upper Rotliegend GroupZechstein Fm (carbonate members)
Slochteren Fm (sandstone members)
0.2
1
Limburg Group 0.4
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

Heidarabad, R.G.; Shin, K. Carbon Capture and Storage in Depleted Oil and Gas Reservoirs: The Viewpoint of Wellbore Injectivity. Energies 2024, 17, 1201. https://doi.org/10.3390/en17051201

AMA Style

Heidarabad RG, Shin K. Carbon Capture and Storage in Depleted Oil and Gas Reservoirs: The Viewpoint of Wellbore Injectivity. Energies. 2024; 17(5):1201. https://doi.org/10.3390/en17051201

Chicago/Turabian Style

Heidarabad, Reyhaneh Ghorbani, and Kyuchul Shin. 2024. "Carbon Capture and Storage in Depleted Oil and Gas Reservoirs: The Viewpoint of Wellbore Injectivity" Energies 17, no. 5: 1201. https://doi.org/10.3390/en17051201

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