Next Article in Journal
Analysis of a High-Voltage Room Quasi-Smoke Gas Explosion
Next Article in Special Issue
Analysis of the Influencing Factors of Regional Carbon Emissions in the Chinese Transportation Industry
Previous Article in Journal
Time-Sharing Control Strategy for Multiple-Receiver Wireless Power Transfer Systems
Previous Article in Special Issue
The Relationship between Carbon Dioxide Emissions, Economic Growth and Agricultural Production in Pakistan: An Autoregressive Distributed Lag Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of CO2 Storage in View of Safety and Cost-Effectiveness

1
Research Center of Energy Storage Technologies, Clausthal University of Technology, 38640 Goslar, Germany
2
Institute of Petroleum Engineering, Clausthal University of Technology, 38678 Clausthal-Zellerfeld, Germany
3
State Key Laboratory of Geomechanics and Geotechnical Engineering, Institute of Rock and Soil Mechanics, Chinese Academy of Sciences, Wuhan 430071, China
4
Department of Petroleum & Gas Engineering, University of Engineering & Technology Lahore, Lahore 54890, Pakistan
*
Authors to whom correspondence should be addressed.
Submission received: 21 November 2019 / Revised: 15 January 2020 / Accepted: 19 January 2020 / Published: 29 January 2020

Abstract

:
The emissions of greenhouse gases, especially CO2, have been identified as the main contributor for global warming and climate change. Carbon capture and storage (CCS) is considered to be the most promising strategy to mitigate the anthropogenic CO2 emissions. This review aims to provide the latest developments of CO2 storage from the perspective of improving safety and economics. The mechanisms and strategies of CO2 storage, focusing on their characteristics and current status, are discussed firstly. In the second section, the strategies for assessing and ensuring the security of CO2 storage operations, including the risks assessment approach and monitoring technology associated with CO2 storage, are outlined. In addition, the engineering methods to accelerate CO2 dissolution and mineral carbonation for fixing the mobile CO2 are also compared within the second section. The third part focuses on the strategies for improving economics of CO2 storage operations, namely enhanced industrial production with CO2 storage to generate additional profit, and co-injection of CO2 with impurities to reduce the cost. Moreover, the role of multiple CCS technologies and their distribution on the mitigation of CO2 emissions in the future are summarized. This review demonstrates that CO2 storage in depleted oil and gas reservoirs could play an important role in reducing CO2 emission in the near future and CO2 storage in saline aquifers may make the biggest contribution due to its huge storage capacity. Comparing the various available strategies, CO2-enhanced oil recovery (CO2-EOR) operations are supposed to play the most important role for CO2 mitigation in the next few years, followed by CO2-enhanced gas recovery (CO2-EGR). The direct mineralization of flue gas by coal fly ash and the pH swing mineralization would be the most promising technology for the mineral sequestration of CO2. Furthermore, by accelerating the deployment of CCS projects on large scale, the government can also play its role in reducing the CO2 emissions.

1. Introduction

In the last few centuries, the CO2 concentration in the atmosphere has already risen above 410 ppm from a level of below 300 ppm in pre-industrial times [1,2]. As shown in Figure 1, the continuous rise in the Earth’s surface temperature appears to be strongly linked with atmospheric concentration of CO2, which suggests that CO2 may be the main contributor to global warming and climate change. In addition, it makes up an estimated 77% of greenhouse gases [3,4].
In addition, the CO2 emission may increase the frequency of extreme extratropical cyclones. Unless the greenhouse gas emissions are efficiently mitigated, the extratropical cyclones are projected to more than triple in number by the end of this century in both Europe and North America [5]. In response to such intense global climate change, the Intergovernmental Panel on Climate Change’s (IPCC) assessment suggested that the average global warming should be limited less than 2 °C within this century on the basis of estimated results from integrated assessment models (IAMs) [6].
To achieve this goal, carbon capture and storage (CCS) needs to be promoted, as it is currently the most effective technology for slowing down atmospheric CO2 enrichment and extenuating associated potential climate problems [7]. According to the estimation by the IEA [8], CCS alone could undertake almost a 19% reduction in global CO2 emissions by 2050 (Figure 2). Furthermore, the overall cost of achieving the same emission reduction targets will increase by 70% without CCS [9], which highlights the importance of CCS on the mitigation of CO2 emission from the economic point of view as well.
There are 51 CCS engineering projects across the world, mainly scheduled in North America, Australia, China, and Western Europe, although only 19 CCS projects are currently in operation (Figure 3) [10]. Although CCS has been proven to be technically feasible, the risks and economics associated with CCS challenges its large-scale application. Consequently, the contribution of CCS is still very limited in attenuating climate change because of the high cost and safety risk associated with CO2 leakage [11,12]. More efforts on improving the safety and economics are required to develop the CCS technology, gaining support from government, and improving public acceptance to accelerate the application of large-scale CCS.
In the last decade, nearly every aspect of CCS has been discussed extensively [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44], see Table 1. However, the strategies for improving the safety and economics have not been reviewed in detail. In addition, the CCS technology is developing rapidly, and the recent progress needs to be reviewed and analyzed.
This review attempts to discuss the most recent developments on addressing challenges associated with assessing and decreasing the risks of leakage, cutting the cost of CO2 storage and promoting the developments of commercial-scale CCS projects. Firstly, different mechanisms and strategies of CO2 storage are summarized and discussed. Then, the risk assessment of CO2 storage and strategies for accelerating CO2 dissolution and mineral carbonation are reviewed. Finally, the strategies for improving cost-effectiveness, including enhanced industrial production with CO2 storage and co-injection of CO2 with impurities, are examined.

2. Mechanisms and Strategies of CO2 Storage

2.1. Mechanisms of CO2 Storage

As shown in Figure 4a, there are four main trapping mechanisms of CO2 storage involving (1) structural and stratigraphic, (2) residual, (3) solubility, and (4) mineral trapping [42]. With structural and stratigraphic trapping, which is the most dominant trapping mechanism, once CO2 is injected subsurface, it will rise to the top of geological structures due to the buoyancy effect but stay below the impermeable caprock. In residual trapping, the injected CO2 displaces formation fluid when it flows through the formation rock. The displaced fluid disconnects and traps the remaining CO2 within the pores of rocks due to the capillary force [45]. In the residual trapping mechanism, the saturation of trapped CO2 is at least 10% and can reach more than 30% of the pore volume in some reservoir rocks [46,47]. In solubility trapping, CO2 dissolves in formation fluids and becomes immobile, thereby decreasing the volume of free CO2 [48]. This dissolved CO2 will slightly increase the density of formation fluid by around 1%. It is sufficient to promote the convection flow with such a small density difference [49], which is also in favor of the trapping of CO2. The solubility of CO2 in groundwater ranges from 2% to 6%, and it decreases with the rising temperature and salinity [47]. For the mineral trapping mechanism, CO2 is trapped by geochemical reactions in reservoir, usually precipitating as carbonate, which can trap the CO2 in immobile secondary phases effectively [50].
As shown in Figure 4b, these trapping mechanisms play different roles on CO2 storage in the time scale from 1 to 10,000 years. Clearly, structural trapping plays a vital role in the initial stage of CO2 storage, and its effect weakens gradually. The residual trapping and solubility trapping show a significant impact in tens of years and lock up a certain amount of CO2 for thousands of years. With respect to mineral trapping, it begins to show a significant effect almost around a hundred years and plays a key role at a geological timescale.

2.2. Strategies of CO2 Storage

2.2.1. CO2 Storage in Saline Aquifers

CO2 storage in saline aquifers is one of the most important options due to the huge amount of storage capacity, which is estimated to be sufficient for the sequestration of 10,000 Gt of CO2, namely the emissions from large stationary sources for more than 100 years [32,53,54]. Compared with the other storage sites, the saline aquifers usually possess more wide distribution and greater regional coverage. Therefore, it has a better chance to be located near the CO2 emission sources, which could reduce the cost of CO2 transportation [47,55]. The most crucial issue brought by the sequestration of CO2 in saline aquifers is pressure build up and CO2 plume migration in formation, which has the potential to lead to the fracturing of formation and reactivation of faults and leakage of CO2, which should receive more attention [56]. Birkholzer et al. [57] conducted a numerical simulation to determine the influence of large-scale CO2 storage with an injection rate of 1.52 million tons per year (Mtpa) in an open saline aquifer. Their results indicated that there is significant pressure build up in the formation more than 100 km away from the injection zone, but the CO2 plume migration is rather small—that is, around 2 km—and is concentrated on the top of saline aquifer due to the buoyancy effect. They also showed that the pressure perturbation may reach shallow groundwater formation when there is a caprock with relatively high permeability (higher than 10−18 m2) between the saline aquifer and the shallow layers. However, the migration of reservoir fluids into groundwater formation is extremely unlikely. This demonstrates the safety of large-scale CO2 storage in saline aquifers.
There are mainly five commercial-scale CCS projects in saline aquifers, including the Sleipner project [58,59,60], the Snøhvit project [61], the In Salah project [62,63], the Gorgon project [64], and the Quest project [65]. Detailed geological and engineering data are shown in Table 2.
Regarding the Sleipner project, CO2 was separated from the methane produced at the Sleipner field in the North Sea. Then, the CO2 was injected into a regional saline aquifer within the Utsira Sand formation, and a total of 18 million tons were injected by 2018 since its initiation in 1996 [60]. Based on the engineering experiences from the Sleipner project, the CO2 separated from the liquefied natural gas (LNG) project was injected to the deeper Tubåen Formation at a rate of 2000 tons per day in the Snøhvit CCS project, which is located in the Barents Sea. The project was launched in 2008 with a total amount of 1600 ktons of CO2 injected until August 2012, and it is expected that about 23 million tons of CO2 will be stored there based on the projected lifetime of the Snøhvit LNG project [61,69].
The CCS project at In Salah, Algeria, is one of the world’s pioneering CCS projects. More than 3.8 million tons of CO2 have been injected to the Carboniferous sandstone at the Krechba field since 2004 [62]. This CCS project is unique due to the diversity of monitoring methods, including satellite monitoring and 4D seismic data, which have been used to monitor the response of formation to CO2 injection. Meanwhile, the accessibility of these monitoring data to the public is very high [62,63,70,71,72,73,74,75,76,77], so it could be a commendable case to study the CCS in saline aquifers.
The Gorgon CCS project located in the northwest of Australia, and it owns a Jurassic saline reservoir in the Dupuy Formation. In the lifetime of the Gorgon project, more than 120 million tons of CO2 is planned to be injected at a rate of approximately 3.8 Mtpa [64].
The Quest CCS project began in 2015 and is designed to store the CO2 from an existing facility for upgrading heavy oil in Scotford of Alberta, Canada. It is expected that approximately 27 million tons of CO2 can be injected to the Basal Cambrian Sands formation at an injection rate of 1.08 Mtpa through three to eight vertical wells [65].
Aside from the previously mentioned large-scale CCS projects, there are some small-scale projects as well. These include the Ketzin pilot site [78,79], the Illinois Basin-Decatur Project [80], and the Shenhua CCS demonstration project [81]. These projects have been conducted with detailed modeling and monitoring during operation, which demonstrates the safety of this technology and helps increase the public acceptance.
Although the storage capacity of saline aquifers is huge, the overall progress of CO2 storage in such aquifers throughout the world is still slow due to the lack of financial incentives. Therefore, some policies related to the taxes on carbon emission with a higher price might need to be formulated, which highlights the important role of the government on the application of CCS at a large scale.

2.2.2. CO2 Storage in Depleted Oil and Gas Reservoirs

There are many advantages of CO2 storage in depleted oil and gas reservoirs. Firstly, oil and gas reservoirs have a large amount of existing equipment installed on the surface and underground, which could be reused for CO2 storage with only minor modification. Secondly, the seal quality and integrity of the caprock are guaranteed and have been comprehensively characterized during the exploration and production process [56]. Thirdly, the extent of pressure perturbations and induced stress changes is much smaller in depleted oil and gas reservoirs compared with aquifers because of the long-term extraction of oil and gas [56]. Compared with depleted oil reservoirs, the depleted gas reservoirs are more favorable for CCS due to higher ultimate recovery and compressibility of gas, a larger storage capacity per pore volume is available [82,83,84]. Comparing the types of reservoirs used in this form of storage, condensate gas reservoirs are more advantageous over wet and dry gas reservoirs resulting from the little remaining gas, the phase behavior of the mixture of condensate gas and CO2, as well as the good injectivity of it [85]. Furthermore, the sequestrated CO2 per pore volume in depleted condensate reservoirs is very high: approximately 13 times higher than that of the equivalent aquifer [82]. However, attention should be paid as the phase change may occur in depleted condensate reservoirs, while not in dry and wet gas reservoirs.
There are some traits associated with the long-term trapping mechanisms of CO2 in natural gas fields. The noble gas and carbon isotope traces results show that solubility trapping in formation water is dominant, while mineral trapping is limited in the natural gas reservoirs with siliciclastic or carbonate lithologies [86]. It should be mentioned that the residual gas in the depleted reservoirs has an effect on the CO2 sequestration. Generally, capillary trapping capacity exhibits a positive relation with the remaining gas, while structural trapping capacity, dissolution trapping capacity, and the total storage capacity are inversely related with it [87].
The most important issue related to CO2 storage in depleted gas reservoirs is the low reservoir pressure, it’s sometimes below 20 bar at the initial stage of injection, which may lead to a strong Joule–Thomson cooling effect, probably reducing the reservoir temperature, further forming hydrate, freezing the residual water, and even compromising the well injectivity, especially when cold CO2 is injected [88,89,90]. When the temperature of the reservoir is over 40 °C, the Joule–Thomson cooling effect is not noticeable in permeable reservoirs, even though the initial formation pressure is as low as 2 MPa. However, the Joule–Thomson cooling may lead to the formation of hydrate, as the initial formation pressure is 6 MPa and the reservoir temperature is less than 20 °C [88]. In order to avoid the Joule–Thomson cooling and ensure a relatively high temperature at a low reservoir pressure, a high temperature of injected CO2 or a high mass flow rate should be applied. It should be noted that the high mass flow rate may lead to other problems at the beginning and shut-in of the injection. The system in the ROAD project connects a CO2 capture system at the Masvlakte Power Plant with an offshore depleted gas field that has a depletion pressure below 2 MPa [91]. It is a single source and sink system that allows pressure and temperature control at the shoreline inlet of the offshore pipeline by adjusting the level of after cooling at the compressor. It is used to ensure a high downhole temperature and ease the Joule–Thomson cooling effect. That is, a high temperature of injected CO2 is applied in the reservoir with low pressure, whereas a low temperature of injected fluids can be acceptable in the reservoir with higher pressure, to keep the injection pressure requirement at a low level [90]. Furthermore, the co-injection of SO2 and CO2 is an alternative method to suppress Joule–Thomson cooling and shows a beneficial thermal consequence [92]. In addition, the presence of methane can potentially reduce the Joule–Thomson cooling effect [93].
A few projects dedicated to CO2 storage have been implemented in depleted gas reservoirs. The first demonstration project in Australia named the CO2CRC Otway Project is well-known [94], in which the CO2 was injected into the Waarre C Formation at a depth of about 2050 m. This project commenced in March 2008 and ended on August 2009, with a total storage capacity of 65,445 tons [95]. It is worth mentioning that a community led “stakeholder reference group” has been set up in this project to communicate with the public and help increase their acceptance about CCS technology, which could be a demonstration for other CCS projects. Overall, the CO2CRC Otway Project demonstrates that the sequestration of CO2 in depleted gas fields can be achieved safely [95], and it provides a basis for the large-scale CO2 sequestration in depleted oil and gas fields. According to the experience gained in this project, the suitability and storage capacity of similar depleted gas reservoirs has been evaluated. For example, the depleted P18-4 gas field on the offshore of Netherlands [96] and the DF-1 South China Sea Gas field [97], owning a potential capacity of 1 Gt and 8 Mt respectively, are identified as suitable sites for the sequestration of CO2.
Generally, due to the advantages of low risk and cost-effectiveness, CCS in depleted reservoirs can play an important role in the mitigation of global warming, before the wide application of large-scale CCS in saline aquifers [98].

2.2.3. CO2 Storage in Coal Beds

CO2 injection into coal beds is another attractive strategy for CO2 storage. Most of the suitable coal beds for CO2 storage are located at a depth ranging from 300 to 900 m [99]. The sequestration of CO2 in coal beds possesses the major advantage that the potential coal beds are usually located nearby the existing or planned coal-fired power plants. Therefore, the transportation cost could be reduced significantly. However, CO2 storage in coal beds is still an immature technology, and only some pilot studies have been conducted on its suitability and storage capacity. The evaluated effective storage capacity of Cretaceous–Tertiary coal beds in Alberta, Canada is 6.4 Gt [99], and the potential storage capacity for the coal beds in China is about 142.67 Gt [100], which signifies the potential contribution of coal beds on the mitigation of CO2 emissions.

2.2.4. CO2 Storage in Deep Ocean

The CO2 can also be directly injected into the deep ocean at water depth of more than 2700 m [101,102], where the liquid CO2 can sink downward to the seafloor, because the CO2 is denser than seawater under high pressure and low temperature [102,103]. The storage capacity is extremely large due to the enormous volume of the ocean. However, this CCS technology cannot be applied widely because it may affect the marine environment.

2.2.5. CO2 Storage in Deep-Sea Sediments

The option of CO2 storage in deep-sea sediments not only combines the merits of geologic storage and ocean storage, but it also avoids many shortcomings [104,105,106,107]. For example, it is free from the potential harm to the ocean ecosystems as the CO2 is injected into the sediment deep beneath the ocean rather than directly into ocean. The storage mechanisms in terrestrial sequestration such as dissolution trapping, residual trapping, and mineral trapping still play a positive role. In addition, new storage mechanisms, including gravitational trapping and hydrate trapping, also work in the sequestration. The gravitational trapping comes from the fact that the higher density drives the CO2 into the deep sea [103] to the so-called negative buoyancy zone (NBZ). The depth at which the density of CO2 is identical to the salinity and temperature-dependent density of seawater is approximately 2700 m [52]. The hydrate trapping works because of the formation of CO2 hydrate under the condition of high pressure and low temperature [104]. Figure 5 shows the long-term evolution of injected CO2 in the deep-sea sediments. At the initial stage of injection, a little of hydrates form at the bottom of hydrate formation zone (HFZ), which is beneficial to reduce the permeability of the caprock. The area of hydrate caprock expands along with more CO2, reaching the bottom of the HFZ and limiting the CO2 below it. Meanwhile, the aqueous saturated with CO2 will sink downward because of the buoyancy-driven advection. Finally, the hydrate CO2 and liquid CO2 will dissolve in seawater and change into CO2 aqueous solution through diffusion, and permanent storage occurs.
Despite the enormous capacity and feasibility this technology shares, it is still in the formulation technology readiness level. In addition, CO2 storage in deep-sea sediments is far more expensive than onshore methods. In addition, it may take a long time to increase the public acceptance of this method [106,108].
In a summary of this part, there are several options for the underground CO2 storage, including the saline aquifers, depleted oil and gas reservoirs, coal beds, deep ocean, and the deep-sea sediments. The pros and cons of these storage strategies are summarized in Table 3.

3. Security Assessment of Underground CO2 Storage

3.1. Numerical Methods for the Security Assessment of CO2 Storage

The brine migration caused by CO2 injection may affect the ground water resources [57]. In addition, the chemical reaction between CO2 with the cement and well string [79,109,110], and the formation response such as the reactivation of faults and the shear failure of caprock, may lead to the failure of well integrity and caprock integrity, resulting in the leakage of CO2 [11,42]. Therefore, before the implementation of a CCS project, it is very important to assess risks through predicting the CO2 injection-induced formation responses, including the formation pressure change, formation deformation, and migration of CO2 plume, etc. Generally, such temporal and spatial responses of formation can be predicted both analytically and numerically. However, due to the physical and geological complexities in CCS projects, only a few semi-analytical models have been developed to estimate risks related with the migration of CO2 plume and leakage along abandoned wells [111,112,113,114,115]. For example, Nordbotten et al. [112] derived the solution for CO2 plume evolution during injection in dimensionless form as follows:
( Q well   ( p ( r well   , t ) p init   ) + Δ E p ) ( 2 π λ w k B Q well   2 ) = 0 1 ( λ 1 ) ln r ( b ) ( ( λ 1 ) b + 1 ) 2 d b + Γ 0 1 b [ r ( b ) ] 2 d b + 1 2 λ 1 λ ln ( π B φ V ( t ) )
where Qwell presents the volumetric injection rate; p(rwell,t) denotes the fluid pressure at the injection well; pinit is the initial pressure; ΔEp presents the potential energy required to submerge the CO2 into the denser water; λw presents the mobility of water; k is the permeability; B is the reservoir thickness; λ denotes the total mobility; φ is the porosity; V(t) denotes the total injected volume; and b presents the thickness in the CO2 plume profile, which is a function of radial distance from the injection well (r) and time (t).
The numerical simulation methods are more popular in assessment of the risks associated with CO2 storage. Many thermal–hydraulic (TH) coupled simulators have been developed for multi-component and multi-phase flow in CO2 storage, such as TOUGH [57,116,117], ECLIPSE [118,119], CMG-GEM [120,121,122,123], STOMP [124,125], MRST [126], COMET3 [127], IPARS [128], MUFTE-UG [129], FLUENT [130], and Tempest [131,132]. The hydro–thermal–mechanical (THM) coupled simulators are mostly constructed based on the coupling framework of fluid flow simulator (TH) and mechanical simulator (M). Specifically, the THM simulators used in CCS mainly include TOUGH-FLAC3D [63,133,134], TOUGH2-RDCA [135], TOUGH-RBSN [136], Sierra Arpeggio (Aria-Adagio) [73], OpenGeoSys-ECLIPSE [137], ABAQUS-ECLIPSE [138], and ECLIPSE-VISAGE [76]. Among them, the TOUGH-FLAC has been well tested and applied in many simulations of CCS [63,139], and it was developed by the Lawrence Berkeley National Laboratory [140,141].
In recent years, some software has been developed for the risk assessment in CCS, such as the National Risk Assessment Partnership (NRAP) Toolset and Leakage Assessment and Cost Estimation (PyLACE). The NRAP is designed to evaluate the environmental risks associated with CCS operation by the U.S. Department of Energy’s National Energy Technology Laboratory [142]. In NARP, the geological system of CCS is divided into several subsystems firstly, and each subsystem is characterized by a reduced-order model [143]. Then, the reduced-order models are linked by an integrated assessment model based on the system modeling approach [142]. Finally, the whole system model can be used to evaluate the risk performance. The framework of the NARP is shown in Figure 6.
By using NARP, two major types of risk, including CO2 leakage and induced seismicity, can be simulated. Additionally, the behavior of several important components in the CCS systems, including reservoirs, seals, wells, and ground water aquifers, could be modeled using corresponding tools. For example, the Wellbore Leakage Analysis Tool (WLAT) could be used for the assessment of the leakage potential of existing wells [145], and the Design for Risk Evaluation and Monitoring (DREAM) is developed for the evaluation and optimization of monitoring designs for long-term CO2 storage operation.
The Python-based web application PyLACE is designed to quantify the financial risks associated with potential CO2 leakage in a CCS system [146]. There are two major functional blocks in PyLACE: one of them is metamodel development, and the other one is metamodel-based decision support. It can convert the process-level risk assessment models into high-fidelity metamodels for the purpose of online assessment by using the high-performance computing and cloud computing infrastructures.
Recently, a deep neural network inversion was applied on 4D seismic data for estimating saturation and pressure [147], proving the availability of deep neutral network on the application of data-based inversion. To make the assessment of CCS more efficient and effective, the machine learning technology is encouraged to be used. It can be forecasted that the evaluation of CCS will be more and more intelligent with the development of machine learning technology.

3.2. Monitoring Technologies in the Assessment of the CCS Risks

The injected CO2 will be retained in underground for a long period, and the CO2 plume may affect the surrounding environment and the groundwater in particular [148]. Since it is difficult to predict with reasonable accuracy the key issues or risks in CCS by utilizing only simulation tools [149], the monitoring and history matching is very important in CCS assessment. The monitoring is used in most of the field development plans and routine field operations [62]. The most common monitoring technology used in CCS includes 3D seismic, 4D seismic, microseismic, vertical seismic profiling, gravimetry, cross-hole electromagnetic, pressure and temperature monitoring, geochemical sampling, soil and gas sampling analysis, tracers, atmospheric monitoring, microbiology, core analysis, satellite monitoring, and distributed temperature sensing technology.
The 3D seismic can provide a tri-dimensional image of the formation structures and the CO2 plume. The quality of the 3D seismic is affected by the medium. In off-shore monitoring, the 3D seismic monitoring data with high quality could be obtained, and the CO2 bodies above 106 kg at the depth of 1–2 km could identified due to the enhanced penetration of seismic waves in water [148].
The 4D seismic involves repeating 3D seismic in time-lapse mode to image the CO2 plume in the reservoir over time, which is beneficial for the monitoring of the migration process of CO2. A major challenge for the 4D signal to reflect the field data with high accuracy is the non-repeatable noise level in the data. This is on account of that the seismic imaging experiment is difficult to be repeated from one survey to the next due to the variations in the sources’ receiver positioning and geometry, the soil moisture content, and the formation water properties [150].
The microseismic activity levels show a correlation with the CO2 injection periods. Thus, it is beneficial for the understanding of the subsurface CO2 injection and migration process. In this procedure, the one-dimensional array that consists of several three-component downhole geophones will be deployed in the vertical well. Afterwards, the waveform data will be detected by the geophones and then transferred to the digitizers for recording [151].
Apart from conventional surface seismic acquisition, the hydrophone arrays, buried geophone, and the fiber optic cables are permanently installed in the vertical seismic profiling systems to achieve long-term monitoring and obtain the geological structure details [152]. This monitoring technology has been successfully used in the Ketzin pilot site and the MRCSP project.
Gravimetry testing can detect the variations of fluid density due to CO2 injection and thus provide the information of the location of CO2. The limit of this method is that the testing result is affected by the shape of the CO2 plume [148,153].
Cross-hole electromagnetic is a non-invasive method on the determination of the subsurface physical and chemical properties. It can also provide the information for the detection and monitoring of the location of CO2. In this domain, the electrical conductivity before and after the CO2 injection is obtained firstly. Then, it can be converted to the CO2 saturation with the help of inversion algorithms and appropriate rock-physics models [154,155]. It should be mentioned that this monitoring technology is only suitable for small-scale areas such as the area between wells [154].
Monitoring the formation fluid pressure and the variation of temperature can help in evaluating the risks associated with the failure of the caprock integrity and identifying the flow path of the injected CO2, respectively. It is important to point out that the wellhead pressure and temperature cannot provide enough information for the CO2 injection process, which has been certified at the Ketzin pilot site [156]. Therefore, the downhole pressure and temperature monitoring are recommended in the CO2 storage operation.
The geochemical sampling analysis could reveal the chemical variations, such as the drop of pH and natural variations in water chemistry, which is crucial for establishing a useful baseline for groundwater hydrology [157]. In addition, it can provide the information of the variation of the concentration of minerals that may be induced by the dissolution of carbonates and precipitation of anhydrite. It should be mentioned that the chemical reaction is a slow process in the sandstone formation [158], which is difficult to be detected in a short-term CO2 storage process.
Soil and gas monitoring can provide the data of CO2 concentration, which is beneficial for the definition of the baseline before the injection [148]. In addition, it can provide more data on natural CO2 variations in different environments and associated seasonal fluctuations.
Tracers monitoring is a cost-effective method for monitoring the origin of CO2 observations at wells and in the storage complex. The mechanism involves the co-injection of some specific compounds that can be detected in a very small concentration such as SF6, SF5CF3, and the isotope 14C together with CO2 [95,159]. Thus, the trail of the injected CO2 could be reflected by the traces.
Atmospheric monitoring means detecting the atmospheric concentration of CO2 that may be changed by the leakage of CO2 from the underground, which is beneficial for the identification of the anomalies above the natural base line [160]. The reliability of this technology may be affected by the significant natural variation of atmospheric concentration of CO2 induced by the organic matter decomposition and the soil respiration [148].
Microbiology monitoring can be conducted on the samples of reservoir fluids and minerals before and after the CO2 injection, which could be defined as a baseline and the modification caused by the presence of CO2, respectively [161]. Specifically, the biocenosis such as the sulfate-reducing bacteria (SRB) in the rock substrate and fluid samples could be analyzed by the molecular biological method, polymerase chain reaction–single strand conformation polymorphism (PCR-SSCP) method, and fluorescent in situ hybridization method [162]. This information related to the microbiology is valuable for the identification of biogeochemical process that affect the diffusion of CO2 in the reservoirs.
Core analysis is essential for the acquisition of the petrophysical and rock mechanical properties. Some measuring methods including the SEM imaging, XRD, and X-ray elemental analysis are usually included to obtain the micro morphological and mineralogical properties of the core [62].
The satellite-borne synthetic aperture radar (SAR) monitoring can provide the information related to the change of the ground surface caused by CO2 injection operation, which can be used to modify the model of CO2 distribution in the underground. By using SAR, the amplitude and the phase can be obtained firstly. Then, the phase difference between two observations can be converted into the displacement of the surface through the platform altitude and look angle [163]. Comparing the geophysical surveys methods, the satellite-borne SAR monitoring is considered as a cost-effective monitoring tool, and it has been successfully used in the In Salah project [163].
The distributed temperature sensing (DTS) technology is developed to overcome the limitations of conventional temperature monitoring that cannot provide the high vertical spatial resolution and real-time data. The fiber optic cable is used as a distributed sensor in DTS, which offers the possibility to measure the temperature from the surface to the bottomhole along the extension of the fiber [164,165].
The advantages and applications of the monitoring technologies in the CCS demonstration projects are summarized in Table 4 and Table 5, respectively. It shows that the geochemical sampling analyses, 3D seismic, microseismic, and pressure and temperature logs have gained the most popularity among the monitoring technologies due to their high performance in acquiring the characteristics of geological structures and formation fluids.

3.3. Generating CO2-in-Water Foams

The injected CO2 exhibits much lower viscosity and density compared with oil and brine in the reservoir conditions and leads to a poor displacement efficiency. Generating high viscosity CO2-in-water foams with large gas volume fraction would be an effective strategy to address this issue [166], which has some advantages for both oil reservoirs and saline aquifers. For oil reservoirs, CO2-in-water foams can improve oil recovery and the economics of CCUS in petroleum systems [167]. Guo and Aryana [168] used a glass microfluidic device to investigate the flow behavior of foam in oil-saturated heterogeneous porous medium, and it resulted that the foam injection can improve the oil recovery through the improvement of the sweep efficiency. A field test of CO2 foam flooding was conducted in the North Ward-Estes field in Texas, demonstrating that the foam can notably improve the sweep efficiency and be economically successful [169]. For saline aquifers, CO2-in-water foams can also improve the sweep efficiency, storage capacity, and economics. Guo et al. [170] used a glass-fabricated microfluidic device to investigate the effect of various factors on the CO2 storage capacity in aquifers. Their results showed that the CO2 storage capacity would be increased by over 30% with the foam injection compared with CO2 injection cases, demonstrating the superiority of CO2 foam on the improvement of CO2 storage capacity. It should be mentioned that the foam can also reduce the risk of leakage in the underground CO2 storage unit result from the reduction of fluid mobility. For instance, due to the significant shear rates differences between the flowing in the reservoir rock and leakage pathway, the foam may become gel inside the leak and reduce the leakage through the shear-induced gelation, i.e., the particles break the interaction barrier of forming clusters in high shear stress condition [171].
To achieve stable CO2 foams, the surfactants are used traditionally. The hydrophilic/CO2-philic balance (HCB) is considered as the principle for the designing of surfactants for CO2 foams, which can characterize the balance of surfactant and solvent interactions [172]. In recent years, the nanoparticles have been used combining with the surfactant solutions to improve the stability of CO2 foams, and Worthen et al. developed the concept of HCB to be nanoparticle HCB [173]. Worthen et al. [174] generated viscous and stable CO2-in-water foams with the mixture of nanoparticles (bare colloidal silica) and surfactant (caprylamidopropyl betaine). They suggested that the formation of foams was caused by the reduction of interfacial tension through the surfactant, while the stability of foams may be improved by the adsorption of nanoparticles at the CO2–water interface. The behavior of silica nanoparticles on the reduction of carbon footprint was also demonstrated by Rognmo et al. [167]. In addition, some other nanoparticles such as nano lauramidopropyl betaine with alpha-olefin sulfonate also have been demonstrated to perform well in maintaining a high foam quality [170], showing the superiority of nanoparticles on the stabilizing of CO2 foams. Overall, the generating of CO2-in-water foams especially combining with the nanoparticles are supposed to be a candidate in the designing of CO2 storage, while the overall cost should be considered.

3.4. Accelerating CO2 Dissolution Process

As mentioned in above sections, the free CO2 can remain more than 1000 years underneath the caprock, which increases the uncertainties in the long-term fate of injected CO2 and increases the cost of long-term monitoring operation. To address this issue, accelerating the dissolution process of CO2 and minimizing the free CO2 in the underground is an effective strategy [175]. Cameron and Durlofsky [176] used the Hooke–Jeeves direct search algorithm to optimize the locations of CO2 injection wells and the injection rate to minimize the mobile CO2 in the CCS system within 1000 years, and the results showed that the fraction of mobile CO2 would decrease from 0.220 to 0.072 in the optimal case, highlighting the importance of well location optimization. Anchliya et al. [175] proposed an engineered injection method to promote the dissolution and trapping of CO2. Figure 7 shows the schematic of the engineered injection process. Compared with conventional CO2 injection scenarios, there is another brine injection well located exactly over the horizontal CO2 injection well and near the top of reservoir. In the engineered injection system, another two brine production wells are placed at either side of the CO2 injection well. The brine injection well is used to limit the upward movement and impel the horizontal flows of CO2 under a lateral pressure gradient provided by brine injection, which increases the sweep efficiency and enhances the CO2 dissolution and trapping [175]. Numerical simulation studies show that about 90% of the injected CO2 could be immobilized within 20 years after CO2 injection ceases due to the fast dissolution trapping and residual trapping. It should be pointed out that the potential risk of pressurization due to CO2 injection could also be addressed by the engineered injection method through controlling the brine injection and production rates. In addition, it can enhance the storage capacity of CO2 in a bounded aquifer formation. Additional drilled wells are needed for the engineered injection system, which increases the cost of this CCS operation. Consequently, the adaptability of the engineered injection method should be quantified and site specific.
Another injection scheme called water-alternating-gas (WAG) injection was firstly proposed in the petroleum industry in the late 1950s to improve the sweep efficiency of reservoirs. The different types of WAG injection are illustrated in Figure 8. Zhang and Agarwal investigated the potential of the WAG injection scheme with the goal of improving the efficiency of CO2 storage [177,178]. The results showed that the optimized WAG scheme could accelerate CO2 dissolution and decrease the impact zone up to 14% comparing with that of the constant gas injection scheme. However, the WAG injection scheme may decrease the total storage capacity of the reservoirs due to the large amount of injected water. In addition, it may increase the cost in the injection process, thus it has not been applied widely in the CCS operation.
To eliminate the adverse effects of WAG injection schemes on storage capacity, an injection scheme combining the intermittent injection method and brine production was proposed by Tanaka et al. [180]. The schematic diagram of intermittent injection is shown in Figure 9, which suggests that a diagonal pair of wells were used for CO2 injection alternately. In this injection scheme, a diagonal pair of wells (Well 1 and Well 3) are used for CO2 injection, while another pair of wells (Well 2 and Well 4) are used for producing brine, which will be re-injected to the reservoir through Well 1 and Well 3. Numerical simulation results show that both the dissolved and residual CO2 are increased. Specifically, the ratio of the trapped CO2 increases by 20% compared with the base case, which has only one well with continuous injection. Another advantage of this injection scheme is that it could mitigate the pressure buildup through intermittent injection and water production, which highlights the importance of the CO2 injection design and brine management.
In terms of the above methods for accelerating the dissolution of CO2, they may increase the cost in the injection process due to water injection. In addition, additional wells are needed for the engineered injection and intermittent injection with four wells, which results high costs on the operation. However, the cost on the monitoring would be reduced due to the relative low leakage risk caused by the rapid dissolution of CO2. The overall economic costs of the CCS unit and the storage capacity are supposed to be taken into account for optimization when considering the utilization of these technologies.

3.5. Accelerating Mineral Carbonation Process

The mineralization of CO2 is an effective way to fix the injected CO2 and guarantee its security permanently. However, for conventional CCS in saline aquifers, it takes tens of thousands of years for the mineral trapping because of the low reactivity of silicate minerals in sedimentary rocks with CO2 [52]. To accelerate the mineral trapping of CO2, some novel methods have been developed, such as CO2 storage in basalt rock formation [12], CO2 storage in peridotite formation [181], direct mineralization of flue gas by coal fly ash [182], direct aqueous mineral carbonation [29], and pH swing mineralization [183].

3.5.1. CO2 Storage in Basalt Rock Formation

Applying the sequestration of CO2 in basalt formations is proposed by Gislason and Oelkers [12]. Basalt contains approximately 25% of magnesium, calcium, and iron oxides, and it is far more reactive with carbonic water compared with sedimentary silicate rock [184]. Another merit of this method is that abounding basaltic rocks exist on the Earth’s surface [185], which offers the possibility of large application of CCS in basalt formation. The field test of CCS in basalt was conducted in the CarbFix pilot project in Iceland in 2012 [159]. In this project, 175 tons of pure CO2 were injected with water firstly. Then, 73 tons of CO2–H2S mixture (55 tons CO2) were fully dissolved in water and injected. The measured dissolved inorganic carbon of 14C and the monitoring well shows that more than 95% of injected CO2 was mineralized to carbonate minerals within two years, demonstrating the efficiency of mineral trapping of CO2 in basalt. However, large-scale application of this technology requires substantial quantities of water during the CO2 injection process. In addition, the cost of storing and transporting CO2 for the CarbFix project is about twice that of storage in typical sedimentary basins. Therefore, the application prospect of this technology is not very promising in the near future.

3.5.2. CO2 Storage in Peridotite Formation

The mantle peridotite is mainly composed of olivine and pyroxene, which can react with CO2 and H2O to form hydrous silicate, Fe-oxides, and carbonates. Isotope analysis and reconnaissance mapping indicate that about 104 to 105 tons of CO2 per year are trapped to solid minerals via peridotite weathering effect in Oman [181]. The main reaction can be formulated as:
2 Mg 2 SiO 4   +   Mg 2 Si 2 O 6 + 4 H 2 O   =   2 Mg 3 Si 2 O 5 ( OH ) 4
Mg 2 SiO 4 + 2 CO 2   =   2 MgCO 3   +   SiO 2
Mg 2 SiO 4   +   CaMgSi 2 O 6 + 2 CO 2 + 2 H 2 O   =   Mg 3 Si 2 O 5 ( OH ) 4   +   CaCO 3   +   MgCO 3 .
As shown in Figure 10, the carbonation rate could be enhanced more than 1 million times compared with the natural rate with the three-step operation process, beginning with drilling and fracturing, followed by injection of hot CO2 (approximately 185 °C) at a rapid rate to heat the fractured peridotite, and then injecting CO2 with normal temperature. In this case, the system maintains a temperature of 185 °C and high carbonation rate due to the exothermic carbonation of peridotites. It is estimated that the peridotite in Oman alone could trap more than 1 billion tons of CO2 per year into carbonate minerals, showing a huge amount of capacity.
However, except for the peridotite exposing through large thrust faults, the mantle peridotite is generally more than 6 km below the seafloor and 40 km below the land surface [181], making it difficult for the sequestration of CO2 in these formations. Thus, the peridotite in shallow areas could be used effectively for the sequestration of CO2.

3.5.3. Direct Mineralization of Flue Gas by Coal Fly Ash

Reddy et al. [182] conducted a preliminary experiment to study the reaction between flue gas and coal fly ash in a fluidized bed reactor. This experimental setup is shown in Figure 11. The results show that the concentrations of CO2 and SO2 in flue gas dropped from 13.0% to 9.6%, from 107.8 to 15.1 ppmv, respectively in 2 min. In addition, the Hg in flue gas was also mineralized by fly ash particles. Reddy et al. [182] conducted a pilot-scale study with a fly ash content of 100–300 kg to investigate the feasibility of this technology. In the pilot studies, the fly ash particles were fluidized by the flowing of flue gas in a fluidized bed reactor to ensure sufficient mixing and contact between them, and the reaction occurred under a fixed pressure of 115.1 kPa. Experimental results show that the content of CaCO3 produced by the reaction of flue gas and fly ash ranged from 2.5% to 4% in 10 min. Meanwhile, the contents of S and Hg in the fly ash increased from non-detectable to 0.45 and 0.5 mg/kg, respectively. This confirmed that the flue gas components can be captured without separation and mineralized by fly ash particles using an accelerated mineral carbonation process. However, the treatment of the carbonated fly ash produced by using this method is still an important problem that needs be addressed [30].

3.5.4. Direct Aqueous Mineral Carbonation

Direct aqueous mineral carbonation is a process that uses a bicarbonate-bearing solution mixed with reactant minerals such as magnesium and calcium silicate rocks to convert gaseous CO2 into solid form [186]. The magnesium and calcium silicate rocks are distributed all over the world. For instance, the magnesium silicate rock in Eastern Finland has the ability to store 10 million tons of CO2 per year for 200 to 300 years [187].
The overall chemical reaction of the carbonation of serpentine can be described as:
Mg 3 Si 2 O 5 ( OH ) 4 + 3 CO 2 =   3 MgCO 3 + 2 SiO 2 + 2 H 2 O .
In the process, the serpentine was treated at 630 °C to attain an active mineral. Then, the stoichiometric conversion of 78% of the silicate to carbonate was observed by using the bicarbonate-bearing solution at the conditions of 150 °C and 18.5 MPa, with 15% solids. This occurred within 30 min, showing an extremely high reaction rate for the carbon mineralization. To reduce the energy consumption on the heat reactivation and increase the effective reaction area, mechanical activation was proposed by adding an attrition grinding step. In this case, the reaction condition was optimized to 25 °C and 1 MPa, which still led to up to 65% carbonation within 1 h [186]. For instance, the combination of grinding and heat activation was used to attain a better carbonation mineralization performance, as shown in Figure 12 [29]. Based on this concept, the feasibility of this technology in the mineralization of flue gas was investigated by Verduyn et al. [29]. As shown in Figure 12, the relative high pH due to the lower solubility of flue gas makes it difficult for the leaching of cations. To eliminate this, the contact of the mineral and flue gas is done in a slurry mill and a leaching basin.

3.5.5. pH Swing Mineralization

To increase the conversion efficiency in the CO2 mineral trapping process, the pH swing approach was proposed by Park and Fan [188]. They combined the internal grinding in acidic solvent for a rapid dissolution of a serpentine sample. Three solid products including SiO2-rich solids, iron oxide, and magnesium carbonate were produced by controlling the pH. Teir et al. [187] used HCl and HNO3 to dissolute serpentinite and then transformed the serpentinite to hydromagnesite with the help of CO2. In their results, the pure hydromagnesite, which is thermally stable at 300 °C, was produced. However, the additional amount of chemicals used in this work increase the costs and make it infeasible for CCS operations. To reduce the cost, a pH swing CO2 mineralization method with a recyclable reaction solution was proposed by Kodama et al. [189]. They selectively extracted the alkaline-earth metal from steelmaking slag in an ammonium chloride solution. The reacted solution was used for CO2 absorbent and produced ammonium carbonate. Then, the calcium carbonate was precipitated in another reactor with the recovery of ammonium chloride. The results showed that the selectivity of the calcium extraction reaction reached 60%, and pure CaCO3 was produced with an energy consumption of 300 kWh/t-CO2. Wang and Maroto-Valer [183] proposed a modified carbon mineralization process as shown in Figure 13. Through experimental studies, they concluded that (NH4)2CO3 is more beneficial for increasing the efficiency of carbon fixation compared with NH4HCO3, and the optimal efficiency of CO2 mineralization reaches 46.6%. Furthermore, the pH swing mineralization process was optimized by Sanna et al. [190] under different temperature conditions. The results showed that the total CO2 trapping efficiency was 62.6% at the condition of 80 °C, with the molar ratio of 1:4:3 for Mg:NH4 salts:NH3. However, the energy consumption and overall economic cost need to be lowered before any large-scale application.
To summarize this section, CO2 storage in basalt rock formation and peridotite formation is limited by the distribution of the particular rock. In addition, the substantial quantities of water and energy consumption for heating increase the cost a lot. The direct mineralization of flue gas by coal fly ash would be a promising technology result since the gas could be mineralized without separation, which decreases the overall cost a lot. The pH swing mineralization may be another promising technology for the sequestration of CO2 due to the sustainability. The recyclable and cheap chemical reagents ought to be introduced into the mineralization process to make this technology more cost-effective.

4. Strategies for Improving the Cost-Effectiveness of CO2 Storage

4.1. Enhanced Industrial Production with CO2 Storage

Resources production during CO2 storage is an effective method to partly cover the cost of CCS and obtain additional economic benefits, which is called CO2 capture, utilization, and storage (CCUS) [19]. In the process of CCUS, CO2 usually works as a working fluid to enhance the recovery of underground resources through displacement, dissolution, thermal conductivity, and reactive transport. The potential geological formations for CCUS include oil reservoirs, gas reservoirs, saline aquifers, shale formation, un-mineable coal seams, hot dry rock, uranium deposit formation, and natural gas hydrate reservoirs [191]. The corresponding CCUS technologies are CO2-enhanced oil recovery (CO2-EOR), CO2-enhanced gas recovery (CO2-EGR), CO2-enhanced water recovery (CO2-EWR), CO2-enhanced shale gas recovery (CO2-ESGR), CO2-enhanced coal bed methane recovery (CO2-ECBM), CO2-enhanced geothermal systems (CO2-EGS), CO2-enhanced in situ uranium leaching (CO2-IUL), and CH4–CO2 replacement from natural gas hydrates, respectively [191,192].

4.1.1. CO2-EOR

In CO2-EOR technology, the CO2 is injected into oil reservoirs to enhance the recovery of crude oil, which is the most successful and promising technology combining the utilization and sequestration of CO2 [27,193,194]. The displacement of oil by CO2 can be classified as multicontact miscible and immiscible processes, depending on the properties of CO2 and the reservoir fluids at the condition of reservoir pressure and temperature [195,196]. In the multicontact miscible displacement procedure, the minimum miscibility pressure (MMP) is required for multicontact miscible displacement. The immiscible displacement occurs when the pressure is lower than the MMP, with less components exchange between CO2 and oil in the reservoir [197].
There are three CO2 injection schemes for the operation of CO2-EOR, including continuous injection, water-alternating-gas (WAG) injection, and cyclic injection. In the continuous injection scheme, CO2 injection and oil production are running continuously, which has been applied in the North Cross Devonian Unit for enhanced oil recovery [198]. The multicontact process can be achieved through vaporizing and condensing [199]. However, this injection scheme has not gained much popularity in the field application compared with the WAG injection and cyclic injection. WAG injection is the widely used form, because it can decrease the mobility ratio between the injection fluids with oil and lead to late gas breakthrough and high oil recovery. In the design of WAG injection, the optimization algorithm such as the Lagrangian and stochastic simplex approximate gradient algorithm could be used to obtain the maximum net present value [200]. Although the WAG injection is beneficial for improving the oil recovery, it may cause the gas to flow upward, while the water and oil flow downward due to the large density differences, resulting in early gas breakthrough, especially in the reservoirs with highly permeable channels and large vertical heterogeneity [199]. To address this issue, the cyclic injection process, i.e., gas huff-n-puff process was proposed, which is composed of three stages, including the gas injection stage, well shutting state, and the oil production stage.
For conventional oil reservoirs, CO2 flow dominated throng the rock matrix. The mechanism of CO2-EOR is due to the solubility of CO2 in oil under the supercritical phase condition, which can decrease the oil density and viscosity, leading to enhancing the oil recovery [199]. For the unconventional tight oil reservoirs such as shale oil reservoirs, fracturing is an essential technic for the exploitation. In this scenario, CO2 flowing is dominated by fracture flow instead of rock matrix flow. The process of the CO2-EOR in fractured oil reservoirs can be divided into four steps, as shown in Figure 14. In the initial stage (Step 1), the injected CO2 flows rapidly through the fracture. Then, the CO2 starts to permeate into the rock matrix under the displacement effect (Step 2). During this stage, the permeating CO2 may carry oil into the rock and decrease the oil production. Simultaneously, the permeated CO2 would lead to the swelling of oil and then mitigating out of the matrix, which is beneficial for the oil production. The oil continues to swell and lower viscosity by the permeated CO2, and moves to the fracture in the follow stage (Step 3), which corresponds to the well shutting stage in the huff-n-puff process. Finally, the pressure equilibrium inside of the matrix is approached, thus the migrating of the miscible or immiscible oil from the matrix to the fracture is dominated by diffusion effects. The oil in the bulk CO2 is swept through the fractures to the production well by the production pressure [201]. The cyclic injection scheme can also promote the propagating of reservoir pressure due to CO2 injection near the injection well, especially for the reservoir with ultralow permeability. Specifically, the CO2 huff-n-puff performs better on the oil recovery when the reservoir permeability is lower than 0.03 mD [202]. Characterization of the flow behavior of CO2 and oil in the low permeability formation with complex natural and hydraulically created fractures under the in situ conditions are supposed to be emphasized, which is beneficial for improving the efficiency of CO2-EOR.
Apart from increasing oil recovery, CO2-EOR provides an additional advantage of CO2 sequestration, which could be an important economic incentive for early CO2 storage projects [203]. Typically, 3 tons of CO2 injection can produce approximately 1 bbl of incremental oil. It is shown that about a 5% to 15% enhancement in oil production can be achieved by using CO2-EOR [204]. In the largest discovered fields over the world, it is estimated that approximately 470 billion barrels of incremental oil could be produced simultaneously with 140 billion metric tons of stored CO2 by using CO2-EOR [205].
The first CO2-EOR pilot project was implemented at the SACROC oil field in 1972 [206], in which CO2 foam was implemented to alter the mobility and improve the sweep efficiency [207,208]. At present, CO2-EOR is a relatively mature technology that has been widely used in the petroleum industry to enhance oil recovery for tens of years, with the capacity of more than 1000 million tons of CO2 stored subsurface [209,210]. This technology has gained great success in North America. In the USA, more than 260,000 bb/d are produced due to the application of CO2-EOR technology [211]. In the Weyburn oilfield in Canada, the CO2-EOR project was conducted to extend the life of an oilfield. About 20 million tons of CO2 is planned to be sequestrated in the oil reservoir [212,213]. In recent years, the feasibility of CO2-EOR in China has been massively studied. The first CO2-EOR project in China, i.e., the Jilin Oilfield, injected nearly 217,000 tons of CO2 with a storage efficiency of 96% by April 2013 [214], and the CO2 capacity is about 600,000 tons [10]. The technology of CO2-EOR has application prospects in the Shengli Oilfield and Bohai Bay Basin, with 6.7% incremental oil recovery and 683 million tons of incremental oil production, respectively [215,216]. CO2-EOR also attracted much attention in Europe. In Poland, the potential utilization of anthropogenic CO2 for CO2-EOR was studied in the B8 oilfields in Baltic Sea and Brage on the Norwegian Continental Shelf [217], which is a part of the ongoing PRO_CCS project funded by Norway Grants. The simulation results showed that the total storage capacity of Brage and B8 oilfields are 33 and 4.8 million tons in 17 years of injection, with an expected incremental oil production of 98 and 14.6 million bbls, respectively.
To co-optimize the CO2 storage and enhanced oil recovery, Ampomah et al. [218] proposed an objective function (Equation (6)) considering both CO2 storage and oil production, which can be optimized by the neural network and genetic algorithm. In the optimal case of the Farnsworth field unit, more than 94% of CO2 could be sequestrated with approximately 80% of the oil produced, which provides a guideline for the co-optimization of CO2 storage and EOR.
f = w 1 × F O P T + w 2 × F G I T
where w is weight assigned to vector, FOPT is the cumulative produced gas, and FGIT is the cumulative injected gas.
Similarly, a framework to co-optimize the oil production and CO2 storage was developed by Jahangiri and Zhang [219]. In the framework, the net present value (NPV) is treated as the optimization objection function, which was solved by the ensemble-based optimization algorithm.
NPV = t = 1 T C ( 1 + r ) t C 0
where t is the time step, T is the operation period, r is the periodic discount rate, C is the cash flow in the time step that is determined by the price, injection, and production volume of CO2 and oil, and C0 is initial investment.
By using this method, the well injection patterns and injection rates for the maximum NPV can be determined. In addition, the discrete time optimization model could be used for maximizing the total profit in CO2-EOR operations, with consideration of both enhanced oil recovery and geological CO2 sequestration [220].
Artificial neural network models can be used to predict and optimize the performances of CO2-EOR, as shown in Figure 15 during the multi-cycled water-alternating-gas process [221]. There are four neurons in the input layer corresponding to water-to-gas injection time ratios (WAG), temperature, permeability ratio, and initial water saturation, respectively. Subsequently, the oil recovery, oil production rate, gas and oil ratio (GOR), and net CO2 storage amounts are set as the targets, which correspond with four neurons in the output layer and with 10 neurons in the hidden layer. The oil recovery and net CO2 storage can be accurately predicted in this framework. For instance, the optimal injection scheme could be obtained for a maximum economic profit in various reservoir conditions by using this method.
The machine learning approach can also be applied to optimize oil recovery as well as CO2 storage [222]. As shown in Figure 16, a history matching model was developed based on production history data in the CO2-EOR process under this optimization framework. Then, the hybridized multi-layer and radial basis function neural network method were utilized to train a proxy model, which is beneficial for increasing computational effectiveness in the optimization process. After a proxy model with reliable accuracy was realized, the machine learning optimization algorithm was used to obtain the optimal solution of the objective function that incorporates the role of parameters such as oil recovery and CO2 storage. This work highlights the adaptability of a robust machine learning approach for optimizing the CO2-EOR process. Considering the mature technology and huge market demand, it can be concluded that CO2-EOR may play a more important role on the mitigation of CO2 than other strategies of CO2 utilization in the next few years.

4.1.2. CO2-EGR

The technology of CO2-EGR means that it enhances the gas recovery by CO2 injection. Enhanced gas recovery is realized by both the displacement and re-pressurization of the remaining gas in a depleting or depleted reservoir [223], especially for sour gas reservoirs in which CO2 is produced mixed with the natural gas. The separated CO2 from the produced gas could be injected back into the reservoir to enhance gas recovery. Additionally, CO2 has the potential to reduce the dew point pressure of reservoir fluids in wet gas reservoirs, which is favorable for eliminating condensate blockage and improving CH4 production [224,225]. It is estimated that up to 11% incremental gas recovery can be achieved by CO2 displacement [223].
The feasibility of CO2-EGR has been investigated by many experimental and numerical simulation studies [83,131,132,211,224,226,227,228,229,230]. Some typical displacement experiments in a variety of temperature and pressure conditions, revealing the mechanism of CO2-EGR and providing a guide line for the application of this technology, are summarized in Table 6.
The most critical hurdle in CO2-EGR is the breakthrough of CO2 in the reservoir producing CO2-contaminated gas [131,237]. Actually, the preferential pathway has a significant impact on CO2 breakthrough and ultimate CH4 recovery [238], so the geological formations for CO2-EGR, especially the microstructures, are supposed to be characterized in detail. Irreducible water in reservoirs also has an impact on the mixing of CO2 and CH4 [239]. The dispersion increases with irreducible water, because the pores occupied by irreducible water lead to much narrower pores and more tortuous flow paths.
In addition to the above-mentioned geological parameters, engineering parameters also have a significant influence on gas mixing and CO2-EGR performance [131,240,241]. CO2 injection with a horizontal well in the lower parts and CH4 production in the upper parts of reservoirs could mitigate the breakthrough of CO2 at the production well [131,242]. CO2 injection during the early decline phase of natural gas production is beneficial for ensuring the displacement in supercritical phase and achieving a high CH4 recovery [131]. On the contrary, it may cause the trapping of CH4 in unswept areas under high pressure. CO2 injection in the late phase could avert this shortcoming and improve the performance of CCS, which is more attractive when the CO2 sequestration is considered [131,243]. On the whole, it can be concluded that the time of CO2 injection to obtain a maximum incremental recovery is highly affected by the allowable produced CO2 concentration at the production well, which is determined by the economics of CCS projects [244].
Whether the CO2 is injected in the early or later stage, it is recommended to inject CO2 at a relatively high pressure to ensure the supercritical phase in the displacement process. In this case, the distribution of CO2 is dominated by gravity forces [245]. As the CO2 is much more dense than CH4, CO2 will occupy the smaller space and spread at a slower rate, which could mitigate CO2 breakthrough. However, if the injected CO2 is in gas phase in the reservoir, the CO2 will occupy a large volume and mix with the CH4 more easily, which can lead to an early CO2 breakthrough [245].
Regarding the injection rate, of course, a high injection rate can increase the gas recovery [241]. However, a high injection rate also brings excessive gas mixing, which is harmful for methane production. It is suggested that the injection rate should be lower than the production rate to avoid an early breakthrough of CO2 [237]. In Al-Hasami’s study [223], 9% incremental methane recovery can be achieved when the CO2 injection rate is only 13% of the production rate. Instead of the constant injection rate, a constant pressure injection scheme was proposed to avoid potential risks related to high pressure [246]. The optimal injection strategy could be achieved by an optimization code based on genetic algorithm and multi-phase simulator TOUGH2 (GA-TOUGH2).
Geological parameters also greatly affect the performance of CO2-EGR. The viscous and gravity force affecting parameters e.g., permeability, formation dip, and thickness, play a vital role in the stability of displacement. The fluid properties such as the diffusion coefficient and water salinity take second place on affecting the CO2 breakthrough [244]. Specifically, the connate water in reservoirs has a positive impact on CO2-EGR performance. As a result, the dissolution of CO2 in reservoir fluids is favorable for enhancing the storage capacity and mitigating the CO2 breakthrough in the production well [223,237,247].
There are several CO2-EGR projects around the world, including the Alberta gas field project, the K12-B field project, and the CLEAN project. The Alberta gas field project is located in Canada. In this project, impure CO2 with less than 2% of H2S has been injected into the depleted Long Coulee Glauconite F gas Pool in southeastern Alberta since 2002, but the operation was terminated in 2005 because of the breakthrough of acid gas [245]. The K12-B gas field is located in the Dutch continental shelf in the North Sea, with a reservoir depth of about 3800 m below the sea level. The reservoir pressure has dropped from 40 MPa to 4 MPa with a production of 90% of the initial gas in place. The initial reservoir temperature is 128 °C [248]. Over 0.1 million tons of CO2, which is separated from the produced gas directly at the offshore platform, has been injected over a period of 13 years since 2004. Monitoring data shows that the well integrity has remained stable [249]. Furthermore, no major complications occurred in the lifetime of this project, which proves that the safety of CO2-EGR can be ensured [250]. The CLEAN project was conducted between 2008 and 2011 to inject CO2 into the Altmark natural gas field in Germany. The risk assessment of this project has been conducted based on digital databases. The results showed that the safety and efficiency of EGR technology based on CO2 injection could be achieved. Meanwhile, the borehole integrity could be achieved without any intervention, providing a guideline on CO2-EGR [251].
Generally, the technology of CO2-EGR is still immature and needs more efforts to address the problems, such as mitigating the CO2 breakthrough and achieving a favorable performance in both enhancing CH4 recovery and CO2 sequestration. The economic success is largely dependent on the political developments in the next years and decades [251].

4.1.3. CO2-EWR

Similar to CO2-EOR and CO2-EGR, CO2-EWR is a methodology combining CO2 sequestration and saline water production [252,253,254], which is developed from the technology of CCS in saline aquifers. Figure 17 shows a diagram of CO2-EWR technology. The operation of CO2-EWR could decrease formation pressure and avoid potential leakage through extracting formation water, thus it could further improve the storage efficiency and achieve higher security and stability of large-scale geological CO2 sequestration [255]. Besides, the produced saline water could be used for drinking, industrial, and agricultural utilization after desalination treatment such as using a high-efficiency reverse osmosis system [252]. Meanwhile, the deep brine resources obtained through the cascade extraction may create economic profit and fill the cap of cost in the operation of CCS technology [253]. Kobos et al. [252] proposed a numerical simulation model to investigate the feasibility of CO2-EWR based on a hypothetical case study from a representative power plant and saline formation in the southwestern part of the United States. In their work, the extracted saline water was treated with a high-efficiency reverse osmosis system and then used as power plant cooling water. The results showed that the coupled technology of CO2 storage and saline water extraction and treatment is feasible for tens to hundreds of years.
Unfortunately, the added cost of extraction wells is considered a shortcoming of CO2-EWR [252]. Furthermore, the production of brine must be ceased once the breakthrough of CO2 occurs [256]. In general, under effective engineering design, the CO2-EWR technology has application prospects.

4.1.4. CO2-ESGR

In regard to CO2-ESGR technology, the CO2 is injected into shale gas reservoirs to replace and displace shale gas, for the ultimate goal of enhancing the shale gas recovery, with a side benefit of CO2 sequestration synchronously [257]. The dominate mechanism of CO2-ESGR is the competitive absorption of CO2 by shale matrix [199], such as with a CO2 sorption capacity up to 1 mmol per gram for the Muderong Shale [258]. In addition, the pressure gradient displacement plays an important role [259]. Liu et al. [259] conducted a numerical simulation to evaluate the feasibility of CO2-ESGR, and results showed that over 95% of the injected CO2 was instantaneously adsorbed and sequestrated in the reservoirs. However, only limited ESGR performance was detected due to the limited communication between the wells in this study. The feasibility of CO2-ESGR on the Devonian Gas Shale Play of eastern Kentucky was investigated by Schepers et al. [127]. They found that the huff-n-puff scenario was not suitable, while the full-field continuous CO2 injection was a good option. About 300 tons of CO2 were injected within one and half months. A significantly increased recovery was attained, and approximately half of the injected CO2 was sequestrated. Therefore, it can be concluded that there remains a long way before the application of CO2-ESGR, and the contribution to the mitigation of CO2 emission is still limited.

4.1.5. CO2-ECBM

In CO2-ECBM, the CO2 is injected into un-mineable coal seams to displace and replace coal bed methane, simultaneously achieving CO2 sequestration in the coal seams. Similar to the CO2-ESGR, CO2 works as displacing fluid and is competitive in the process of CO2-ECBM [260,261]. The potential ECBM recovery in China is estimated to be over 3.751 Tm3 [100], highlighting the superiority of this technology. However, the injected CO2 in CO2-ECBM projects is usually less than 1 million tons per year, and many coal seams usually with low permeability such as those in Western Europe are not suitable for the application of this technology [148]. Herein, the role of CO2-ECBM on mitigating the emission of CO2 is limited.

4.1.6. CO2-EGS

Geothermal energy is regarded as a clean, renewable, and reliable energy for its advantages of sustainability and environment-friendly characteristics [262]. It is extracted through water traditionally. Brown [263] firstly proposed the concept of using supercritical CO2 instead of water as the heat exchange fluid in an EGS. It has been proven that the heat extraction efficiency of CO2-based systems is superior to water-based systems. If this concept is popularized, more regions worldwide with relatively low temperature can be used for electricity production in an economically beneficial manner [254,262,264]. In addition, the mobility of CO2 is better than that of water, which is beneficial for the production of fluids and the extraction of geothermal energy.
In recent years, the technology combining geothermal extraction and CO2 storage has gained more attention [264,265], which can achieve an efficient geothermal energy extraction as well as mitigating CO2 emission. For example, the regional energy deficit could decrease by 22.1% and the CO2 emissions could decrease by 31.3% in the Latium Region in Central Italy if the CO2-EGS was applied [266]. However, the technology of the CO2 plume geothermal system is still at the conceptual stage and pre-feasibility studies phase, and many efforts need be devoted toward its study before its application.

4.1.7. CO2-IUL

CO2-IUL is a technology that injects CO2 and leaches uranium ore out of geological formation through reaction with ore and minerals in the ore deposits [191]. CO2-IUL could increase the recovery of uranium and simultaneously be favorable for CO2 storage, especially for sandstone-type uranium mining [191]. However, the global annual demand of natural uranium is only around 0.1 million tons [267], thus it may be difficult for CO2-IUL to reduce CO2 emission significantly due to its limits and demands.

4.1.8. CH4-CO2 Replacement from Natural Gas Hydrates

The technology of CH4-CO2 replacement from natural gas hydrates (NGH) is regarded as a win–win method for exploring NGH and simultaneously storing CO2 in the form of CO2 hydrates formation [268,269,270,271]. As shown in Figure 18, the mechanisms of this technology can be divided into four steps. Firstly, the CO2 molecule diffuses into the surface of CH4 hydrate and decreases the stability of CH4 hydrate structure (Figure 18a). Secondly, due to CH4 hydrate dissociation, the CH4 molecule escapes from the hydrate cage (Figure 18b). In the next stage, the hydrate is re-formed. As shown in Figure 18c, the CO2 molecules mainly occupy the large cage, while CH4 molecules occupy the small cage. Finally, the CH4 molecules diffuse from the surface of hydrate and change into gas, while the CO2 molecule diffuse into deeper hydrate layer to continue replacing the CH4 in hydrate (Figure 18d) [270]. To improve the performance of CH4–CO2 replacement from natural gas hydrates, a thermal stimulation approach was proposed [271]. By using this approach, the CH4 replacement exhibits an upper limit of 64.63%, and the maximum CO2 storage efficiency can reach up to 78.40%–96.73% [271].
Based on the concept of thermal stimulation to CH4–CO2 replacement, a geothermal-assisted CO2 replacement method (GACR) was proposed by Liu et al. [272]. As described in Figure 19, the CO2 with ambient temperature was injected into geothermal reservoir for heating, then the heated CO2 flows upward into the hydrate-bearing layer (HBL) to enhance the NGH dissociation. Numerical simulation results showed that the GACR method can significantly accelerate NGH dissociation and increase CH4 recovery. However, the application of this method is limited by a strict precondition that a thermal reservoir exists below the methane hydrate reservoir.
The research on CH4–CO2 replacement from natural gas hydrates is still in the preliminary experimental and numerical study phase [271,273,274]. However, it is expected that great progress will be made in the near future under the stimulation of methane hydrate production and CO2 sequestration.
In a short summary of this part, the utilization of CO2 for resources production and CO2 storage are encouraged to be designed for the whole process of engineering operation to optimize its performance. To achieve this goal, artificial intelligence may play an important role. Although CO2-EOR has been used commercially, the other technologies are in the pilot plant phase in terms of technology readiness level [204]. CO2-EOR would be the most promising technology combining the utilization and sequestration of CO2 in the next few years. CO2-EGR is another promising technology, whose application is limited by the mixing of CO2 and CH4 in the reservoir. Considering that the mixing behavior is affected by the geological parameters, i.e., porosity, permeability, residual water saturation, and engineering parameters, i.e., injection rate, injection pressure, production rate, a site selection system for CO2-EGR project is encouraged to be developed.

4.2. Co-Injection of CO2 with Impurities

The biggest obstacle for large-scale CO2 storage is the lack of financial incentives [12]. Most strategies of CO2 storage could not generate profit, so the measures to decrease the cost of CO2 storage is very important and beneficial for the large-scale application of this technology. It is estimated that the cost of carbon capture and storage is dominated by the process of capture and gas separation, which costs $55 to $112 per ton of CO2 [12]. Therefore, co-injection of CO2 with impurities can be a cost-effective option for CO2 sequestration [275].
The impurities may be co-injected with CO2 including CH4, H2S, SO2, N2, and O2. Among them, H2S and CH4 are usually mixed with CO2 in produced acid gas, which can be used in CO2 storage. Other gases are the main components of flue gas, which is captured from the major CO2 emission sources such as power plants [276].
The impure CO2 can also be used for CCUS especially for CO2-EOR with multicontact miscible CO2 flooding [195]. In the multicontact miscible displacement procedure, the MMP is a key control variable due to it having a notable impact on the design and development of assets and being closely related to the economically feasibility, which is affected by the impurities in CO2 a lot. In general, the presence of H2S and SO2 in CO2 can reduce the MMP [195,277], while the presence of CH4, N2, and O2 can increase the MMP of CO2 [278,279], which is disadvantageous for the CO2-EOR operation and may increase the risks for the fracturing of the formation due to the higher injection pressure required.
The presence of impurities may change the other CO2 thermophysical properties and phase behavior [280], and affect the performance of CCS. N2 would lead to a delay for CO2 breakthrough when it is co-injected, because the solubility of CO2 in irreducible water is much higher than that of N2 [281]. However, the N2 would decrease the density of the dissolved phase and increase the risk in the long term [282]. Generally, the storage capacity of reservoirs decreases proportionally to the concentration and the compressibility factor of impurities when N2 is co-injected with CO2 [82]. For instance, the reduced storage capacity may be even higher than the volume fraction of impurities when O2 is included. However, the negative impact of impurities on capacity could be alleviated by storing the impure CO2 in a reservoir with high temperature [276].
There is no significant effect of H2S, with a fraction of less than 30%, as impurity on the dissolution of CO2 [275]. However, when it was co-injected with CO2 under the condition of 20 MPa and 45 °C, the H2S with a concentration over 20% has a potential to decrease the interfacial tension and increase the contact angle, leading to a low capillary force [283]. This means that H2S may increase the risks of gas leakage, which should receive attention. The impure CO2 with H2S can be trapped by hematite even in a dry system driven by the reduction of ferric iron in hematite by sulfide species, verifying the feasibility of co-injection of CO2 with H2S [284].
CH4 produced from an acid gas reservoir may also serve as impure gas and be used in CCS projects. There is no significant negative influence of CH4 on the interfacial tension and wettability even with the concentration of up to 20% [283]. The storage capacity of reservoirs also decreases proportionally to the concentration and compressibility factor of CH4 [82]. As the concentration of CH4 in injected CO2 is very low, its effect on the CCS is minor and can be neglected.
The presence of SO2 usually controls the acid-induced reactions with calcium-rich minerals when it is co-injected into reservoirs as an impurity, but the quantitative effect is very minor and could be negligible based on the German Bunter Sandstone [285]. Generally, the porosity in sandstone increases under the impact of SO2, while the porosity of the intermediate shale layer decreases because of the conversion of dominant calcite to anhydrite [286]. For instance, the conversion of Ca2+ bearing carbonate to anhydrite is observed when the SO2 was co-injected with CO2 into the German Bunter Sandstone [287]. A field study was conducted to investigate the geochemical impacts of SO2 and O2 as impurities on the reactions of minerals and fluids in a siliciclastic reservoir [288]. The CO2-saturated water with impurities was injected into reservoirs and allowed to interact with minerals for three weeks. The results showed that the pyrite dissolved due to the O2 acting as an oxidizing agent. However, the concentrations of SO2 and O2 are 67 ppm and 6150 ppm respectively, which is too low to lead to a significant effect on fluid–rock interaction. It could be inferred that the impact of impurities on the interaction with formation rock is highly dependent on the composition of minerals, which should be analyzed site specifically. It should be mentioned that the co-injection of SO2 and CO2 could suppress Joule–Thomson cooling, which is a beneficial thermal consequence for CCS [92].
In short, co-injection of CO2 with impurities is an effective strategy to reduce the cost of CO2 storage. However, the interaction of the impurities with formation rock, and the effect of impurities on thermophysical properties of reservoir fluids need to be further studied to reduce the uncertainties in CO2 storage process.

4.3. Prospects of CCS/CCUS Technologies

The economic factor for the CCS projects is believed to be one of the most important incentives for the industry. The price for CO2 emissions at the first major carbon market and also the biggest one, i.e., the European Union Emission Trading System, is approximately $7 per ton of CO2, which is much lower than the cost of CCS [12]. Therefore, there are no financial incentives of CCS for industries unless a higher price of carbon emission is set, demonstrating the important role that should be played by the government in the mitigation of CO2 emission.
Table 7 shows the large-scale CCS projects (more than 0.4 Mtpa) throughout the world from 1972 until the end of the 2020s. It can be seen that the CCUS for EOR and CCS in saline formations have made major contributions toward CO2 storage, which is in accordance with the prediction model formulated by Mac Dowell et al. [203]. Nearly half of 51 large-scale CCS projects scheduled by the end of the 2020s are designed for EOR, which shows its economic viability in EOR operations. In addition, 21 projects of CCS in saline formations are also planned, since its CO2 storage capacity may up to 4 Mtpa. Moreover, the average CO2 capture capacity of CCS in depleted gas fields is much greater than that of EOR and may reach to 2.8 Mtpa, showing the potential of the mitigating of CO2 emissions. However, on account of the great extent of the mixing between CO2 and CH4, which is an obstacle for enhancing additional recovery of CH4, there are only three large-scale CCS projects in depleted gas fields in the near future. With the development of technology to address this issue, CCS in depleted gas fields can play a more important role in CO2 storage.

5. Conclusions

The status of the strategies for CO2 storage has been discussed in view of assessing the security as well as improving the cost-effectiveness. In addition, the role of CCS technologies and their potential contribution on the mitigation of CO2 emissions in future are summarized. Based on the studies carried out in this review, the following conclusions have been obtained.
Firstly, the sequestration of CO2 in depleted oil and gas reservoirs could play an important role in reducing CO2 emissions in the near future, because the existing installed equipment and comprehensively characterized reservoir integrity will significantly reduce the cost of CCS. The leakage of CO2 through abandoned wells is an obstacle for the application of this technology. To address this issue, the long-term experiments and molecular dynamic simulations are needed to figure out the kinetics between CO2 with the well string, cement, as well as formation minerals under the relevant conditions. Secondly, if implemented on a large scale, CO2 storage in saline aquifers may make the biggest contribution in reducing CO2 emissions due to its huge storage capacity. Moreover, the scientifically proven technologies such as CO2 storage in coal beds, deep ocean, and deep-sea sediments are still immature technologies and do not appear to be capable of making a great contribution to the mitigation of CO2 emissions in the foreseeable future.
Another point is the need to investigate accurate risk assessment associated with CO2 storage and provide a guideline for the design of CCS projects. Attempting to make the CCS assessment more intelligential, the machine learning technology ought to be used.
It has also been demonstrated that the direct mineralization of flue gas by coal fly ash would be a promising technology result since the gas could be mineralized without separation. In addition, the pH swing mineralization may be another promising technology for the sequestration of CO2 due to the sustainability. The recyclable and cheap chemical reagents ought to be introduced into the mineralization process to make this technology more cost-effective.
Among the variety of CCUS strategies, CO2-EOR followed by CO2-EGR is supposed to play the most important role in the mitigation of CO2 in the next few years. The utilization of other strategies seems to be negligible in the near future. The co-injection of impurities with CO2 is an effective methodology to decrease the overall cost of CO2 storage. The physical and chemical effects of the impurities on reservoir fluids and formation rock should be studied site specific, to reduce the uncertainties in CO2 storage.
The government is supposed to play a major role in mitigating CO2 emission, a higher tax on CO2 emissions and financial subsidy on CO2 storage is encouraged to accelerate the deployment of CCS projects at a large-scale.

Author Contributions

Conceptualization, C.C., H.L. and J.L.; methodology, Z.H. and F.M.; data, C.C. and W.F.; writing—original draft preparation, C.C.; writing—review and editing, C.C., H.L., J.L. and F.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC), grant number 51809259; the CAS Pioneer Hundred Talents Program in China; the China Scholarship Council, grant number 201808080067, 201708080145.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Met Office. Our Changing World—Global Indicators. Available online: https://www.metoffice.gov.uk/climate-guide (accessed on 9 June 2019).
  2. Scripps CO2 Program. CO2 Concentration at Mauna Loa Observatory, Hawaii. Available online: http://scrippsco2.ucsd.edu (accessed on 9 June 2019).
  3. MacDowell, N.; Florin, N.; Buchard, A.; Hallett, J.; Galindo, A.; Jackson, G.; Adjiman, C.S.; Williams, C.K.; Shah, N.; Fennell, P. An overview of CO2 capture technologies. Energy Environ. Sci. 2010, 3, 1645–1669. [Google Scholar] [CrossRef] [Green Version]
  4. Rahman, F.A.; Aziz, M.M.A.; Saidur, R.; Bakar, W.A.W.A.; Hainin, M.R.; Putrajaya, R.; Hassan, N.A. Pollution to solution: Capture and sequestration of carbon dioxide (CO2) and its utilization as a renewable energy source for a sustainable future. Renew. Sustain. Energy Rev. 2017, 71, 112–126. [Google Scholar] [CrossRef]
  5. Hawcroft, M.; Walsh, E.; Hodges, K.; Zappa, G. Significantly increased extreme precipitation expected in Europe and North America from extratropical cyclones. Environ. Res. Lett. 2018, 13, 124006. [Google Scholar] [CrossRef]
  6. Edenhofer, O.; Pichs-Madruga, R.; Sokona, Y.; Kadner, S.; Minx, J.; Brunner, S. Change 2014: Mitigation of Climate Change; Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK and New York, USA, 2014. [Google Scholar]
  7. Brinckerhoff, P. Accelerating the Uptake of CCS: Industrial Use of Captured Carbon Dioxide; Global CCS Institute, 2011; Available online: https://www.globalccsinstitute.com/resources/publications-reports-research/?search=Accelerating+the+Uptake+of+CCS (accessed on 11 June 2019).
  8. IEA. Energy Technology Perspectives, Scenarios and Strategies to 2050. Available online: https://www.iea.org/publications/freepublications/publication/etp2010.pdf (accessed on 9 June 2019).
  9. IEA. Technology Roadmap Carbon Capture and Storage. Available online: https://insideclimatenews.org/sites/default/files/IEA-CCS%20Roadmap.pdf (accessed on 13 June 2019).
  10. GCCSI. Status CCS Project Database. Available online: https://co2re.co/FacilityData (accessed on 15 November 2019).
  11. Pawar, R.J.; Bromhal, G.S.; Carey, J.W.; Foxall, W.; Korre, A.; Ringrose, P.S.; Tucker, O.; Watson, M.N.; White, J.A. Recent advances in risk assessment and risk management of geologic CO2 storage. Int. J. Greenh. Gas Control 2015, 40, 292–311. [Google Scholar] [CrossRef] [Green Version]
  12. Gislason, S.R.; Oelkers, E.H. Carbon storage in basalt. Science 2014, 344, 373–374. [Google Scholar] [CrossRef]
  13. Atia, A.; Mohammedi, K. A review on the application of enhanced oil/gas recovery through CO2 sequestration. In Carbon Dioxide Chemistry, Capture and Oil Recovery, 1st ed.; Karamé, I., Shaya, J., Srour, H., Eds.; IntechOpen: London, UK, 2018; pp. 241–253. [Google Scholar]
  14. Koytsoumpa, E.I.; Bergins, C.; Kakaras, E. The CO2 economy: Review of CO2 capture and reuse technologies. J. Supercrit. Fluids 2018, 132, 3–16. [Google Scholar] [CrossRef]
  15. Li, L.; Zhao, N.; Wei, W.; Sun, Y. A review of research progress on CO2 capture, storage, and utilization in Chinese Academy of Sciences. Fuel 2013, 108, 112–130. [Google Scholar] [CrossRef]
  16. 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]
  17. Boot-Handford, M.E.; Abanades, J.C.; Anthony, E.J.; Blunt, M.J.; Brandani, S.; Mac Dowell, N.; Fernández, J.R.; Ferrari, M.-C.; Gross, R.; Hallett, J.P.; et al. Carbon capture and storage update. Energy Environ. Sci. 2014, 7, 130–189. [Google Scholar] [CrossRef]
  18. Godec, M.; Koperna, G.; Gale, J. CO2-ECBM: A review of its status and global potential. Energy Procedia 2014, 63, 5858–5869. [Google Scholar] [CrossRef] [Green Version]
  19. Liu, H.J.; Were, P.; Li, Q.; Gou, Y.; Hou, Z. Worldwide status of CCUS technologies and their development and challenges in China. Geofluids 2017, 1–25. [Google Scholar] [CrossRef]
  20. Pires, J.C.M.; Martins, F.G.; Alvim-Ferraz, M.C.M.; Simões, M. Recent developments on carbon capture and storage: An overview. Chem. Eng. Res. Des. 2011, 89, 1446–1460. [Google Scholar] [CrossRef]
  21. Aminu, M.D.; Nabavi, S.A.; Rochelle, C.A.; Manovic, V. A review of developments in carbon dioxide storage. Appl. Energy 2017, 208, 1389–1419. [Google Scholar] [CrossRef] [Green Version]
  22. Kemper, J. Biomass and carbon dioxide capture and storage: A review. Int. J. Greenh. Gas Control 2015, 40, 401–430. [Google Scholar] [CrossRef]
  23. Michael, K.; Golab, A.; Shulakova, V.; Ennis-King, J.; Allinson, G.; Sharma, S.; Aiken, T. Geological storage of CO2 in saline aquifers—A review of the experience from existing storage operations. Int. J. Greenh. Gas Control 2010, 4, 659–667. [Google Scholar] [CrossRef]
  24. Oh, T.H. Carbon capture and storage potential in coal-fired plant in Malaysia—A review. Renew. Sustain. Energy Rev. 2010, 14, 2697–2709. [Google Scholar] [CrossRef]
  25. Riaz, A.; Cinar, Y. Carbon dioxide sequestration in saline formations: Part I—Review of the modeling of solubility trapping. J. Petrol. Sci. Eng. 2014, 124, 367–380. [Google Scholar] [CrossRef]
  26. Sanna, A.; Uibu, M.; Caramanna, G.; Kuusik, R.; Maroto-Valer, M.M. A review of mineral carbonation technologies to sequester CO2. Chem. Soc. Rev. 2014, 43, 8049–8080. [Google Scholar] [CrossRef] [Green Version]
  27. Singh, P.; Haines, M. A review of existing carbon capture and storage cluster projects and future opportunities. Energy Procedia 2014, 63, 7247–7260. [Google Scholar] [CrossRef] [Green Version]
  28. Tang, Y.; Yang, R.; Bian, X. A review of CO2 sequestration projects and application in China. Sci. World J. 2014, 381854. [Google Scholar] [CrossRef] [Green Version]
  29. Verduyn, M.; Geerlings, H.; Mossel, G.V.; Vijayakumari, S. Review of the various CO2 mineralization product forms. Energy Procedia 2011, 4, 2885–2892. [Google Scholar] [CrossRef] [Green Version]
  30. Wee, J.-H. A review on carbon dioxide capture and storage technology using coal fly ash. Appl. Energy 2013, 106, 143–151. [Google Scholar] [CrossRef]
  31. Burnside, N.M.; Naylor, M. Review and implications of relative permeability of CO2/brine systems and residual trapping of CO2. Int. J. Greenh. Gas Control 2014, 23, 1–11. [Google Scholar] [CrossRef] [Green Version]
  32. De Silva, G.P.D.; Ranjith, P.G.; Perera, M.S.A. Geochemical aspects of CO2 sequestration in deep saline aquifers: A review. Fuel 2015, 155, 128–143. [Google Scholar] [CrossRef]
  33. Pan, P.; Wu, Z.; Feng, X.; Yan, F. Geomechanical modeling of CO2 geological storage: A review. J. Rock Mech. Geotech. Eng. 2016, 8, 936–947. [Google Scholar] [CrossRef]
  34. Abidoye, L.K.; Khudaida, K.J.; Das, D.B. Geological carbon sequestration in the context of two-phase flow in porous media: A review. Crit. Rev. Environ. Sci. Technol. 2015, 45, 1105–1147. [Google Scholar] [CrossRef] [Green Version]
  35. Abid, K.; Gholami, R.; Choate, P.; Nagaratnam, B.H. A review on cement degradation under CO2-rich environment of sequestration projects. J. Nat. Gas Sci. Eng. 2015, 27, 1149–1157. [Google Scholar] [CrossRef] [Green Version]
  36. Li, Q.; Liu, G. Risk assessment of the geological storage of CO2: A review. In Geologic Carbon Sequestration; Vishal, V., Singh, T.N., Eds.; Springer: Basel, Switzerland, 2016; pp. 249–284. [Google Scholar] [CrossRef]
  37. Mayer, B.; Humez, P.; Becker, V.; Dalkhaa, C.; Rock, L.; Myrttinen, A.; Barth, J.A.C. Assessing the usefulness of the isotopic composition of CO2 for leakage monitoring at CO2 storage sites: A review. Int. J. Greenh. Gas Control 2015, 37, 46–60. [Google Scholar] [CrossRef]
  38. Zhang, M.; Bachu, S. Review of integrity of existing wells in relation to CO2 geological storage: What do we know? Int. J. Greenh. Gas Control 2011, 5, 826–840. [Google Scholar] [CrossRef]
  39. Bai, M.; Zhang, Z.; Fu, X. A review on well integrity issues for CO2 geological storage and enhanced gas recovery. Renew. Sustain. Energy Rev. 2016, 59, 920–926. [Google Scholar] [CrossRef]
  40. Song, J.; Zhang, D. Comprehensive review of caprock-sealing mechanisms for geologic carbon sequestration. Environ. Sci. Technol. 2013, 47, 9–22. [Google Scholar] [CrossRef]
  41. Zahid, U.; Lim, Y.; Jung, J.; Han, C. CO2 geological storage: A review on present and future prospects. Korean J. Chem. Eng. 2011, 28, 674–685. [Google Scholar] [CrossRef]
  42. Shukla, R.; Ranjith, P.; Haque, A.; Choi, X. A review of studies on CO2 sequestration and caprock integrity. Fuel 2010, 89, 2651–2664. [Google Scholar] [CrossRef]
  43. Bachu, S. Review of CO2 storage efficiency in deep saline aquifers. Int. J. Greenh. Gas Control 2015, 40, 188–202. [Google Scholar] [CrossRef]
  44. Carroll, A.G.; Przeslawski, R.; Radke, L.C.; Black, J.R.; Picard, K.; Moreau, J.W.; Haese, R.R.; Nichol, S. Environmental considerations for subseabed geological storage of CO2: A review. Cont. Shelf Res. 2014, 83, 116–128. [Google Scholar] [CrossRef]
  45. Bradshaw, J.; Bachu, S.; Bonijoly, D.; Burruss, R.; Holloway, S.; Christensen, N.P.; Mathiassen, O.M. CO2 storage capacity estimation: Issues and development of standards. Int. J. Greenh. Gas Control 2007, 1, 62–68. [Google Scholar] [CrossRef] [Green Version]
  46. Krevor, S.; Blunt, M.J.; Benson, S.M.; Pentland, C.H.; Reynolds, C.; Al-Menhali, A.; Niu, B. Capillary trapping for geologic carbon dioxide storage—From pore scale physics to field scale implications. Int. J. Greenh. Gas Control 2015, 40, 221–237. [Google Scholar] [CrossRef] [Green Version]
  47. Zhang, Z.; Huisingh, D. Carbon dioxide storage schemes: Technology, assessment and deployment. J. Clean. Prod. 2017, 142, 1055–1064. [Google Scholar] [CrossRef]
  48. Bian, X.Q.; Xiong, W.; Kasthuriarachchi, D.T.K.; Liu, Y.B. Phase equilibrium modeling for carbon dioxide solubility in aqueous sodium chloride solutions using an association equation of state. Ind. Eng. Chem. Res. 2019, 58, 10570–10578. [Google Scholar] [CrossRef]
  49. Zhang, K.; Wu, Y.S.; Pruess, K. User’s Guide for TOUGH2-MP-a Massively Parallel Version of the TOUGH2 Code (No. LBNL-315E); Lawrence Berkeley National Laboratory: Berkeley, CA, USA, 2008. [Google Scholar]
  50. Sundal, A.; Hellevang, H.; Miri, R.; Dypvik, H.; Nystuen, J.P.; Aagaard, P. Variations in mineralization potential for CO2 related to sedimentary facies and burial depth—A comparative study from the North Sea. Energy Procedia 2014, 63, 5063–5070. [Google Scholar] [CrossRef] [Green Version]
  51. Zhao, X.; Liao, X.; Wang, W.; Chen, C.; Rui, Z.; Wang, H. The CO2 storage capacity evaluation: Methodology and determination of key factors. J. Energy Inst. 2014, 87, 297–305. [Google Scholar] [CrossRef]
  52. Metz, B.; Davidson, O.; De Coninck, H.; Loos, M.; Meyer, L. Carbon Dioxide Capture and Storage; IPCC Special Report. New York, NY, USA, 2005. Available online: https://www.researchgate.net/publication/239877190_IPCC_Special_Report_on_Carbon_dioxide_Capture_and_Storage (accessed on 15 June 2019).
  53. Celia, M.A.; Bachu, S.; Nordbotten, J.M.; Bandilla, K.W. Status of CO2 storage in deep saline aquifers with emphasis on modeling approaches and practical simulations. Water Resour. Res. 2015, 51, 6846–6892. [Google Scholar] [CrossRef]
  54. Davison, J.; Freund, P.; Smith, A. Putting Carbon Back in the Ground; IEA Greenhouse Gas R & D Programme. Available online: https://www.osti.gov/etdeweb/biblio/20204888 (accessed on 22 June 2019).
  55. Cooper, C. A technical basis for carbon dioxide storage. Energy Procedia 2009, 1, 1727–1733. [Google Scholar] [CrossRef] [Green Version]
  56. 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]
  57. Birkholzer, J.T.; Zhou, Q.; Tsang, C.F. Large-scale impact of CO2 storage in deep saline aquifers: A sensitivity study on pressure response in stratified systems. Int. J. Greenh. Gas Control 2009, 3, 181–194. [Google Scholar] [CrossRef] [Green Version]
  58. Audigane, P.; Gaus, I.; Pruess, K.; Xu, T. A long term 2D vertical modelling study of CO2 storage at Sleipner (North Sea) using TOUGHREACT. In Proceedings of theTOUGH Symposium, Berkeley, CA, USA, 15–17 May 2006. [Google Scholar]
  59. Audigane, P.; Gaus, I.; Czernichowski-Lauriol, I.; Pruess, K.; Xu, T. Two-dimensional reactive transport modeling of CO2 injection in a saline aquifer at the Sleipner site, North Sea. Am. J. Sci. 2007, 307, 974–1008. [Google Scholar] [CrossRef] [Green Version]
  60. Williams, G.; Chadwick, A. Chimneys and channels: History matching the growing CO2 plume at the Sleipner storage site. In Proceedings of the Fifth CO2 Geological Storage Workshop, Utrecht, The Netherlands, 21–23 November 2018. [Google Scholar]
  61. Hansen, O.; Gilding, D.; Nazarian, B.; Osdal, B.; Ringrose, P.; Kristoffersen, J.B.; Eiken, O.; Hansen, H. Snøhvit: The history of injecting and storing 1 Mt CO2 in the fluvial Tubåen Fm. Energy Procedia 2013, 37, 3565–3573. [Google Scholar] [CrossRef] [Green Version]
  62. Ringrose, P.S.; Mathieson, A.S.; Wright, I.W.; Selama, F.; Hansen, O.; Bissell, R.; Saoula, N.; Midgley, J. The In Salah CO2 storage project: Lessons learned and knowledge transfer. Energy Procedia 2013, 37, 6226–6236. [Google Scholar] [CrossRef] [Green Version]
  63. Rutqvist, J.; Vasco, D.W.; Myer, L. Coupled reservoir-geomechanical analysis of CO2 injection and ground deformations at In Salah, Algeria. Int. J. Greenh. Gas Control 2010, 4, 225–230. [Google Scholar] [CrossRef] [Green Version]
  64. Flett, M.A.; Beacher, G.J.; Brantjes, J.; Burt, A.J.; Dauth, C.; Koelmeyer, F.M.; Lawrence, R.; Leigh, S.; McKenna, J.; Gurton, R.; et al. Gorgon project: Subsurface evaluation of carbon dioxide disposal under Barrow Island. In Proceedings of the SPE Asia Pacific Oil and Gas Conference and Exhibition, Perth, Australia, 20–22 October 2008. [Google Scholar]
  65. Bourne, S.; Crouch, S.; Smith, M. A risk-based framework for measurement, monitoring and verification of the Quest CCS Project, Alberta, Canada. Int. J. Greenh. Gas Control 2014, 26, 109–126. [Google Scholar] [CrossRef]
  66. Arts, R.J.; Chadwick, A.; Eiken, O.; Thibeau, S.; Nooner, S. Ten years’ experience of monitoring CO2 injection in the Utsira Sand at Sleipner, offshore Norway. First Break 2008, 26, 65–72. [Google Scholar]
  67. Huang, X.; Bandilla, K.W.; Celia, M.A.; Bachu, S. Basin-scale modeling of CO2 storage using models of varying complexity. Int. J. Greenh. Gas Control 2014, 20, 73–86. [Google Scholar] [CrossRef]
  68. Bachu, S. Drainage and imbibition CO2/brine relative permeability curves at in situ conditions for sandstone formations in western Canada. Energy Procedia 2013, 37, 4428–4436. [Google Scholar] [CrossRef] [Green Version]
  69. Hermanrud, C.; Eiken, O.; Hansen, O.R.; Nordgaard Bolaas, H.M.; Simmenes, T.; Teige, G.M.G.; Hansen, H.; Johansen, S. Importance of pressure management in CO2 storage. In Proceedings of the Offshore Technology Conference, Houston, TX, USA, 6–9 May 2013. [Google Scholar]
  70. Bjørnarå, T.I.; Bohloli, B.; Park, J. Field-data analysis and hydromechanical modeling of CO2 storage at In Salah, Algeria. Int. J. Greenh. Gas Control 2018, 79, 61–72. [Google Scholar] [CrossRef]
  71. Eiken, O.; Ringrose, P.; Hermanrud, C.; Nazarian, B.; Torp, T.A.; Høier, L. Lessons learned from 14 years of CCS operations: Sleipner, In Salah and Snøhvit. Energy Procedia 2011, 4, 5541–5548. [Google Scholar] [CrossRef] [Green Version]
  72. Gemmer, L.; Hansen, O.; Iding, M.; Leary, S.; Ringrose, P. Geomechanical response to CO2 injection at Krechba, In salah, Algeria. First Break 2012, 30, 79–84. [Google Scholar]
  73. Newell, P.; Yoon, H.; Martinez, M.J.; Bishop, J.E.; Bryant, S.L. Investigation of the influence of geomechanical and hydrogeological properties on surface uplift at In Salah. J. Petrol. Sci. Eng. 2017, 155, 34–45. [Google Scholar] [CrossRef]
  74. Rinaldi, A.P.; Rutqvist, J. Modeling of deep fracture zone opening and transient ground surface uplift at KB-502 CO2 injection well, In Salah, Algeria. Int. J. Greenh. Gas Control 2013, 12, 155–167. [Google Scholar] [CrossRef]
  75. Rinaldi, A.P.; Rutqvist, J.; Finsterle, S.; Liu, H.H. Inverse modeling of ground surface uplift and pressure with iTOUGH-PEST and TOUGH-FLAC: The case of CO2 injection at In Salah, Algeria. Comput. Geosci. 2017, 108, 98–109. [Google Scholar] [CrossRef] [Green Version]
  76. Shi, J.Q.; Smith, J.; Durucan, S.; Korre, A. A coupled reservoir simulation-geomechanical modelling study of the CO2 injection-induced ground surface uplift observed at Krechba, in Salah. Energy Procedia 2013, 37, 3719–3726. [Google Scholar] [CrossRef] [Green Version]
  77. Stork, A.L.; Verdon, J.P.; Kendall, J.M. The microseismic response at the In Salah Carbon Capture and Storage (CCS) site. Int. J. Greenh. Gas Control 2015, 32, 159–171. [Google Scholar] [CrossRef] [Green Version]
  78. Martens, S.; Kempka, T.; Liebscher, A.; Lüth, S.; Möller, F.; Myrttinen, A.; Norden, B.; Schmidt-Hattenberger, C.; Zimmer, M.; Kühn, M. Europe’s longest-operating on-shore CO2 storage site at Ketzin, Germany: A progress report after three years of injection. Environ. Earth Sci. 2012, 67, 323–334. [Google Scholar] [CrossRef] [Green Version]
  79. Opedal, N. Ensuring integrity of CO2 storage: An overview of ongoing experimental activity. In Proceedings of the Fifth CO2 Geological Storage Workshop, Utrecht, The Netherlands, 21–23 November 2018. [Google Scholar]
  80. Finley, R.J.; Frailey, S.M.; Leetaru, H.E.; Senel, O.; Couëslan, M.L.; Scott, M. Early operational experience at a one-million tonne CCS demonstration project, Decatur, Illinois, USA. Energy Procedia 2013, 37, 6149–6155. [Google Scholar] [CrossRef] [Green Version]
  81. Yang, G.; Li, Y.; Atrens, A.; Liu, D.; Wang, Y.; Jia, L.; Lu, Y. Reactive transport modeling of long-term CO2 sequestration mechanisms at the Shenhua CCS demonstration project, China. J. Earth Sci. 2017, 28, 457–472. [Google Scholar] [CrossRef]
  82. 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]
  83. Mamora, D.D.; Seo, J.G. Enhanced gas recovery by carbon dioxide sequestration in depleted gas reservoirs. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 29 September–2 October 2002. [Google Scholar]
  84. Stein, M.H.; Ghotekar, A.L.; Avasthi, S.M. CO2 sequestration in a depleted gas field: A material balance study. In Proceedings of the SPE EUROPEC/EAGE Annual Conference and Exhibition, Barcelona, Spain, 14–17 June 2010. [Google Scholar]
  85. 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] [Green Version]
  86. Gilfillan, S.M.; Lollar, B.S.; Holland, G.; Blagburn, D.; Stevens, S.; Schoell, M.; Cassidy, M.; Ding, Z.; Zhou, Z.; Lacrampe-Couloume, G.; et al. Solubility trapping in formation water as dominant CO2 sink in natural gas fields. Nature 2009, 458, 614–618. [Google Scholar] [CrossRef] [Green Version]
  87. 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]
  88. Mathias, S.A.; Gluyas, J.G.; Oldenburg, C.M.; Tsang, C.F. Analytical solution for Joule–Thomson cooling during CO2 geo-sequestration in depleted oil and gas reservoirs. Int. J. Greenh. Gas Control 2010, 4, 806–810. [Google Scholar] [CrossRef] [Green Version]
  89. Oldenburg, C.M. Joule-Thomson cooling due to CO2 injection into natural gas reservoirs. Energy Convers. Manag. 2007, 48, 1808–1815. [Google Scholar] [CrossRef] [Green Version]
  90. Twerda, A.; Belfroid, S.; Neele, F. CO2 injection in low pressure depleted reservoirs. In Proceedings of the Fifth CO2 Geological Storage Workshop, Utrecht, The Netherlands, 21–23 November 2018. [Google Scholar]
  91. Böser, W.; Belfroid, S. Flow assurance study. Energy Procedia 2013, 37, 3018–3030. [Google Scholar] [CrossRef] [Green Version]
  92. 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]
  93. Loeve, D.; Hofstee, C.; Maas, J.G. Thermal effects in a depleted gas field by cold CO2 injection in the presence of methane. Energy Procedia 2014, 63, 5378–5393. [Google Scholar] [CrossRef] [Green Version]
  94. Sharma, S.; Cook, P.; Berly, T.; Lees, M. The CO2CRC Otway Project: Overcoming challenges from planning to execution of Australia’s first CCS project. Energy Procedia 2009, 1, 1965–1972. [Google Scholar] [CrossRef] [Green Version]
  95. Jenkins, C.R.; Cook, P.J.; Ennis-King, J.; Undershultz, J.; Boreham, C.; Dance, T.; de Caritat, P.; Etheridge, D.M.; Freifeld, B.M.; Hortle, A.; et al. Safe storage and effective monitoring of CO2 in depleted gas fields. Proc. Natl. Acad. Sci. USA 2012, 109, E35–41. [Google Scholar] [CrossRef] [Green Version]
  96. Arts, R.J.; Vandeweijer, V.P.; Hofstee, C.; Pluymaekers, M.P.D.; Loeve, D.; Kopp, A.; Plug, W.J. The feasibility of CO2 storage in the depleted P18-4 gas field offshore the Netherlands (the ROAD project). Int. J. Greenh. Gas Control 2012, 11, S10–S20. [Google Scholar] [CrossRef]
  97. Zhang, L.; Niu, B.; Ren, S.; Zhang, Y.; Yi, P.; Mi, H.; Ma, Y. Assessment of CO2 storage in DF1-1 gas field South China Sea for a CCS demonstration. J. Can. Pet. Technol. 2010, 49, 9–14. [Google Scholar] [CrossRef]
  98. 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]
  99. Bachu, S. Carbon dioxide storage capacity in uneconomic coal beds in Alberta, Canada: Methodology, potential and site identification. Int. J. Greenh. Gas Control 2007, 1, 374–385. [Google Scholar] [CrossRef]
  100. Yu, H.; Zhou, G.; Fan, W.; Ye, J. Predicted CO2 enhanced coalbed methane recovery and CO2 sequestration in China. Int. J. Coal Geol. 2007, 71, 345–357. [Google Scholar] [CrossRef]
  101. Brewer, P.G.; Friederich, G.; Peltzer, E.T.; Orr, F.M. Direct experiments on the ocean disposal of fossil fuel CO2. Science 1999, 284, 943–945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Fer, I.; Haugan, P.M. Dissolution from a liquid CO2 lake disposed in the deep ocean. Limnol. Oceanogr. 2003, 48, 872–883. [Google Scholar] [CrossRef] [Green Version]
  103. Levine, J.S.; Matter, J.M.; Goldberg, D.; Cook, A.; Lackner, K.S. Gravitational trapping of carbon dioxide in deep sea sediments: Permeability, buoyancy, and geomechanical analysis. Geophy. Res. Lett. 2007, 34. [Google Scholar] [CrossRef] [Green Version]
  104. House, K.Z.; Schrag, D.P.; Harvey, C.F.; Lackner, K.S. Permanent carbon dioxide storage in deep-sea sediments. Proc. Natl. Acad. Sci. USA 2006, 103, 12291–12295. [Google Scholar] [CrossRef] [Green Version]
  105. Koide, H.; Shindo, Y.; Tazaki, Y.; Iijima, M.; Ito, K.; Kimura, N.; Omata, K. Deep sub-seabed disposal of CO2—The most protective storage. Energy Convers. Manag. 1997, 38, S253–S258. [Google Scholar] [CrossRef]
  106. Schrag, D.P. Storage of carbon dioxide in offshore sediments. Science 2009, 325, 1658–1659. [Google Scholar] [CrossRef] [Green Version]
  107. Teng, Y.; Zhang, D. Long-term viability of carbon sequestration in deep-sea sediments. Sci. Adv. 2018, 4, eaao6588. [Google Scholar] [CrossRef] [Green Version]
  108. Adams, E.E.; Caldeira, K. Ocean Storage of CO2. Elements 2008, 4, 319–324. [Google Scholar] [CrossRef]
  109. 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] [Green Version]
  110. Gawel, K.; Todorovic, J.; Liebscher, A.; Wiese, B.; Opedal, N. Study of materials retrieved from a Ketzin CO2 monitoring well. Energy Procedia 2017, 114, 5799–5815. [Google Scholar] [CrossRef]
  111. Celia, M.A.; Nordbotten, J.M.; Court, B.; Dobossy, M.; Bachu, S. Field-scale application of a semi-analytical model for estimation of CO2 and brine leakage along old wells. Int. J. Greenh. Gas Control 2011, 5, 257–269. [Google Scholar] [CrossRef]
  112. Nordbotten, J.M.; Celia, M.A.; Bachu, S. Injection and storage of CO2 in deep saline aquifers: Analytical solution for CO2 plume evolution during injection. Transp. Porous Media 2005, 58, 339–360. [Google Scholar] [CrossRef]
  113. Xu, Z.; Fang, Y.; Scheibe, T.D.; Bonneville, A. A fluid pressure and deformation analysis for geological sequestration of carbon dioxide. Comput. Geosci. 2012, 46, 31–37. [Google Scholar] [CrossRef]
  114. Wang Huimin, W.J.G. A fully coupled mathematical model with geochemical reaction for caprock sealing efficiency in CO2 geosequestration. In Proceedings of the 15th China Rock Mechanics and Engineering Academic Annual Meeting, Beijing, China, 19–22 September 2018. [Google Scholar]
  115. Bao, J.; Chu, Y.; Xu, Z.; Tartakovsky, A.M.; Fang, Y. Uncertainty quantification for the impact of injection rate fluctuation on the geomechanical response of geological carbon sequestration. Int. J. Greenh. Gas Control 2014, 20, 160–167. [Google Scholar] [CrossRef]
  116. Lengler, U.; De Lucia, M.; Kühn, M. The impact of heterogeneity on the distribution of CO2: Numerical simulation of CO2 storage at Ketzin. Int. J. Greenh. Gas Control 2010, 4, 1016–1025. [Google Scholar] [CrossRef]
  117. Wasch, L.J.; Wollenweber, J.; Tambach, T.J. Intentional salt clogging: A novel concept for long-term CO2 sealing. Greenh. Gases Sci. Technol. 2013, 3, 491–502. [Google Scholar] [CrossRef]
  118. Iogna, A.; Guillet-Lhermite, J.; Wood, C.; Deflandre, J.P. CO2 storage and enhanced gas recovery: Using extended black oil modelling to simulate CO2 injection on a North Sea depleted gas field. In Proceedings of the SPE Europec featured at 79th EAGE Conference and Exhibition, Paris, France, 12–15 June 2017. [Google Scholar]
  119. Rafiee, M.M.; Ramazanian, M. Simulation study of enhanced gas recovery process using a compositional and a black oil simulator. In Proceedings of the SPE Enhanced Oil Recovery Conference, Kuala Lumpur, Malaysia, 19–21 July 2011. [Google Scholar]
  120. Jia, W.; McPherson, B.J.; Pan, F.; Xiao, T.; Bromhal, G. Probabilistic analysis of CO2 storage mechanisms in a CO2-EOR field using polynomial chaos expansion. Int. J. Greenh. Gas Control 2016, 51, 218–229. [Google Scholar] [CrossRef] [Green Version]
  121. Mishra, S.; Ganesh, P.R.; Schuetter, J. Developing and validating simplified predictive models for CO2 geologic sequestration. Energy Procedia 2017, 114, 3456–3464. [Google Scholar] [CrossRef]
  122. Ren, B. Local capillary trapping in carbon sequestration: Parametric study and implications for leakage assessment. Int. J. Greenh. Gas Control 2018, 78, 135–147. [Google Scholar] [CrossRef]
  123. Wriedt, J.; Deo, M.; Han, W.S.; Lepinski, J. A methodology for quantifying risk and likelihood of failure for carbon dioxide injection into deep saline reservoirs. Int. J. Greenh. Gas Control 2014, 20, 196–211. [Google Scholar] [CrossRef]
  124. Bao, J.; Hou, Z.; Fang, Y.; Ren, H.; Lin, G. Uncertainty quantification for evaluating impacts of caprock and reservoir properties on pressure buildup and ground surface displacement during geological CO2sequestration. Greenh. Gases Sci. Technol. 2013, 3, 338–358. [Google Scholar] [CrossRef]
  125. Hou, Z.; Bacon, D.H.; Engel, D.W.; Lin, G.; Fang, Y.; Ren, H.; Fang, Z. Uncertainty analyses of CO2 plume expansion subsequent to wellbore CO2 leakage into aquifers. Int. J. Greenh. Gas Control 2014, 27, 69–80. [Google Scholar] [CrossRef]
  126. Allen, R.; Nilsen, H.M.; Lie, K.A.; Møyner, O.; Andersen, O. Using simplified methods to explore the impact of parameter uncertainty on CO2 storage estimates with application to the Norwegian Continental Shelf. Int. J. Greenh. Gas Control 2018, 75, 198–213. [Google Scholar] [CrossRef]
  127. Schepers, K.C.; Nuttall, B.C.; Oudinot, A.Y.; Gonzalez, R.J. Reservoir modeling and simulation of the Devonian gas shale of eastern Kentucky for enhanced gas recovery and CO2 storage. In Proceedings of the SPE International Conference on CO2 Capture, Storage, and Utilization, San Diego, CA, USA, 10–11 November 2009. [Google Scholar]
  128. Jung, H.; Singh, G.; Espinoza, D.N.; Wheeler, M.F. Quantification of a maximum injection volume of CO2 to avert geomechanical perturbations using a compositional fluid flow reservoir simulator. Adv. Water Resour. 2018, 112, 160–169. [Google Scholar] [CrossRef]
  129. Ebigbo, A.; Class, H.; Helmig, R. CO2 leakage through an abandoned well: Problem-oriented benchmarks. Comput. Geosci. 2006, 11, 103–115. [Google Scholar] [CrossRef]
  130. Luo, F.; Xu, R.N.; Jiang, P.X. Numerical investigation of the influence of vertical permeability heterogeneity in stratified formation and of injection/production well perforation placement on CO2 geological storage with enhanced CH4 recovery. Appl. Energy 2013, 102, 1314–1323. [Google Scholar] [CrossRef]
  131. Khan, C.; Amin, R.; Madden, G. Economic modelling of CO2 injection for enhanced gas recovery and storage: A reservoir simulation study of operational parameters. Energy Environ. Res. 2012, 2. [Google Scholar] [CrossRef] [Green Version]
  132. Khan, C.; Amin, R.; Madden, G. Carbon dioxide injection for enhanced gas recovery and storage (reservoir simulation). Egyp. J. Pet. 2013, 22, 225–240. [Google Scholar] [CrossRef] [Green Version]
  133. Rinaldi, A.P.; Rutqvist, J. Modeling ground surface uplift during CO2 sequestration: The case of in Salah, Algeria. Energy Procedia 2017, 114, 3247–3256. [Google Scholar] [CrossRef]
  134. Rutqvist, J. Status of the TOUGH-FLAC simulator and recent applications related to coupled fluid flow and crustal deformations. Comput. Geosci. 2011, 37, 739–750. [Google Scholar] [CrossRef] [Green Version]
  135. Pan, P.Z.; Rutqvist, J.; Feng, X.T.; Yan, F. An approach for modeling rock discontinuous mechanical behavior under multiphase fluid flow conditions. Rock Mech. Rock Eng. 2014, 47, 589–603. [Google Scholar] [CrossRef]
  136. Liu, H.H.; Houseworth, J.; Rutqvist, J.; Zheng, L.; Asahina, D.; Li, L.; Vilarrasa, V.; Chen, F.; Nakagawa, S.; Finsterle, S.; et al. Report on THMC Modeling of the Near Field Evolution of a Generic Clay Repository: Model Validation and Demonstration; Lawrence Berkeley National Laboratory: Berkeley, CA, USA, 2013. [Google Scholar]
  137. Benisch, K.; Graupner, B.; Bauer, S. The coupled OpenGeoSys-eclipse simulator for simulation of CO2 storage—Code comparison for fluid flow and geomechanical processes. Energy Procedia 2013, 37, 3663–3671. [Google Scholar] [CrossRef] [Green Version]
  138. Fei, W.B.; Li, Q.; Liu, X.H.; Wei, X.C.; Jing, M.; Song, R.R.; Li, X.C.; Wang, Y.S. Coupled analysis for interaction of coal mining and CO2 geological storage in Ordos Basin, China. In Proceedings of the 8th Asian Rock Mechanics Symposium, Sapporo, Japan, 14–16 October 2014. [Google Scholar]
  139. Gou, Y.; Hou, Z.; Liu, H.; Zhou, L.; Were, P. Numerical simulation of carbon dioxide injection for enhanced gas recovery (CO2-EGR) in Altmark natural gas field. Acta Geotechnica 2014, 9, 49–58. [Google Scholar] [CrossRef]
  140. Rutqvist, J.; Tsang, C.F. TOUGH-FLAC: A numerical simulator for analysis of coupled thermal-hydrologic-mechanical processes in fractured and porous geological media under multi-phase flow conditions. In Proceedings of the TOUGH Symposium 2003, Berkeley, CA, USA, 12–14 May 2003. [Google Scholar]
  141. Rutqvist, J.; Tsang, C.F. A study of caprock hydromechanical changes associated with CO2-injection into a brine formation. Environ. Geol. 2002, 42, 296–305. [Google Scholar] [CrossRef]
  142. Pawar, R.J.; Bromhal, G.S.; Chu, S.; Dilmore, R.M.; Oldenburg, C.M.; Stauffer, P.H.; Zhang, Y.; Guthrie, G.D. The National Risk Assessment Partnership’s integrated assessment model for carbon storage: A tool to support decision making amidst uncertainty. Int. J. Greenh. Gas Control 2016, 52, 175–189. [Google Scholar] [CrossRef] [Green Version]
  143. Namhata, A.; Zhang, L.; Dilmore, R.M.; Oladyshkin, S.; Nakles, D.V. Modeling changes in pressure due to migration of fluids into the Above Zone Monitoring Interval of a geologic carbon storage site. Int. J. Greenh. Gas Control 2017, 56, 30–42. [Google Scholar] [CrossRef] [Green Version]
  144. NETL. U.S. DOE’s National Risk Assessment Partnership: Assessing Carbon Storage Risk Performance to Support Decision Making Amidst Uncertainty. Available online: https://www.cslforum.org/cslf/sites/default/files/documents/AbuDhabi2017/Bromhal-NRAP-TG-AbuDhabi0517.pdf (accessed on 12 May 2019).
  145. Doherty, B.; Vasylkivska, V.; Huerta, N.J.; Dilmore, R. Estimating the leakage along wells during geologic CO2 Storage: Application of the Well Leakage Assessment Tool to a hypothetical storage scenario in Natrona County, Wyoming. Energy Procedia 2017, 114, 5151–5172. [Google Scholar] [CrossRef]
  146. Sun, A.Y.; Jeong, H.; González-Nicolás, A.; Templeton, T.C. Metamodeling-based approach for risk assessment and cost estimation: Application to geological carbon sequestration planning. Comput. Geosci. 2018, 113, 70–80. [Google Scholar] [CrossRef]
  147. Dramsch, J.S.; Corte, G.; Amini, H.; Lüthje, M.; MacBeth, C. Deep learning application for 4D pressure saturation inversion compared to Bayesian inversion on North Sea data. In Proceedings of the Second EAGE Workshop Practical Reservoir Monitoring, Amsterdam, The Netherlands, 1–4 April 2019. [Google Scholar]
  148. Leung, D.Y.C.; Caramanna, G.; Maroto-Valer, M.M. An overview of current status of carbon dioxide capture and storage technologies. Renew. Sustain. Energy Rev. 2014, 39, 426–443. [Google Scholar] [CrossRef] [Green Version]
  149. Nordbotten, J.M.; Flemisch, B.; Gasda, S.E.; Nilsen, H.M.; Fan, Y.; Pickup, G.E.; Wiese, B.; Celia, M.A.; Dahle, H.K.; Eigestad, G.T.; et al. Uncertainties in practical simulation of CO2 storage. Int. J. Greenh. Gas Control 2012, 9, 234–242. [Google Scholar] [CrossRef]
  150. Lumley, D. 4D seismic monitoring of CO2 sequestration. Lead. Edge 2010, 29, 150–155. [Google Scholar] [CrossRef]
  151. Oye, V.; Aker, E.; Daley, T.M.; Kühn, D.; Bohloli, B.; Korneev, V. Microseismic monitoring and interpretation of injection data from the in Salah CO2 storage site (Krechba), Algeria. Energy Procedia 2013, 37, 4191–4198. [Google Scholar] [CrossRef] [Green Version]
  152. Götz, J.; Lüth, S.; Henninges, J.; Reinsch, T. Vertical seismic profiling using a daisy-chained deployment of fibre-optic cables in four wells simultaneously—Case study at the Ketzin carbon dioxide storage site. Geophys. Prospect. 2018, 66, 1201–1214. [Google Scholar] [CrossRef]
  153. Kabirzadeh, H.; Sideris, M.G.; Shin, Y.J.; Kim, J.W. Gravimetric Monitoring of confined and unconfined geological CO2 reservoirs. Energy Procedia 2017, 114, 3961–3968. [Google Scholar] [CrossRef]
  154. Böhm, G.; Carcione, J.M.; Gei, D.; Picotti, S.; Michelini, A. Cross-well seismic and electromagnetic tomography for CO2 detection and monitoring in a saline aquifer. J. Petrol. Sci. Eng. 2015, 133, 245–257. [Google Scholar] [CrossRef]
  155. Carcione, J.M.; Gei, D.; Picotti, S.; Michelini, A. Cross-hole electromagnetic and seismic modeling for CO2 detection and monitoring in a saline aquifer. J. Petrol. Sci. Eng. 2012, 100, 162–172. [Google Scholar] [CrossRef]
  156. Liebscher, A.; Möller, F.; Bannach, A.; Köhler, S.; Wiebach, J.; Schmidt-Hattenberger, C.; Weiner, M.; Pretschner, C.; Ebert, K.; Zemke, J. Injection operation and operational pressure–temperature monitoring at the CO2 storage pilot site Ketzin, Germany—Design, results, recommendations. Int. J. Greenh. Gas Control 2013, 15, 163–173. [Google Scholar] [CrossRef] [Green Version]
  157. Boreham, C.; Underschultz, J.; Stalker, L.; Kirste, D.; Freifeld, B.; Jenkins, C.; Ennis-King, J. Monitoring of CO2 storage in a depleted natural gas reservoir: Gas geochemistry from the CO2CRC Otway Project, Australia. Int. J. Greenh. Gas Control 2011, 5, 1039–1054. [Google Scholar] [CrossRef]
  158. Gaus, I. Role and impact of CO2–rock interactions during CO2 storage in sedimentary rocks. Int. J. Greenh. Gas Control 2010, 4, 73–89. [Google Scholar] [CrossRef]
  159. Matter, J.M.; Stute, M.; Snæbjörnsdottir, S.Ó.; Oelkers, E.H.; Gislason, S.R.; Aradottir, E.S.; Sigfusson, B.; Gunnarsson, I.; Sigurdardottir, H.; Gunnlaugsson, E.; et al. Rapid carbon mineralization for permanent disposal of anthropogenic carbon dioxide emissions. Science 2016, 352, 1312–1314. [Google Scholar] [CrossRef] [Green Version]
  160. Etheridge, D.; Luhar, A.; Loh, Z.; Leuning, R.; Spencer, D.; Steele, P.; Zegelin, S.; Allison, C.; Krummel, P.; Leist, M.; et al. Atmospheric monitoring of the CO2CRC Otway Project and lessons for large scale CO2 storage projects. Energy Procedia 2011, 4, 3666–3675. [Google Scholar] [CrossRef] [Green Version]
  161. Morozova, D.; Zettlitzer, M.; Let, D.; Würdemann, H. Monitoring of the microbial community composition in deep subsurface saline aquifers during CO2 storage in Ketzin, Germany. Energy Procedia 2011, 4, 4362–4370. [Google Scholar] [CrossRef] [Green Version]
  162. Schilling, F.; Borm, G.; Würdemann, H.; Möller, F.; Kühn, M. Status report on the first European on-shore CO2 storage site at Ketzin (Germany). Energy Procedia 2009, 1, 2029–2035. [Google Scholar] [CrossRef] [Green Version]
  163. Onuma, T.; Okada, K.; Otsubo, A. Time series analysis of surface deformation related with CO2 injection by satellite-borne SAR interferometry at In Salah, Algeria. Energy Procedia 2011, 4, 3428–3434. [Google Scholar] [CrossRef] [Green Version]
  164. Mawalkar, S.; Brock, D.; Burchwell, A.; Kelley, M.; Mishra, S.; Gupta, N.; Pardini, R.; Shroyer, B. Where is that CO2 flowing? Using Distributed Temperature Sensing (DTS) technology for monitoring injection of CO2 into a depleted oil reservoir. Int. J. Greenh. Gas Control 2019, 85, 132–142. [Google Scholar] [CrossRef] [Green Version]
  165. Shatarah, I.S.; lbrycht, R. Distributed temperature sensing in optical fibers based on Raman scattering: Theory and applications. Meas. Autom. Monit. 2017, 63, 41–44. [Google Scholar]
  166. Worthen, A.J.; Parikh, P.S.; Chen, Y.; Bryant, S.L.; Huh, C.; Johnston, K.P. Carbon dioxide-in-water foams stabilized with a mixture of nanoparticles and surfactant for CO2 storage and utilization applications. Energy Procedia 2014, 63, 7929–7938. [Google Scholar] [CrossRef] [Green Version]
  167. Rognmo, A.U.; Heldal, S.; Fernø, M.A. Silica nanoparticles to stabilize CO2-foam for improved CO2 utilization: Enhanced CO2 storage and oil recovery from mature oil reservoirs. Fuel 2018, 216, 621–626. [Google Scholar] [CrossRef]
  168. Guo, F.; Aryana, S.A. Improved sweep efficiency due to foam flooding in a heterogeneous microfluidic device. J. Petrol. Sci. Eng. 2018, 164, 155–163. [Google Scholar] [CrossRef]
  169. Chou, S.I.; Vasicek, S.L.; Pisio, D.L.; Jasek, D.E.; Goodgame, J.A. CO2 foam field trial at north Ward-Estes. In Proceedings of the SPE Annual Technical Conference and Exhibition, Washington, DC, USA, 4–7 October 1992. [Google Scholar]
  170. Guo, F.; Aryana, S.A.; Wang, Y.; McLaughlin, J.F.; Coddington, K. Enhancement of storage capacity of CO2 in megaporous saline aquifers using nanoparticle-stabilized CO2 foam. Int. J. Greenh. Gas Control 2019, 87, 134–141. [Google Scholar] [CrossRef]
  171. Pizzocolo, F.; Peters, E.; Loeve, D.; Hewson, C.W.; Wasch, L.; Brunner, L.J. Feasibility of novel techniques to mitigate or remedy CO2 leakage. In Proceedings of the SPE Europec featured at 79th EAGE Conference and Exhibition, Paris, France, 12–15 June 2017. [Google Scholar]
  172. Johnston, K.P.; Rocha, S.R.P.d. Colloids in supercritical fluids over the last 20 years and future directions. J. Supercrit. Fluids 2009, 47, 523–530. [Google Scholar] [CrossRef]
  173. Worthen, A.J.; Bagaria, H.G.; Chen, Y.; Bryant, S.L.; Huh, C.; Johnston, K.P. Nanoparticle-stabilized carbon dioxide-in-water foams with fine texture. J. Colloid Interface Sci. 2013, 391, 142–151. [Google Scholar] [CrossRef]
  174. Worthen, A.J.; Bryant, S.L.; Huh, C.; Johnston, K.P. Carbon dioxide-in-water foams stabilized with nanoparticles and surfactant acting in synergy. AIChE J. 2013, 59, 3490–3501. [Google Scholar] [CrossRef]
  175. Anchliya, A.; Ehlig-Economides, C.A.; Jafarpour, B. Aquifer management to accelerate CO2 dissolution and trapping. SPE J. 2012, 17, 805–816. [Google Scholar] [CrossRef]
  176. Cameron, D.A.; Durlofsky, L.J. Optimization of well placement, CO2 injection rates, and brine cycling for geological carbon sequestration. Int. J. Greenh. Gas Control 2012, 10, 100–112. [Google Scholar] [CrossRef]
  177. Zhang, Z.; Agarwal, R.K. Numerical simulation and optimization of CO2 sequestration in saline aquifers for vertical and horizontal well injection. Comput. Geosci. 2012, 16, 891–899. [Google Scholar] [CrossRef]
  178. Zhang, Z.; Agarwal, R. Numerical simulation and optimization of CO2 sequestration in saline aquifers. Comput. Fluids 2013, 80, 79–87. [Google Scholar] [CrossRef] [Green Version]
  179. Harris, J.; Kovscek, A.R.; Orr, F.M.; Zoback, M.D. Geologic Storage of CO2 in Coal Beds; Global Climate and Energy Project (GCEP) Technical Report; Stanford University: Stanford, CA, USA, 2009. [Google Scholar]
  180. Tanaka, K.; Vilcáez, J.; Sato, K. Improvement of CO2 geological storage efficiency by injection and production well design. Energy Procedia 2013, 37, 4591–4597. [Google Scholar] [CrossRef] [Green Version]
  181. Kelemen, P.B.; Matter, J. In situ carbonation of peridotite for CO2 storage. Proc. Natl. Acad. Sci. USA 2008, 105, 17295–17300. [Google Scholar] [CrossRef] [Green Version]
  182. Reddy, K.J.; John, S.; Weber, H.; Argyle, M.D.; Bhattacharyya, P.; Taylor, D.T.; Christensen, M.; Foulke, T.; Fahlsing, P. Simultaneous capture and mineralization of coal combustion flue gas carbon dioxide (CO2). Energy Procedia 2011, 4, 1574–1583. [Google Scholar] [CrossRef] [Green Version]
  183. Wang, X.; Maroto-Valer, M.M. Optimization of carbon dioxide capture and storage with mineralisation using recyclable ammonium salts. Energy 2013, 51, 431–438. [Google Scholar] [CrossRef]
  184. Schaef, H.T.; McGrail, B.P.; Owen, A.T. Carbonate mineralization of volcanic province basalts. Int. J. Greenh. Gas Control 2010, 4, 249–261. [Google Scholar] [CrossRef]
  185. Goldberg, D.S.; Takahashi, T.; Slagle, A.L. Carbon dioxide sequestration in deep-sea basalt. Proc. Natl. Acad. Sci. USA 2008, 105, 9920–9925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. O’Connor, W.K.; Dahlin, D.C.; Nilsen, D.N.; Gerdemann, S.J.; Rush, G.E.; Walters, R.P.; Turner, P.C. Research status on the sequestration of carbon dioxide by direct aqueous mineral carbonation. In Proceedings of the 18th Annual International Pittsburgh Coal Conference, Newcastle, Australia, 3–7 December 2001. [Google Scholar]
  187. Teir, S.; Eloneva, S.; Fogelholm, C.J.; Zevenhoven, R. Fixation of carbon dioxide by producing hydromagnesite from serpentinite. Appl. Energy 2009, 86, 214–218. [Google Scholar] [CrossRef]
  188. Park, A.H.A.; Fan, L.S. CO2 mineral sequestration: Physically activated dissolution of serpentine and pH swing process. Chem. Eng. Sci. 2004, 59, 5241–5247. [Google Scholar] [CrossRef]
  189. Kodama, S.; Nishimoto, T.; Yamamoto, N.; Yogo, K.; Yamada, K. Development of a new pH-swing CO2 mineralization process with a recyclable reaction solution. Energy 2008, 33, 776–784. [Google Scholar] [CrossRef]
  190. Sanna, A.; Dri, M.; Maroto-Valer, M. Carbon dioxide capture and storage by pH swing aqueous mineralisation using a mixture of ammonium salts and antigorite source. Fuel 2013, 114, 153–161. [Google Scholar] [CrossRef]
  191. Wei, N.; Li, X.; Fang, Z.; Bai, B.; Li, Q.; Liu, S.; Jia, Y. Regional resource distribution of onshore carbon geological utilization in China. J. CO2 Util. 2015, 11, 20–30. [Google Scholar] [CrossRef]
  192. Burton, E.; Beyer, J.; Bourcier, W.; Mateer, N.; Reed, J. Carbon utilization to meet California’s climate change goals. Energy Procedia 2013, 37, 6979–6986. [Google Scholar] [CrossRef] [Green Version]
  193. Azzolina, N.A.; Peck, W.D.; Hamling, J.A.; Gorecki, C.D.; Ayash, S.C.; Doll, T.E.; Nakles, D.V.; Melzer, L.S. How green is my oil? A detailed look at greenhouse gas accounting for CO2—nhanced oil recovery (CO2-EOR) sites. Int. J. Greenh. Gas Control 2016, 51, 369–379. [Google Scholar] [CrossRef]
  194. Bian, X.Q.; Han, B.; Du, Z.M.; Jaubert, J.N.; Li, M.J. Integrating support vector regression with genetic algorithm for CO2-oil minimum miscibility pressure (MMP) in pure and impure CO2 streams. Fuel 2016, 182, 550–557. [Google Scholar] [CrossRef]
  195. Metcalfe, R.S. Effects of impurities on minimum miscibility pressures and minimum enrichment levels for CO2 and rich-gas displacements. SPE J. 1982, 22, 219–225. [Google Scholar] [CrossRef]
  196. Brush, R.M.; Davitt, H.J.; Aimar, O.B.; Arguello, J.; Whiteside, J.M. Immiscible CO2 flooding for increased oil recovery and reduced emissions. In Proceedings of the SPE/DOE Improved Oil Recovery Symposium, Tulsa, OK, USA, 3–5 April 2000. [Google Scholar]
  197. Sahin, S.; Kalfa, U.; Celebioglu, D. Bati Raman field immiscible CO2 application-status quo and future plans. In Proceedings of the Latin American and Caribbean Petroleum Engineering Conference, Buenos Aires, Argentina, 15–18 April 2007. [Google Scholar]
  198. Aryana, S.A.; Barclay, C.; Liu, S. North cross devonian unit - a mature continuous CO2 flood beyond 200% HCPV injection. In Proceedings of the SPE Annual Technical Conference and Exhibition, Amsterdam, The Netherlands, 27–29 October 2014. [Google Scholar]
  199. Jia, B.; Tsau, J.S.; Barati, R. A review of the current progress of CO2 injection EOR and carbon storage in shale oil reservoirs. Fuel 2019, 236, 404–427. [Google Scholar] [CrossRef]
  200. Chen, B.; Reynolds, A.C. Optimal control of ICV’s and well operating conditions for the water-alternating-gas injection process. J. Petrol. Sci. Eng. 2017, 149, 623–640. [Google Scholar] [CrossRef]
  201. Hawthorne, S.B.; Gorecki, C.D.; Sorensen, J.A.; Steadman, E.N.; Harju, J.A.; Melzer, S. Hydrocarbon mobilization mechanisms from upper, middle, and lower Bakken reservoir rocks exposed to CO2. In Proceedings of the SPE Unconventional Resources Conference Canada, Calgary, AB, Canada, 5–7 November 2013. [Google Scholar]
  202. Zuloaga, P.; Yu, W.; Miao, J.; Sepehrnoori, K. Performance evaluation of CO2 huff-n-puff and continuous CO2 injection in tight oil reservoirs. Energy 2017, 134, 181–192. [Google Scholar] [CrossRef]
  203. Mac Dowell, N.; Fennell, P.S.; Shah, N.; Maitland, G.C. The role of CO2 capture and utilization in mitigating climate change. Nat. Clim. Chang. 2017, 7, 243–249. [Google Scholar] [CrossRef] [Green Version]
  204. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon capture and storage (CCS): The way forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef] [Green Version]
  205. Carpenter, S.M.; Koperna, G. Development of the first internationally accepted standard for geologic storage of carbon dioxide utilizing Enhanced Oil Recovery (EOR) under the International Standards Organization (ISO) Technical Committee TC-265. Energy Procedia 2014, 63, 6717–6729. [Google Scholar] [CrossRef] [Green Version]
  206. Kane, A.V. Performance Review of a large-scale CO2-WAG enhanced recovery project, SACROC Unit Kelly-Snyder field. J. Petrol. Technol. 1979, 31, 217–231. [Google Scholar] [CrossRef]
  207. Sanders, A.W.; Jones, R.M.; Linroth, M.A.; Nguyen, Q.P. Implementation of a CO2 foam pilot study in the SACROC field: Performance evaluation. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 8–10 October 2012. [Google Scholar]
  208. Langston, M.V.; Hoadley, S.F.; Young, D.N. Definitive CO2 flooding response in the SACROC unit. In Proceedings of the SPE/DOE Enhanced Oil Recovery Symposium, Tulsa, OK, USA, 16–21 April 1988. [Google Scholar]
  209. Gozalpour, F.; Ren, S.R.; Tohidi, B. CO2 EOR and storage in oil reservoir. Oil Gas Sci. Technol. 2006, 60, 537–546. [Google Scholar] [CrossRef] [Green Version]
  210. Marston, P.M. Incidentally speaking: A systematic assessment and comparison of incidental storage of CO2 during EOR with other Near-term storage options. Energy Procedia 2017, 114, 7422–7430. [Google Scholar] [CrossRef]
  211. Clemens, T.; Secklehner, S.; Mantatzis, K.; Jacobs, B. Enhanced gas recovery-challenges shown at the example of three gas fields. In Proceedings of the SPE EUROPEC/EAGE Annual Conference and Exhibition, Barcelona, Spain, 14–17 June 2010. [Google Scholar]
  212. Hattenbach, R.P.; Wilson, M.; Brown, K.R. Capture of carbon dioxide from coal combustion and its utilization for enhanced oil recovery. In Proceedings of the GHGT-4 Conference, Interlaken, Switzerland, 30 August–2 September 1998. [Google Scholar]
  213. Preston, C.; Monea, M.; Jazrawi, W.; Brown, K.; Whittaker, S.; White, D.; Law, D.; Chalaturnyk, R.; Rostron, B. IEA GHG Weyburn CO2 monitoring and storage project. Fuel Process. Technol. 2005, 86, 1547–1568. [Google Scholar] [CrossRef]
  214. Lui, L.C.; Leamon, G. Developments towards environmental regulation of CCUS projects in China. Energy Procedia 2014, 63, 6903–6911. [Google Scholar] [CrossRef] [Green Version]
  215. Lv, G.; Li, Q.; Wang, S.; Li, X. Key techniques of reservoir engineering and injection–production process for CO2 flooding in China’s SINOPEC Shengli Oilfield. J. CO2 Util. 2015, 11, 31–40. [Google Scholar] [CrossRef]
  216. Yang, W.; Peng, B.; Liu, Q.; Wang, S.; Dong, Y.; Lai, Y. Evaluation of CO2 enhanced oil recovery and CO2 storage potential in oil reservoirs of Bohai Bay Basin, China. Int. J. Greenh. Gas Control 2017, 65, 86–98. [Google Scholar] [CrossRef]
  217. Mathisen, A.; Skagestad, R. Utilization of CO2 from Emitters in Poland for CO2 -EOR. Energy Procedia 2017, 114, 6721–6729. [Google Scholar] [CrossRef]
  218. Ampomah, W.; Balch, R.S.; Cathar, M.; Will, R.; Lee, S.Y.; Dai, Z. Performance of CO2-EOR and storage processes under uncertainty. In Proceedings of the SPE Europec featured at 78th EAGE Conference and Exhibition, Vienna, Austria, 30 May–2 June 2016. [Google Scholar]
  219. Jahangiri, H.R.; Zhang, D. Ensemble based co-optimization of carbon dioxide sequestration and enhanced oil recovery. Int. J. Greenh. Gas Control 2012, 8, 22–33. [Google Scholar] [CrossRef]
  220. Tapia, J.F.D.; Lee, J.Y.; Ooi, R.E.H.; Foo, D.C.Y.; Tan, R.R. CO2 allocation for scheduling enhanced oil recovery (EOR) operations with geological sequestration using discrete-time optimization. Energy Procedia 2014, 61, 595–598. [Google Scholar] [CrossRef] [Green Version]
  221. Le Van, S.; Chon, B.H. Evaluating the critical performances of a CO2–Enhanced oil recovery process using artificial neural network models. J. Petrol. Sci. Eng. 2017, 157, 207–222. [Google Scholar] [CrossRef]
  222. You, J.; Ampomah, W.; Kutsienyo, E.J.; Sun, Q.; Balch, R.S.; Aggrey, W.N.; Cather, M. Assessment of enhanced oil recovery and CO2 storage capacity using machine learning and optimization framework. In Proceedings of the SPE Europec featured at 81st EAGE Conference and Exhibition, London, UK, 3–6 June 2019. [Google Scholar]
  223. 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]
  224. Odi, U. Analysis and potential of CO2 huff-n-puff for near wellbore condensate removal and enhanced gas recovery. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 8–10 October 2012. [Google Scholar]
  225. Odi, U. Optimal Process Design for Coupled CO2 Sequestration and Enhanced Gas Recovery in Carbonate Reservoirs. Ph.D. Thesis, Texas A&M University, College Station, TX, USA, 2013. [Google Scholar]
  226. Eliebid, M.; Mahmoud, M.; Shawabkeh, R.; Elkatatny, S.; Hussein, I.A. Effect of CO2 adsorption on enhanced natural gas recovery and sequestration in carbonate reservoirs. J. Nat. Gas Sci. Eng. 2018, 55, 575–584. [Google Scholar] [CrossRef]
  227. Klimkowski, Ł.; Nagy, S.; Papiernik, B.; Orlic, B.; Kempka, T. Numerical simulations of enhanced gas recovery at the Załęcze gas field in Poland confirm high CO2 storage capacity and mechanical integrity. Oil Gas Sci. Technol. 2015, 70, 655–680. [Google Scholar] [CrossRef] [Green Version]
  228. Narinesingh, J.; Alexander, D. CO2 enhanced gas recovery and geologic sequestration in condensate reservoir: A simulation study of the effects of injection pressure on condensate recovery from reservoir and CO2 storage efficiency. Energy Procedia 2014, 63, 3107–3115. [Google Scholar] [CrossRef] [Green Version]
  229. Seo, J.G.; Mamora, D.D. Experimental and simulation studies of sequestration of supercritical carbon dioxide in depleted gas reservoirs. J. Energy Resour. Technol. 2005, 127, 1–6. [Google Scholar] [CrossRef] [Green Version]
  230. Zangeneh, H.; Safarzadeh, M.A. Enhanced gas recovery with carbon dioxide sequestration in a water-drive gas condensate reservoir: A case study in a real gas Field. J. Petrol. Sci. Eng. 2017, 7, 3–11. [Google Scholar]
  231. Seo, J.G. Experimental and Simulation Studies of Sequestration of Supercritical Carbon Dioxide in Depleted Gas Reservoirs. Ph.D. Thesis, Texas A&M University, College Station, TX, USA, 2004. [Google Scholar]
  232. Abba, M.K.; Abbas, A.J.; Nasr, G.G. Enhanced gas recovery by CO2 injection and sequestration: Effect of connate water salinity on displacement efficiency. In Proceedings of the SPE Abu Dhabi International Petroleum Exhibition & Conference, Abu Dhabi, UAE, 13–16 November 2017. [Google Scholar]
  233. Abdoulghafour, H.; Gouze, P.; Luquot, L.; Leprovost, R. Characterization and modeling of the alteration of fractured class-G Portland cement during flow of CO2 -rich brine. Int. J. Greenh. Gas Control 2016, 48, 155–170. [Google Scholar] [CrossRef]
  234. Liu, S.; Song, Y.; Zhao, C.; Zhang, Y.; Lv, P.; Jiang, L.; Liu, Y.; Zhao, Y. The horizontal dispersion properties of CO2-CH4 in sand packs with CO2 displacing the simulated natural gas. J. Nat. Gas Sci. Eng. 2018, 50, 293–300. [Google Scholar] [CrossRef]
  235. Liu, K.; Yu, Z.; Saeedi, A.; Esteban, L. Effects of permeability, heterogeneity and gravity on supercritical CO2 displacing gas under reservoir conditions. In Proceedings of the SPE Enhanced Oil Recovery Conference, Kuala Lumpur, Malaysia, 11–13 August 2015. [Google Scholar]
  236. Abba, M.; Abbas, A.; Saidu, B.; Nasr, G.; Al-Otaibi, A. Effects of gravity on flow behaviour of supercritical CO2 during enhanced gas recovery (EGR) by CO2 injection and sequestration. In Proceedings of the Fifth CO2 Geological Storage Workshop, Utrecht, The Netherlands, 21–23 November 2018. [Google Scholar]
  237. Zangeneh, H.; Jamshidi, S.; Soltanieh, M. Coupled optimization of enhanced gas recovery and carbon dioxide sequestration in natural gas reservoirs: Case study in a real gas field in the south of Iran. Int. J. Greenh. Gas Control 2013, 17, 515–522. [Google Scholar] [CrossRef]
  238. Wang, J.G.; Liu, J.; Liu, J.; Chen, Z. Impact of rock microstructures on the supercritical CO2 enhanced gas recovery. In Proceedings of the CPS/SPE International Oil and Gas Conference and Exhibition in China, Beijing, China, 8–10 June 2013. [Google Scholar]
  239. Honari, A.; Zecca, M.; Vogt, S.J.; Iglauer, S.; Bijeljic, B.; Johns, M.L.; May, E.F. The impact of residual water on CH4-CO2 dispersion in consolidated rock cores. Int. J. Greenh. Gas Control 2016, 50, 100–111. [Google Scholar] [CrossRef]
  240. Dou, X.; Liao, X.; Wang, H.; Zhao, T.; Cheng, Z.; Ren, W.; Zhang, R. The study of CO2 flooding and sequestration in tight gas reservoir: Different completion measures. In Proceedings of the SPE Nigeria Annual International Conference and Exhibition, Lagos, Nigeria, 4–6 August 2015. [Google Scholar]
  241. Khan, C.M.H. Techno-Economic Reservoir Simulation Model for CO2 Sequestration Evaluation. Ph.D. Thesis, Curtin University, Perth, Australia, 2013. [Google Scholar]
  242. Narinesingh, J.; Alexander, D. Injection well placement analysis for optimizing CO2 enhanced gas recovery coupled with sequestration in condensate reservoirs. In Proceedings of the SPE Trinidad and Tobago Section Energy Resources Conference, Port of Spain, Trinidad and Tobago, 13–15 June 2016. [Google Scholar]
  243. Feather, B.; Archer, R. Enhanced natural gas recovery by carbon dioxide injection for storage purposes. In Proceedings of the 17th Australasian Fluid Mechanics Conference, Auckland, New Zealand, 5–9 December 2010. [Google Scholar]
  244. Regan, M.L.M. A Numerical Investigation into the Potential to Enhance natural Gas Recovery in Water-Drive Gas Reservoirs Through the Injection of CO₂. Ph.D. Thesis, The University of Adelaide, Adelaide, Australia, 2010. [Google Scholar]
  245. Pooladi-Darvish, M.; Hong, H.; Theys, S.O.P.; Stocker, R.; Bachu, S.; Dashtgard, S. CO2 injection for enhanced gas recovery and geological storage of CO2 in the long Coulee Glauconite, F. pool, Alberta. In Proceedings of the SPE Annual Technical Conference and Exhibition, Denver, CO, USA, 21–24 September 2008. [Google Scholar]
  246. Biagi, J.; Agarwal, R.; Zhang, Z. Simulation and optimization of enhanced gas recovery utilizing CO2. Energy 2016, 94, 78–86. [Google Scholar] [CrossRef]
  247. Patel, M.J.; May, E.F.; Johns, M.L. Inclusion of connate water in enhanced gas recovery reservoir simulations. Energy 2017, 141, 757–769. [Google Scholar] [CrossRef] [Green Version]
  248. Geel, C.R.; Arts, R.J.; Van Eijs, R.M.H.E.; Kreft, E.; Hartman, J.; D’Hoore, D. Geological site characterization of the nearly depleted K12-B gas field, offshore the Netherlands. In Proceedings of the International Symposium on Site Characterization for CO2 Geological Storage, Berkeley, CA, USA, 20–22 March 2006. [Google Scholar]
  249. Vandeweijer, V.; van der Meer, B.; Hofstee, C.; Mulders, F.; D’Hoore, D.; Graven, H. Monitoring the CO2 injection site: K12-B. Energy Procedia 2011, 4, 5471–5478. [Google Scholar] [CrossRef] [Green Version]
  250. Vandeweijer, V.; Hofstee, C.; Graven, H. 13 Years of safe CO2 injection at K12-B. In Proceedings of the Fifth CO2 Geological Storage Workshop, Utrecht, The Netherlands, 21–23 November 2018. [Google Scholar]
  251. Kühn, M.; Tesmer, M.; Pilz, P.; Meyer, R.; Reinicke, K.; Förster, A.; Kolditz, O.; Schäfer, D. CLEAN: Project overview on CO2 large-scale enhanced gas recovery in the Altmark natural gas field (Germany). Environ. Earth Sci. 2012, 67, 311–321. [Google Scholar] [CrossRef]
  252. Kobos, P.H.; Cappelle, M.A.; Krumhansl, J.L.; Dewers, T.A.; McNemar, A.; Borns, D.J. Combining power plant water needs and carbon dioxide storage using saline formations: Implications for carbon dioxide and water management policies. Int. J. Greenh. Gas Control 2011, 5, 899–910. [Google Scholar] [CrossRef] [Green Version]
  253. Li, Q.; Wei, Y.N.; Liu, G.; Shi, H. CO2-EWR: A cleaner solution for coal chemical industry in China. J. Clean. Prod. 2015, 103, 330–337. [Google Scholar] [CrossRef]
  254. Liu, H.; Hou, Z.; Were, P.; Sun, X.; Gou, Y. Numerical studies on CO2 injection–brine extraction process in a low-medium temperature reservoir system. Environ. Earth Sci. 2015, 73, 6839–6854. [Google Scholar] [CrossRef]
  255. Liu, G.; Gorecki, C.D.; Saini, D.; Bremer, J.M.; Klapperich, R.J.; Braunberger, J.R. Four-site case study of water extraction from CO2 storage reservoirs. Energy Procedia 2013, 37, 4518–4525. [Google Scholar] [CrossRef] [Green Version]
  256. Dewers, T.; Eichhubl, P.; Ganis, B.; Gomez, S.; Heath, J.; Jammoul, M.; Kobos, P.; Liu, R.; Major, J.; Matteo, E.; et al. Heterogeneity, pore pressure, and injectate chemistry: Control measures for geologic carbon storage. Int. J. Greenh. Gas Control 2018, 68, 203–215. [Google Scholar] [CrossRef]
  257. Dahaghi, A.K. Numerical simulation and modeling of enhanced gas recovery and CO2 sequestration in shale gas reservoirs: A feasibility study. In Proceedings of the SPE International Conference on CO2 Capture, Storage, and Utilization, New Orleans, LA, USA, 10–12 November 2010. [Google Scholar]
  258. Busch, A.; Alles, S.; Gensterblum, Y.; Prinz, D.; Dewhurst, D.; Raven, M.; Stanjek, H.; Krooss, B. Carbon dioxide storage potential of shales. Int. J. Greenh. Gas Control 2008, 2, 297–308. [Google Scholar] [CrossRef]
  259. Liu, F.; Ellett, K.; Xiao, Y.; Rupp, J.A. Assessing the feasibility of CO2 storage in the New Albany Shale (Devonian–Mississippian) with potential enhanced gas recovery using reservoir simulation. Int. J. Greenh. Gas Control 2013, 17, 111–126. [Google Scholar] [CrossRef]
  260. Baran, P.; Zarębska, K.; Krzystolik, P.; Hadro, J.; Nunn, A. CO2-ECBM and CO2 sequestration in Polish Coal Seam—Experimental study. J. Sustain. Mining 2014, 13, 22–29. [Google Scholar] [CrossRef] [Green Version]
  261. Fang, Z.; Li, X.; Hu, H. Gas mixture enhance coalbed methane recovery technology: Pilot tests. Energy Procedia 2011, 4, 2144–2149. [Google Scholar] [CrossRef] [Green Version]
  262. Randolph, J.B.; Saar, M.O. Impact of reservoir permeability on the choice of subsurface geothermal heat exchange fluid: CO2 versus water and native brine. In Proceedings of the Annual Meeting of the Geothermal Council (Geothermal 2011), San Diego, CA, USA, 23–26 October 2011. [Google Scholar]
  263. Brown, D.W. A hot dry rock geothermal energy concept utilizing supercritical CO2 instead of water. In Proceedings of the twenty-fifth workshop on geothermal reservoir engineering, Stanford, CA, USA, 24–26 January 2000. [Google Scholar]
  264. Randolph, J.B.; Saar, M.O. Combining geothermal energy capture with geologic carbon dioxide sequestration. Geophy. Res. Lett. 2011, 38, L10401. [Google Scholar] [CrossRef] [Green Version]
  265. Liu, H.; Hou, Z.; Li, X.; Wei, N.; Tan, X.; Were, P. A preliminary site selection system for a CO2-AGES project and its application in China. Environ. Earth Sci. 2015, 73, 6855–6870. [Google Scholar] [CrossRef]
  266. Procesi, M.; Cantucci, B.; Buttinelli, M.; Armezzani, G.; Quattrocchi, F.; Boschi, E. Strategic use of the underground in an energy mix plan: Synergies among CO2, CH4 geological storage and geothermal energy. Latium Region case study (Central Italy). Appl. Energy 2013, 110, 104–131. [Google Scholar] [CrossRef]
  267. Underhill, D.H. Analysis of Uranium Supply to 2050; No. IAEA-SM-362; 2000; Available online: https://inis.iaea.org/search/search.aspx?orig_q=RN:31054412 (accessed on 16 October 2019).
  268. Chong, Z.R.; Yang, S.H.B.; Babu, P.; Linga, P.; Li, X.S. Review of natural gas hydrates as an energy resource: Prospects and challenges. Appl. Energy 2016, 162, 1633–1652. [Google Scholar] [CrossRef]
  269. Ota, M.; Abe, Y.; Watanabe, M.; Smith, R.L.; Inomata, H. Methane recovery from methane hydrate using pressurized CO2. Fluid Phase Equilib. 2005, 228–229, 553–559. [Google Scholar] [CrossRef]
  270. Yuan, Q.; Sun, C.Y.; Yang, X.; Ma, P.C.; Ma, Z.W.; Liu, B.; Ma, Q.L.; Yang, L.Y.; Chen, G.J. Recovery of methane from hydrate reservoir with gaseous carbon dioxide using a three-dimensional middle-size reactor. Energy 2012, 40, 47–58. [Google Scholar] [CrossRef]
  271. Zhang, L.; Yang, L.; Wang, J.; Zhao, J.; Dong, H.; Yang, M.; Liu, Y.; Song, Y. Enhanced CH4 recovery and CO2 storage via thermal stimulation in the CH4/CO2 replacement of methane hydrate. Chem. Eng. J. 2017, 308, 40–49. [Google Scholar] [CrossRef]
  272. Liu, Y.; Hou, J.; Zhao, H.; Liu, X.; Xia, Z. A method to recover natural gas hydrates with geothermal energy conveyed by CO2. Energy 2018, 144, 265–278. [Google Scholar] [CrossRef]
  273. Lee, Y.; Choi, W.; Shin, K.; Seo, Y. CH4-CO2 replacement occurring in sII natural gas hydrates for CH4 recovery and CO2 sequestration. Energy Convers. Manag. 2017, 150, 356–364. [Google Scholar] [CrossRef]
  274. Xu, C.G.; Cai, J.; Yu, Y.S.; Chen, Z.Y.; Li, X.S. Research on micro-mechanism and efficiency of CH4 exploitation via CH4-CO2 replacement from natural gas hydrates. Fuel 2018, 216, 255–265. [Google Scholar] [CrossRef]
  275. Jafari Raad, S.M.; Hassanzadeh, H. Prospect for storage of impure carbon dioxide streams in deep saline aquifers—A convective dissolution perspective. Int. J. Greenh. Gas Control 2017, 63, 350–355. [Google Scholar] [CrossRef]
  276. Wang, J.; Ryan, D.; Anthony, E.J.; Wigston, A.; Basava-Reddi, L.; Wildgust, N. The effect of impurities in oxyfuel flue gas on CO2 storage capacity. Int. J. Greenh. Gas Control 2012, 11, 158–162. [Google Scholar] [CrossRef]
  277. Sayegh, S.G.; Krause, F.F.; Fosti, J.E. Miscible displacement of crude oil by CO2/SO2 mixtures. SPE Reserv. Eng. 1987, 2, 199–208. [Google Scholar] [CrossRef]
  278. Zhang, P.Y.; Huang, S.; Sayegh, S.; Zhou, X.L. Effect of CO2 impurities on gas-injection EOR processes. Proceedings of SPE/DOE Fourteenth Symposium on Improved Oil Recovery, Tulsa, OK, USA, 17–21 April 2004. [Google Scholar]
  279. Yang, F.; Zhao, G.B.; Adidharma, H.; Towler, B.; Radosz, M. Effect of oxygen on minimum miscibility pressure in carbon dioxide flooding. Ind. Eng. Chem. Res. 2007, 46, 1396–1401. [Google Scholar] [CrossRef]
  280. Hajiw, M.; Corvisier, J.; El Ahmar, E.; Coquelet, C. Impact of impurities on CO2 storage in saline aquifers: Modelling of gases solubility in water. Int. J. Greenh. Gas Control 2018, 68, 247–255. [Google Scholar] [CrossRef]
  281. Turta, A.T.; Sim, S.S.; Singhal, A.K.; Hawkins, B.F. Basic investigations on enhanced gas recovery by gas-gas displacement. In Proceedings of the Petroleum Society’s 8th Canadian International Petroleum Conference, Calgary, AB, Canada, 12–14 June 2007. [Google Scholar]
  282. Li, D.; Jiang, X.; Meng, Q.; Xie, Q. Numerical analyses of the effects of nitrogen on the dissolution trapping mechanism of carbon dioxide geological storage. Comput. Fluids 2015, 114, 1–11. [Google Scholar] [CrossRef]
  283. Chen, C.; Chai, Z.; Shen, W.; Li, W. Effects of impurities on CO2 sequestration in saline aquifers: Perspective of interfacial tension and wettability. Ind. Eng. Chem. Res. 2017, 57, 371–379. [Google Scholar] [CrossRef]
  284. Alpermann, T.; Dietrich, M.; Ostertag-Henning, C. Mineral trapping of a CO2/H2S mixture by hematite under initially dry hydrothermal conditions. Int. J. Greenh. Gas Control 2016, 51, 346–356. [Google Scholar] [CrossRef]
  285. Fischer, S.; Wolf, L.; Fuhrmann, L.; Gahre, H.; Rütters, H. Simulated fluid-rock interactions during storage of temporally varying impure CO2 streams. In Proceedings of the Fifth CO2 Geological Storage Workshop, Utrecht, The Netherlands, 21–23 November 2018. [Google Scholar]
  286. Wolf, J.L.; Niemi, A.; Bensabat, J.; Rebscher, D. Benefits and restrictions of 2D reactive transport simulations of CO2 and SO2 co-injection into a saline aquifer using TOUGHREACT V3.0-OMP. Int. J. Greenh. Gas Control 2016, 54, 610–626. [Google Scholar] [CrossRef]
  287. Wolf, J.L.; Fischer, S.; Rütters, H.; Rebscher, D. Reactive transport simulations of impure CO2 injection into saline aquifers using different modelling approaches provided by TOUGHREACT V3.0-OMP. Proc. Earth Planet. Sci. 2017, 17, 480–483. [Google Scholar] [CrossRef]
  288. 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]
Figure 1. Correlation between atmospheric concentration of CO2 and the global temperature since 1850s (Data from [1,2]).
Figure 1. Correlation between atmospheric concentration of CO2 and the global temperature since 1850s (Data from [1,2]).
Energies 13 00600 g001
Figure 2. International Energy Agency (IEA) forecasts of key technologies for CO2 emission reductions (modified from [8]).
Figure 2. International Energy Agency (IEA) forecasts of key technologies for CO2 emission reductions (modified from [8]).
Energies 13 00600 g002
Figure 3. Commercial-scale integrated carbon capture and storage (CCS) projects around the world. Circle size is proportional to the CO2 capture capacity, and the color indicates different stages of the lifecycle of the project (data from [10]).
Figure 3. Commercial-scale integrated carbon capture and storage (CCS) projects around the world. Circle size is proportional to the CO2 capture capacity, and the color indicates different stages of the lifecycle of the project (data from [10]).
Energies 13 00600 g003
Figure 4. (a) The four main CO2 trapping mechanisms [51]; (b) the contribution of four CO2 trapping mechanisms with time (modified from [52]).
Figure 4. (a) The four main CO2 trapping mechanisms [51]; (b) the contribution of four CO2 trapping mechanisms with time (modified from [52]).
Energies 13 00600 g004
Figure 5. The long-term evolution of the injected CO2 in deep-sea sediments (modified from [104]).
Figure 5. The long-term evolution of the injected CO2 in deep-sea sediments (modified from [104]).
Energies 13 00600 g005
Figure 6. The framework of the National Risk Assessment Partnership (NARP) [144].
Figure 6. The framework of the National Risk Assessment Partnership (NARP) [144].
Energies 13 00600 g006
Figure 7. Engineered injection method to accelerate CO2 dissolution and trapping (adapted from [175]).
Figure 7. Engineered injection method to accelerate CO2 dissolution and trapping (adapted from [175]).
Energies 13 00600 g007
Figure 8. Schematic of various schemes of water-alternating-gas (WAG) injection [179].
Figure 8. Schematic of various schemes of water-alternating-gas (WAG) injection [179].
Energies 13 00600 g008
Figure 9. Schematic of the intermittent injection method (modified from [180]).
Figure 9. Schematic of the intermittent injection method (modified from [180]).
Energies 13 00600 g009
Figure 10. Calculated temperature and the carbonation rate relative to the rate for CO2 in surface water at 25 °C and 0.1 MPa in the three-step injection operation (adapted from [181]).
Figure 10. Calculated temperature and the carbonation rate relative to the rate for CO2 in surface water at 25 °C and 0.1 MPa in the three-step injection operation (adapted from [181]).
Energies 13 00600 g010
Figure 11. Preliminary experimental setup for CO2 capture and mineralization (modified from [182]).
Figure 11. Preliminary experimental setup for CO2 capture and mineralization (modified from [182]).
Energies 13 00600 g011
Figure 12. Mineralization process concept for (a). pure CO2, (b). flue gas (modified from [29]).
Figure 12. Mineralization process concept for (a). pure CO2, (b). flue gas (modified from [29]).
Energies 13 00600 g012
Figure 13. The schematic of carbon mineralization process using recyclable ammonium salts (modified from [183]).
Figure 13. The schematic of carbon mineralization process using recyclable ammonium salts (modified from [183]).
Energies 13 00600 g013
Figure 14. Conceptual steps of CO2-enhanced oil recovery (CO2-EOR) in fractured tight oil reservoirs ([201]).
Figure 14. Conceptual steps of CO2-enhanced oil recovery (CO2-EOR) in fractured tight oil reservoirs ([201]).
Energies 13 00600 g014
Figure 15. Artificial neural network structure of the models, Y represents oil recovery, oil production rate, gas and oil ratio (GOR), and net CO2 storage amount [221].
Figure 15. Artificial neural network structure of the models, Y represents oil recovery, oil production rate, gas and oil ratio (GOR), and net CO2 storage amount [221].
Energies 13 00600 g015
Figure 16. Flowchart of the optimization framework (modified from [222]).
Figure 16. Flowchart of the optimization framework (modified from [222]).
Energies 13 00600 g016
Figure 17. Depiction of the CO2-enhanced water recovery (CO2-EWR) technology [253].
Figure 17. Depiction of the CO2-enhanced water recovery (CO2-EWR) technology [253].
Energies 13 00600 g017
Figure 18. Schematic diagram for CH4–CO2 replacement in hydrates [270].
Figure 18. Schematic diagram for CH4–CO2 replacement in hydrates [270].
Energies 13 00600 g018
Figure 19. Schematic well group configuration diagram of the geothermal-assisted CO2 replacement method (GACR) [272].
Figure 19. Schematic well group configuration diagram of the geothermal-assisted CO2 replacement method (GACR) [272].
Energies 13 00600 g019
Table 1. Summary of review literature on CCS technology.
Table 1. Summary of review literature on CCS technology.
Research FieldsRef.Review Scope
CO2 capture and utilization[13]Review of the application of CO2 for enhanced oil and gas recovery
[14]Review of CO2 capture and reuse technologies, highlighting the strategies of CO2 capture in variety of scenarios, and the state of the art for CO2 utilization
[15]Review of CO2 capture, utilization, and storage (CCUS) in Chinese Academy of Sciences, highlighting the strategies for CCUS in China
[16]Review of the property impacts of CCS, highlighting the effect of uncertainties in thermal–physical properties on the design of components and processes in CCS
[17]Review of CCS highlighting the CO2 capture technologies, the pilot plants, and the economic and legal aspects of CCS
[18]Review of CO2 enhanced coal bed methane recovery, highlighting the CO2 storage trials in the San Juan Basin in USA, and the estimation of CO2 storage capacity in coal seams
[19]Review of CCUS technologies highlighting the engineering projects and their developments in China
[20]Review of CCS highlighting the findings obtained in CCS operational projects including the technologies of CO2 capture, separation, transport, and storage
Options for CO2 storage and CCS projects[21]Review of CCS highlighting the options for CO2 storage, the evaluation criteria for CO2 storage sites, and the major CO2 storage projects
[22]Review of biomass with CCS (Bio-CCS), highlighting the economics and global status of Bio-CCS, and the role of Bio-CCS in the food–water–energy–climate nexus
[23]Review of CO2 storage in saline aquifers, highlighting the geological and operation parameters, and the monitoring technologies for existing saline aquifers storage operations
[24]Review of the CCS in a coal-fired plant in Malaysia, highlighting the choices of coal plants and the capture technologies
[25]Review of CO2 storage in saline formations, highlighting the modeling of solubility trapping
[26]Review of mineral carbonation (MC) technologies for CO2 sequestration, highlighting the mechanisms of MC technologies and their contribution in decreasing the cost of CCS
[27]Review of CCS projects and future opportunities, highlighting the technical details and business plan for CCS projects
[28]Review of CO2 storage projects in China, highlighting the CO2 source, and CO2 storage strategies in China
[29]Review of CO2 mineralization product forms, highlighting the mineralization process for CO2 storage
[30]Review of CCS by using coal fly ash, highlighting the feasibility and prospects of CCS using coal fly ash
CO2-brine-rock systems[31]Review of the relative permeability and residual trapping in CO2 storage systems, highlighting the estimating and measuring methods
[32]Review of the geochemical aspects of CO2 storage in saline aquifers, highlighting the advantages of CO2 storage in saline aquifers, and the CO2–brine–rock interactions in the aquifers
[33]Review of geomechanical modeling of CO2 storage, highlighting the numerical methods and their application in the modeling of ground deformation, faults, and fracture propagation
[34]Review of CO2 sequestration highlighting the trapping mechanisms and the flow of CO2 brine in porous media system
Well integrity and risk assessment[35]Review of the cement degradation in CO2-rich conditions of CCS projects, highlighting the degradation of Portland cement
[36]Review of the risk assessment of CO2 storage, highlighting the regulations and strategies of risk assessment for CO2 storage
[37]Review of the isotopic composition of CO2 for leakage monitoring in CCS project, highlighting the stable isotopes as a tracer for injected CO2
[38]Review of the integrity of existing wells for CCS, highlighting the mechanical well failure and chemical issue due to cement carbonation
[39]Review of well integrity of CCS, highlighting the corrosion of metallic and cement, and the remedial measures
[40]Review of caprock sealing mechanisms for CO2 storage, highlighting the problems associated with CO2 leakage, the leakage paths, and the factors that affect leakage
[41]Review of CO2 storage highlighting the capacity estimation of storage sites, the monitoring technologies, and the simulation tools for CCS
[42]Review of CO2 storage and caprock integrity, highlighting the major CCS project in operation and CO2 migration in the reservoirs
Storage efficiency and environmental considerations[43]Review of CO2 storage efficiency in saline aquifers, highlighting the factors that affect CO2 plume migration and the methods to estimate the storage capacity
[44]Review of environmental considerations for CO2 storage in a sub-seabed, highlighting the potential ecological impacts
Table 2. Large-scale CCS project in saline aquifers.
Table 2. Large-scale CCS project in saline aquifers.
Num.ProjectInjection Rate (t/d)Permeability (mD)Depth (m)Thickness of Reservoir (m)Thickness of Caprock (m)Reservoir Temperature (°C)Reservoir Pressure (MPa)Ref.
1Snøhvit2000450255060309528.5[23,61]
2Sleipner270030001000250753710.3[23,58,66]
3In Salah3500131800209009017.9[23,63]
4Gorgon10,410 25230028025010022[23,64]
5Quest2960100200040705518.9[65,66,67,68]
Table 3. The pros and cons of a variety of storage strategies.
Table 3. The pros and cons of a variety of storage strategies.
Option ProsCons
Saline aquifersHuge amount of storage capacity, wide distribution, commercial technology readiness levelNo economic benefit
Depleted oil and gas reservoirsExisting installed equipment, guaranteed caprock integrity, characterized geological conditions, small pressure perturbations and induced stress changes, additional oil and gas recoveryDemonstration technology readiness level
Coal bedsLow transportation cost due to its potential location near the coal-fired power plants, additional coal bed methane recoveryPilot plant technology readiness level
Deep oceanLarge storage capacityFormulation technology readiness level, no economic benefit, may affect the marine environment
Deep-sea sedimentsEnormous storage capacity, free from the potential harm to the ocean ecosystemsFormulation technology readiness level, no economic benefit, far more expensive than onshore methods
Table 4. Main monitoring technologies in CCS.
Table 4. Main monitoring technologies in CCS.
Monitoring TechnologyAdvantagesRef.
3D seismic Provides a tridimensional image of geological structures and the plume migration of CO2.[62]
4D seismicSignificant benefits for overburden imaging and time-lapse responses with improved acquisition plan.[62]
Microseismic It is very useful for monitoring geomechanical response to injection.[151]
Vertical seismic profilingValuable information on the geological structure details.[152]
Gravimetry Beneficial for the evaluation of formation fluids density and CO2 plume.[153]
Cross-hole electromagneticAdvantageous for the detection and monitoring of the location of CO2.[155]
Pressure and temperature monitoringDirect information for the evaluation of the stability of the reservoir.[156]
Geochemical samplingNatural variations in water chemistry are crucial for establishing a useful baseline for groundwater hydrology.[157]
Soil and gas samplingMore data on natural CO2 variations in different environments and associated seasonal fluctuations is needed.[62]
TracersValuable and cost-effective method for monitoring the origin of CO2 observations at wells and in the storage complex.[62]
Atmospheric monitoringUseful data to identity the anomalies above the natural baseline.[160]
MicrobiologyValuable data to identify biogeochemical process that affect the diffusion of CO2 in the reservoirs.[161]
Core analysisGood petrophysical data and rock mechanical properties are essential.[62]
Satellite monitoringValuable and cost-effective monitoring data for onshore CO2 injection operation.[163]
Distributed temperature sensing technologyIt can provide high-resolution information on the migration of CO2 in the reservoir. [164]
Table 5. Application of the main monitoring technologies in some CCS demonstration projects (modified from [148]).
Table 5. Application of the main monitoring technologies in some CCS demonstration projects (modified from [148]).
Monitoring TechnologySleipnerFrioNagaokaKetzinIn-SalahOtwayWeyburnMRCSP
3D seismic × ××××
4D seismic ××
Microseismic × × × ××
Vertical seismic profiling × ×
Gravimetry × × ××
Cross-hole electromagnetic × ××
Pressure and temperature logs ×××× ×
Geochemical sampling ×× ××××
Soil and gas sampling × × ×
Tracers × ××
Atmospheric monitoring ×
Microbiology ×
Core analysis × ×
Satellite monitoring × ×
Distributed temperature sensing technology ×
Table 6. Typical displacement experiments on the CO2-enhanced gas recovery (CO2-EGR) process.
Table 6. Typical displacement experiments on the CO2-enhanced gas recovery (CO2-EGR) process.
Rock TypeSaturated FluidsT (°C)P (MPa)Key ObservationsRef.
Carbonate coreCH420–603.55–20.79Whether CO2 is in the gas, liquid, or supercritical phase, it could enhance the recovery of CH4.[83]
Carbonate coreSaturated with methane with or without water20–803.55–20.79The coefficient of CO2 increases with temperature and decreases with pressure.[231]
Berea sandstone coreDry core, initial saturation of 10% water, and initial saturation of 10% brine (20 wt %), respectively408.96The salinity of connate water will decrease the dispersion of CO2 in CH4.[232]
Sandstone and carbonate coreCH460–8010–12The residual water narrows the pore and consequently increases the dispersion of supercritical CO2 and CH4.[233]
Sandstone coreCH4 and simulate natural gas (90% CH4 + 10% CO2) respectively40–5510–14The dispersion coefficient of CO2 in the simulate natural gas is larger than that of CH4.[234]
Sandstone coreFormation water and N25021The gravity segregation effect is notable in the porous and permeable core, while the heterogeneity effect becomes dominant in the low permeability of the core.[235]
Bandera sandstone coreCH4508.96The gravity has significant effects on the flow behavior of SCO2 at lower flow rates.[236]
Table 7. Large-scale CCS projects (more than 0.4 Mtpa) from 1972 to the end of the 2020s (data from [10]).
Table 7. Large-scale CCS projects (more than 0.4 Mtpa) from 1972 to the end of the 2020s (data from [10]).
Strategy Under EvaluationEORCCSTotal
Saline FormationDepleted Gas Fields
Quality of project24321351
Capture capacity (Mtpa) 42.11–43.418.1–8.640.35–85.17.5–8.598.06–145.61
Average capture capacity (Mtpa) 1.75–1.812.7–2.871.92–4.052.5–2.831.92–2.86

Share and Cite

MDPI and ACS Style

Cao, C.; Liu, H.; Hou, Z.; Mehmood, F.; Liao, J.; Feng, W. A Review of CO2 Storage in View of Safety and Cost-Effectiveness. Energies 2020, 13, 600. https://0-doi-org.brum.beds.ac.uk/10.3390/en13030600

AMA Style

Cao C, Liu H, Hou Z, Mehmood F, Liao J, Feng W. A Review of CO2 Storage in View of Safety and Cost-Effectiveness. Energies. 2020; 13(3):600. https://0-doi-org.brum.beds.ac.uk/10.3390/en13030600

Chicago/Turabian Style

Cao, Cheng, Hejuan Liu, Zhengmeng Hou, Faisal Mehmood, Jianxing Liao, and Wentao Feng. 2020. "A Review of CO2 Storage in View of Safety and Cost-Effectiveness" Energies 13, no. 3: 600. https://0-doi-org.brum.beds.ac.uk/10.3390/en13030600

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