Jinshuo Zou†
,
Divyani Gupta† and
Gemeng Liang
*
School of Chemical Engineering, The University of Adeliade, Adelaide, South Australia, Australia. E-mail: Gemeng.liang@adelaide.edu.au
First published on 22nd August 2025
The world is facing the pressing dual challenge of an escalating energy crisis and the excessive accumulation of greenhouse gases. Current researchers are now reimagining carbon dioxide (CO2) not simply as a waste to be captured, but as a valuable energy carrier to be utilized in next-generation energy systems. However, a comprehensive understanding of how CO2 participates in energy conversion and storage remains limited. This review addresses this critical knowledge gap by systematically examining four innovative types of CO2-based energy technologies. These include metal-CO2 batteries and CO2 capture-integrated storage systems that leverage the electrochemical activity of CO2 for efficient and sustainable energy storage, as well as molten carbonate fuel cells and CO2-based electricity generators that generate electricity energy in innovative approaches. These emerging systems leverage the physicochemical properties of CO2 to create transformative energy solutions. This review highlights the pivotal role of CO2, examines critical challenges, and proposes forward-looking strategies for future energy development.
The ongoing energy crisis poses a significant threat to modern society, affecting nearly every aspect of daily life. Driven by the depletion of fossil fuel resources, geopolitical instability, and the urgent need for decarbonization, the crisis has led to high energy prices, disrupted supply chains, and increased pressure on energy security. These challenges not only hinder economic growth but also limit access to affordable energy, and impede the global transition toward sustainable development. In response to the energy crisis, significant efforts have been devoted to advancing renewable energy sources, enhancing energy efficiency, and developing advanced energy producing and storage technologies.5
Among emerging solutions, the integration of CO2 utilization into energy conversion systems has opened new paradigms, where CO2 is captured and utilized as a functional reactant contributing to energy storage and producing.6 This approach leverages the electrochemical reactivity of CO2 to participate in redox reactions, enabling the design of novel energy storage cells and electricity generation devices. Furthermore, coupling CO2 conversion with renewable electricity provides a sustainable pathway to close the carbon loop, reduce reliance on fossil fuels, and enhance overall energy system efficiency.7,8 As such, CO2-based energy producing and storage technologies represent a promising intersection of carbon management and clean energy innovation. Despite rapid advancements in CO2-utilizing energy technologies, the field still lacks a comprehensive and cohesive review that connects scattered breakthroughs into a clear and guiding framework, leaving a critical gap in the literature that must be addressed to drive future innovation.
This timely review critically examines the emerging strategy of integrating CO2 utilization into energy producing and storage, an area at the nexus of carbon mitigation and clean energy innovation. We focus on four major approaches: metal–CO2 batteries, CO2 capture-based energy storage systems, molten carbonate fuel cells, and CO2-based electricity generators (Fig. 1). Their working principles, material design strategies, technical challenges, and application prospects are systematically discussed. By consolidating recent progress and outlining future directions, this work provides essential insights to advance this rapidly evolving and impactful field.
![]() | ||
Fig. 1 Summary of CO2 utilization in energy storage and producing. Four systems including metal–CO2 batteries, CO2 capture-based energy storage, molten carbonate fuel cells, and CO2-based electricity generators are illustrated. The mechanism schematic of CO2-based electricity generators is reproduced with permission.9 Copyright 2024, Springer Nature. |
CO2 conversion products | Battery type | Cathode catalysts | Electrolyte | Discharge/charge voltage | Energy efficiency | Max discharge capacity | Stability | Ref. |
---|---|---|---|---|---|---|---|---|
Carbonate and carbon | Li–CO2 | NiCo2S4 | 1 M lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) in tetraethylene glycol dimethyl ether (TEGDME) | 2.67/3.82 V at 20 μA cm−2 | 88.7% | 5528 μA h cm−2 | 600 h at 20 μA cm−2 | 13 |
Ru/CNTs | 1 M LiTFSI in dimethyl sulfoxide (DMSO) | 2.8/3.75 V at 100 mA g−1 | — | 11 888 mA h g−1 at 100 mA g−1 | 2000 h at 100 mA g−1 | 12 | ||
Fe2O3@CoS | 1 M LiTFSI in TEGDME | 2.9/3.1 V 20 μA cm−2 | 92.4% | 4900 μA h cm−2 | 550 h at 20 μA cm−2 | 14 | ||
NiFeCoCuRu high entropy alloys | 1 M LiTFSI in DMSO | 2.85/3.72 V at 100 mA g−1 | — | — | 2900 h at 100 mA g−1 | 15 | ||
Hydrogen-bonded organic framework-FJU-1-Ru@CNT | 1 M LiTFSI in TEGDME | 2.79/3.89 V | 65.77% | 24 245.3 mA h g−1 at 10 mA g−1 | 200 cycles at 1 A g−1 with 1000 mA h g−1 | 16 | ||
Na–CO2 | Tetraethylene glycol dimethyl-treated multi-wall carbon nanotube | 1 M NaClO4 in TEGDME | 2.6/3.2 V at 1 A g−1 | 81% | 60 000 mA h g−1 at 1 A g−1 | 200 cycles with 2000 mA h g−1 | 17 | |
Metal-free porous carbon | 1 M NaClO4 in TEGDME | 2.28/3.32 V | 71.2% at 10 μA cm−2 | 2422 μA h cm−2 | 1000 h at 10 μA cm−2 | 18 | ||
Activated multiwall carbon nanotubes | [Poly(vinylidene fluoride-co-hexafluoropropylene)]-4% SiO2/NaClO4–TEGDME | 2.75/3.49 V | — | 5000 mA h g−1 at 50 mA g−1 | 400 cycles at 500 mA g−1 | 19 | ||
K–CO2 | Bamboo-like N-doped carbon nanotube | 1.0 M potassium bis(trifluoromethanesulfonyl)imide (KTFSI) in TEGDME | 2.40/4.01 V | 53–62% | 9436 mA h g−1 at 200 mA g−1 | 450 cycles or 3100 h with capacity of 500 mA h g−1 | 20 | |
Carboxyl-containing multi-walled carbon nanotubes | 0.3 M KPF6 in TEGDME | 2.2/3.0 V | — | 3681 mA h g−1 at 100 mA g−1 | 400 cycles | 21 | ||
Mg–CO2 | CNTs | MgCl2 and Mg(TFSI)2 in DME | Discharge at 1 V at 200 mA g−1 | — | — | 250 h | 22 | |
Al–CO2 | NPG@Pd | AlCl3 and 1-ethyl-3-methylimidazolium chloride with a mole ratio of 1.3 | 0.72/0.74 V | 87.7% | 638 mA h g−1 | 30 cycles at 333 mA g−1 | 23 | |
Oxalate | Li–CO2 | Cu(II) MOF (benzene-1,3,5-tricarboxylic acid)/CNTs | 1 M LiTFSI in TEGDME | 2.8/3.7 V | — | 9040![]() |
400 cycles at 500 mA g−1 | 24 |
Mo2C/carbon nanotube | 1 M LiCF3SO3/TEGDME | 2.8/3.5 V | 77% | 1150 mA h at 20 mA | 40 cycles at 20 mA | 25 | ||
MoN nanofibers on carbon-cloth | 1 M LiTFSI/TEGDME | 2.83/3.19 V | 88% | 6542.9 μA h cm−2 | 86 cycles at 100 μA cm−2 | 26 | ||
Na–CO2 | Co-encapsulated N-doped carbon framework | Gel electrolyte | Voltage gap: 1.7 V | — | 3094 mA h g−1 | 366 cycles (2200 h) at 0.1 mA cm−2 | 27 | |
Ni0·95Fe0·05(OH)2 | Sodium ion conducting solid-state electrolyte | 3.35/3.5 V | — | 1166 mA h g−1 | 100 h | 28 | ||
Mg–CO2 | Nitrogen-doped carbon nanotubes | 0.25 M Mg(TFSI)2-TEGDME with 1,3-propylene amine additives | 0.75/2.25 V | — | 600 mA h g−1 at 200 mA g−1 | 70 cycles, more than 400 h at 200 mA g−1 | 29 | |
Al–CO2 | Ketjen black | AlCl3: 1-ethyl-3-methylimidazolium chloride ionic liquid electrolyte | 1.12/1.17 V | — | 9394 mA h gcarbon−1 | 1000 h at 0.2 mA cm−2 | 30 | |
Carbonate and CO | Li–CO2 | Porous fractal Zn | 1 M LiTFSI/TEGDME | 1.91 V/– | 67% | — | 1500 min during discharge | 31 |
Carbon-containing compound products | Zn–CO2 | BiOF | 0.8 M KHCO3 saturated with CO2 | 1.2 V/– | 70.74% | — | 200 cycles and 68 h | 32 |
3D porous Pd interconnected nanosheets | Catholyte: 1 M NaCl; anolyte: 1 M KOH added with 0.02 M Zn(CH3COO)2 | Charge voltage: 1 V | 81.2![]() |
— | Over 100 cycles | 33 | ||
Ni Nanoclusters and single atom site | Catholyte: 1 M KOH; anolyte: 6 M KOH and 0.2 M Zn(CH3COO)2 | 0.65 V | — | — | 1200 cycles and 420 h at 5 mA cm−2 | 34 | ||
Ir@Au bimetal nanomaterial | Catholyte: 0.8 M KHCO3; anolyte: 0.8 M KOH with 0.02 M Zn(CH3COO)2 | Charge voltage: 2 V | 68% | — | 90 cycles at 0.9 mA (5 mA cm−2) | 35 |
Anode: M → Mn+ + e− | (1) |
Cathode: Mn+ + CO2 + e− → M2(CO3)n + C | (2) |
Among these MCBs, Li–CO2 battery has emerged as one of the most extensively studied configurations due to its large theoretical specific energy (∼1876 W h kg−1), high cell voltage of 2.8 V, and the largest theoretical specific capacity of Li (3861 mA h g−1). During discharge, the valence state of carbon in CO2 shifts from +4 to 0 (eqn (3)), which proceeds through multiple intermediates and reaction steps. The CO2 reduction starts with the adsorption of CO2 on the catalytic sites at the cathodes. Upon accepting two electrons, the CO2 molecules are converted into oxalate anions (C2O42−), as described by eqn (4). The C2O42− is usually unstable and undergoes a disproportion reaction to form CO22− and CO2 through eqn (5). This is followed by the coupling step of C2H42− with CO22− to produce CO32− and amorphous carbon (C) via eqn (6). Finally, the CO32− species readily combine with lithium ions (Li+) derived from the anode (eqn (7)), resulting in the formation of lithium carbonate (Li2CO3) products. The nucleation and growth of Li2CO3 and carbon in Li–CO2 batteries are fundamentally governed by the interaction strength between reaction intermediates and the catalyst surface.10 There are two distinct growth mechanisms that critically influence the morphology of discharge products and the overall battery performance. When the binding energy between intermediates and the catalyst (e.g., Co3O4-based cathodes) is weak, product formation proceeds via a solvation-mediated growth mechanism.37 In this case, Li2CO3 tends to nucleate in the electrolyte near the cathode surface, leading to the development of bulky, flower-like structures. These discharge products can deliver high capacity but often compromise charge efficiency due to their large size and sluggish decomposition. In contrast, strong adsorption between intermediates and the cathode (e.g., MnOOH arrays on stainless-steel (SS) mesh cathodes) drives a surface adsorption growth mechanism, resulting in thin, uniform layers of Li2CO3 directly on the electrode.38 This morphology is more conducive to efficient charge reactions, enhancing reversibility and energy efficiency, though it may limit the achievable capacity. Balancing these two mechanisms is therefore key to optimizing the electrochemical performance of Li–CO2 batteries.
4Li+ + 3CO2 + 4e− → 2Li2CO3 + C | (3) |
2CO2 + 2e− → C2O42− | (4) |
C2O42− → CO22− + CO2 | (5) |
C2O42− + CO22− → 2CO32− + C | (6) |
CO32− + 2Li+ → 2Li2CO3 | (7) |
A critical limitation to the rechargeability of Li–CO2 batteries lies in the decomposition of discharge products during the charging process. Li2CO3, a wide-bandgap insulating compound with inherently low solubility in typical organic electrolytes, limit the battery reversibility and rechargeability.39 Its decomposition typically demands a high overpotential, largely determined by the electrochemical properties at the CO2/electrolyte/catalyst interface. Under optimized conditions, where a catalytically active interface is established, Li2CO3 and the accompanying amorphous carbon can be decomposed simultaneously, thereby closely reversing the discharge reaction (eqn (7)). Achieving such favorable conditions typically requires the integration of highly active catalysts, such as Ru, Ir, Ni, or Mn, with specially designed electrolytes (e.g., water-in-salt systems or ionic liquids) that support interfacial reactivity and Li+ ion transport.40 However, when the interfacial environment is suboptimal with inactive catalysts, incompatible electrolytes, or ineffective surface contact between solid catalysts and discharge products, carbon fails to participate in the decomposition, and only Li2CO3 undergoes direct electrochemical decomposition at elevated voltages. This alternative pathway often triggers the formation of reactive oxygen species such as singlet oxygen (1O2), as described in eqn (8) and (9).36,41,42 These species are known to aggressively degrade organic electrolytes and carbon cathode structures, accelerating parasitic reactions and compromising cell stability. Although Li–CO2 batteries are technically rechargeable along both pathways, the latter route is plagued by irreversible side reactions, accumulation of unwanted by-products, and poor cycling performance. Therefore, elucidating the fundamental discharge–charge pathways, particularly the formation and decomposition mechanisms of Li2CO3, carbon, and associated intermediates, is critical for the rational design of cathode materials and electrolytes.
2Li2CO3 → 2CO2 + O2˙− + 4Li+ + 3e− | (8) |
2Li2CO3 → 2CO2 + 1O2 + 4Li+ + 4e− | (9) |
The formation and deposition of carbonate and carbon as discharge products have been confirmed through a range of advanced characterization techniques. Since Li2CO3 typically grows partially or entirely in crystalline form on the cathode surface, its characteristic peaks are prominently observed in the XRD patterns.12 On the other hand, the carbonate ion is Raman- and IR-active, making both spectroscopic techniques effective tools for confirming the presence of lithium carbonate. However, one main challenge lies in verifying carbon (C) as a genuine discharge product, as this requires ruling out interference from the commonly used carbon-based gas diffusion electrodes, which can contribute background carbon signals. To address this, Yang et al. employed a Ru catalyst nanoparticle-decorated Ni foam (Ru@Ni) cathode to specifically identify carbon products.43 Their ex situ Raman measurements revealed clear D and G bands of carbon on the fully discharged cathode, thereby providing direct evidence of carbon formation during discharge. Qiao et al. employed in situ surface-enhanced Raman spectroscopy (SERS) to study the CO2 conversion pathways in an in situ Raman Li–CO2 cells (0.5 M LiClO4 in DMSO, saturated with CO2) (Fig. 2a).44 To construct the catalytic cathode for SERS analysis, a thin Ru film was deposited onto the Au substrate supported on the Al mesh via radio frequency magnetron sputtering. By replacing the conventional carbon-based gas diffusion layers, this electrode configuration effectively eliminates interference from pre-existing carbon species, thereby ensuring in situ identification of newly formed carbonaceous products. The Raman spectra exhibited a distinct, sharp peak at 1085 cm−1, corresponding to the symmetric stretching vibration of Li2CO3, alongside two broader bands at 1319 cm−1 and 1587 cm−1, attributable to the disordered (D-band) and graphitic (G-band) vibrations of the amorphous carbon, respectively, confirming the co-formation of Li2CO3 and amorphous C. The two discharge products show synchronously increasing trend (Fig. 2b). To provide more quantitative information on the discharge products, the H2SO4 is injected to the cathode after discharge with 0.6 mA h capacity, followed by gas analysis using gas chromatography-mass spectrometry (GC-MS) (Fig. 2c). The inset pie chart presents the proportion of formed Li2CO3, determined from the integrated area under the CO2 evolution rate relative to the total amount of CO2 reduced, as estimated from the discharge capacity. Moreover, the in situ differential electrochemical mass spectrometry (DEMS) serves as another powerful technique to monitor real-time CO2 consumption and gaseous product evolution, providing critical insights into reaction stoichiometry and product formation. In the proposed reaction pathway of eqn (3), the theoretical charge-to-mass ratio (e−/CO2) is 1.33, reflecting the number of electrons required per CO2 molecule involved. Wang et al. also employed DEMS to quantify both electron transfer and CO2 conversion during the operation of a Li–CO2 cell.45 Their results revealed that a discharge and charge capacity of 0.2 mA h corresponded to approximately 5.6 μmol of CO2 consumption or release, coupled with 7.48 μmol of electron transfer. Specifically, during the discharge process, 5.3 μmol of CO2 was consumed, yielding a calculated e−/CO2 ratio of 1.41. Upon charging, 5.1 μmol of CO2 was released, corresponding to a slightly higher e−/CO2 value of 1.47. These results align well with the theoretical prediction for the reaction pathway yielding Li2CO3 and carbon as the main discharge products, thereby offering strong empirical support for the proposed mechanism. Additionally, the integration of in situ GC-MS with isotope tracing techniques enables precise monitoring of gas evolution and unambiguous identification of the carbon origin in the evolved CO2 during the charging process. Zhou et al. preloaded the cathode with Li2CO3 and 13C for subsequent charging, employing a carbon-only cathode and a LiCF3SO3/tetraglyme electrolyte (molar ratio 1 : 4).46 During the charge process, CO2 with a mass-to-charge ratio (m/z) of 44 was identified as the dominant gaseous product, indicating that carbon did not participate in the charge reaction under these conditions. However, the information from in situ experiment is still limited by current instrumentation and analytical techniques, such as inadequate temporal and spatial resolution. Furthermore, the typically low concentrations of intermediates and products in Li–CO2 batteries result in weak signal intensities and significant spectral overlap, which further complicate the identification of CO2-involved reaction pathways.
![]() | ||
Fig. 2 Overview of MCBs with carbonate and carbon as discharge products. (a) Conceptual illustration of an energy storage system based on Li–CO2 electrochemistry technology. (b) Galvanostatic discharge profiles and corresponding Raman signal intensity of a Li–CO2 cell comprising a sputtered gold cathode, a CO2-saturated 0.5 M LiClO4-DMSO electrolyte, and a lithium metal foil. The cell was tested under a current of 5 μA with a capacity cutoff of 10 μAh. (c) CO2 gas evolution rate during the quantification of carbonate species on a cathode previously discharged to 0.6 mA h. The blue arrow denotes the injection of H2SO4, which decomposes Li2CO3 to release CO2. The inset pie chart shows the yield of fixed Li2CO3 relative to the total reduced CO2, as determined by integrating the CO2 evolution curve and correlating it with the discharge capacity. (a)–(c) Are reproduced with permission.44 Copyright 2017, Cell Press. (d) Schematic of p-band center modulation in nitrogen and sulfur co-doped carbon catalysts, along with the catalyst synthesis strategy for Na–CO2 batteries. (e) Rate performance of batteries with the four cathodes, including N, S co-doped porous carbon on carbon paper (NS-PC/CP), N-doped porous carbon on carbon paper (N-PC/CP), S-doped porous carbon on carbon paper (S-PC/CP), and the porous carbon on carbon paper (PC/CP). (d) and (e) Are reproduced with permission.18 Copyright 2024, Elsevier. (f) Ex situ Raman spectra of a K–CO2 battery during discharge and charge. (g) In situ DEMS results for the K–CO2 battery with a B-NCNT cathodes under discharge current of 200 μA within a capacity limit of 1000 μAh. The green, orange, and purple dotted lines represent theoretical gas evolution rates for CO2, CO, and O2, respectively, corresponding to potential reaction pathways. (f) and (g) Are adapted with permission.20 Copyright 2024, Wiley. |
In summary, Li–CO2 batteries exhibit several notable advantages: (1) a high theoretical energy density (∼1876 W h kg−1), which makes them highly attractive for long-duration and high-capacity energy storage applications; (2) the direct utilization of CO2 as a reactant, offering a promising route for carbon mitigation and contributing to carbon neutrality goals; and (3) the use of a lightweight lithium metal anode, which helps reduce overall system weight while enhancing the energy output per unit mass. However, despite these merits, the practical application of Li–CO2 batteries is currently hindered by several critical challenges: (1) poor rechargeability arising from the formation of electrochemically inert and insulating Li2CO3 discharge products, which impede reversible reactions; (2) low energy efficiency and short cycle life, mainly due to sluggish reaction kinetics and side reactions; and (3) significant safety concerns, including dendrite growth on the lithium anode and its interaction with electrolytes and impurities (such as moisture or oxygen). Therefore, overcoming these limitations through material innovation, electrolyte engineering, and system design will be essential for enabling Li–CO2 batteries to emerge as a viable and sustainable alternative to conventional lithium-ion battery technologies.
In comparison to Li–CO2 batteries, Na–CO2 batteries offer a compelling alternative due to the abundance and low cost of Na metal. In Na–CO2 batteries, the formation and decomposition of carbonate and carbon represent the predominant and most widely reported electrochemical reactions.17,47 A representative reaction, 4Na + 3CO2 → 2Na2CO3 + C (ΔrG° = −905.6 kJ mol−1), exhibits a lower Gibbs free energy compared to the analogous reaction in Li–CO2 batteries (eqn (3), ΔrG° = −1081 kJ mol−1). This reduced reaction Gibbs free energy in Na–CO2 systems suggests the potential for achieving lower charge voltages, thereby minimizing the risk of electrolyte decomposition during cycling. Hu et al. developed a rechargeable room-temperature Na–CO2 battery, composed of a sodium metal anode, a glass fiber separator, an ether-based electrolyte, and a tetraethylene glycol dimethyl ether-treated multi-walled carbon nanotube (t-MWCNT) cathode supported on a binder-free Ni mesh.17 The battery demonstrated an impressive reversible capacity of 60 000 mA h g−1 at 1 A g−1 (corresponding to approximately 1000 W h kg−1), and maintained stable cycling over 200 cycles with a controlled capacity of 2000 mA h g−1 and a charge voltage of below 3.7 V. Recently, Wang et al. developed a nitrogen and sulfur co-doped porous carbon cathode featuring high heteroatom content and a moderately positioned p-band center for application in Na–CO2 batteries.18 The active sites of the carbon catalyst are engineered through the incorporation of nitrogen or sulfur heteroatoms, which modulate the p-band center to tailor orbital hybridization, thereby enhancing the kinetics of both CO2 reduction and evolution reactions (Fig. 2d and e). This system demonstrated exceptional cycling performance of over 1000 hours with a small voltage gap of just 1.04 V and a high energy efficiency of 71.2% at a current density of 10 μA cm−2. These results represent a remarkable improvement over most Li–CO2 batteries that operate via the carbonate and carbon formation pathway, underscoring the potential of Na–CO2 systems for CO2 utilization in energy storage. Despite a relatively lower theoretical energy density of 1125 W h kg−1, their economic viability and material sustainability have drawn significant interest from both scientific and industrial communities.
K–CO2 batteries have recently emerged as a promising alternative to Li–CO2 and Na–CO2 systems. Similarly, K–CO2 batteries utilize the electrochemical conversion of CO2 to K2CO3 and carbon.21 The standard redox potential of K (K/K+ = −2.92 V vs. SHE) enables a high theoretical discharge voltage of 2.48 V (based on K2CO3 formation), surpassing that of Na–CO2 batteries (2.35 V). Beyond its favorable thermodynamics, K+ offers additional intrinsic advantages. In comparision to Li+ and Na+, K+ exhibits weaker Lewis acidity and a smaller Stokes radius (3.6 Å in propylene carbonate), which facilitating faster ion migration in the electrolyte and across the electrode–electrolyte interface.48 This kinetic advantage, coupled with the earth abundance and low cost of K, positions K–CO2 batteries as a promising solution for scalable and economically viable CO2-based energy storage system. Zhang et al. developed a K–CO2 battery using bifunctional metal-free 3D porous carbon electrocatalysts (N doped carbon nanotubes (N-CNTs)) for CO2 reduction and evolution.49 The 3D carbon network cathode effectively prevents graphene restacking, enhances active site exposure, offers space for discharge product deposition, and facilitates efficient electron, ion, and CO2 transport. As a result, the battery exhibits high reversibility over 250 cycles at the limited specific capacity of 300 mA hg−1. Li et al. reported a K–CO2 battery involving passivated K anodes with artificial surface film and N-doped carbon nanotubes cathodes.20 As shown in the Raman spectra (Fig. 2f), the D and G bands of carbon at 1335 and 1560 cm−1 are consistently observed throughout the redox process. A distinct peak at 1030 cm−1, attributed to the vibrational modes of K2CO3, appears during discharge. In situ DEMS analysis (Fig. 2g) confirms that CO2 consumption begins with discharge and closely follows the theoretical reaction: 4K+ + 3CO2 + 4e− → 2K2CO3 + C. Notably, CO evolution is negligible, indicating that CO2 reduction proceeds predominantly via K2CO3 and C formation. This battery achieves a high specific discharge capacity (9436 mA h g−1), maintains excellent rate performance with a small voltage gap (0.81 V at 50 mA g−1), and demonstrates excellent durability, lasting 450 cycles or 3100 hours under a limited capacity of 500 mA h g−1. Despite these achievements, K–CO2 systems are still in the early stages of development. Limited by the lack of efficient bifunctional catalysts, their electrochemical performance remains unsatisfactory, and fundamental reaction mechanisms are yet to be fully understood. Nevertheless, the K–CO2 battery shows strong potential as a future low-cost and sustainable energy storage system.
Mg–CO2 batteries offer a competitive advantage due to the high theoretical volumetric capacity of Mg (3833 mA h cm−3) and abundant resources, allowing for more CO2 fixation per cubic meter at a lower cost. Additionally, the dendrite-free nature of magnesium alleviates safety concerns. In Mg–CO2 batteries, CO2 can be converted into carbonate and carbon in organic electrolyte systems. For example, Zhang et al. developed a moisture-assisted Mg–CO2 battery that showed over 250 hours of cycle life and maintained a discharge voltage above 1 V at 200 mA g−1.22 Theoretical calculations indicate that MgCO3·3H2O forms and decomposes more easily than MgCO3, supported by negative adsorption energy and stretched bonds. Raman spectroscopy confirmed the presence of CO32− at 1120 cm−1 and carbon D and G bands at 1350 cm−1 and 1580 cm−1, respectively, validating the reaction pathway in eqn (10). However, research on Mg–CO2 batteries is still limited, and improving the cycling efficiency and energy density remains one of the key challenges in designing high-performance Mg–CO2 batteries.
Reversible Al–CO2 batteries based on the reaction in eqn (11) have demonstrated the feasibility of converting CO2 into carbonate and carbon. Al not only exhibits a much higher intrinsic safety than Li, Na, and K, but also offers a high theoretical specific capacity of 2978 mA h g−1, positioning it as a promising candidate for metal-CO2 battery systems. Ma et al. designed a rechargeable Al–CO2 battery employing Al foil as the anode, an ionic liquid electrolyte, and Pd-coated nanoporous gold (NPG@Pd) as the cathode catalyst.23 This system delivered an exceptionally low voltage gap of 0.091 V and a high energy efficiency of 87.7%, indicating remarkable electrochemical performance. Nonetheless, its practical application remains hindered by limited cycling stability and an unclear reaction mechanism, with insufficient mechanistic insight from in situ or operando characterizations. Therefore, the rational design of high-performance bifunctional catalysts and the elucidation of fundamental reaction pathways will be critical to unlocking the full potential of Al–CO2 batteries.
2Mg + 3CO2 + 6H2O ↔ 2MgCO3·3H2O + C | (10) |
4Al + 9CO2 ↔ 2Al2(CO3)3 + 3C | (11) |
In Li–CO2 batteries, since the initial report by Hou et al. in 2017, which demonstrated that Mo-based catalysts enabled the formation of Li2C2O4 as the sole discharge product, significant efforts have been devoted to stabilizing oxalate intermediates to modulate the reaction pathway and enhance battery performance.25 To date, researchers have successfully employed strategies involving catalyst design, redox mediators (RMs), and electrolyte engineering to benefit the selective formation of Li2C2O4, thereby achieving substantial improvements in energy efficiency, reversibility, and cycling stability.
Recent studies have demonstrated that Mo-based catalysts, such as Mo2C, MoN, and Mo3P, are highly effective in stabilizing Li2C2O4 as the primary discharge product in Li–CO2 batteries. The formation of a Mo–O coupling bridge plays a critical role in preserving Li2C2O4 and enhancing reaction kinetics. For instance, Mo2C catalysts have been extensively investigated through experimental and computational approaches, showing exclusive Li2C2O4 generation, a reduced charging voltage of 3.5 V, and an energy efficiency of 77%. DFT calculations by Yang et al. revealed that the disproportionation of Li2C2O4 is thermodynamically unfavorable on Mo2C, ensuring Li2C2O4 stability.51 Similarly, Cheng et al. developed MoN nanofibers on carbon cloth (CC@MoN NFs), where the delocalized electrons of Mo3+ stabilize Li2C2O4 through weak Mo–O bonds, achieving a low charge potential of 3.19 V, 88% energy efficiency, and >500 hours of cycle life.26 Moreover, a Mo3P/Mo Mott–Schottky heterojunction catalyst reported by Wang et al. delivered a high specific capacity of 10 577 mA h g−1, a low overpotential of 0.15 V, and high energy efficiency of 94.7% by promoting reversible Li2C2O4 conversion and accelerating interface kinetics.52 These findings collectively highlight the critical role of Mo-based catalysts in effectively stabilizing oxalate intermediates and advancing the development of high-performance Li–CO2 batteries.
The use of RMs has emerged as another effective strategy to promote Li2C2O4 formation and enhance the electrochemical performance of Li–CO2 batteries. In a typical RM-assisted discharge process, the RM first reacts with CO2 to form an RM-CO2 adduct, which subsequently undergoes reduction to regenerate the RM and yield Li2C2O4. Acting as a liquid catalyst, the RM enables multidirectional interaction with CO2, thereby significantly accelerating cathode reaction kinetics. Typical RMs, such as tris(2,2′-bipyridyl)-dichloro-ruthenium(II) (Ru-(bpy)3Cl2), binuclear copper(I) complexes, cupric chloride (CuCl2), and 2,2,6,6-tetramethylpiperidoxyl, have been demonstrated to stabilize Li2C2O4.53–56 The advantages of RM-assisted discharge/charge process include ultra-low overpotentials and exceptionally high energy efficiencies. For instance, Wang et al. demonstrated that the 2,2,6,6-tetramethylpiperidoxyl RM enabled the preferential formation of Li2C2O4, delivering an outstanding discharge voltage of 2.97 V and a low charge voltage of approximately 3.15 V across a wide current range (100 to 800 mA g−1), indicating excellent rate capability.56 However, the application of soluble RMs also faces challenges, such as the shuttle effect and parasitic reactions with the anode, which can deteriorate battery stability over time. To overcome the shuttle effect of soluble RMs and enhance battery stability, Li et al. developed a solid-state redox mediator anchored onto the cathode via a Cu(II) coordination compound with benzene-1,3,5-tricarboxylic acid (Fig. 3a).24 This design enables efficient CO2 capture and promotes Li2C2O4 formation through a dimerization mechanism (Fig. 3b). Because of the solid nature of the RMs, no shuttle effect occurs in this system. The resulting Li–CO2 battery exhibits low overpotential (0.9 V), and significantly improved cycling stability of over 400 cycles, advancing metal–CO2 batteries toward practical application.
![]() | ||
Fig. 3 Overview of oxalate-based discharge products in MCBs. (a) Schematic of a Li–CO2 battery employing a solid RM(II)-BTC cathode. The solid-state nature of RM(II)-BTC suppresses the shuttle effect, while its high CO2 adsorption capacity, fast redox kinetics, and reversibility enable efficient battery performance. (b) Proposed reaction pathway for oxalate (Li2C2O4) formation mediated by RM(II)-BTC. Upon discharge, RM(II)-BTC is first reduced to Li2RM(I)-BTC, which subsequently reacts with CO2 to form Li2RM(II)-BTC-(CO2)2−. This intermediate then decomposes to yield Li2C2O4 and regenerate RM(II)-BTC. The dashed frame illustrates the transition of RM(II)-BTC into its reduced lithium-containing form. Atom colours: gray (C), red (O), cyan (Cu2+), green (Cu+), purple (Li+), blue (e−). (a) and (b) Are adapted with permission.24 Copyright 2024, Springer Nature. (c) Electrochemical reaction mechanism of CO2RR in a quasi-solid-state Na–CO2 battery with Co-encapsulated N-doped carbon framework (Co-NCF) catalysts and gel electrolyte. (d) In situ Raman spectra of the quasi-solid-state Na–CO2 batteries, revealing the generation of C2O42−. (c) and (d) Are adapted with permission.27 Copyright 2022, Cell Press. (e) Schematic of a Mg–CO2 battery utilizing 1,3-propylene diamine (PDA) as a bifunctional additive, facilitating both CO2 capture and conversion. (f) Raman spectra of the pristine and discharged Ni@Ru cathodes. (e) and (f) Are adapted with permission.29 Copyright 2024, Wiley. |
The stabilization of oxalate can also be achieved by electrolyte engineering. In a high-concentration LiTFSI/DMSO-based electrolyte, solid Li2CO3 and carbon products are formed. Wang et al. optimized the electrolyte composition to a low-concentration dimethyl sulfoxide (LC-DMSO) electrolyte (LiTFSI:
DMSO = 1
:
12), resulting in switching the discharge products to oxalate.57 Spectroscopic and electrochemical analyses demonstrated that the LC-DMSO electrolyte facilitates the formation of C2O42− intermediates through interactions with dissolved CO2 molecules. Owing to the high donor number of DMSO (29.8), a stronger solvation sheath around Li+ ions is constructed, which effectively prevents direct association between Li+ and the soluble C2O42− intermediates. As a result, a reduced charge potential of 3.4 V and sustained cycling of over 260 hours are achieved. This study provides valuable insights for electrolyte design in regulating the discharge products in metal–CO2 batteries.
Oxalate discharge products have also been reported in Na–CO2 batteries, although relatively less frequently. This is likely due to the limited and still-developing research on Na–CO2 batteries. Compared to conventional discharge products such as sodium carbonate and carbon, sodium oxalate is more readily decomposed, leading to a significant reduction in the charging voltage. Amin et al. employed a bimetallic layered double hydroxide catalyst on the cathode of Na–CO2 batteries and observed the formation of sodium oxalate after discharge.28 The cell demonstrated stable operation for up to 100 hours before degradation occurred. Furthermore, by adopting either a prolonged charging strategy or an asymmetric cycling protocol, the cell could be effectively revived from its degraded state. Xu et al. further advanced the development of Na–CO2 batteries by constructing a reversible quasi-solid-state system capable of forming stable Na2C2O4 discharge products.27 Using a Co-encapsulated N-doped carbon framework catalyst and a gel electrolyte containing an imidazole-based organic cation, the system effectively stabilized the thermodynamically unstable Na2C2O4 (Fig. 3c). The electron-agglomeration effect of Co nanoparticles enhanced CO2 adsorption and promoted Na2C2O4 formation, which is confirmed by the in situ Raman results (Fig. 3d). As a result, the battery delivered a high discharge capacity of 3094 mA h g−1, maintained 1777 mA h g−1 at 0.5 mA cm−2, and achieved excellent cycling stability of over 366 cycles (2200 hours). These works highlight the critical value of designing Na–CO2 battery systems capable of producing sodium oxalate, offering new avenues for improving battery reversibility and efficiency.
Oxalates have also been reported in Mg–CO2 and Al–CO2 batteries, though their formation requires specific conditions. For instance, Peng et al. designed a highly reversible and high-rate Mg–CO2 battery using a liquid 1,3-propylene diamine (PDA) additive in conventional electrolytes(Fig. 3e and f).29 At the cathode, PDA enhances Mg2+ de-solvation, CO2RR kinetics, and the formation of the decomposable discharge product MgC2O4. At the anode, it promotes the formation of a Mg2+-conductive solid electrolyte interphase (SEI), enabling highly reversible Mg plating/stripping. This approach results in a Mg–CO2 battery with over 400 hours lifespan at 200 mA g−1 and excellent rate performance from 100 to 2000 mA g−1. Fetrow et al. developed a rechargeable Al–CO2 battery that produces oxalate discharge products by using aluminium iodide as a redox mediator.30 This innovation enables the battery to achieve an ultralow overpotential of 0.05 V. Additionally, the Al–CO2 cell maintains a high discharge voltage of 1.12 V and achieves a capacity of 9394 mA h g−1. The aluminium oxalate, confirmed by the nuclear magnetic resonance (NMR) analysis, contributes to the improvement of battery reversibility. These findings highlight the potential of oxalate-based discharge products in improving the performance and reversibility of Mg–CO2 and Al–CO2 batteries.
CO2 + e− → CO2− | (12) |
CO2− + CO2 → C2O42− | (13) |
2Li+ + 2CO2 + 2e− → Li2CO3 + CO(g) | (14) |
Based on the nature of the charge–discharge reactions on the cathode, Zn–CO2 batteries can be categorized into reversible and non-reversible types. In reversible Zn–CO2 batteries, the discharge reaction involves the reduction of CO2 into value-added chemicals, with formic acid (HCOOH) being one of the most successfully demonstrated products to date. During discharge, HCOOH is generated at the cathode and dissolves into the aqueous electrolyte, following eqn (15). During the subsequent charging process, HCOOH undergoes oxidation back to CO2 (the reverse reaction of eqn (15)), thereby completing a reversible cycle. Wang et al. developed a Zn–CO2 battery using a dual-anion regulated Bi electrocatalyst (Fig. 3a).59 The BiOF catalysts achieved highest FEHCOO- value at −1.8 V is 97% (Fig. 3b) and exhibited stability against high and low current densities (Fig. 3c). Moreover, Xie et al. developed a reversible aqueous Zn–CO2 battery using a bifunctional Pd cathode with a 3D porous nanosheet structure.33 The system achieved 90% FE for CO2 reduction to HCOOH and stable battery performance, delivering high energy efficiency (81.2%), a low charge voltage (1 V), and excellent cycling stability over 100 cycles, demonstrating the feasibility and outstanding of this Zn–CO2 system.
CO2 + 2H+ + 2e− → HCOOH | (15) |
In non-reversible Zn–CO2 batteries, the discharge process still involves the CO2 reduction to valuable chemicals or fuels, but these chemicals and fuels can be separated timely to avoid being consumed during charging process. In this circumstance, the water molecules can be oxidized to form oxygen gas at the cathode surface (eqn (16)).35 This reaction pathway design removes the need for re-oxidizing discharge products, allowing the generation and separation of value-added products, thereby positioning the Zn–CO2 cell as both an energy storage device and a chemical synthesis platform. Miao et al. developed bifunctional electrocatalysts comprising Ni clusters (Nix) coupled with single Ni sites (Ni–N4/Nix), for CO2 reduction and oxygen evolution in aqueous rechargeable Zn–CO2 batteries (Fig. 4d).34 Theoretical calculations (Fig. 4e) and in situ FTIR results (Fig. 4f) reveal that the *COOH intermediate of the electrochemical CO2 reduction exhibits a synergistic binding with both Nix clusters and Ni–N4–C single-atom sites. This dual–site interaction facilitates CO2 activation and effectively lowers the energy barrier of the potential-determining step. This catalyst achieved nearly 100% FE for CO production from −0.4 to −0.8 V vs. RHE, outperforming Ni–N4–C (55.0%). In addition, this bifunctional catalyst exhibited OER activities superior to commercial RuO2, while exceeding that of Ni–N4–C. This Zn–CO2 battery delivered a peak power density of 11.7 mW cm−2 and stable cycling over 1200 cycles (420 h), highlighting the importance of rational bifunctional catalyst design for enhancing non-reversible Zn–CO2 battery performance.
H2O → 1/2O2 + 2H+ + 2e− | (16) |
![]() | ||
Fig. 4 Summary of the MCBs producing value-added carbon-containing compounds. (a) Schematic of a rechargeable aqueous Zn–CO2 batteries with a BiOF-based cathode. (b) Faradaic efficiencies for formate production on various catalysts, including BiOF, BiF3, and Bi2O3 catalysts. (c) Galvanostatic charge–discharge profiles at various current densities. (a)–(c) Are adapted with permission.59 Copyright 2024, Wiley. (d) Schematic of an alkaline Zn–CO2 battery employing a Ni–N4/Nix cathode and 1 M KOH catholyte. (e) Calculated free energy diagrams for electrochemical CO2 reduction to CO reaction on Ni–N4, Ni–N4/Ni5, Ni–N4/Ni8, and N4/Ni5 catalyst at U = 0 V vs. RHE. (f) In situ FTIR spectra of Ni–N4/Ni5 catalysts collected across a potential range from open circuit potential (OCP) to −0.8 V vs. RHE. (d)–(f) Are adapted with permission.34 Copyright 2023, Wiley. |
At the anode, the anodic reactions during both discharge and charge are governed by the Zn/Zn2+ redox couple, with the specific processes influenced by the electrolyte pH. In near-neutral electrolytes, Zn is oxidized to Zn2+ (eqn (17)), where Zn2+ acts as the charge carrier. In alkaline electrolytes, Zn is first oxidized to Zn(OH)42− (eqn (18)), followed by the decomposition of Zn(OH)42− into ZnO, OH−, and H2O. These anodic processes are critical for maintaining charge balance and enabling reversible energy storage.
Zn → Zn2+ + 2e−, E0 = −0.76 V vs. SHE | (17) |
Zn + 4OH− → Zn(OH)42− + 2e−, E0 = −1.22 V vs. SHE | (18) |
In systems that generate solid-state discharge products, such as Li–CO2 batteries, the formation of carbonate and carbon results in exceptionally high theoretical energy density (1876 W h kg−1). These systems typically exhibit notable discharge overpotentials, but also possess the potential for reversible cycling, especially when supported by advanced catalysts and tailored electrolytes. The combination of high energy output and rechargeability positions these batteries as strong candidates for electric vehicles (EVs), stationary grid storage, military power supplies, and long-duration applications such as space exploration or underwater systems. Crucially, the ability to store energy while consuming CO2 in these systems highlights their dual value: CO2 utilization and enabling high-performance energy solutions.
In contrast, MCBs that produce liquid-phase or gas-phase discharge products, such as Zn–CO2 batteries, often follow irreversible reaction pathways, yet offer considerable advantages in terms of safety in aqueous electrolyte and effective CO2 conversion. Their operational characteristics make them especially well-suited for low-power and intermittent-use applications, including wearable electronics, remote communication devices, environmental monitoring sensors, and portable control systems. Importantly, the CO2 consumed in these batteries is not only utilized but also transformed into valuable chemical or fuel products, contributing directly to the development of technologies that reduce more CO2 than they produce. By designing systems that allow the efficient separation of liquid products from electrolytes or the capture of gaseous products from flowing streams, these batteries can function as both energy reservoirs and chemical production units. This dual functionality significantly enhances the economic value proposition, enabling the recovery of value-added fuels or industrial feedstocks directly from waste CO2.
In summary, MCBs not only offer a versatile platform for energy storage but also present a transformative pathway for harnessing CO2 as a resource, aligning with global efforts toward net-zero carbon technologies and sustainable energy systems.
Limited electrocatalytic activity and the need for advanced bifunctional catalysts. A major limitation of current MCBs lies in the low catalytic activity of cathode for both CO2 reduction and the reverse oxidation of discharge products, leading to sluggish kinetics and large overpotentials. To overcome this, there is an urgent need to develop highly active and durable bifunctional catalysts that can efficiently facilitate both the discharge and charge processes. Recent advances in artificial intelligence (AI), particularly machine learning (ML), offer a powerful approach to accelerate catalyst discovery. By integrating computational screening, data-driven modeling, and experimental validation, machine learning can help identify optimal catalyst compositions, structures, and surface properties, greatly enhancing the design efficiency and performance predictability of next-generation electrocatalysts.
Interface issues associated with solid-state discharge products. In systems like Li–CO2 batteries, the discharge products (e.g., Li2CO3) are typically solid and electronically insulating. Their deposition on the cathode surface leads to poor solid–solid interfacial contact, severely impeding reaction kinetics and reversibility. To address this, effective mobile species, such as redox mediators should be developed. These mediators can transfer electrons between the electrode and the discharge product, facilitating decomposition at lower overpotentials. Moreover, they can help alleviate interfacial resistance and improve charge transport in systems dominated by solid-phase reactions, offering a promising route toward more reversible and efficient MCBs.
Low productivity of chemical/fuel-producing MCBs and product separation challenges. In MCBs that can convert CO2 into valuable chemicals or fuels (such as formate, methane, or syngas), the inherently low current densities limit the total amount of products generated. This increases the relative cost of product separation and collection, particularly in flow systems. To address this bottleneck, it is critical to develop flow-type cells with optimized design parameters, including current density, electrolyte flow rate, and gas flow rate, to ensure high conversion efficiency and selective separation of products without affecting the stable energy storage role.
Energy storage system | Electrodes | Mechanism | CO2 adsorption capacity | Energy consumption | Performance | Ref. | ||
---|---|---|---|---|---|---|---|---|
No. of cycles | Current/voltage | Coulombic efficiency | ||||||
Battery | Cathode: CO2-binding polyanthraquinone-carbon nanotube (Q-CNT) coated substrate | Electro-swing for CO2 capture/release during battery charging/discharging | 7.4 μmol of CO2 (at 1.3 V) | 43 kJ mol−1 (at 60% bed utilization) | 7000 (<30% capacity loss) | −1.8 V | Faradaic efficiency ∼90% | 62 |
Anode: ferrocene-CNT (Fc-CNT) coated substrate sandwiched between two cathodes | ||||||||
Supercapacitor | YP80F, a biowaste-derived activated carbon electrode | pH swing for CO2 capture/release | 170 mmol CO2 per kg of electrode (30 mA g−1, pure CO2) | 18 kJ molCO2−1 (300 mA g−1) | 12 000 (>2500 h, pure CO2) | 150 mA g−1 | > 99.8% (under 15% O2) | 63 |
Positive charging mode for supercapacitor operation | ||||||||
Aqueous flow-battery | Negolyte: sodium 2,2′-(phenazine-1,8-diyl)bis(ethane-1-sulfonate) (1,8-ESP) | pH swing for CO2 capture/release | 1.41 mol L−1 (10 mA cm−2) | 55 kJ molCO2−1 (10 mA cm−2) | 220 cycles for CO2 capture/release (18 days, no capacity decay) | 20 mA cm−2 | 95–82% (under 3–20% O2) | 64 |
Battery charging: CO2 capture and electricity storage | ||||||||
Posolyte: ferricyanide/ferrocyanide redox couple (Fe(CN)63−/Fe(CN)64−) | Battery discharging: CO2 release and electricity delivery | |||||||
Vanadium redox-flow battery | Posolyte: vanadium redox couple (VO2+/VO2+) | pH swing for CO2 capture/release | — | 54 kJ mol CO2−1 | 10 day–night cycles for CO2 capture/release | 0.5 V (daytime) and −0.5 V (night time) | — | 61 |
Negolyte: ferricyanide/ferrocyanide redox couple (Fe(CN)63−/Fe(CN)64−) | CO2 release: oxidation of VO2+ to VO2+ in the posolyte, reduction of Fe(CN)63− to Fe(CN)64− in the negolyte and electricity storage (daytime) | |||||||
CO2 sorbent: potassium carbonate (K2CO3) in a separate absorption column | CO2 capture: reduction of VO2+ back to VO2+ in the posolyte, oxidation of Fe(CN)64− to Fe(CN)63− in the negolyte and electricity delivery (nighttime) |
The pH-swing CO2 capture can be integrated with existing energy storage technologies to improve both the carbon capture and grid storage (Fig. 5a). For example, the process of charging and discharging the redox-flow battery can align with the pH swing cycle i.e. storing energy during the CO2 release phase and releasing energy during the CO2 capture phase.69 Also, supercapacitors, which store energy via electrostatic charge, could be used in combination with pH-swing systems to serve this purpose.63 In this configuration, a supercapacitor could be used to control the electrochemical reactions driving the pH swing, while simultaneously providing fast energy storage or supply to the grid.
The electrochemical cell used for CO2 capture can serve as an energy buffer, storing renewable energy during off-peak hours and releasing energy during peak demand.66 This buffering function enables the decoupling of CO2 capture from the grid and allows the system to operate in a more flexible and efficient manner. For example, integrating the electro-swing CO2 capture system with redox flow batteries presents an innovative approach. In this system, the electrochemical sorbents could be used as the electrolyte in the flow battery, allowing the system to function as both a CO2 capture device and an energy storage system. This would allow for continuous operation where energy is stored and released in a closed loop. Supercapacitors can also be used in combination with the electrochemical sorbent regeneration system.
The CO2-capture integrated energy storage systems have already started to progress in grid-scale storage applications. For instance, pH swing-based CO2 capture integrated energy storage is reported using an aqueous flow cell system.64 A redox-active phenazine derivative (negolyte) on the negative side of cell (sodium 2,2′-(phenazine-1,8-diyl)bis(ethane-1-sulfonate) (1,8-ESP)) allowed PCET for pH swing-based CO2 capture. When paired with ferrocyanide (posolyte) on the positive side, this cell enabled energy storage function. The negolyte pH was increased during charging, resulting in reaction between CO2 and OH− ions, enabling CO2 capture (1.4 molCO2 L−1 at 10 mA cm−2) and energy storage. This process is reversed during discharging due to the decreased pH, aiding CO2 release and energy delivery. In simple terms, deacidification during charging results in CO2 capture and energy storage, while acidification during discharging releases pure CO2 and stored energy (Fig. 6a and b). This system could flexibly work as a pure energy storage device (open-circuit voltage: 1.1 V; coulombic efficiency: 99.9%; round-trip energy efficiency: 78.5%; capacity retention: 95% for 170 days (fade rate ∼0.05% per day)) by isolating the electrolyte from CO2, depending on the market requirements.
![]() | ||
Fig. 6 (a) CO2 capture-energy storage and (b) CO2 release-energy delivery in phenazine-based electrochemical cell during charging and discharging process respectively. Reproduced with permission.64 Copyright 2023, Springer Nature. Schematic representation of vanadium-based redox flow battery for integrated CO2-capture and renewable energy storage showing (c) energy storage and CO2 release during daytime and (d) energy supply to grid and CO2 capture during night. Reproduced with permission.61 Copyright 2025, American Chemical Society. (e) Challenges for current performance and projected applications in CO2-capture integrated energy storage technology. |
However, improving the stability of organic redox molecules in presence of O2 (present in flue gas ∼3–5%) or (air ∼20%) remains crucial, since O2 can oxidise these molecules during prolonged cycling. In this case, the chemical oxidation of negolyte (1,8-ESP) due to O2 was reversed by applying reduction potential. But it caused the oxidation of posolyte (ferrocyanide), requiring electrochemical rebalancing by applying reverse potentials.86
Lately, vanadium redox flow batteries have also been employed for CO2 capture-based energy storage (Fig. 6c and d).61 It establishes (a) absorbent and pH-based CO2 capture/release, (b) battery charging during daylight hours (when renewable electricity is available) with CO2 release and (c) CO2 capture and electricity delivery to the grid during night (when solar energy is inaccessible). The membrane-separated electrochemical cell contained vanadium redox (VO2+/VO2+) posolyte and ferricyanide/ferrocyanide redox (Fe(CN)63−/Fe(CN)64−) negolyte. An absorption column containing potassium carbonate (K2CO3) solution can capture CO2 in the form of dissolved HCO3− or CO32− ions, which is fed to negolyte. During daytime i.e. CO2 release and battery charging, VO2+ and H2O undergo oxidation to generate VO2+, H+ and e− which migrate towards negolyte (reduction), decreasing its pH and thereby releasing absorbed CO2. Upon reversing the potential, VO2+ is reduced to VO2+ in posolyte due to negolyte oxidation, while simultaneous increase in pH results in CO2 capture and delivery of the stored energy to the grid during night. It is noteworthy to mention that the vanadium redox couple in this system is oxygen insensitive while the decoupling of negolyte and CO2 absorption column suppressed the adverse effects of O2 impurities.
For real-time applications, such as grid-scale energy storage, the main challenge lies in the intermittent nature of renewables (e.g., solar, wind) which needs to be balanced for energy demand and supply.88 Further, the CO2 release process also require continuous energy supply.89 Both challenges necessitate an efficient and long-cycling energy storage device which can provide grid stability. On the other hand, application in smart buildings equipped with HVAC systems require sophisticated control systems, large space, high efficiency and durability sustainability and economic feasibility of such systems. As far as the energy-intensive industries are concerned, practical concerns regarding infrastructure, space requirements and effective storage and transport of CO2 needs to be considered. Overall, the integrated CO2-capture and energy storage system represents a highly complex system with high projected capital costs, necessitating improved energy efficiency and CO2 capture-release performance with minimal energy consumption. Continued research in materials development and performance engineering is the need of the hour.90 The use of solid-state CO2 capture materials with improved electrochemical stability over a wide range of applied potentials show promise while extending the research for other potential energy storage technologies (e.g., batteries, supercapacitors) can improve the commercial viability. Technical challenges including CO2 management and possible safety issues also require advanced monitoring and safety protocols, but at a later stage.
CH4 + H2O → CO + 3H2 | (19) |
CO + H2O → CO2 + H2 | (20) |
These reactions are endothermic, particularly in the steam reforming step (+206 kJ molCH4−1), but the required heat is partially offset by the exothermic electrochemical oxidation of H2. This thermal balance helps maintain the high operating temperature of the cell (600–650 °C) without external heating. The H2 produced is partly used in the electrochemical reaction to generate electricity, while the CO2 remains in the anode stream for downstream capture. This coupling of reforming and electrochemical processes allows MCFCs to generate power and provide enrich CO2 simultaneously. By adjusting conditions such as fuel utilization and temperature, the anode acts as an efficient zone for both fuel conversion and CO2 management, giving MCFCs a distinct advantage over conventional power or capture systems.
H2 + CO32− → H2O + CO2 + 2e− | (21) |
CO2 + ½O2 + 2e− → CO32− | (22) |
CO2 (cathode) + H2 (anode) + ½O2 → H2O + CO2 (anode) + electricity + heat | (23) |
MCFCs capture CO2 as an intrinsic part of the electricity generation process. For each methane molecule consumed in the anode, four CO2 molecules are drawn into the system at the cathode, and five are released at the anode.93 This results in a net transfer of CO2 from the external flue gas into a concentrated stream suitable for sequestration or utilization. Different from traditional carbon capture methods that separate CO2 post-combustion with a parasitic energy penalty, MCFCs utilize CO2 as an electrochemical reactant, enabling both energy production and CO2 concentration in a single step.
A typical molten carbonate fuel cell (MCFC) consists of a porous anode, a porous cathode, and a porous matrix that contains the molten carbonate electrolyte. The system relies on capillary pressure equilibrium among the three porous components. As illustrated in Fig. 7b, this equilibrium condition requires that the capillary pressures in the anode, cathode, and matrix be equal.92 When this balance is achieved, the distribution of the electrolyte becomes stable for a given total volume of melt. This distribution is governed by the total pore volume and pore size distribution of each component. Moreover, the most electrochemically active region within the anode and cathode of an MCFC is the three-phase boundary (TPB), where the gas, electrolyte, and solid phase electrodes coexist.92 For example, in porous nickel anodes in Fig. 7c, for instance, the TPB is located near the top of the electrolyte meniscus within the pores. The contact angle (θ) at this interface is a critical parameter, as it determines the extent to which reactant gases can penetrate into the pores and participate in electrochemical reactions. Ideally, the entire pore wall would be uniformly covered by a thin film of molten carbonate, allowing full utilization of the internal surface for reaction. Achieving this requires the contact angle θ to approach zero. Consequently, under practical conditions, only a portion of the internal surface area of the MCFC anode is electrochemically active. Realizing such optimal wetting conditions calls for careful interfacial engineering of the electrode materials to ensure a stable TPB and consistent electrochemical performance under high-temperature operating conditions.
![]() | ||
Fig. 7 Summary of the MCFC systems. (a) Schematic of MCFC system, adapted with permission.91 Copyright 2020, Elsevier. (b) Schematic illustration of in-pore meniscus formation and wetting behavior, highlighting the capillary equilibrium condition characterized by the balance of interfacial tensions among the three porous components. (c) Details of the three phase boundary (TPB). (b) and (c) Are adapted with permission.92 Copyright 2020, Elsevier. Sankey-style diagrams illustrating the molar flow distributions for high (d) and low (e) fuel utilization conditions. The width of each arrow corresponds to the respective molar flow rate, with the exception of hydrogen (H2), whose flow width has been scaled down by a factor of 3.3 to place it on a comparable chemical energy basis with methane (CH4), enabling clearer visual comparison. The data in (d) and (e) are reproduced with permission.93 Copyright 2022 Elsevier. |
The energy consumption is an important component of the overall value proposition of the MCFC systems. Fig. 7d and e compared the fuel consumption and the corresponding CO2 for sequestered and H2 for recycling.93 In Fig. 7d, the MCFC captures CO2 through two sequential processes, with the co-produced H2 recycled back into the fuel system. In the first stage, 100 mol of CH4 is consumed, generating 100 mol of CO2 for capture. The MCFC is assumed to capture 90% of this CO2, although the capture rate can be adjusted depending on system design. To achieve this, an additional 32 mol of CH4 is processed, all of which is fully captured.
By contrast, the configuration in Fig. 7e adopts a low fuel utilization strategy during the initial stage to maximize H2 production (300 mol) while simultaneously capturing CO2. After separation of H2 and CO2, the resulting hydrogen, characterized by minimal CO2 emissions, is employed as a clean fuel in a secondary process. This approach offers a significant advantage in terms of energy efficiency. The integrated design not only enhances overall hydrogen output but also consolidates CO2 capture in a single intensified unit, comprising the MCFC and H2/CO2 separation modules, thereby reducing system complexity, capital investment, and operational overhead. While the energy penalty may not be lower than that of the configuration in Fig. 7d, the ability to produce low-carbon H2 and streamline process integration positions the design in Fig. 7e as a more energy- and cost-effective solution for sustainable fuel and CO2 management.
Systems | Capacity (MW) | Efficiency | CO2 capture ratio | CO2 emission coefficient (kg of CO2 kW h−1) | Ref. |
---|---|---|---|---|---|
Coal-fired power plant (CFPP)-MCFC | CFPP (450)-MCFC (198) | 45.77% | 76.9% | 0.148 | 102 |
CFPP-MCFC | — | 40% | 39% | 0.253 | 103 |
Direct internal reforming (DIR)-MCFC-steam turbine (ST) | DIR-MCFC (0.5)-MGT (0.1)-ST | 0.52% | — | 0.101 | 104 |
Reciprocating engine (RE)-MCFC | Reciprocating engine (6)-MCFC (1.8) | 41% | 76% | 0.121 | 105 |
Integrated gasification (IG)-MCFC | IG-MCFC (0.2) | 44.7% | — | 0.751 | 106 |
Natural gas combined cycle (NGCC)-MCFC (89) | NGCC (GT: 271, ST: 191)-MCFC (89) | 58% | 80% | 0.071 | 98 |
NGCC-MCFC | NGCC (800)-MCFC (212) | 52.5% | 58.1% | 0.149 | 107 |
MCFC-steam gas turbine (SGT) | MCFC (2.4)-SGT (0.33) | 48.2% | 583 t CO2/y | 0.751 | 108 |
In recent years, the development of MCFCs using green ammonia as fuel has attracted attention due to its potential to reduce fossil fuel dependence. Lu et al. built a single-cell MCFC system to study the electrochemical performance of direct ammonia-fed MCFCs, focusing on key factors such as the NH3/H2 volume flow ratio and the operation temperature.109 Operating at 680 °C, the cell attained its best performance, with a current density of 0.059 A cm−2 and a high ammonia conversion rate of 91.6%. When using NH3/H2 mixed gas, the best results (0.057 A cm−2, 84.9% conversion) were achieved at a volume ratio of 9, indicating that higher H2 content can lower ammonia conversion. SEM analysis confirmed that ammonia does not damage the porous structure of the electrodes, supporting its direct use as a reliable fuel in MCFCs. Another promising fuel for MCFCs is the deployment of bio-based fuels. Dybiński et al. developed a molten carbonate fuel cell (MCFC) directly fuelled by raw bioethanol derived from fruit fermentation without pre-treatment.110 However, solid particle contamination in the fermentation product prevents stable long-term operation of the fuel cell, making purification of the bioethanol necessary before use.
MCFCs can also be integrated with other technologies such as biomass gasification. Tawakoli et al. proposed using sewage sludge as a biomass feedstock to address waste management and reduce CO2 emissions, while supporting sustainable energy production.111 They simulated a biomass gasification system combined with a MCFC in Aspen Plus software, using air and steam as gasifying agents. The study optimized key parameters including the equivalence ratio, steam to biomass ratio, gasification temperature, fuel cell temperature, inlet air flow rate, and system pressure to minimize the levelized cost of electricity and maximize power output and efficiency. The genetic algorithm-based optimization showed that the system performed best at an equivalence ratio of 0.102, steam to biomass ratio of 0.24, gasification temperature of 894 °C, fuel cell temperature of 776 °C, air flow rate of 24.8 kmol h−1, and system pressure of 1.26 bar. Under these parameters, the system delivered a levelized cost of electricity of 0.42 USD kW h−1, a power output of 0.224 W cm−2, and an overall efficiency of 60.1%.
Moreover, a successful example of the potential application of MCFCs is their integration into a hybrid renewable energy system for a four-story building in Bandar Dayyer, southern Iran.112 In this system, biogas produced from sewage digestion was used to fuel a MCFC, which operated at 650 °C and generated 98 kW of electricity. The system also included water desalination, producing over 1100 kg of fresh water per day, meeting both domestic and process demands. With overall energy and exergy efficiencies of 54% and 52.58% respectively, and relatively low CO2 emissions, this application highlights the capability of MCFCs to provide clean electricity, water, heating, and cooling in a sustainable and environmentally friendly manner.
The first challenge in MCFCs is the material degradation. MCFCs operate at elevated temperatures (typically 600–700 °C), which accelerates the degradation of cell components, including electrodes, current collectors, and the electrolyte/electrode interface. Corrosion, thermal stress, and chemical instability significantly limit the system's lifespan and reliability. To address this, the development of advanced materials is essential. Research focuses on corrosion-resistant alloys, thermally stable ceramic matrices, and robust electrode coatings that can withstand harsh conditions. Implementing protective layers and optimizing material selection can significantly enhance the durability and operational life of MCFCs.
When using bio-derived or unprocessed fuels such as biogas or raw bioethanol, impurities like sulfur compounds, tars, or particulates can poison the anode catalyst or clog system components. This leads to rapid performance degradation and operational instability. Incorporating pre-treatment units to remove harmful contaminants and designing reformers that convert complex fuels into clean hydrogen-rich gas ensure stable cell performance. In cases where raw fuels are used, minimal purification steps must be introduced to maintain system feasibility without excessive cost.
Maintaining the integrity of the molten carbonate electrolyte is critical. Issues such as electrolyte leakage, evaporation, or compositional changes can disrupt ionic conductivity and reduce overall efficiency. Stable electrolyte matrices and improved sealing techniques can mitigate these problems. Using lithium/sodium/potassium carbonate mixtures with optimized ratios enhances thermal stability. Additionally, advanced cell designs that confine the electrolyte and reduce evaporation can improve electrolyte retention and system performance.
In summary, while MCFCs offer compelling advantages in terms of efficiency, fuel flexibility, and integrated CO2 capture, their large-scale deployment is still constrained by several interrelated technical and economic challenges. A multidisciplinary approach that combines materials science, electrochemical engineering, and system integration will be essential to unlock the full potential of MCFCs as a cornerstone of future low-carbon energy infrastructure.
Systems | Maximum output voltage | Energy conversion efficiency | Cycling life | Mechanism | Ref. |
---|---|---|---|---|---|
Photosynthesis-based electricity generator | 2.4 V | 38% | 2 h | Coupling natural photosynthesis with mechanically induced electromagnetic generation | 113 |
Nanosheet-agarose hydrogel (NAH)-based electricity generator | 5 V (by connecting 50 of the generator) | 0.6% | 60 h | A spatial nanoconfinement-based ion separation strategy to convert CO2 adsorption into electrical energy | 9 |
The first system achieves CO2-to-electricity conversion by coupling natural photosynthesis with mechanically induced electromagnetic generation.113 In this design, CO2 is dissoved in water at the first step. Then the aquatic plant, Ceratophyllum demersum, performs photosynthesis under illumination, converting dissolved CO2 into O2 bubbles (Fig. 8a). These bubbles accumulate beneath a superhydrophobic cone-shaped cap affixed to a buoyant (Fig. 8b). As buoyancy increases, the device ascends through a surrounding magnetic field. Upon reaching the air–water interface, gas release leads to a rapid density increase and initiates a downward motion. This cyclic surfacing-diving behavior, sustained by continuous photosynthetic activity, produces a time-varying magnetic flux that drives electricity generation via Faraday's law.
![]() | ||
Fig. 8 Summary of CO2-based electricity generation systems. (a) Atmospheric CO2 dissolves in water under ambient conditions, serving as a carbon source for subsequent energy conversion. (b) A sunlight-driven mini-generator utilizes the harvested CO2 to power multiple small electronic devices. (c)–(e) Photographs of the sunlight-powered generator system and its operation. (f) Time-dependent potential output curve of the solar-powered mini-generator. (a)–(f) Are adapted with permission.113 Copyright 2023, Wiley. (g) Schematic diagram of a prototype NAH (nano-adsorbent hydrogel) electricity generator, designed with a symmetric architecture between two electrodes to maintain balanced initial chemical potentials. (h) Mechanism of ion-separation-driven electricity generation from CO2 adsorption: (i) polyethyleneimine (PEI) is grafted onto boron nitride nanosheets to form amine-functionalized adsorbents (h-BN-NH2); (ii) upon CO2 adsorption, charged species including cations (h-BN-NH3+) and anions (HCO3−) are released; (iii) the cations are spatially confined within nanostructured domains, while the anions diffuse through the hydrogel matrix, enabling efficient ion separation and electricity generation. (g) and (h) Are adapted with permission.9 Copyright 2024, Springer Nature. |
The system achieves a maximum output voltage of 2.34 V and an energy conversion efficiency of up to 38%, sufficient to operate six LEDs for over two hours under ambient sunlight and using only naturally dissolved CO2 (Fig. 8c–f). Notably, the process does not interfere with the intrinsic CO2-recycling function of photosynthesis, as it simply harnesses the byproduct oxygen bubbles without introducing external energy inputs. This represents a rare example of energy harvesting from the thermodynamic minimum state of carbon. By integrating living biological components with engineered functional materials, the device offers a compelling strategy for sustainable power generation in environmentally benign, self-sustaining systems.
Overall, this technology enables CO2-powered electricity generation without the need for external energy input. Its potential applications can be extended to weather monitoring, environmental sensing, and powering microelectronic devices. However, a key challenge is the strong dependence of photosynthesis on environmental conditions such as weather, which causes fluctuations in bubble generation and leads to unstable power output. To overcome this issue, it is necessary to optimize factors including light intensity and temperature to maintain a steady rate of oxygen production. In addition, the limited efficiency of electromagnetic induction results in relatively low voltage output. For example, a single surfacing-diving cycle generates only 0.32 to 2.34 V, with power output in the microwatt range restricting the system's practical applicability. Therefore, increasing the number of coil turns and adjusting the density of the movable device along with the amount of collected gas bubbles may help optimize the surfacing-diving motion, thereby improving voltage output and energy conversion efficiency.
The second system employs a spatial nanoconfinement-based ion separation strategy to convert CO2 adsorption into electrical energy.9 Central to this approach is a nanosheet-agarose hydrogel (NAH) composite generator, which enables highly selective ion transport within its internal channels, thus facilitating direct electricity generation from CO2 uptake (Fig. 8g). To address the inherent challenge of ion size similarity encountered with conventional adsorbents, a nanocomposite adsorbent was synthesized by grafting polyethyleneimine (PEI) with rich –NH/–NH2 functional groups, onto two-dimensional hexagonal boron nitride (h-BN) nanosheets, forming PEI-functionalized h-BN (h-BN-NH2). Upon exposure to CO2, these nanocomposites undergo chemisorption and subsequently release large NH3+ cations (100–300 nm) and angstrom-scale HCO3− anions, establishing a size disparity exceeding two orders of magnitude (Fig. 8h).
This significant size difference enables effective ion discrimination: the bulky NH3+ ions, associated with the extended h-BN skeleton, are sterically confined within the agarose hydrogel matrix, while the smaller HCO3− anions freely diffuse through interconnected water-filled channels. As a result, a highly asymmetric ion transport is achieved, with a reported diffusion coefficient ratio (D−/D+) greater than 106. This imbalance in ion migration creates a charge separation across the generator, giving rise to a measurable voltage. By connecting 50 of these generators in a vertically stacked configuration, the system delivers a stable output voltage of 5 V, which can power standard low-power devices such as light-emitting diodes.
Despite its promising performance and novel mechanism, low power density and limited output stability remain key challenges for NAH-based electricity generators. Although integrating multiple units can achieve open-circuit voltages up to 5 V, the output current (∼1 mA) is insufficient for high-power applications. Unlike conventional batteries that rely on rapid redox reactions, these generators produce electricity through ion migration driven by CO2-induced chemical gradients. As a result, their voltage response is relatively slow, with noticeable time lag. Under continuous load, the voltage tends to drop and only recovers after the load is removed, indicating poor real-time performance. To address these challenges, it is essential to develop hydrogels with improved ion mobility and conductivity. Incorporating capacitive elements into hybrid designs can help buffer power fluctuations, while the introduction of micro- and nano-scale ion channels is suggested to enable more precise control over ion diffusion, thereby enhancing both response time and power output.
Together, these two systems highlight the diversity and potential of CO2-based electricity generation. Despite being in the early stages of development, both systems underscore the transformative potential of converting CO2 into a direct energy source, providing a foundation for future carbon-negative technologies.
Key technologies | Mechanisms | Advantages | Limitations |
---|---|---|---|
Metal-CO2 batteries | Cathode: CO2 reduction to various products during discharge and CO2/O2 evolution during charge | High theoretical energy density, particularly for Li–CO2 and Na–CO2 systems | Poor cycle life & stability because of the carbonate deposition, side reactions, and anode dendrite growth |
The CO2 utilization contributing to carbon neutrality | Scalability & safety concerns because of the gas management | ||
Anode: metal oxidation during discharge and reduction during charge | Generation of value-added chemicals/fuels, particularly in Zn–CO2 batteries | High cost of the anode metals | |
CO2 capture-integrated storage systems | CO2 capture and release mediated by energy storage device (e.g., supercapacitors, batteries) involving electro- and pH-swing mechanisms | Dual advantage of CO2 capture/release and energy storage | Early-stage and limited understanding |
Low cell voltage | |||
Extended applications in grid-storage | Poor cycling stability mainly due to reaction with atmospheric O2 | ||
Complex system design and high maintenance | |||
Molten carbonate fuel cells | Electricity generation through electrochemical reactions between fuel and oxidant, with carbonate ions serving as the charge carriers | Direct internal reforming of hydrocarbon fuels | High operating temperature accelerating material degradation |
Suitable for combined heat and power | High system cost and maintenance demand | ||
Integration with large stationary power applications | |||
CO2-based electricity generators | Photosynthesis-generated O2 drives device motion to induce electricity via electromagnetic induction | Zero or low external energy input | Low and discontinuous voltage output |
Green and sustainable energy conversion | Slow energy conversion rate | ||
Carbon dioxide adsorption by spatially nanoconfined ion separation | Potential for integration with other devices for specific functions | Limited scalability | |
Stability and durability challenges |
For energy storage systems including MCBs and capture integrated energy storage, performance issues such as low energy efficiency, poor reversibility, and limited scalability hinder practical deployment. For CO2 based electricity generation devices, the efficiency of CO2 utilization, long-term operational stability, and cost-effectiveness remain significant barriers. To address these challenges and advance CO2 utilization in energy systems, coordinated efforts are required across the following key areas.
• Performance engineering: materials design and optimization should prioritize the development of highly active, selective, and durable components, including electrodes, electrolytes, and separators. In MCBs, the development of bifunctional cathode catalysts remains critical to improve CO2 conversion/utilization and energy storage. Besides, the challenges including poor interfacial contact due to solid-state discharge products necessitate the use of redox mediators to enhance charge transfer and battery reversibility. For integrated CO2 capture-energy storage systems, redox active materials must endure prolonged cycling and resist oxidative degradation. Further, flow-cell designs must also be optimized to overcome low product yield and separation inefficiencies in MCBs and CO2-capture integrated energy storage devices. The designing of oxygen insensitive and stable organic redox-active materials is the need of the hour. In MCFCs, material degradation is the main drawback caused by high-temperature operation resulting in corrosion and excessive thermal stress. This necessitates corrosion-resistant electrodes, thermally stable electrolytes and protective coatings to ensure enhanced durability and practical performance.
• Fundamental insights: complex interfacial mechanisms in MCBs need major emphasis to develop high-performance MCBs by employing advanced operando-techniques and correlating the inferences with theoretical and experimental analysis. Further, the exploration of new CO2 conversion pathways is particularly important for CO2-capture-based energy storage and electricity generation. The currently reported systems involve a variety of physical and chemical mechanisms by which CO2 participates in energy-related reactions. However, these mechanisms remain largely underexplored or poorly understood, indicating substantial room for improvement and innovation. Interdisciplinary collaboration among researchers in physics, chemistry, biology, and materials science is essential to refine these pathways. By deepening our mechanistic understanding and optimizing reaction processes, it becomes possible to harness CO2 more efficiently for power generation and energy storage. Lastly, in MCFCs, optimizing operation conditions and electrochemical activity require interfacial engineering at this stage to ensure practical viability.
• Device engineering: device architecture optimization plays a critical role in enhancing system stability, scalability, and real-world applicability. For example, CO2-utilizing energy storage systems can be integrated with renewable energy and CO2 capture technologies to build more sustainable and cost-effective platforms. Electricity generation devices should be scaled up through modular stacking designs to meet practical energy demands. MCFCs should be designed to efficiently integrate with power plants to enhance their practical significance.
• Furthermore, the current diversity of CO2-utilizing systems is accompanied by a lack of standardized performance metrics, making it difficult to compare and evaluate different technologies. Therefore, the development of standardized testing protocols for system benchmarking is urgently needed to guide future research and enable fair comparisons across platforms.
By addressing these challenges through interdisciplinary innovation, CO2-utilizing energy systems have the potential to evolve from conceptual technologies into practical solutions that support the global clean energy transition. This transformation could redefine CO2 not merely as a greenhouse gas, but as a key resource for carbon mitigation and sustainable power generation.
Footnote |
† Equal contribution. |
This journal is © The Royal Society of Chemistry 2025 |