Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report (2024)

Chapter: 6 Chemical CO2 Conversion to Elemental Carbon Materials

Previous Chapter: 5 Mineralization of CO2 to Inorganic Carbonates
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

6

Chemical CO2 Conversion to Elemental Carbon Materials

6.1 INTRODUCTION TO ELEMENTAL CARBON PRODUCTS

Carbon accounts for ~27.3 wt% of the total mass of CO2. Captured CO2 could supply the elemental carbon needed for products in a net-zero future. These elemental carbon products can be divided into two categories according to the need for additional reactants and subsequent processes: directly and indirectly derived elemental carbon materials. The products in each category are described in Section 6.2. Elemental carbon materials can be directly derived from CO2 noncatalytically or catalytically by CO2 decomposition or by reacting with a reducing agent. Other elemental carbon materials are produced indirectly with CO2 as their carbon source via multiple steps (e.g., reduction or decomposition of CO2-derived chemicals). There are several CO2 reduction reaction pathways by which captured CO2 can result in useful elemental carbon materials, as discussed in Section 6.3.

As introduced in Chapter 2, the market for elemental carbon materials could increase by 400 percent from 2020 to 2050 (see Table 2-1), owing to increased demand for materials with novel structural and electronic properties. Increased understanding of structures and characteristics of such materials has expanded potential applications to include energy conversion (e.g., supercapacitors [Luo et al. 2023]), novel chemical and material syntheses (e.g., nonprecious-metal electrocatalysts [Collins et al. 2023]), construction materials, health care (e.g., for bioimaging), and environmental protection (e.g., photocatalytic degradation of pollutants [Yao et al. 2023]), among others (Dabees et al. 2023; Malode et al. 2023; Sasikumar et al. 2023; Son et al. 2023; Zhang et al. 2023a). As discussed in Section 2.2.5.5 of this report, some elemental carbon products of CO2 utilization can provide durable storage of carbon, can replace highly carbon-emitting processes via material substitution, or can produce high value products, although most products in this class are likely to be small-volume and thus will not utilize large amounts of CO2. Status and research needs for CO2 conversion to carbon nanotubes (CNTs) were discussed in Chapter 4, Chemical Utilization of CO2 into Chemicals and Fuels, in the 2019 National Academies report Gaseous Carbon Waste Streams Utilization: Status and Research Needs (NASEM 2019), although other classes of elemental carbon materials were not addressed. They were briefly discussed as long-lived CO2-derived products, or Track 1 products, in Section 3.3.4 of the first report of this committee, Carbon Dioxide Utilization Markets and Infrastructure: Status and Opportunities: A First Report (NASEM 2023). Chapter 2 of the present report provides market information on elemental carbon materials, and this chapter provides an up-to-date review of research and development (R&D) on CO2 conversion to elemental carbon materials.

CO2-derived carbon materials can adopt multidimensional (zero- to three-dimensional [0D–3D]) structures from among more than 1500 hypothetical 3D-periodic allotropes of carbon found so far (Hoffman et al. 2016). Zero-dimensional (0D) nanocarbon materials are those having three dimensions only at the nanoscale, with no

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

dimension larger than ~100 nm—for example, carbon quantum dots (CQDs) in sphere form and with crystal lattice or cross-linked network structures consisting of sp2 and sp3 hybridized1 carbon and heteroatoms, such as O and N; graphene quantum dots (GQDs) in the form of a single truncated atomic layer of graphite (Bacon et al. 2014); and fullerene, representing a class of carbon allotropes in the form of spherical, cage molecules with carbon atoms located at the surface and vertices of a polyhedral structure consisting of pentagons and hexagons (Dhall 2023). One-dimensional (1D) nanocarbon materials are those with only one dimension beyond the nanoscale—that is, larger than ~100 nm; for example, carbon nanorods (CNRs), CNTs, carbon nanowires (CNWs), and carbon tubular clusters (CTCs). Despite their common 1D structures, CNRs, CNTs, and CNWs each possess distinguishing features. For example, CNRs have aspect ratios (length/width) of 3–5 with lengths typically of 10–120 nm (Abraham et al. 2021), whereas CNW aspect ratios can be higher than 103, while CNT aspect ratios can be >108. Thus, their aspect ratios are significantly different. Also, CNWs have very high specific surface areas. Moreover, CNTs are chiral materials, which governs their metallic or semiconducting character. Among CNTs are novel CTCs, which are very stable. Interestingly, the electronic properties of metallic CTCs are not affected by their diameters, number of walls, or chirality. Two-dimensional (2D) nanocarbon materials have only two dimensions beyond the nanoscale—that is, larger than ~100 nm. These include carbon nanofilms with thin layers of material spanning from a fraction of a nanometer to several micrometers in thickness (Ranzoni and Cooper 2017); carbon nanolayers with monolayer, bilayers, and multilayers (Schaefer 2010); graphene; and nanocoatings. Three-dimensional (3D) carbon nanomaterials (CNMs) are those with three dimensions beyond nanoscale—that is, all larger than ~100 nm; for example, bulk carbon powders, graphite, carbon fibers (CFs), carbon foams, carbon-carbon composites (CCCs), bundles of CNWs and CNTs, as well as multinanolayers, and dispersions of nanoparticles (Chung 2002; Park 2015; Spradling and Guth 2003; Terrones et al. 1998; Windhorst and Blount 1997; Wissler 2006; Zhao et al. 2023).

Each type of 0D–3D carbon material consists of atoms with unique electronic orbital hybridization state(s), and thus specific characteristics and applications. As shown in Figure 6-2, the carbon materials with sp, sp2, and sp3 hybridizations, created intrinsically and extrinsically via different defect engineering approaches, typically have different applications in energy conversion and storage areas—for example, Li-ion, Na-ion, and K-ion batteries (Rajagopalan et al. 2020), supercapacitors, the hydrogen evolution reaction, and the oxygen reduction reaction (Luo et al. 2023). CQDs typically have sp2 carbon cores and sp3 carbon shells terminated with –OH, –COOH, –NH2, and other functional groups resulting from their preparation processes, which determine the properties and applications of CQDs. Also, the amount of sp2 carbon and the ratio of sp2 to sp3 carbon can significantly affect the photoluminescence properties, and thus the bioimaging ability, of CQDs (Jana and Dev 2022; Jhonsi 2018). Furthermore, the amount of sp2 + sp3 bonds in the carbon shell or on the surface of CQDs has a substantial effect on the solubilities of CQDs in water and other solvents (e.g., ethanol). The water solubilities of CQDs determine their various applications in many fields, including solar cells (Kim et al. 2021), drug delivery (Jana and Dev 2022; Zoghi et al. 2023), and gene delivery (Rezaei and Hashemi 2021). The carbon atom hybridization structures of 0D–3D carbon materials can change significantly upon chemical modification—for example, ozone and high-temperature oxidation and irradiation. For example, all carbon atoms in pristine or pure graphite have sp2 hybridization. Pristine graphene with its sp2 hybridization, although extremely interesting owing to its large specific surface area, unusual physicochemical properties and extraordinary anisotropic mechanical strength, and exceptional thermal and electronic conductivity, is a zero-band-gap semiconductor (a semi-metal), which makes it uninteresting for a number of device applications (Gui et al. 2008; Mbayachi et al. 2021; Rani and Jindal 2013). To make graphene widely useful in the electronics space, its band gap needs to be opened and tunable according to application requirements. Pristine graphene therefore needs to be modified in various ways (e.g., UV-light-assisted oxidation [Güneş et al. 2011]) to enable its carbon atom hybridization structures to change from 100 percent sp2 to mixtures of sp2 and sp3 with different sp2/sp3 ratios to achieve improved properties and more applications, including in electronic, electromagnetic, and optical devices and for catalysis. Representative 0D–3D carbon materials, their typical carbon hybridization characteristics, and their major properties are summarized in Table 6-1.

___________________

1 Orbital hybridization of carbon atoms in different types of elemental carbon materials are described as sp, sp2, and sp3, describing the amount of s- and p-type orbital character of the carbon atoms. The amount of s or p character impacts the bonding between the carbon atoms, with sp having linear, sp2 having trigonal planar, and sp3 tetrahedral character. See Figure 6-2 for images showing the orbital hybridization and illustration of the impact on bonding in elemental carbon materials.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

TABLE 6-1 0D–3D Carbon Materials and Their Structural Characteristics and Major Properties

0D 1D 2D 3D
Representative carbon materials
  • Carbon quantum dots
  • Graphene quantum dots
  • Fullerenes
  • CNTs
  • CNWs
  • CNRs
  • CTCs
  • Carbon nanofibers
  • Graphene
  • Carbon nanofilms
  • Carbon nanolayers
  • Bulk carbon powders
  • Graphite
  • CFs
  • Carbon foams
  • CCCs
Typical orbital hybridization characteristics of their carbon atoms
  • sp2 (e.g., fullerene and the aromatic domain or core of CQDs)
  • sp2 + sp3 (e.g., CQDs and GQD)
  • sp + sp2 (e.g., CNWs)
  • sp2 (e.g., CNFs, CNTs and CTCs)
  • sp2 +sp3 (e.g., CNTs)
  • sp2 (e.g., pristine graphene)
  • sp2 + sp3 (e.g., carbon nanolayers, carbon nanofilms)
  • sp2 (e.g., graphite)
  • sp3 (e.g., diamond)
  • sp2 +sp3 (e.g., CFs)
Major properties
  • Low toxicity
  • Biocompatibility
  • Outstanding photostability
  • Tunable
  • Multicolored emission
  • High-water solubility
  • Good dispersibility
  • Easy surface grafting for different applications
  • Good conductivity
  • Abundant reactive hydroxyl groups distributed on surface of major 1D product—CNTs
  • More active sites, resulting from surface modification
  • Hydrogen storage capacity
  • Thermodynamically favorable for mass transfer and diffusion of reaction substances
  • Diathermancy
  • High mechanical strength
  • High surface area
  • Superior mechanical flexibility
  • High electronic mobility
  • Abundant photo-electrochemically reactive sites
  • Low recombination of photogenerated charges
  • Good mechanical properties
  • High surface area
  • Designability
  • Well-developed porous channels for ion transfer
  • Large specific surface area
  • Good reaction micro-environment
  • More active sites
  • Potentially very high mechanical strength

NOTES: Orbital hybridization patterns of carbon atoms in different types of materials are described as sp, sp2, and sp3, where the s and p describes the amount of s- or p-type orbital character around the carbon atoms. The amount of s or p character impacts the bonding around the carbon atoms, with sp having linear, sp2 having trigonal planar, and sp3 having tetrahedral character which in turn affects the material properties. See Figure 6-2 for images showing the orbital hybridization and illustration of the impact on bonding.

SOURCES: Based on data from Budyka et al. (2017); Cartwright et al. (2014); He et al. (2021); Khan and Alamry (2022); Lee et al. (2020); Lesiak et al. (2018); Liu et al. (2020a, 2020b); Tsang et al. (2006); Van Tran et al. (2022); Zhang et al. (2023b). Zhao et al. (2003); Zhou and Zhang (2021).

Carbon materials with different hybridization structures sometimes have been (at laboratory scale) utilized similarly, as evidenced in Figure 6-2. For example, both sp- and sp2-hybridized carbon materials could be used for synthesizing supercapacitors owing to their similar structures and properties, such as high conductivity and stability, and thus can provide the same function (Luo et al. 2023). Both CQDs and CNRs can be used as sensing materials because both contain sp2 hybridized structures; their sensing capabilities increase with the amount of sp2 hybridized carbon present. CQDs are being used as components of light-emitting diodes, luminescent solar concentrators, and photovoltaic cells; they also can be functionalized to act as catalysts (Zhou et al. 2024).

As indicated in Figure 6-1, a variety of pathways exist for producing 0D–3D carbon materials from CO2. The major challenge of converting CO2 into elemental carbon materials is to reduce the formal oxidation state of C from +4 to 0 via various reduction reactions, which may be realized with thermochemical, electrochemical, photochemical, plasmachemical, or hybrid/tandem processes. All of these technologies are still in the R&D phase at the present time, and each has its advantages and disadvantages from energy consumption and environmental impact perspectives, as discussed in the following sections of this chapter.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Summary of the feedstock inputs, processes, products, and applications for 0D–3D carbon materials
FIGURE 6-1 Summary of the feedstock inputs, processes, products, and applications for 0D–3D carbon materials.
SOURCE: Icons from the Noun Project, https://thenounproject.com. CC BY 3.0.
Applications of differently hybridized and defective carbon materials, as shown, in energy conversion and storage
FIGURE 6-2 Applications of differently hybridized and defective carbon materials, as shown, in energy conversion and storage.
SOURCES: Adapted from X. Luo, H. Zheng, W. Lai, P. Yuan, S. Li, D. Li, and Y. Chen, 2023, “Defect Engineering of Carbons for Energy Conversion and Storage Applications,” Energy and Environmental Materials, Wiley. Orbitals sourced from University of Saskatchewan – Hybrid Orbitals, https://openpress.usask.ca/intro-organic-chemistry/chapter/1-5. CC BY NC-SA 4.0. Icons from the Noun Project, https://thenounproject.com. CC BY 3.0.
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

6.2 CO2-DERIVED ELEMENTAL CARBON MATERIALS

6.2.1 Directly Derived Elemental Carbon Materials

6.2.1.1 Fullerenes

Fullerenes consist of sp2 hybridized carbon atoms linked by single and double bonds, forming a spherical- or cylindrical-shaped closed structure (Ramazani et al. 2021). The most famous of the closed spherical fullerenes is C60, for the 60 carbon atoms in the molecule, also known as buckminsterfullerene or a buckyball, for the molecule’s resemblance to both the geodesic domes of Buckminster Fuller and a standard soccer ball. Fullerene C60 has a van der Waals diameter of about 1.1 nm and a nucleus-to-nucleus diameter of about 0.71 nm. Open-ended cylindrical fullerenes are known as CNTs; single-walled CNTs typically have 0.5–2 nm diameters. Owing to their unique structure and electronic characteristics, the applications of fullerenes are extensive, including photodynamic therapy, drug and gene delivery, nano-sensors, battery electrodes, and organic solar cells, covering many industries (Anctil et al. 2011; Ramazani et al. 2021). Hence, there is significant interest among chemical and material scientists in developing fullerene synthesis methods. General strategies for fullerene synthesis involve generation of fullerene-rich soot through arc discharges, combustion, laser ablation, or microwaves, and then purification of the soot by toluene or benzene washing, Soxhlet extraction, or active carbon filtering (Keypour et al. 2013; Komatsu et al. 2004; Parker et al. 1992; Ramazani et al. 2021; Taylor et al. 1993).

The conventional carbon sources for fullerenes are graphite, aliphatic and olefinic hydrocarbons, chloroform, and aromatics (e.g., naphthalene) (Ramazani et al. 2021); however, there is increasing interest in using captured CO2 as a feedstock for fullerene synthesis (Chen and Lou 2009; Motiei et al. 2001). Chen and Lou (2009) reported that CO2 can be successfully reduced to C60 by metallic lithium at 700°C and 100 MPa, as confirmed by ultraviolet (UV)-visible absorption spectroscopy, matrix-assisted laser desorption/ionization (MALDI) time-of-flight mass (TOF) mass spectrometry (MS), and high-performance liquid chromatography. Higher fullerenes (C70, C78, etc.) were not detected using this synthetic approach. The authors postulated that CO2 radical anion (CO2•−) or other single carbon radicals, resulting from the transfer of an electron from elemental Li to supercritical CO2 (whose polarity increases with its density, thus facilitating electron transfer [Tucker 1999]), are possible key intermediates during the reduction of CO2 to C60. The generation of the CO2/CO2•− couple in the reaction system is the key step for the formation of C60 owing to the fact that the standard potential of the redox CO2/CO2•− couple in an aprotic solvent such as N,N-dimethylformamide (DMF) containing a counter-cation (tetraethylammonium, NEt4+) can be as negative as −2.2 V versus saturated calomel electrode (Bhugun et al. 1996) and the standard potential of the redox couple Li+/Li, 3.04 V, although the potentials of the redox CO2/CO2•− and Li+/Li couples in this reaction system were not measured (Chen and Lou 2009; Motiei et al. 2001). The finding by Motiei et al. (2001) was used by Chen and Lou to explain the feasibility of C60 formation.

6.2.1.2 Hollow Carbon Spheres

Unlike 0D fullerene, hollow carbon spheres (HCSs) can have either 2D or 3D structure (Deshmukh et al. 2010; Liu et al. 2019) and both sp2 and sp3 hybridized carbon atoms on their surface (Deshmukh et al. 2010). HCSs have diameters between 2 nm and several microns. HCSs have unique properties, including encapsulation ability, controllable permeability, surface functionality, high surface-to-volume ratios, and excellent chemical and thermal stabilities (Li et al. 2016). HCSs can be synthesized with hard-templating, soft-templating, and template-free processes. The properties of HCSs vary, depending on the raw carbon materials and synthesis conditions used. For example, the surface areas of HCS can change from a few m2/g to more than 1000 m2/g, which determines their potential applications, especially those of functionalized HCSs.

Synthesis of HCSs from CO2 has been demonstrated using the microbubble-effect-assisted electrolytic method in a CaCO3-containing LiCl–KCl melt electrolyte at 450°C (Deng et al. 2017). The authors employed precise control of the electrode potential to tune the electrochemical reduction rate of carbonate ions and the CO microbubble effect to shape the hollow spheres within the resultant carbon sheets. The produced HCSs exhibited good plasticity and capacitance, which are desirable properties of HCSs used for battery, capacitor, and fuel cell

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

applications, and composite materials. Li et al. (2018) also synthesized HCSs from CO2 using molten carbonate electrolyzers and proposed the following mechanism for the conversion:

Cathode reaction: CO32− + 4e → C + 3O2− (R6.1)
Anode reaction: 2O2− − 4e → O2 (R6.2)
Electrolyte reproduction: O2− + CO2 → CO32− (R6.3)
6.2.1.3 Carbon Nanofiber

1D carbon nanofibers (CNFs) with sp2 hybridized carbon atoms (Deng et al. 2013; Zhou et al. 2020) can be used for energy storage (Zhang et al. 2016), electrochemical catalysis (Shakoorioskooie et al. 2018), sensor manufacturing (Sengupta et al. 2020), and high-strength building material development (Ren et al. 2015). Traditional CNF synthesis methods include electrospinning/carbonization, arc/plasma, and chemical vapor deposition (CVD), with electrospinning/carbonization being the principal approach. Conventional raw materials for CNF preparation include polyacrylonitriles, pitch, acetone, and hydrocarbon gases (Ren et al. 2015) but recently, renewable feedstocks (e.g., lignin and cellulose) for CNF manufacturing have received considerable attention (Lallave et al. 2007; Wu et al. 2013).

Several research groups have demonstrated that CO2 could be a good candidate for CNF synthesis (Lau et al. 2016; Novoselova et al. 2007; Ren et al. 2015; Xie et al. 2024). Ren et al. (2015) reported a one-pot synthesis method for synthesizing CNFs via electrolytic conversion of CO2. The technology, based on Li2O looping for CO2 reduction, employs low-cost, scalable nickel and steel electrodes to decompose CO2 into CNFs and O2, producing CNFs with diameters of 200–300 nm and lengths of 20–200 μm. The Coulombic efficiency is higher than 80 percent and can be close to 100 percent (i.e., complete decomposition) if all the products can be collected. Lau et al. (2016) developed a system integrating Li2CO3 electrolysis with a combined cycle natural gas power plant to produce CNFs and pure O2. The system consumes 219 kJ to convert 1 mol of CO2 to 1 mol of carbon, while the pure O2 generated from CO2 decomposition is sent back to the gas turbine, improving electricity generation efficacy. Xie et al. (2024) reported an electrochemical–thermochemical tandem catalysis system to convert CO2 to CNF using renewable hydrogen, which achieved an average yield of 2.5 gcarbon gmetals−1h−1 (“metals” here refers to the catalyst used for conversion; in Xie et al. [2024] this was an FeCo alloy and extra metallic Co). In summary, CO2 conversion to CNFs shows promise as a method to reduce CO2 emissions while generating high-value CNFs with market potential in nanoelectronics, energy storage, and construction materials.

6.2.1.4 Carbon Nanotubes

Carbon nanotubes (CNTs) are rolled graphene sheets. CNTs are 1D carbon materials with sp2 or sp2 + sp3 hybridized carbon atoms and are classified as single-walled nanotubes (SWNTs) or multiwalled nanotubes (MWNTs). SWNTs are open-ended cylinders with only one wrapped graphene sheet, while MWNTs are an assembly of homocentric SWNTs. The dimensions of SWNTs and MWNTs differ in both length and diameter, leading to considerably different properties (Kozinsky and Mazari 2006; Rathinavel et al. 2021). For example, unlike SWNTs, the mechanical properties (e.g., Young’s modulus) of MWNTs depend not only on diameter and chirality but also on the number of sidewalls (Rathinavel et al. 2021). Various methods have been developed for preparing CNTs, including arc discharge, laser ablation, CVD, and injection of carbon atoms into metal particles (Rodríguez-Manzo et al. 2007). Factors affecting CNT syntheses include the carbon source (e.g., hydrocarbons, alcohols), catalyst (e.g., Al2O3), temperature, pressure, and flow of gases. CNTs have been explored for use in many applications, including sensing (Wang et al. 2023), cancer therapy (Mishra et al. 2023), preparation of biological fuel cells (ul Haque et al. 2023), light-weight reinforced high-strength materials (Hong et al. 2023), Li/Na-ion batteries (Qu et al. 2024), and others (Kordek-Khalil et al. 2024).

CO2 can be used as a carbon source for CNT synthesis. Research to-date primarily has explored electrochemical conversion processes (Douglas et al. 2018; Li et al. 2018; Licht et al. 2016; Moyer et al. 2020). For example,

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Licht et al. (2016) reported that ambient CO2 could be successfully captured and converted to CNTs and CNFs in molten lithium carbonates at high yield via electrolysis using inexpensive steel electrodes, with the resultant carbon materials exhibiting good, stable capacities for energy storage. Douglas et al. (2018) synthesized CNTs with small diameters (~10 nm) using ambient CO2 and Fe catalysts. They concluded that (1) the energy input costs for the conversion of CO2 into CNTs are $50/kgCNT and $5/kgCNT using Al2O3 and ZrO2 as thermal insulation materials, respectively, and (2) the CO2 to small-diameter CNTs technology is superior to other CO2 conversion technologies with lower-value materials as their products. Li et al. (2018) found that the electrolyte composition in a molten carbonate electrolyzer that captures CO2 from air and converts it to CNTs and HCSs plays a key role in the selectivity toward CNTs, as well as in determining the diameter of the synthesized CNTs.

6.2.1.5 Graphene

2D graphene, with its sp2 hybridized carbon atoms and very stable structure, is the thinnest (sheet thickness of 0.34 nm) and strongest nanomaterial known (Yu et al. 2020). Graphene itself has limited applications owing to its easy agglomeration, and difficult processing (Yu et al. 2020), and it requires modification and functionalization to increase its potential applications. Graphene synthesis techniques include exfoliation of graphite, reduction of graphene oxide, thermal and plasma CVD of hydrocarbons, thermal decomposition of silicon carbide, and unzipping of CNTs. Graphene functionalization processes are based on (1) the formation of covalent bonds between graphene and introduced functional groups (e.g., −OH and −COOH); (2) the formation of non-covalent bonds (e.g., π–π interactions, hydrogen, ionic, and dative bonding); and (3) element doping (Yu et al. 2020). The primary challenges facing production and use of graphene and functionalized graphene are high costs and carbon/environmental footprint, determined by the characteristics of typical synthesis methods. For example, graphene oxide (GO) reduction requires the use of highly corrosive agents and thus a long washing process after reduction, energy-intensive CVD, and low quality of large-scale production (Liu et al. 2020c; Urade et al. 2023). Nonetheless, graphene and functionalized graphene could have a wide range of applications—including in supercapacitors, solar cells, electrodes, and e-textiles—owing to their many desirable properties (Su et al. 2020; Urade et al. 2023; Yu et al. 2024), including rich functional group variability and density, environmental stability and compactness (Su et al. 2020). For example, the deleterious ion migration that reduces operational stability of iodide perovskite solar cells synthesized with organic–inorganic halide perovskite materials, resulting from the weak Coulomb interactions in the perovskite lattice, can be suppressed by using graphene, which has a lattice parameter (0.246 nm) smaller than the radius of I (0.412 nm) (Su et al. 2020).

CO2 also has been explored as a feedstock for graphene synthesis (Chakrabarti et al. 2011; Hu et al. 2016; Liu et al. 2020c; Strudwick et al. 2015; Wei et al. 2016). For example, Liu et al. (2020c) used molten carbonate electrolysis to synthesize graphene, where the conversion occurs via carbonate formation in Li2O, electrolysis of Li2CO3, and then exfoliation of the resultant carbon nanoplatelets:

Chemical dissolution and carbonate formation: CO2 + Li2O → Li2CO3 (R6.4)
Electrolysis: Li2CO3 → Cplatelets + O2 + Li2O (R6.5)
Exfoliation (DC voltage): Cplatelets → Cgraphene (R6.6)

Addition of zinc and increased electrolysis current led to the selective (more than 95 percent yield) formation of high-purity carbon nanoplatelets rather than CNTs, and exfoliation of the carbon nanoplatelets produced graphene in 83 percent yield by mass of the original carbon nanoplatelets (Liu et al. 2020c).

6.2.1.6 Graphite

Graphite is a 3D material with a stacked planar sp2-hybridized C6 fused ring structure—that is, stacked layers of graphene with AB stacking (Jara et al. 2019). It is known for its high specific surface area, thermal conductivity,

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

fracture strength, and special charge transport phenomena. It can be obtained as naturally occurring graphite or produced by synthesis (Surovtseva et al. 2022). Natural occurring graphite is mined but requires energy- and chemical-intensive beneficiation and purification thereafter, especially for its use in batteries (Surovtseva et al. 2022). Synthetic graphite is tunable in microstructure and morphology. The synthesis typically includes two sequential steps: formation of amorphous carbon via carbonization of a carbon precursor and subsequent graphitization of the amorphous carbon. Both steps occur at high temperature and are therefore energy-intensive and generally CO2-emitting. Synthetic graphite can be prepared by various methods, including graphitizing nongraphitic carbons (e.g., cokes [Gharpure and Vander Wal 2023]), processing hydrocarbons (e.g., agricultural wastes or biomass materials [Yap et al. 2023]), CVD at >2500°C, and decomposing unstable carbides. Graphite is used in batteries (Kim et al. 2024), refractories (Chandra and Sarkar 2023), metallurgical processing (Li et al. 2023a), and other fields.

Some researchers have synthesized graphite from CO2 (Chen et al. 2017a; Hu et al. 2015, 2019; Hut et al. 1986; Liang et al. 2021; Ognibene et al. 2003; Yu et al. 2021). For example, Liang et al. (2021) prepared synthetic graphite submicroflakes by heating CO2 in the presence of lithium aluminum hydride (LiAlH4) at 126°C. This synthetic graphite was compared to commercial graphite to test its potential application as an anode for lithium storage materials, and both showed stable reversible capacities around 320 mAh g–1 from the 1st to 100th cycles at a current density of 0.1 A g–1. After 100 cycles, the synthetic graphite and commercial graphite achieved 99 percent and 95.4 percent retention efficiencies, respectively, suggesting that the synthetic graphite prepared with CO2 is superior to its commercial counterpart, especially considering that it was generated without a separate graphitization step. Electrochemical methods also can be used to produce graphite from CO2, with Chen et al. (2017a) demonstrating that CO2 captured from synthetic flue gas by a molten salt (Li2CO3 – Na2CO3 – K2CO3 – Li2SO4) at as low as 775°C (the electrolysis temperature) without the use of any catalyst can produce nano-structured graphite.

6.2.2 Indirectly Derived Carbon-Rich Carbon Materials

6.2.2.1 Carbon Fiber

Carbon fibers (CFs), with a $7.1 billion market in 2023 and annual growth rate of 12.6 percent (marketsandmarkets 2024), have extremely useful properties, including high elastic moduli, compressive and tensile strengths, and thermal and electrical conductivities, as well as low coefficients of thermal expansion (Hiremath et al. 2017). CFs are reinforcing materials widely used in airplanes, cars, and wind turbine blade manufacturing (Liu and Kumar 2012). The most common carbon source for CF synthesis is polyacrylonitrile (Le and Yoon 2019). Polyacrylonitrile-based CF manufacturing includes fiber spinning, stabilization, carbonization, and graphitization steps (Kaur et al. 2016). Synthesis of polyacrylonitrile is based on the free-radical polymerization of acrylonitrile (Pillai et al. 1992), which is produced via catalytic ammoxidation of propylene. Propylene, in turn, is generally produced from the reaction of ethylene with 2-butene, where the latter is synthesized via ethylene dimerization (Pillai et al. 1992). Thus, ethylene is a critical intermediate in CF production.

As discussed in Chapter 7 of this report, ethylene can be produced via thermochemical, electrochemical, or potentially photochemical conversion of CO2, and thus, CO2 can play an indirect role in CF syntheses (Li et al. 2020; Pappijn et al. 2020). The development of highly active, selective, and stable CO2 to ethylene catalysts would facilitate the CO2-to-CF pathway.

6.2.2.2 Carbon-Carbon Composites

As noted in Section 6.2.2.1, CO2 can be a raw material for producing the critical precursor to CFs, which then could be used subsequently to produce CCCs. CCCs are lightweight, high-strength materials with good electrical properties, making them attractive for a wide spectrum of applications. The CCC manufacturing process involves saturation (impregnation) of other materials into carbon matrices, followed by graphitization or carbonization (a pyrolysis process) to form a graphitic structure (Windhorst and Blount 1997). During pyrolysis, voids form because of volatilization, which is deleterious for the mechanical properties of the CCCs. Repeated impregnation

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

and carbonization can address this problem, but repetition increases manufacturing time and thus cost. Additionally, as-made CCCs may not possess adequate microstructure, porosity, interlaminar shear strength, flexural, ultrasonic and vibration damping behavior for certain applications (Bansal et al. 2013). Further modifications may be necessary to achieve the desired qualities, including physical treatments (e.g., plasma-based surface changes) and chemical treatments or use of additives.

Among the possible additives are carbon nanomodifiers (CNMOs), or the carbon nanostructure materials introduced in Table 6-1. Some CNMOs, such as nanographene, have been synthesized from CO2, as discussed in Section 6.2.1. CNMOs might be able to overcome the abovementioned challenges faced during CCC manufacturing, combining with other matrices, such as pitch, to enhance the properties of CCCs. CNMOs can mitigate shrinkage and tailor the properties of CCCs. For example, Bansal et al. (2013) introduced nanographene platelets (NGPs) to fill defects such as pores and cracks during manufacturing of CCCs. When the NGP-to-CCC ratio was 1.5 wt%, the interlaminar shear strength, flexural strength, and Young’s modulus of the CCCs increased by 22 percent, 27 percent, and 15 percent, respectively, compared to CCCs that did not contain NGPs. Meanwhile, the porosity of the modified CCCs was reduced by 17.5 percent. Eslami et al. (2015) filled carbon-fiber/phenolic composites with MWNTs. When the sample was modified by 1 wt% MWNTs, the thermal stability of the CCCs increased, according to thermogravimetric analysis, and the linear and mass ablation rates decreased by about 80 percent and 52 percent, respectively. Scanning electron microscopy showed the formation of a strong carbon network in CCCs resulting from the addition of MWNTs. These examples suggest that CO2-derived CNMs can play an important role in the development of high-quality CCCs.

6.3 EMERGING TECHNOLOGIES TO REDUCE CO2 TO ELEMENTAL CARBON

6.3.1 Introduction

As introduced above, the reduction of CO2 to elemental carbon (CTEC) can be performed by four major chemical reaction processes: thermochemical, photochemical, electrochemical, and plasmachemical reduction (see Figure 6-1). Table 6-2 summarizes the strengths and shortcoming of these four chemical conversion pathways for CTEC. The four pathways could be coordinated to develop potentially more efficient and less expensive CTEC processes by combining their strengths and overcoming their shortcomings, as listed in Table 6-2. For example, the temperature required for thermochemical CO2 conversion to CNTs can be as high as 700°C (Lou et al. 2006), however, a combined photo-thermochemical CO2-to-CNT process can proceed at as low as 80°C (Duan et al. 2013).

The majority of CTEC conversion technologies are still in the lab-scale study phase, and thus their development status can be described as “emerging,” although the specific technology readiness levels (TRLs) of different conversion technologies vary. Initial research on CTEC processes examined thermochemical pathways, and as early as 1978, cation-excess magnetite was used to convert CO2 to carbon via CTEC with an efficiency of nearly 100 percent at 290°C, although the structure of the generated carbon wasn’t reported by the researchers (Tamaura and Tahata 1990). Thermochemical CTEC processes are at higher TRLs than other pathways, with bench-scale2 and pilot-scale3 conversions being successfully demonstrated. As of April 2024, no CTEC technology has been commercialized.

Some CTEC processes are effective at forming specific high-value carbon materials but are very energy-intensive. For example, high-quality CNMs, such as CNFs and CNTs, can be produced by CVD, but this method requires very high temperatures and low pressures over long periods of time and is not easily scalable. As a result, this method is estimated to have an unusually large carbon footprint of up to 600 tonnes of CO2 emitted per tonne of CNM produced (Wang et al. 2020). Alkali and alkaline earth metals, such as lithium, sodium, magnesium, and calcium, can be used as reductants to reduce CO2 to various carbon products, including carbon spheres, graphene, and CNTs. However,

___________________

2 For example, 100 percent direct conversion of CO2 to C has been demonstrated at the bench scale with Ni0.39Fe2.6O4 (reaction conditions: flow rate of simulated flue gas: 9 dm3/h; composition of the simulated flue gas flue gas: 20% CO2 and 80% N2) (Taylor et al. 1993).

3 The bench-scale result described in Taylor et al. (1993) was tested at pilot scale by designing a system with the capacity to treat 1000 Nm3 h−1 flue gas (composition: 9.4% CO2, 74.8% N2, 0.8% O2, 15% H2O, 50 ppm NOx) from a liquefied natural gas combustion boiler (Yoshida et al. 1997).

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

TABLE 6-2 Comparison of the Strengths and Weaknesses of the Four Major State-of-the-Art (SOA) Chemical Reduction Processes Being Explored for CTEC Material Conversion

Thermochemical Reduction Photochemical Reduction Electrochemical Reduction Plasmachemical Reduction
Strengths
  • High conversion rate
  • Easy scale up
  • Relatively mature in terms of material preparation and regeneration, as well as equipment manufacturing and operation
  • When combined with photochemical reduction, relatively easy to synergize, disperse, and activate catalysts
  • Moderate reaction conditions (e.g., temperature as low as 80°C)
  • Relatively low energy requirements for reaction process
  • Easy to avoid side reactions and thus by-products with photocatalyst defect engineering
  • Relatively easy to recycle spent catalysts
  • Relatively easy to realize high selectivity for desired product(s)
  • Environmentally benign reaction process
  • Short starting time
  • Relatively less expensive
  • Quick to reach reaction conditions
  • Easy to increase internal energy of reactants
Weaknesses
  • Relatively high temperature requirements for reaction, which could deactivate materials via coking, and thus decrease CO2 conversion rates
  • Relatively low overall energy utilization efficiency
  • Suboptimal conversion of electric energy into radiation energy of desired wavelengths of light-emitting diodes (LEDs)
  • Relatively high loss of the heat generated from some light sources
  • Difficult to achieve high conversion and scale up owing to limited direct light access (surface area)
  • Slow mass transfer of reactants to the active surface area of electrode
  • Not cost-effective owing to the use of precious metal electrocatalysts
  • Frequently requires expensive product separation methods (e.g., membranes)
  • Generally low product selectivity
  • Energy generation rate of plasma system is much higher than the total energy consumption rate of reactions
  • Relatively high heat loss owing to conduction leads to low energy efficiency
  • Low catalyst target selectivity

SOURCES: Based on data from the following: Thermochemical: Álvarez et al. (2017); Kondratenko et al. (2013); Kosari et al. (2022); Ye et al. (2019); Zuraiqi et al. (2022). Photochemical: Duan et al. (2013); Han et al. (2023a); Li et al. (2014); Yaashikaa et al. (2019). Electrochemical: Lu and Jiao (2016); Overa et al. (2022); Pérez-Gallent et al. (2020); Sajna et al. (2023); Spinner et al. (2012); Tackett et al. (2019). Plasmachemical: George et al. (2021); Lerouge et al. (2001); Martirez et al. (2021); Snoeckx and Bogaerts (2017).

regeneration of these reductants is also very energy intensive. For CTEC processes to be competitive against other alternative carbon sources or processes, energy efficiency must be maximized and low-carbon energy sources used.

The distribution percentages of the 161 journal papers found with CO2 being the sole carbon source in the four CTEC research areas is shown in Figure 6-3. Clearly, research work in the thermochemical area dominates all those four areas, accounting for 70 percent of the total work reported in published papers. The thermochemical research largely involves CTEC studies of decomposition-based reactions between CO2 and cation-excess materials and reactions between CO2 and strong reducing agents. Note that only published journal papers as of January 2024 were collected in the analysis given in Figure 6-3.

The history of the annual publications resulting from global CTEC R&D efforts is presented in Figure 6-4, where it is evident that this is an understudied phenomenon, yet to take off. The first CTEC paper was published in 1978, followed by a drought of CTEC research for 12 years. CTEC research productivity was stable from 1990 to 2001, during which about three papers were published annually. The average quantity of annual CTEC publications has tripled since 2015, a significant increase. However, the pace of CTEC R&D activities has been much slower than other CO2 utilization technologies. Figure 6-4 reveals that thermochemical methods have dominated CTEC technology history, while photochemical and plasmachemical approaches have only been studied occasionally. Note that researchers are increasingly interested in developing electrochemical CTEC technologies owing to their advantages compared to thermochemical ones, as given in Appendix K.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
The distributions of published papers versus the carbon products generated in the four CTEC areas
FIGURE 6-3 The distributions of published papers versus the carbon products generated in the four CTEC areas.
NOTES: CNF = carbon nanofiber; CNM = carbon nanomaterial; CNO = carbon nano-onion; CNT = carbon nanotube; CNW = carbon nanowire; HCS = hollow carbon sphere. See Appendix K for the full literature review informing this figure.
CTEC publications within four technological areas in different years
FIGURE 6-4 CTEC publications within four technological areas in different years.
NOTE: See Appendix K for the full literature review informing this figure.
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

6.3.2 Reaction Processes

6.3.2.1 Thermochemical Conversion Pathways

Thermochemical CTEC can involve CO2 decomposition or CO2 reaction with strong reducing agents—for example, H2 or alkali and alkaline earth metals, with and without the use of other reactants or catalysts. The strengths and weaknesses of these two types of CTEC technologies are described below.

6.3.2.1.1 Decomposition-Based CTEC

The CO2 decomposition into C can be via the R6.7 pathway:

CO2 ↔ C + O2 ∆G0R6.7 = 394.4 kJ/mol (R6.7)

Synthesizing elemental carbon from CO2 via R6.7 is difficult because of its large, positive ΔG° value. Research efforts have aimed to identify materials that can enable more favorable reduction pathways and thus allow lower operating temperatures to be used (Kim et al. 2001, 2019; Kormarneni et al. 1997; Lin et al. 2011; Sim et al. 2020; Tamaura and Tahata 1990; Tsuji et al. 1996a; Yoshida et al. 1997). The strategy of exposing CO2 to a cation-excess metal oxide works because the material is oxygen-deficient and therefore at much lower temperatures than in the gas phase it is possible to strip oxygen from CO2 by absorbing that oxygen in the metal oxide lattice, filling oxygen vacancy sites. For example, Tamaura and Tahata (1990) found that reacting CO2 with cation-excess magnetite (Fe3+δO4, δ = 0.127) can reduce CO2 to carbon with an efficiency of nearly 100 percent at 290°C. During CO2 reduction, all of the oxygen in CO2 transfers as described above, in the form of O2− to the cation-excess magnetite, because only carbon and no CO was detected (Tamaura and Tahata 1990).

Tsuji et al. (1996a) investigated the reactivity toward CO2 decomposition of metallic iron formed on oxygen-deficient Ni(II)-bearing ferrite. 86 percent CO2 conversion was observed, yielding 97 percent elemental carbon and 3 percent CO. The same group also achieved excellent elemental carbon selectivity with an ultrafine Ni(II) ferrite prepared with coprecipitation of Ni2+, Fe2+, and Fe3+ at 60°C with 36 percent Ni2+ substitution for Fe2+ in magnetite at 300°C (Tsuji et al. 1996b). The associated CO2 decomposition mechanism can be written as:

MxFe3−xO4−δ + (δ − δ')/2 CO2 → MxFe3−xO4−δ' + (δ − δ')/2 C(δ > δ') (R6.8)

where M represents divalent metals, and δ and δ' are the initial oxygen deficiency and the oxygen deficiency at any reaction time (t), respectively (Tsuji et al. 1996b). The change in oxygen deficiency (δ − δ') directly reflects the degree of CO2 conversion to elemental carbon.

Kim et al. (2019) examined SrFeCo0.5Ox for its ability to reduce CO2 to elemental carbon. The highest CO2 decomposition efficiency achieved with SrFeCo0.5Ox reached ~90 percent, while decomposition efficiencies of ≥80 percent at 550°C to 750°C lasted for more than 60 minutes. The reaction mechanism of the interaction between SrFeCo0.5Ox and CO2, proposed by Kim et al. (2019), is shown in Figure 6-5.

The same research group examined another reactant, SrFeO3–δ, for CO2 reduction under the same test conditions, achieving a CO2 decomposition efficiency of ≥90 percent, with decomposition ≥80 percent lasting for ~170 min, 70 percent longer than that realized with SrFeCo0.5Ox (Sim et al. 2020). However, the elemental carbon production selectivity achieved with SrFeCo0.5Ox and SrFeO3–δ, despite their activity, stability, and reproducibility, are not as good as that obtained with other solid oxide reactants—for example, Ni(II) ferrites (Tsuji et al. 1996b). In other words, unlike Ni(II) ferrites, both SrFeCo0.5Ox and SrFeO3–δ based CTEC processes have a common shortcoming—that is, generation of CO as a by-product, which needs to be overcome when elemental carbons are the targeted products. One possible method is to add another reactor to consequently split the CO generated in Step I in Figure 6-5 subsequently into elemental carbon. Another difference between Kim et al. (2019) and previously reported work in this area is that a continuous CO2 decomposition system was used, which is beneficial to eventual commercialization of the cation-excess materials based catalytic CTEC technology.

Note that use of these solid-oxide reactants to decompose CO2 is closely related to research and development being pursued for solar thermochemical syngas production, in which oxygen-deficient metal oxides at high

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
The suggested mechanism of CO2 decomposition with SrFeCo0.5Ox
FIGURE 6-5 The suggested mechanism of CO2 decomposition with SrFeCo0.5Ox.
SOURCE: Reprinted from S.-H. Kim, J.T. Jang, J. Sim, J.-H. Lee, S.-C. Nam, and C.Y. Park, 2019, “Carbon Dioxide Decomposition Using SrFeCo0.5Ox, a Nonperovskite-Type Metal Oxide,” Journal of CO2Utilization 34:709–715, Copyright (2019), with permission from Elsevier.

temperature (heated by concentrated sunlight) strip oxygen from water and/or carbon dioxide to produce hydrogen and carbon monoxide (Wexler et al. 2023a; Zhai et al. 2022). Some of the same materials as mentioned above have been explored for this purpose (Gautam et al. 2020; Wexler et al. 2021, 2023b). As reduction to elemental carbon by these materials therefore likely involves stepwise stripping of oxygen from carbon dioxide to form carbon monoxide (the end goal in solar thermochemical syn gas production) first, it is unsurprising that it is even more difficult to form elemental carbon, given that the second C-O chemical bond in carbon monoxide is a stronger triple bond whereas the first C-O bond to be stripped is a double bond (see Figure 6-5).

6.3.2.1.2 CTEC with Strong Reducing Agents

In this type of CO2 to elemental carbon conversion, H2, alkali metals, and alkaline earth metals are used as reductants—for example,

CO2 + 2H2 ↔ C + 2H2O G0 = −79.8 kJ (R6.9)
CO2 + 2Mg → C + 2MgO G0 = −743.6 kJ (R6.10)

All CTEC reactions of this type are thermodynamically favorable under standard conditions. The temperatures required for R6.9 and R6.10 are lower or much lower than those for the above-mentioned decomposition-based reduction reactions, depending on the reducing agent used. Magnesium (Mg) is the primary metal used as a strong reducing agent to react with CO2 and produce CNMs in this type of CTEC process (R6.10). The major elemental carbon product of such processes is graphene, which can combine with Mg to form Mg matrix composites (Li et al. 2022c, 2023b; Samiee and Goharshadi 2014; Wei et al. 2022; Zhang et al. 2014). R6.10 was first reported in 1978 (Driscoll 1978). In recent years, multiple research groups have demonstrated its use for simultaneous reduction of CO2 and generation of elemental carbon materials (Li et al. 2022c, 2023b; Samiee and Goharshadi 2014; Wei et al. 2022; Zhang et al. 2014).

Ideally, the products of R6.9 and R6.10 are elemental carbon and H2O or metal oxides. However, completely avoiding the generation of by-products is difficult. For example, R6.11 could occur simultaneously with R6.9 (Pease and Chesebro 1928)

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
CO2 + 4H2 ↔ CH4 + 2H2O G0 = −113.3 kJ, (R6.11)

and form an undesirable product, CH4. Also, MgO generated in R6.10 can further react with CO2 to form MgCO3 via

3/2 CO2 + Mg → 3/2 C + MgCO3 G0 = −743.6 kJ. (R6.12)

The formation of undesired products (e.g., CH4, CO, carbonates) in R6.11 and R6.12 decreases elemental carbon selectivity and yield and complicates the separation of elemental carbon during decomposition-based strong reducing agents based CTEC processes. Carbon materials—for example, graphene, produced via the metal reduction CTEC process can exhibit excellent practical properties. For example, Wei et al. (2022) reported successful preparation of few-layer graphene through the reaction between molten Mg and CO2 gas at 750°C, 100°C above the Mg melting temperature. The produced graphene, exhibiting a high degree of graphitization and nanoscale thickness, served as a lithium storage material and achieved excellent rate capability and cycling performance with a reversible capacity of 130 m Ah g−1 after 1,000 cycles at a current density of 1.0 A g−1.

Compared to decomposition-based CTEC technologies, CTEC using strong reducing agents has several notable advantages. First, the reactions of CO2 with strong reducing agents typically have negative Gibbs free energy (ΔG0) and enthalpy (ΔH0) changes, indicating their thermodynamic feasibility under standard conditions. Second, the very negative ΔG0 and ΔH0 values for reactions of CO2 with strong reducing agents to form elemental carbon materials imply that 100 percent CO2 conversion efficiency potentially can be achieved, a precondition for reaching a high yield of elemental carbon. Also, this class of CTEC processes does not require catalysts because moderate temperature elevation alone can significantly increase the rates of the reactions between CO2 and strong reducing agents (e.g., H2, alkali or alkaline earth metals). Moreover, the reaction set-ups and operation are relatively simple, which will reduce costs of the CTEC processes.

Despite having advantages of thermodynamic feasibility, high conversion efficiency, and less dependence on catalysts, CTEC processes using strong reducing agents are not without shortcomings. Selecting an appropriate reactive metal or reducing agent for the CTEC process is crucial, as the choice will affect CO2 conversion efficiency and elemental carbon selectivity. When a metal is used instead of H2 as the reducing agent, the reaction has to proceed at temperatures above the melting point of that metal. These temperatures are typically higher than those needed for the thermal-decomposition-based CTEC (~300°C). Moreover, achieving precise control over the morphology and properties of the elemental carbon produced is not easy, especially when specific structures of elemental carbon materials are desired—for example, 2D graphene and graphite with sp2 carbon hybridization, and 0D CQDs with both sp2 core and sp3 shell carbon hybridizations. However, as noted above, the metal reduction technology generates metal oxides and sometimes by-products like carbonates that are not only long-lived materials to store carbon but that also could be sold for industrial applications owing to their high purities after they are separated from the elemental carbon materials and the metal oxides. Separation of carbon and noncarbon materials is relatively easy. The resulting metal oxides can be reduced to elemental metals (e.g., Mg) for cyclic CTEC conversion but this requires the use of energy-intensive processes. Also, as with any technology, transitioning from the current laboratory-scale experiments to large-scale production for practical applications could present additional challenges.

Last, although using CO2 as a feedstock to produce long-lived elemental carbon materials could provide environmental benefits (Cossutta and McKechnie 2021; Goerzen et al. 2024; NASEM 2023), the overall environmental impact of thermochemical CTEC conversion needs to be assessed holistically, considering requirements for temperature, pressure, product separation, and reactant material recycling, if applicable.

6.3.2.2 Electrochemical and Electrically Driven Thermal Conversion Pathways
6.3.2.2.1 Electrochemical

The use of high-temperature molten carbonate electrolysis to produce elemental carbon products from CO2 has advanced considerably over the past decade. Derived from earlier solar thermal electrochemical processes (Ren et al. 2019), these systems use metal carbonate electrolytes, which have high melting points, so the reactions

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

are typically run at 450°C–700°C. High-temperature molten carbonate electrolysis can form a large spectrum of carbon products, including carbon powder, nanoplatelets, graphene, nano-onions, high-quality nanotubes, and other nanocarbon allotropes. Additives such as calcium or lithium chloride or fluorides can improve the performance of these systems by increasing the solubility of CO2 and the calcium oxide (Tomkute et al. 2013). The mechanism of this reaction is currently unclear (Han et al. 2023b).

High-temperature molten carbonate electrolysis is scalable because all the elemental components are abundant and economical, although high temperatures are required to prevent the molten carbonate electrolyte from solidifying. The molten carbonate precursor, often an alkali metal oxide, reacts with CO2 selectively at low concentrations, including from air, to generate the electrolyte. Thus, this method can be used to integrate CO2 capture from atmospheric or flue gas streams for conversion. Such integration of molten salt CO2 capture and electrochemical transformation eliminates the need for distinct separations processes and the accompanying energy and capital requirements. These systems might be able to operate on waste heat from high-temperature flue gas exhaust in a fossil fuel plant, and the oxygen generated at the anode could be used to facilitate combustion. It has been suggested that this method might be able to achieve a 100× price reduction for the production of CNTs compared to conventional CVD methods (Ren and Licht 2016).

A more recent development is room-temperature CO2 electrolysis into carbon materials, which uses metals that are liquid at room temperature, such as gallium, alloyed with redox-active metals, such as cerium. In one such system, analysis of the liquid revealed a thin film of cerium oxide on the surface, and it is believed redox reactions occur by cycling cerium between the zero and tetravalent oxidation state (Esrafilzadeh et al. 2019). The proposed reactions are shown below, where the R6.13R6.16 reduction reactions occur at the liquid metal working electrode (a cathode made of a gallium-indium-tin alloy, galinstan), and R6.17 is the oxidation reaction at the anode. The carbonaceous nanoflakes have been applied as carbon-based supercapacitor materials (Esrafilzadeh et al. 2019).

(1) 2Ce(galinstan) + 1.5 O2(air) → Ce2O3 (R6.13)
(2) Ce2O3 + 3H2O + 6e → 2 Ce(0) + 6OH (R6.14)
(3) Ce(0) + CO2 → CeO2 + C (R6.15)
(4) CeO2 + 2H2O + 4e → Ce + 4OH (R6.16)
(5) 4OH → O2 + 2 H2O + 4e (R6.17)

This approach has also been reported with other liquid metal electrode formulations (Irfan et al. 2023; Ye et al. 2023). Other examples of CO2 conversion to carbon materials that use liquid metals apply mechanical energy (Tang et al. 2022) or heat (Zuraiqi et al. 2022) as the energy source. The use of liquid metals provides several advantages over CVD and high-temperature molten carbonate electrolysis. The components are abundant and economical and have low toxicity and vapor pressure. Unlike the other electrochemical methods discussed, the process operates at low or room temperature. The high surface tension of the liquid metal also acts as an intrinsic coking-resistant surface. The carbonaceous products spontaneously flake off the surface of the electrode, which facilitates separations and prevents catalyst poisoning or inhibition.

6.3.2.2.2 Electrically Driven Thermal Reduction

Solid oxide electrolyzer cells (SOECs) can be used to combine thermal and electrochemical energy to reduce CO2 to elemental carbon materials. CO2 electrolysis in the SOEC forms carbon monoxide (CO) and solid carbon at the cathode, while oxygen is evolved at the anode. The electrodes again are typically made of oxide perovskites similar to the materials discussed earlier. This method has been applied to produce CNTs, albeit at high temperatures (>800°C) (Tao et al. 2014).

Cathode: CO2 + 2e → CO + O2− (R6.18)
CO + 2e → C + O2− (R6.19)
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Anode: 2O2− → O2 + 4e (R6.20)

A related technology is the use of tandem electro-thermochemical looping to access carbon materials from CO2. In these systems, CO2 is initially reduced to CO in a low-temperature electrolytic cell or a high-temperature SOEC (Luc et al. 2018; Mori et al. 2016). The CO is then heated to high temperatures (500°C–700°C), where it undergoes disproportionation in a Boudouard reaction, forming solid carbon and CO2. This method has been used to produce densely packed and aligned CNTs. Challenges to this approach include integration of the CO stream to the reactor, as well as the purity of the former. Minimizing concomitant evolution of H2 with CO is critical to prevent hydrogen gas accumulation in the system. Additionally, recovered CO2 must be free of carbon materials before being looped back into the electrolysis cell to avoid poisoning the electrodes in the low-temperature electrolytic cell (Luc et al. 2018).

Boudouard Reaction: 2CO ↔ CO2 + C (solid) (R6.21)

When CO2 decomposes to CO and ½ O2, CO can disproportionate via the Boudouard reaction to produce ½ CO2 and ½ C(solid), which, with recycle of CO2, can result in an overall conversion of CO2 → C(solid) + O2 via this two-step pathway. Solid carbon is more readily generated from CO than CO2 below 971K (Chery et al. 2015; Han et al. 2023b). Hence, pathways through CO can provide indirect conversion of CO2 to solid carbon products.

6.3.2.3 Photochemical Conversion Pathways

Photocatalytic CO2 reduction has been widely researched to produce C1 and C2 gas and liquid products (Fu et al. 2019). The direct production of solid carbon materials from CO2 by photocatalysis is challenging due the limited reaction pathways in photochemistry (Li et al. 2019a). However, many photocatalytic products like methane and CO are valuable substrates for downstream carbon material production through chemical synthesis (Duan et al. 2013). A viable approach for carbon material production leveraging photocatalysis is a tandem process, where photocatalytic CO2 reduction first converts CO2 into gaseous products like methane and CO, followed by conversion of methane and CO into carbon materials via (e.g., pyrolytic) decomposition (Anisimov et al. 2010; Shen and Lua 2015).

6.3.2.4 Plasmachemical Conversion Pathways

Plasmachemical processes can facilitate thermodynamically unfavorable reactions, such as CO2 activation, at relatively low temperatures. Several nonthermal plasma sources have been applied to transform CO2, including dielectric barrier discharge, microwave discharge, and gliding arc (Mei et al. 2014). Plasma processing parameters such as discharge power, gas composition, feed flow rate, and dielectric material affect the conversion of reactants. Catalysts can be used to assist plasma reactions, especially when packed-bed reactors are employed, which can integrate catalysts to enhance the synergies between catalysts and plasmas.

Plasma catalysis can be performed via two main types of reactor configurations, in-plasma catalysis and post-plasma catalysis (George et al. 2021; Wang et al. 2018). For the in-plasma catalysis configuration, catalysts are placed in the plasma discharge region, which facilitates a direct reaction between the plasma species and the catalyst surface, potentially enhancing the conversion of reactants to desired products (e.g., nanocarbon materials). In post-plasma catalysis, the reaction occurs in two steps. First, CO2 could participate in in-plasma reactions if plasma is used to split CO2 into CO and O2, while in the second stage, CO and/or CO2 could undergo catalytic surface reactions. The chemical and physical interactions between the nonthermal plasma and catalyst strongly affect the percentage of CO2 transformed and the product selectivity (George et al. 2021). Plasma can impact the catalyst by generating excited species and radicals, lowering the activation barrier and enhancing pathways for surface reactions. At the same time, the catalyst can affect the electric field of the plasma, discharge type, generation of micro discharge in pores, and impurity concentration in the plasma. Consequently, plasma cataly-

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

sis may be able to increase energy efficiency, reaction rate, product yield, and catalyst durability; enhance the concentration of active species; and improve selectivity (Neyts et al. 2015). However, Bogaerts et al. (2020) note that good thermal CO2 activation catalysts might not be good plasma catalysts for CTEC conversion because the plasma could change the properties of the catalyst surface and introduce new species (e.g., excited species and reactive species) for surface reactions, thereby changing surface interactions with the catalyst and thus the reaction pathways.

As of April 2024, research on splitting CO2 into carbon materials and O2 with nonthermal plasma remains at the concept stage. Only one paper is known to have been published in this area, which discusses the feasibility of the concept in detail (Centi et al. 2021). However, no experiments have been performed using plasma with or without the help of catalysts to confirm the reaction.

6.3.3 Potential Alternative or Competitor Routes to Elemental Carbon Materials

In a net-zero future, alternative routes to produce elemental carbon materials include methane pyrolysis and lignocellulosic biomass processing. While a detailed discussion of these approaches is out of scope for this study, brief descriptions are provided below, along with their advantages and disadvantages relative to the production of elemental carbon materials from CO2. Life cycle and techno-economic assessments (see Chapter 3) will be needed to help evaluate and compare the emissions impacts and economic viability of these different approaches for producing elemental carbon materials.

6.3.3.1 Methane Pyrolysis

Methane pyrolysis entails the decomposition of methane to form solid carbon and hydrogen (Amin et al. 2011) and is a potential sustainable pathway to produce both elemental carbon products and clean hydrogen, provided the solid carbon products remain sequestered. The process is endothermic (Lewis et al. 2001) with its enthalpy change being 74.5 kJ/mol-CH4 or 37.3 kJ/mol-H2 produced (Catalan et al. 2023):

CH4 → C(s) + 2 H2 (R6.22)

Methane pyrolysis is attractive as a source of hydrogen because, in principle, solid carbon can be separated from the gas-phase hydrogen product without CO2 capture and sequestration, as would be required for sustainable conventional steam methane reforming (Korányi et al. 2022; Sánchez-Bastardo et al. 2021; Timmerberg et al. 2020). Formation of carbon fibers and nanotubes from methane (fully reduced carbon) may be less difficult than a similar pathway from CO2 (fully oxidized carbon), given that a number of the CNT pathways from CO2 are postulated to proceed through methane or other reduced hydrocarbon intermediates (Kim et al. 2020a, 2020b).

6.3.3.2 Lignocellulosic Biomass Processing

Elemental carbon materials derived from plant lignocellulosic biomass represents a competitive platform to CO2-derived carbon materials. Of course, biomass-based carbon materials also are derived from CO2 ultimately, as plants fix CO2 through photosynthesis and convert it into biopolymers such as cellulose and lignin in the plant cell wall. These biopolymers subsequently can be converted into carbon materials like carbon fiber, graphene, CNTs, carbon foam, and others (Li et al. 2022b). Among different biopolymers, lignin has the greatest potential for carbon material manufacturing owing to its high carbon content and aromatic ring structure (Zhang et al. 2022). Recent studies have advanced fundamental understanding of structure–function relationships of how lignin composition (molecular weight, uniformity, linkage profile, and functional group) can impact its structure and performance. This understanding guides the design of new lignin structures to improve the quality and performance of the resulting carbon materials (Li et al. 2022a, 2022b).

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

6.3.4 Challenges for CO2 Conversion to Elemental Carbon Materials

As noted in Section 6.3.2, except for plasmachemical processes, research into CTEC processes has been ongoing for quite some time, especially thermochemical CTEC research that started as early as 1990 (Tamaura and Tahata 1990). However, CTEC research activities and processes are still limited, despite the large market potential for elemental carbon materials discussed in Chapter 2. All four major types of CTEC technologies (thermochemical, electrochemical, photochemical, and plasmachemical) are still at low TRLs, in concept development stages. The sections below identify challenges common to all four CTEC approaches and those unique to each CTEC approach.

6.3.4.1 Common Challenges for the Four CTEC Approaches
  1. Research efforts have been limited: Although research on CTEC conversions, especially in thermochemical approaches, has been ongoing for several decades, research intensity in the four areas have not been high. The first thermochemical Mg-CO2–to–MgO-C conversion reaction was reported in 1978, and the first thermochemical CO2 splitting with cation-excess magnetite was published in 1990. Nonetheless, only 161 papers have been published on thermochemical, electrochemical, photochemical, and plasmachemical CTEC conversion as of April 2024. Therefore, theoretical and experimental work on CTEC conversions are still limited, especially in the areas of photochemical and plasmachemical CTEC, as illustrated in Figure 6-3.
  2. Difficult to compare CTEC approaches: Research activities on the four CTEC approaches are not balanced; thus, it is difficult to assess the advantages and disadvantages of the different reaction pathways.
  3. Substantial energy requirements: Breaking carbon-oxygen bonds in CO2 is energy intensive; thus, all four CTEC approaches require substantial external energy. For these products and processes to have net-zero emissions, this energy will have to be provided by zero-carbon-emission sources of electricity or heat.
  4. Limited understanding of catalyst/reactant material stability and carbon product selectivity: The stabilities of the catalysts/reactant materials used and carbon materials generated during CTEC processes have not been well characterized.
  5. A grand challenge for all of these CTEC technologies is how to convert CO2selectively at high yield to a particular morphological form of solid carbon.
6.3.4.2 Specific Challenges for Each CTEC Approach
6.3.4.2.1 Thermochemical CTEC

Both the decomposition-based and strong reducing agents–based CTEC have low overall energy utilization efficiencies owing to the low mass and heat transfer efficiencies of fixed bed reactors in SOA thermochemical processes, the need to heat the reaction systems to moderately high temperatures to initiate the reaction, and loss of the heat released during the reaction. The two thermochemical CTEC technologies also have their own specific challenges.

  • Decomposition-based CTEC:
    • Fast reactant material deactivation and the resultant slow reaction kinetics owing to coking via direct active site poisoning and pore plugging from the nanoscale carbon particles generated from CO2 reduction, which lowers the reaction rate.
    • Stabilities of cation-excess reactants during their reactions with CO2 at high temperatures are not well studied.
    • The purity and structure of the carbon materials generated have not been systematically evaluated.
  • CTEC using strong reducing agents:
    • Initiation of the elemental-metals-based CTEC processes are slow owing to the need to reach high temperatures (higher than metal melting points) prior to the initiation of the reactions.
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
6.3.4.2.2 Electrochemical CTEC

Most electrochemical CTEC processes require high temperatures, and therefore greater energy input. Room-temperature electrolysis to produce elemental carbon materials is comparatively nascent and does not generate as high-value materials as the high-temperature systems. Deposition of carbon on electrode surfaces can lead to the significant reduction of active sites and a decrease in electrode activity. It is difficult to successfully convert CO2 from gas to solid carbon owing to the complexity in overcoming kinetic barriers and achieving efficient nucleation and solid carbon structure growth.

6.3.4.2.3 Photochemical CTEC

Only one paper has been published using this CTEC technology (Duan et al. 2013). The research was performed with the help of temperature management. Thus, knowledge of reaction mechanisms for photocatalytic CTEC area is entirely lacking.

6.3.4.2.4 Plasmachemical CTEC

This concept has been proposed but not yet confirmed experimentally.

6.3.5 R&D Opportunities for CO2 Conversion to Elemental Carbon Materials

There are many R&D activities that can advance CTEC processes. Research is needed into how the reaction conditions (e.g., temperature, pressure, composition of CO2-containing feedstock), use of catalyst, and other factors affect the types and qualities of carbon materials produced. Work is also needed to fully characterize the carbon materials generated via CTEC technologies and identify their corresponding markets, including the preparation of organic solar cells and light-emitting diodes, supercapacitors, batteries, sensors, and catalysts. The following lays out specific R&D opportunities across thermochemical, electrochemical, photochemical, and plasmachemical CTEC.

Thermochemical CTEC:

  • Decomposition-based CTEC:
    • There is an opportunity to build on knowledge of solid-oxide reactants, including ferrites and SrFeO3-σ, to develop new types of cation-excess materials with high CO2 conversion activity, ~100 percent carbon formation selectivity, and good stability and regeneration ability for decomposition-based CTEC. Utilizing knowledge already gleaned from the solar thermochemical hydrogen solid-oxide materials research could prove fruitful (Wexler et al. 2021, 2023b). Similarly, exploiting knowledge already gathered from solid oxide fuel cell materials characterization and optimization also could be helpful (Muñoz-García et al. 2014; Ritzmann et al. 2016).
    • The carbon materials obtained with decomposition-based CTEC, mainly CNT and graphene, can differ from those generated by reacting CO2 with strong reducing agents because their reaction temperatures are different. Thus, more R&D is needed to discover how to generate different high-value carbon materials with decomposition-based CTEC processes.
  • Strong-reducing-agents-based CTEC:
    • R&D is needed to diversify the elemental carbon products that can be made from the reaction of Mg and CO2. Current systems primarily yield CNTs and graphene. By changing the reaction conditions, other products such as CNFs and CQDs might be accessible.
    • The primary metal used as a strong reducing agent for CTEC is Mg. R&D is needed to explore the pros and cons of using non-Mg metals for metal reduction CTEC processes.

Electrochemical CTEC:

  • Lowering the operating temperature for molten electrolysis or solid oxide electrolyzer cells or coupling these processes with exothermic reactions could lower the overall energy requirements.
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
  • Low-temperature electrolysis with liquid metals is a promising approach for generating elemental carbon materials, but more research is required to access higher-value carbon products. In the reported electrochemical processes, the mechanism of reduction is not well understood.

Photochemical CTEC:

  • According to Duan et al. (2013), the photochemical method does not work on its own; thermocatalysis needs to be coupled to photocatalysis for the reduction of CO2 to C to occur. Thus, dual-function catalysts that can simultaneously accelerate both thermal- and photo-splitting reactions may play a key role and could be a promising future research direction in this area.

Plasmachemical CTEC:

  • A plasmachemical CTEC scheme (Mei et al. 2014) was proposed about 10 years ago. However, no realizable experimental data have been published to confirm the concept. Accordingly, multiple R&D opportunities in this area exist in this area, including:
    • Theory, modeling, and simulation—While such research has begun for some proposed methods (e.g., methane pyrolysis), none is available yet for understanding plasmachemical CO2 reduction mechanisms.
    • Experiments—Experiments need to be conducted to confirm the feasibility of plasmachemical CTEC, especially regarding the use of catalysts both in situ and post-plasma.

6.4 CONCLUSIONS

6.4.1 Findings and Recommendations

Finding 6-1: CO2 to elemental carbon technologies are far from commercialization—Thermochemical, electrochemical, photochemical, and plasmachemical conversion of CO2 to elemental carbon technologies are generally at technology readiness levels of 3, 2 (for room-temperature electrolysis), 1, and 1, respectively, indicating that all are far from commercialization.

Recommendation 6-1: Support basic research to advance CO2 to elemental carbon technologies—Basic Energy Sciences within the Department of Energy’s Office of Science and the National Science Foundation should invest in building the knowledge foundation and accelerating the maturities of the four CO2 to elemental carbon technology areas: thermochemical, electrochemical, photochemical, and plasmachemical.

Finding 6-2: Demanding materials and energy requirements for CO2 to elemental carbon technologies—All CO2 to elemental carbon (CTEC) technologies need strong reducing agents (e.g., Mg or H2) or very negative electrochemical potentials, oxygen-deficient reactant materials (e.g., cation-excess magnetite), and other materials (e.g., molten salts) as part of their conversion process. To develop CTEC technologies with net-zero or net-negative CO2 footprints, the materials used in the CTEC technologies need to be generated from low-carbon-emission sources. In addition, all CTEC technologies need external energy to initiate and/or maintain the reactions.

Finding 6-3: Challenges with activity, selectivity, and stability of redox-active materials key to CO2 to elemental carbon conversion—Redox-active materials are key to the success of CO2 to elemental carbon technologies (CTEC). However, current catalysts for these technologies lack sufficient activity, selectivity, and stability to achieve high performance. Carbon nanotubes and graphene are the primary carbon materials generated from CO2 via CTEC, with CO2-derived graphene performing better than its commercial counterpart. Several other carbon materials, such as carbon nanofibers, are less frequently produced from CO2.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Recommendation 6-2: Fund research into catalysts and materials for CO2 to elemental carbon conversion—Basic Energy Sciences within the Department of Energy’s Office of Science (DOE-BES) and the National Science Foundation should fund basic research into the discovery of high-performance catalysts that are active, morphologically selective, and robust for low-cost CO2 to elemental carbon (CTEC) conversion. DOE-BES, DOE’s Office of Energy Efficiency and Renewable Energy, and DOE’s Office of Fossil Energy and Carbon Management should, jointly or independently, fund research on materials (e.g., catalysts, reducing agents) used in CO2 to elemental carbon processes. These investigations should aim to discover new, optimal materials for catalysis and separation; understand how to control CTEC reactions to increase the diversity of products and selectively generate desired morphologies; and increase energy efficiency of CTEC reaction processes that can be powered by clean energy.

Finding 6-4: Tandem systems have potential to optimize CO2 to elemental carbon conversion—Combining multiple CO2 to elemental carbon technologies—for example, photo/thermal or electro/thermal combinations, either in one-pot or sequential systems, could be more efficient than any single process alone. These superior efficiencies could include increased carbon yield, optimized systems, minimized energy input, and control of desired carbon material morphology.

Recommendation 6-3: Fund the development of tandem CO2 to elemental carbon technologies to maximize economic and environmental benefits—Basic Energy Sciences within the Department of Energy’s (DOE’s) Office of Science, DOE’s Office of Energy Efficiency and Renewable Energy, and the Advanced Research Projects Agency–Energy should fund independently and/or collectively the development of tandem CO2 to elemental carbon (CTEC) technologies that can combine the advantages of different types of CTEC processes to maximize the economic and environmental benefits of the converted carbon materials.

Finding 6-5: Combined capture and conversion of CO2 to elemental carbon can lead to savings—Combining CO2 capture with CO2 to elemental carbon (CTEC) conversion can lead to reductions in capital and operation costs, and carbon footprints, of CTEC technologies, in contrast to discrete operations of CO2 capture and subsequent CO2 reduction to carbon materials.

Recommendation 6-4: Fund research on integrated CO2 capture and conversion to elemental carbon materials to maximize economic and environmental benefits—The Department of Energy’s Office of Fossil Energy and Carbon Management, Office of Energy Efficiency and Renewable Energy, and the Advanced Research Projects Agency–Energy should fund research on integrated CO2 capture and conversion to elemental carbon materials, with particular consideration of technology integration and economic and environmental benefit enhancement.

6.4.2 Research Agenda for Chemical CO2 Conversion to Elemental Carbon Materials

Table 6-3 presents the committee’s research agenda for chemical CO2 conversion to elemental carbon materials, including research needs (numbered by chapter), and related research agenda recommendations (a subset of research-related recommendations from the chapter). The table includes the relevant funding agencies or other actors; whether the need is for basic research, applied research, technology demonstration, or enabling technologies and processes for CO2 utilization; the research theme(s) that the research need falls into; the relevant research area and product class covered by the research need; whether the relevant product(s) are long- or short-lived; and the source of the research need (chapter section, finding, or recommendation). The committee’s full research agenda can be found in Chapter 11.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

TABLE 6-3 Research Agenda for CO2 Conversion to Elemental Carbon Materials

Research, Development, and Demonstration Need Funding Agencies or Other Actors Basic, Applied, Demonstration, Enabling Research Area Product Class Long- or Short-Lived Research Themes Source
6-A. Foundational knowledge of thermochemical, electrochemical, photochemical, and plasma processes to make elemental carbon products from CO2. DOE-BES
NSF
Basic Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Fundamental knowledge Rec. 6-1
Recommendation 6-1: Support basic research to advance CO2 to elemental carbon technologies—Basic Energy Sciences within the Department of Energy’s Office of Science and the National Science Foundation should invest in building the knowledge foundation and accelerating the maturities of the four CO2 to elemental carbon technology areas: thermochemical, electrochemical, photochemical, and plasmachemical.
6-B. Novel and improved catalysts and low-energy reaction processes to produce elemental carbon products from CO2. DOE-BES
DOE-EERE
DOE-FECM
Basic
Applied
Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Catalyst innovation and optimization Reactor design and reaction engineering. Energy efficiency, electrification, and alternative heating Fin. 6-2
Rec. 6-2
6-C. Catalysts and processes that are selective for particular material morphologies. DOE-BES
DOE-EERE
DOE-FECM
Basic
Applied
Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Catalyst innovation and optimization Fin. 6-2
Rec. 6-2
6-D. Enhanced activity, selectivity, and stability of catalysts to achieve high performance of reactions transforming CO2 to elemental carbon products. DOE-BES
NSF
Basic Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Catalyst innovation and optimization Fin. 6-3
Rec. 6-2
6-E. Understanding and control of processes that produce CO2-derived elemental carbon products. DOE-BES
NSF
Basic Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Fundamental knowledge Fin. 6-3
Rec. 6-2
6-F. Reaction electrification and heat integration including plasma processes (thermochemical, plasmachemical, etc.). DOE-BES
DOE-EERE
DOE-FECM
Basic
Applied
Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Reactor design and reaction engineering Energy efficiency, electrification, and alternative heating Fin. 6-3
Rec. 6-2
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
6-G. Separation of catalyst from solid carbon products, and different elemental carbon materials from each other. DOE-BES
DOE-EERE
DOE-FECM
Basic
Applied
Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Separations Fin. 6-3
Rec. 6-2
Recommendation 6-2: Fund research into catalysts and materials for CO2 to elemental carbon conversion—Basic Energy Sciences within the Department of Energy’s Office of Science (DOE-BES) and the National Science Foundation should fund basic research into the discovery of high-performance catalysts that are active, morphologically selective, and robust for low-cost CO2 to elemental carbon (CTEC) conversion. DOE-BES, DOE’s Office of Energy Efficiency and Renewable Energy, and DOE’s Office of Fossil Energy and Carbon Management should, jointly or independently, fund research on materials (e.g., catalysts, reducing agents) used in CO2 to elemental carbon processes. These investigations should aim to discover new, optimal materials for catalysis and separation; understand how to control CTEC reactions to increase the diversity of products and selectively generate desired morphologies; and increase energy efficiency of CTEC reaction processes that can be powered by clean energy.
6-H. Development of tandem processes to produce elemental carbon products from CO2. DOE-BES
DOE-EERE
DOE-ARPA-E
Basic
Applied
Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Integrated systems Reactor design and reaction engineering Fin. 6-4
Rec. 6-3
Recommendation 6-3: Fund the development of tandem CO2 to elemental carbon technologies to maximize economic and environmental benefits—Basic Energy Sciences within the Department of Energy’s (DOE’s) Office of Science, DOE’s Office of Energy Efficiency and Renewable Energy, and the Advanced Research Projects Agency–Energy should fund independently and/or collectively the development of tandem CO2 to elemental carbon (CTEC) technologies that can combine the advantages of different types of CTEC processes to maximize the economic and environmental benefits of the converted carbon materials.
6-I. Integrated CO2 capture and conversion to elemental carbon materials including improved technology integration and enhanced economic and/or environmental benefits. DOE-FECM
DOE-EERE
DOE-ARPA-E
Applied Chemical Elemental
Carbon
Materials
Long-lived
Short-lived
Integrated systems Fin. 6-5
Rec. 6-4
Recommendation 6-4: Fund research on integrated CO2 capture and conversion to elemental carbon materials to maximize economic and environmental benefits—The Department of Energy’s Office of Fossil Energy and Carbon Management, Office of Energy Efficiency and Renewable Energy, and the Advanced Research Projects Agency–Energy should fund research on integrated CO2 capture and conversion to elemental carbon materials, with particular consideration of technology integration and economic and environmental benefit enhancement.

NOTE: ARPA-E = Advanced Research Projects Agency–Energy; BES = Basic Energy Sciences; DOE = Department of Energy; EERE = Office of Energy Efficiency and Renewable Energy; FECM = Office of Fossil Energy Carbon Management; NSF = National Science Foundation.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

6.5 REFERENCES

Abraham, J., R. Arunima, K.C. Nimitha, S.C. George, and S. Thomas. 2021. “Chapter 3—One-Dimensional (1D) Nanomaterials: Nanorods and Nanowires; Nanoscale Processing.” Pp. 71–101 in Nanoscale Processing, S. Thomas and P. Balakrishnan, eds. Micro and Nano Technologies. Elsevier. https://doi.org/10.1016/B978-0-12-820569-3.00003-7.

Álvarez, A., A. Bansode, A. Urakawa, A.V. Bavykina, T.A. Wezendonk, M. Makkee, J. Gascon, and F. Kapteijn. 2017. “Challenges in the Greener Production of Formates/Formic Acid, Methanol, and DME by Heterogeneously Catalyzed CO2 Hydrogenation Processes.” Chemical Reviews 117(14):9804–9838.

Amin, A.M., E. Croiset, and W. Epling. 2011. “Review of Methane Catalytic Cracking for Hydrogen Production.” International Journal of Hydrogen Energy 36(4):2904–2935. https://doi.org/10.1016/j.ijhydene.2010.11.035.

Anctil, A., C.W. Babbitt, R.P. Raffaelle, and B.J. Landi. 2011. “Material and Energy Intensity of Fullerene Production.” Environmental Science and Technology 45(6):2353–2359.

Anisimov, A.S., A.G. Nasibulin, H. Jiang, P. Launois, J. Cambedouzou, S.D. Shandakov, and E.I. Kauppinen. 2010. “Mechanistic Investigations of Single-Walled Carbon Nanotube Synthesis by Ferrocene Vapor Decomposition in Carbon Monoxide.” Carbon 48(2):380–388. https://doi.org/10.1016/j.carbon.2009.09.040.

Bacon, M., S.J. Bradley, and T. Nann. 2014. “Graphene Quantum Dots.” Particle and Particle Systems Characterization 31(4):415–428. https://doi.org/10.1002/ppsc.201300252.

Bansal, D., S. Pillay, and U. Vaidya. 2013. “Nanographene-Reinforced Carbon/Carbon Composites.” Carbon 55:233–244.

Bhugun, I., D. Lexa, and J.-M. Savéant. 1996. “Catalysis of the Electrochemical Reduction of Carbon Dioxide by Iron (0) Porphyrins. Synergistic Effect of Lewis Acid Cations.” The Journal of Physical Chemistry 100(51):19981–19985.

Bogaerts, A., X. Tu, J.C. Whitehead, G. Centi, L. Lefferts, O. Guaitella, F. Azzolina-Jury, et al. 2020. “The 2020 Plasma Catalysis Roadmap.” Journal of Physics D: Applied Physics 53(44):443001.

Budyka, M.F., E.F. Sheka, and N.A. Popova. 2017. “Graphene Quantum Dots: Theory and Experiment.” Reviews on Advanced Materials Science 51(1).

Cartwright, R.J., S. Esconjauregui, R.S. Weatherup, D. Hardeman, Y. Guo, E. Wright, D. Oakes, S. Hofmann, and J. Robertson. 2014. “The Role of the sp2: sp3 Substrate Content in Carbon Supported Nanotube Growth.” Carbon 75:327–334.

Catalan, L.J.J., B. Roberts, and E. Rezaei. 2023. “A Low Carbon Methanol Process Using Natural Gas Pyrolysis in a Catalytic Molten Metal Bubble Reactor.” Chemical Engineering Journal 462(April 1):142230. https://doi.org/10.1016/j.cej.2023.142230.

Centi, G., S. Perathoner, and G. Papanikolaou. 2021. “Plasma Assisted CO2 Splitting to Carbon and Oxygen: A Concept Review Analysis.” Journal of CO2Utilization 54:101775.

Chakrabarti, A., J. Lu, J.C. Skrabutenas, T. Xu, Z. Xiao, J.A. Maguire, and N.S. Hosmane. 2011. “Conversion of Carbon Dioxide to Few-Layer Graphene.” Journal of Materials Chemistry 21(26):9491–9493.

Chandra, K.S., and D. Sarkar. 2023. “Nanoscale Reinforcement Efficiency Analysis in Al2O3–MgO–C Refractory Composites.” Materials Science and Engineering: A 865:144613.

Chen, C., and Z. Lou. 2009. “Formation of C60 by Reduction of CO2.” The Journal of Supercritical Fluids 50(1):42–45. https://doi.org/10.1016/j.supflu.2009.04.008.

Chen, Z., Y. Gu, L. Hu, W. Xiao, X. Mao, H. Zhu, and D. Wang. 2017a. “Synthesis of Nanostructured Graphite via Molten Salt Reduction of CO2 and SO2 at a Relatively Low Temperature.” Journal of Materials Chemistry A 5(39):20603–20607.

Chery, D., V. Lair, and M. Cassir. 2015. “CO2 Electrochemical Reduction into CO or C in Molten Carbonates: A Thermodynamic Point of View.” Electrochimica Acta 160:74–81. https://doi.org/10.1016/j.electacta.2015.01.216.

Chung, D.D.L. 2002. “Review Graphite.” Journal of Materials Science 37(8):1475–1489. https://doi.org/10.1023/A:1014915307738.

Collins, G., P.R. Kasturi, R. Karthik, J.-J. Shim, R. Sukanya, and C.B. Breslin. 2023. “Mesoporous Carbon-Based Materials and Their Applications as Non-Precious Metal Electrocatalysts in the Oxygen Reduction Reaction.” Electrochimica Acta 439(January):141678. https://doi.org/10.1016/j.electacta.2022.141678.

Cossutta, M., and J. McKechnie. 2021. “Environmental Impacts and Safety Concerns of Carbon Nanomaterials,” Pp. 249–278 in Carbon Related Materials, S. Kaneko, M. Aono, A. Pruna, M. Can, P. Mele, M. Ertugrul, and T. Endo, eds. Singapore: Springer Singapore. https://doi.org/10.1007/978-981-15-7610-2_11.

Dabees, S., T. Osman, and B.M. Kamel. 2023. “Mechanical, Thermal, and Flammability Properties of Polyamide-6 Reinforced with a Combination of Carbon Nanotubes and Titanium Dioxide for Under-the-Hood Applications.” Journal of Thermoplastic Composite Materials 36(4):1545–1575.

Deng, B., X. Mao, W. Xiao, and D. Wang. 2017. “Microbubble Effect-Assisted Electrolytic Synthesis of Hollow Carbon Spheres from CO2.” Journal of Materials Chemistry A 5(25):12822–12827.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Deng, L., R.J. Young, I.A. Kinloch, Y. Zhu, and S.J. Eichhorn. 2013. “Carbon Nanofibres Produced from Electrospun Cellulose Nanofibres.” Carbon 58:66–75.

Deshmukh, A.A., S.D. Mhlanga, and N.J. Coville. 2010. “Carbon Spheres.” Materials Science and Engineering: R: Reports 70(1–2):1–28.

Dhall, S., ed. 2023. Carbon Nanomaterials and Their Nanocomposite-Based Chemiresistive Gas Sensors: Applications, Fabrication and Commercialization. 1st edition. Amsterdam, Netherlands: Elsevier.

Douglas, A., R. Carter, M. Li, and C.L. Pint. 2018. “Toward Small-Diameter Carbon Nanotubes Synthesized from Captured Carbon Dioxide: Critical Role of Catalyst Coarsening.” ACS Applied Materials and Interfaces 10(22):19010–19018.

Driscoll, J.A. 1978. “A Demonstration of Burning Magnesium and Dry Ice.” Journal of Chemical Education 55(7):450.

Duan, Y.Q., T. Du, X.W. Wang, F.S. Cai, Z.H. Yuan. 2013. “Photoassisted CO2 Conversion to Carbon by Reduced NiFe2O4.” Advanced Materials Research 726–731:420–424. https://doi.org/10.4028/www.scientific.net/AMR.726-731.420.

Eslami, Z., F. Yazdani, and M.A. Mirzapour. 2015. “Thermal and Mechanical Properties of Phenolic-Based Composites Reinforced by Carbon Fibers and Multiwall Carbon Nanotubes.” Composites Part A: Applied Science and Manufacturing 72:22–31.

Esrafilzadeh, D., A. Zavabeti, R. Jalili, P. Atkin, J. Choi, B.J. Carey, R. Brkljača, et al. 2019. “Room Temperature CO2 Reduction to Solid Carbon Species on Liquid Metals Featuring Atomically Thin Ceria Interfaces.” Nature Communications 10(1):865. https://doi.org/10.1038/s41467-019-08824-8.

Fu, Z., Q. Yang, Z. Liu, F. Chen, F. Yao, T. Xie, Y. Zhong, et al. 2019. “Photocatalytic Conversion of Carbon Dioxide: From Products to Design the Catalysts.” Journal of CO2Utilization 34(December):63–73. https://doi.org/10.1016/j.jcou.2019.05.032.

Gautam, G.S., E.B. Stechel, and E.A. Carter. 2020. “Exploring Ca-Ce-M-O (M = 3d Transition Metal) Oxide Perovskites for Solar Thermochemical Applications,” Chemistry Mater 32(9964). https://doi.org/10.1021/acs.chemmater.0c02912.

George, A., B. Shen, M. Craven, Y. Wang, D. Kang, C. Wu, and X. Tu. 2021. “A Review of Non-Thermal Plasma Technology: A Novel Solution for CO2 Conversion and Utilization.” Renewable and Sustainable Energy Reviews 135:109702.

Gharpure, A., and R. Vander Wal. 2023. “Improving Graphenic Quality by Oxidative Liberation of Crosslinks in Non-Graphitizable Carbons.” Carbon 209:118010.

Goerzen, D., D.A. Heller, and R. Meidl. 2024. “Balancing Safety and Innovation: Shaping Responsible Carbon Nanotube Policy.” Baker Institute. https://www.bakerinstitute.org/research/balancing-safety-and-innovation-shaping-responsiblecarbon-nanotube-policy.

Gui, G., J. Li, and J. Zhong. 2008. “Band Structure Engineering of Graphene by Strain: First-Principles Calculations.” Physical Review B 78(7):075435.

Güneş, F., G.H. Han, H.-J. Shin, S.Y. Lee, M. Jin, D.L. Duong, S.J. Chae, et al. 2011. “UV-Light-Assisted Oxidative sp3 Hybridization of Graphene.” Nano 6(5):409–418.

Han, G.H., J. Bang, G. Park, S. Choe, Y.J. Jang, H.W. Jang, S.Y. Kim, and S.H. Ahn. 2023a. “Recent Advances in Electrochemical, Photochemical, and Photoelectrochemical Reduction of CO2 to C2+ Products.” Small 19(16):2205765. https://doi.org/10.1002/smll.202205765.

Han, X., K.K. Ostrikov, J. Chen, Y. Zheng, and X. Xu. 2023b. “Electrochemical Reduction of Carbon Dioxide to Solid Carbon: Development, Challenges, and Perspectives.” Energy and Fuels 37(17):12665–12684. https://doi.org/10.1021/acs.energyfuels.3c02204.

He, B., M. Feng, X. Chen, and J. Sun. 2021. “Multidimensional (0D–3D) Functional Nanocarbon: Promising Material to Strengthen the Photocatalytic Activity of Graphitic Carbon Nitride.” Green Energy and Environment 6(6):823–845.

Hiremath, N., J. Mays, and G. Bhat. 2017. “Recent Developments in Carbon Fibers and Carbon Nanotube-Based Fibers: A Review.” Polymer Reviews 57(2):339–368.

Hoffmann, R., A.A. Kabanov, A.A. Golov, and D.M. Proserpio. 2016. “Homo Citans and Carbon Allotropes: For an Ethics of Citation.” Angewandte Chemie International Edition 55(37):10962–10976. Database updated February 15, 2024. https://www.sacada.info/sacada_3D.php.

Hong, S.-H., J.-S. Choi, S.-J. Yoo, and Y.-S. Yoon. 2023. “Structural Benefits of Using Carbon Nanotube Reinforced High-Strength Lightweight Concrete Beams.” Developments in the Built Environment 16:100234.

Hu, L., Y. Song, J. Ge, J. Zhu, and S. Jiao. 2015. “Capture and Electrochemical Conversion of CO2 to Ultrathin Graphite Sheets in CaCl2-based Melts.” Journal of Materials Chemistry A 3(42):21211–21218.

Hu, L., Y. Song, S. Jiao, Y. Liu, J. Ge, H. Jiao, J. Zhu, J. Wang, H. Zhu, and D.J. Fray. 2016. “Direct Conversion of Greenhouse Gas CO2 into Graphene via Molten Salts Electrolysis.” ChemSusChem 9(6):588–594.

Hu, L., W. Yang, Z. Yang, and J. Xu. 2019. “Fabrication of Graphite via Electrochemical Conversion of CO2 in a CaCl2 Based Molten Salt at a Relatively Low Temperature.” RSC Advances 9(15):8585–8593.

Hut, G., H.G. Östlund, and K. van der Borg. 1986. “Fast and Complete CO2-to-Graphite Conversion for 14C Accelerator Mass Spectrometry.” Radiocarbon 28(2A):186–190.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Irfan, M., K. Zuraiqi, C.K. Nguyen, T.C. Le, F. Jabbar, M. Ameen, C.J. Parker, et al. 2023. “Liquid Metal-Based Catalysts for the Electroreduction of Carbon Dioxide into Solid Carbon.” Journal of Materials Chemistry A 11(27):14990–14996. https://doi.org/10.1039/D3TA01379K.

Jana, P., and A. Dev. 2022. “Carbon Quantum Dots: A Promising Nanocarrier for Bioimaging and Drug Delivery in Cancer.” Materials Today Communications 32(August):104068. https://doi.org/10.1016/j.mtcomm.2022.104068.

Jara, A.D., A. Betemariam, G. Woldetinsae, and J.Y. Kim. 2019. “Purification, Application and Current Market Trend of Natural Graphite: A Review.” International Journal of Mining Science and Technology 29(5):671–689.

Jhonsi, M.A. 2018. “Carbon Quantum Dots for Bioimaging.” In State of the Art in Nano-Bioimaging. IntechOpen. https://doi.org/10.5772/intechopen.72723.

Kaur, J., K. Millington, and S. Smith. 2016. “Producing High-Quality Precursor Polymer and Fibers to Achieve Theoretical Strength in Carbon Fibers: A Review.” Journal of Applied Polymer Science 133(38).

Keypour, H., M. Noroozi, and A. Rashidi. 2013. “An Improved Method for the Purification of Fullerene from Fullerene Soot with Activated Carbon, Celite, and Silica Gel Stationary Phases.” Journal of Nanostructure in Chemistry 3:1–9.

Khan, A., and K.A. Alamry. 2022. “Surface Modified Carbon Nanotubes: An Introduction.” Pp. 1–25 in Surface Modified Carbon Nanotubes Volume 1: Fundamentals, Synthesis and Recent Trends. American Chemical Society.

Kim, A., J.K. Dash, P. Kumar, and R. Patel. 2021. “Carbon-Based Quantum Dots for Photovoltaic Devices: A Review.” ACS Applied Electronic Materials 4(1):27–58.

Kim, G.M., W.Y. Choi, J.H. Park, S.J. Jeong, J.-E. Hong, W. Jung, and J.W. Lee. 2020a. “Electrically Conductive Oxidation-Resistant Boron-Coated Carbon Nanotubes Derived from Atmospheric CO2 for Use at High Temperature.” ACS Applied Nano Materials 3(9):8592–8597. https://doi.org/10.1021/acsanm.0c01909.

Kim, G.M., W.-G. Lim, D. Kang, J.H. Park, H. Lee, J. Lee, and J.W. Lee. 2020b. “Transformation of Carbon Dioxide into Carbon Nanotubes for Enhanced Ion Transport and Energy Storage.” Nanoscale 12(14):7822–7833. https://doi.org/10.1039/C9NR10552B.

Kim, J., M. Shin, S.H. So, S. Hong, D.Y. Park, C. Kim, and C.R. Park. 2024. “Correlation Between Structural Characteristics of Edge Selectively-Oxidized Graphite and Electrochemical Performances Under Fast Charging Condition.” Carbon (218):118664.

Kim, J.-S., J.-R. Ahn, C.W. Lee, Y. Murakami, and D. Shindo. 2001. “Morphological Properties of Ultra-fine (Ni, Zn)-Ferrites and Their Ability to Decompose CO2.” Journal of Materials Chemistry 11(12):3373–3376.

Kim, S.-H., J.T. Jang, J. Sim, J.-H. Lee, S.-C. Nam, and C.Y. Park. 2019. “Carbon Dioxide Decomposition Using SrFeCo0. 5Ox, a Nonperovskite-Type Metal Oxide.” Journal of CO2Utilization 34:709–715.

Komarneni, S., M. Tsuji, Y. Wada, and Y. Tamaura. 1997. “Nanophase Ferrites for CO2 Greenhouse Gas Decomposition.” Journal of Materials Chemistry 7(12):2339–2340. https://doi.org/10.1039/a705849g.

Komatsu, N., T. Ohe, and K. Matsushige. 2004. “A Highly Improved Method for Purification of Fullerenes Applicable to Large-Scale Production.” Carbon 42(1):163–167.

Kondratenko, E.V., G. Mul, J. Baltrusaitis, G.O. Larrazábal, and J. Pérez-Ramírez. 2013. “Status and Perspectives of CO2 Conversion into Fuels and Chemicals by Catalytic, Photocatalytic and Electrocatalytic Processes.” Energy and Environmental Science 6(11):3112–3135.

Korányi, T.I., M. Németh, A. Beck, and A. Horváth. 2022. “Recent Advances in Methane Pyrolysis: Turquoise Hydrogen with Solid Carbon Production.” Energies 15(17):6342. https://doi.org/10.3390/en15176342.

Kordek-Khalil, K., A. Moyseowicz, P. Rutkowski, and G. Gryglewicz. 2024. “Excellent Performance of Nitrogen-Doped Carbon Nanofibers as Electrocatalysts for Water Splitting Reactions.” International Journal of Hydrogen Energy 52:494–506.

Kosari, M., A.M.H. Lim, Y. Shao, B. Li, K.M. Kwok, A.M. Seayad, A. Borgna, and H.C. Zeng. 2022. “Thermocatalytic CO2 Conversion by Siliceous Matter: A Review.” Journal of Materials Chemistry A.

Kozinsky, B., and N. Marzari. 2006. “Static Dielectric Properties of Carbon Nanotubes from First Principles.” Physical Review Letters 96(16):166801.

Lallave, M., J. Bedia, R. Ruiz-Rosas, J. Rodríguez-Mirasol, T. Cordero, J.C. Otero, M. Marquez, A. Barrero, and I.G. Loscertales. 2007. “Filled and Hollow Carbon Nanofibers by Coaxial Electrospinning of Alcell Lignin Without Binder Polymers.” Advanced Materials 19(23):4292–4296.

Lau, J., G. Dey, and S. Licht. 2016. “Thermodynamic Assessment of CO2 to Carbon Nanofiber Transformation for Carbon Sequestration in a Combined Cycle Gas or a Coal Power Plant.” Energy Conversion and Management 122:400–410.

Le, T.-H., and H. Yoon. 2019. “Strategies for Fabricating Versatile Carbon Nanomaterials from Polymer Precursors.” Carbon 152:796–817.

Lee, K.H., S.H. Lee, and R.S. Ruoff. 2020. “Synthesis of Diamond-Like Carbon Nanofiber Films.” ACS nano 14(10):13663–13672.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Lerouge, S., M.R. Wertheimer, and L.H. Yahia. 2001. “Plasma Sterilization: A Review of Parameters, Mechanisms, and Limitations.” Plasmas and Polymers 6:175–188.

Lesiak, B., L. Kövér, J. Tóth, J. Zemek, P. Jiricek, A. Kromka, and N.J.A.S.S. Rangam. 2018. “C sp2/sp3 Hybridisations in Carbon Nanomaterials–XPS and (X) AES study.” Applied Surface Science 452:223–231.

Lewis, M.A., M. Serban, C.L. Marshall, and D. Lewis. 2001. “Direct Contact Pyrolysis of Methane Using Nuclear Reactor Heat.” Lemont, IL: Argonne National Laboratory. https://publications.anl.gov/anlpubs/2001/11/41245.pdf.

Li, A., Q. Cao, G. Zhou, B.V.K.J. Schmidt, W. Zhu, X. Yuan, H. Huo, J. Gong, and M. Antonietti. 2019a. “Three-Phase Photocatalysis for the Enhanced Selectivity and Activity of CO2 Reduction on a Hydrophobic Surface.” Angewandte Chemie International Edition 58(41):14549–14555. https://doi.org/10.1002/anie.201908058.

Li, F., A. Thevenon, A. Rosas-Hernández, Z. Wang, Y. Li, C.M. Gabardo, A. Ozden, et al. 2020. “Molecular Tuning of CO2to-Ethylene Conversion.” Nature 577(7791):509–513. https://doi.org/10.1038/s41586-019-1782-2.

Li, J., C. Hu, Y.-Y. Wang, X. Meng, S. Xiang, C. Bakker, K. Plaza, A.J. Ragauskas, S.Y. Dai, and J.S. Yuan. 2022a. “Lignin Molecular Design to Transform Green Manufacturing.” Matter 5(10):3513–3529. https://doi.org/10.1016/j.matt.2022.07.011.

Li, J., C. Hu, J. Arreola-Vargas, K. Chen, and J.S. Yuan. 2022b. “Feedstock Design for Quality Biomaterials.” Trends in Biotechnology 40(12):1535–1549. https://doi.org/10.1016/j.tibtech.2022.09.017.

Li, J., Z. Xu, R. Yang, L. Ren, L. Ma, G. Huang, S. Zhang, and Z. Huang. 2023a. “Carbon Cycle in a Steelmaking Mill: Recycling of Kish Graphite and Its Subsequent Application for Steelmaking Carburant.”

Li, K., X. An, K.H. Park, M. Khraisheh, and J. Tang. 2014. “A Critical Review of CO2 Photoconversion: Catalysts and Reactors.” Catalysis Today 224:3–12.

Li, S., A. Pasc, V. Fierro, and A. Celzard. 2016. “Hollow Carbon Spheres, Synthesis and Applications—A Review.” Journal of Materials Chemistry A 4(33):12686–12713.

Li, X., H. Shi, X. Wang, X. Hu, C. Xu, and W. Shao. 2022c. “Direct Synthesis and Modification of Graphene in Mg Melt by Converting CO2: A Novel Route to Achieve High Strength and Stiffness in Graphene/Mg Composites.” Carbon 186:632–643.

Li, X., X. Wang, X. Hu, C. Xu, W. Shao, and K. Wu. 2023b. “Direct Conversion of CO2 to Graphene via Vapor–Liquid Reaction for Magnesium Matrix Composites with Structural and Functional Properties.” Journal of Magnesium and Alloys 11(4):1206–1212.

Li, Z., D. Yuan, H. Wu, W. Li, and D. Gu. 2018. “A Novel Route to Synthesize Carbon Spheres and Carbon Nanotubes from Carbon Dioxide in a Molten Carbonate Electrolyzer.” Inorganic Chemistry Frontiers 5(1):208–216.

Liang, C., Y. Chen, M. Wu, K. Wang, W. Zhang, Y. Gan, H. Huang, et al. 2021. “Green Synthesis of Graphite from CO2 Without Graphitization Process of Amorphous Carbon.” Nature Communications 12(1):119.

Licht, S., A. Douglas, J. Ren, R. Carter, M. Lefler, and C.L. Pint. 2016. “Carbon Nanotubes Produced from Ambient Carbon Dioxide for Environmentally Sustainable Lithium-ion and Sodium-ion Battery Anodes.” ACS Central Science 2(3):162–168.

Lin, K.-S., A.K. Adhikari, Z.-Y. Tsai, Y.-P. Chen, T.-T. Chien, and H.-B. Tsai. 2011. “Synthesis and Characterization of Nickel Ferrite Nanocatalysts for CO2 Decomposition.” Catalysis Today 174(1):88–96.

Liu, H., S. Wu, N. Tian, F. Yan, C. You, and Y. Yang. 2020a. “Carbon Foams: 3D Porous Carbon Materials Holding Immense Potential.” Journal of Materials Chemistry A 8(45):23699–23723.

Liu, J., R. Li, and B. Yang. 2020b. “Carbon Dots: A New Type of Carbon-Based Nanomaterial with Wide Applications.” ACS Central Science 6(12):2179–2195.

Liu, T., L. Zhang, B. Cheng, and J. Yu. 2019. “Hollow Carbon Spheres and Their Hybrid Nanomaterials in Electrochemical Energy Storage.” Advanced Energy Materials 9(17):1803900.

Liu, X., X. Wang, G. Licht, and S. Licht. 2020c. “Transformation of the Greenhouse Gas Carbon Dioxide to Graphene.” Journal of CO2Utilization 36:288–294.

Liu, Y., and S. Kumar. 2012. “Recent Progress in Fabrication, Structure, and Properties of Carbon Fibers.” Polymer Reviews 52(3):234–258.

Lou, Z., C. Chen, H. Huang, and D. Zhao. 2006. “Fabrication of Y–Junction Carbon Nanotubes by Reduction of Carbon Dioxide with Sodium Borohydride.” Diamond and Related Materials 15(10):1540–1543. https://doi.org/10.1016/j.diamond.2005.12.044.

Lu, Q., and F. Jiao. 2016. “Electrochemical CO2 reduction: Electrocatalyst, Reaction Mechanism, and Process Engineering.” Nano Energy 29:439–456.

Luc, W., M. Jouny, J. Rosen, and F. Jiao. 2018. “Carbon Dioxide Splitting Using an Electro-Thermochemical Hybrid Looping Strategy.” Energy and Environmental Science 11(10):2928–2934. https://doi.org/10.1039/C8EE00532J.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Luo, X., H. Zheng, W. Lai, P. Yuan, S. Li, D. Li, and Y. Chen. 2023. “Defect Engineering of Carbons for Energy Conversion and Storage Applications.” Energy and Environmental Materials 6(3):e12402. https://doi.org/10.1002/eem2.12402.

Malode, S.J., M.M. Shanbhag, R. Kumari, D.S. Dkhar, P. Chandra, and N.P. Shetti. 2023. “Biomass-Derived Carbon Nanomaterials for Sensor Applications.” Journal of Pharmaceutical and Biomedical Analysis 222:115102.

marketsandmarkets. 2022. “Carbon Fiber Market, Industry Share Forecast Trends Report.” https://www.marketsandmarkets.com/Market-Reports/carbon-fiber-396.html.

Martirez, J.M.P., J.L. Bao, and E.A. Carter. 2021. “First-Principles Insights into Plasmon-Induced Catalysis.” Annual Review of Physical Chemistry 72:99–119.

Mbayachi, V.B., E. Ndayiragije, T. Sammani, S. Taj, and E.R. Mbuta. 2021. “Graphene Synthesis, Characterization and Its Applications: A Review.” Results in Chemistry 3:100163.

Mei, D., X. Zhu, Y.-L. He, J.D. Yan, and X. Tu. 2014. “Plasma-Assisted Conversion of CO2 in a Dielectric Barrier Discharge Reactor: Understanding the Effect of Packing Materials.” Plasma Sources Science and Technology 24(1):015011.

Mishra, S., S. Kumari, A.C. Mishra, R. Chaubey, and S. Ojha. 2023. “Carbon Nanotube–Synthesis, Purification and Biomedical Applications.” Current Nanomaterials 8(4):328–335.

Mori, S., N. Matsuura, L.L. Tun, and M. Suzuki. 2016. “Direct Synthesis of Carbon Nanotubes from Only CO2 by a Hybrid Reactor of Dielectric Barrier Discharge and Solid Oxide Electrolyser Cell.” Plasma Chemistry and Plasma Processing 36(1):231–239. https://doi.org/10.1007/s11090-015-9681-2.

Motiei, M., Y. Rosenfeld Hacohen, J. Calderon-Moreno, and A. Gedanken. 2001. “Preparing Carbon Nanotubes and Nested Fullerenes from Supercritical CO2 by a Chemical Reaction.” Journal of the American Chemical Society 123(35): 8624–8625. https://doi.org/10.1021/ja015859a.

Moyer, K., M. Zohair, J. Eaves-Rathert, A. Douglas, and C.L. Pint. 2020. “Oxygen Evolution Activity Limits the Nucleation and Catalytic Growth of Carbon Nanotubes from Carbon Dioxide Electrolysis via Molten Carbonates.” Carbon 165:90–99.

Muñoz-García, A.B., A.M. Ritzmann, M. Pavone, J.A. Keith, and E.A. Carter. 2014. “Oxygen Transport in Perovskite-Type Solid Oxide Fuel Cell Materials: Insights from Quantum Mechanics,” Accounts of Chemical Research 47(3340). https://doi.org/10.1021/ar4003174.

NASEM (National Academies of Sciences, Engineering, and Medicine). 2019. Gaseous Carbon Waste Streams Utilization: Status and Research Needs. Washington, DC: The National Academies Press. https://doi.org/10.17226/25232.

NASEM. 2023. Carbon Dioxide Utilization Markets and Infrastructure: Status and Opportunities: A First Report. Washington, DC: The National Academies Press.

Neyts, E.C., K. Ostrikov, M.K. Sunkara, and A. Bogaerts. 2015. “Plasma Catalysis: Synergistic Effects at the Nanoscale.” Chemical Reviews 115(24):13408–13446.

Novoselova, I.A., N.F. Oliynyk, and S.V. Volkov. 2007. “Electrolytic Production of Carbon Nanotubes in Chloride-Oxide Melts Under Carbon Dioxide Pressure.” Pp. 459–465 in Hydrogen Materials Science and Chemistry of Carbon Nanomaterials. D.V. Schur, B. Baranowski, A.P. Shpak, V.V. Skorokhod, A. Kale, T.N. Veziroglu, S.Y. Zaginaichenko, eds. Berlin: Springer.

Ognibene, T.J., G. Bench, J.S. Vogel, G.F. Peaslee, and S. Murov. 2003. “A High-Throughput Method for the Conversion of CO2 Obtained from Biochemical Samples to Graphite in Septa-Sealed Vials for Quantification of 14C via Accelerator Mass Spectrometry.” Analytical Chemistry 75(9):2192–2196.

Overa, S., B.H. Ko, Y. Zhao, and F. Jiao. 2022. “Electrochemical Approaches for CO2 Conversion to Chemicals: A Journey Toward Practical Applications.” Accounts of Chemical Research 55(5):638–648.

Pappijn, C.A.R., M. Ruitenbeek, M.-F. Reyniers, and K.M. Van Geem. 2020. “Challenges and Opportunities of Carbon Capture and Utilization: Electrochemical Conversion of CO2 to Ethylene.” Frontiers in Energy Research 8:557466.

Park, S.-J. 2015. Carbon Fibers. Vol. 210. Springer Series in Materials Science. Dordrecht: Springer Netherlands. https://doi.org/10.1007/978-94-017-9478-7.

Parker, D.H., K. Chatterjee, P. Wurz, K.R. Lykke, M.J. Pellin, L.M. Stock, and J.C. Hemminger. 1992. “Fullerenes and Giant Fullerenes: Synthesis, Separation, and Mass Spectrometric Characterization.” Carbon 30(8):1167–1182.

Pease, R.N., and P.R. Chesebro. 1928. “Equilibrium in the Reaction CH4 + 2H2O ⇄ CO2 + 4H21.” Journal of the American Chemical Society 50(5):1464–1469. https://doi.org/10.1021/ja01392a035.

Pérez-Gallent, E., C. Sánchez-Martínez, L.F.G. Geers, S. Turk, R. Latsuzbaia, and E.L.V. Goetheer. 2020. “Overcoming Mass Transport Limitations in Electrochemical Reactors with a Pulsating Flow Electrolyzer.” Industrial and Engineering Chemistry Research 59(13):5648–5656. https://doi.org/10.1021/acs.iecr.9b06925.

Pillai, S.M., G.L. Tembe, and M. Ravindranathan. 1992. “Dimerization of Ethene to 2-Butene and Metathesis with 1-Butene by Sequential Use of Homogeneous Catalyst Systems.” Applied Catalysis A: General 81(2):273–278.

Qu, Q., J. Guo, H. Wang, K. Zhang, and J. Li. 2024. “Carbon Nanofibers Implanted Porous Catalytic Metal Oxide Design as Efficient Bifunctional Electrode Host Material for Lithium-Sulfur Battery.” Electrochimica Acta 473:143454.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Rajagopalan, R., Y. Tang, X. Ji, C. Jia, and H. Wang. 2020. “Advancements and Challenges in Potassium Ion Batteries: A Comprehensive Review.” Advanced Functional Materials 30(12):1909486.

Ramazani, A., M.A. Moghaddasi, A.M. Malekzadeh, S. Rezayati, Y. Hanifehpour, and S.W. Joo. 2021. “Industrial Oriented Approach on Fullerene Preparation Methods.” Inorganic Chemistry Communications 125:108442.

Rani, P., and V.K. Jindal. 2013. “Designing Band Gap of Graphene by B and N Dopant Atoms.” RSC Advances 3(3):802–812.

Ranzoni, A., and M.A. Cooper. 2017. “The Growing Influence of Nanotechnology in Our Lives.” Pp. 1–20 in Micro- and Nanotechnology in Vaccine Development, M. Skwarczynski and I. Tóth, eds. Micro and Nano Technologies Series. Amsterdam and Boston: Elsevier/William Andrew.

Rathinavel, S., K. Priyadharshini, and D. Panda. 2021. “A Review on Carbon Nanotube: An Overview of Synthesis, Properties, Functionalization, Characterization, and the Application.” Materials Science and Engineering: B 268:115095.

Ren, J., and S. Licht. 2016. “Tracking Airborne CO2 Mitigation and Low Cost Transformation into Valuable Carbon Nanotubes.” Scientific Reports 6(1):27760. https://doi.org/10.1038/srep27760.

Ren, J., F.-F. Li, J. Lau, L. González-Urbina, and S. Licht. 2015. “One-Pot Synthesis of Carbon Nanofibers from CO2.” Nano Letters 15(9):6142–6148.

Ren, J., A. Yu, P. Peng, M. Lefler, F.-F. Li, and S. Licht. 2019. “Recent Advances in Solar Thermal Electrochemical Process (STEP) for Carbon Neutral Products and High Value Nanocarbons.” Accounts of Chemical Research 52(11):3177–3187. https://doi.org/10.1021/acs.accounts.9b00405.

Rezaei, A., and E. Hashemi. 2021. “A Pseudohomogeneous Nanocarrier Based on Carbon Quantum Dots Decorated with Arginine as an Efficient Gene Delivery Vehicle.” Scientific Reports 11(1):13790.

Ritzmann, A.M., J.M. Dieterich, and E.A. Carter. 2016. “Density Functional Theory + U Analysis of the Electronic Structure and Defect Chemistry of LSCF (La0.5Sr0.5Co0.25Fe0.75O3-δ).” Physical Chemistry Chemical Physics 18:12260–12269. https://doi.org/10.1039/C6CP01720G.

Rodríguez-Manzo, J.A., M. Terrones, H. Terrones, H.W. Kroto, L. Sun, and F. Banhart. 2007. “In Situ Nucleation of Carbon Nanotubes by the Injection of Carbon Atoms into Metal Particles.” Nature Nanotechnology 2(5):307–311.

Sajna, M.S., S. Zavahir, A. Popelka, P. Kasak, A. Al-Sharshani, U. Onwusogh, M. Wang, H. Park, and D. S. Han. 2023. “Electrochemical System Design for CO2 Conversion: A Comprehensive Review.” Journal of Environmental Chemical Engineering 11(5):110467.

Samiee, S., and E.K. Goharshadi. 2014. “Graphene Nanosheets as Efficient Adsorbent for an Azo Dye Removal: Kinetic and Thermodynamic Studies.” Journal of Nanoparticle Research 16:1–16.

Sánchez-Bastardo, N., R. Schlögl, and H. Ruland. 2021. “Methane Pyrolysis for Zero-Emission Hydrogen Production: A Potential Bridge Technology from Fossil Fuels to a Renewable and Sustainable Hydrogen Economy.” Industrial and Engineering Chemistry Research 60(32):11855–11881. https://doi.org/10.1021/acs.iecr.1c01679.

Sasikumar, B., S.A.G. Krishnan, M. Afnas, G. Arthanareeswaran, P.S. Goh, and A.F. Ismail. 2023. “A Comprehensive Performance Comparison on the Impact of MOF-71, HNT, SiO2, and Activated Carbon Nanomaterials in Polyetherimide Membranes for Treating Oil-in-Water Contaminants.” Journal of Environmental Chemical Engineering 11(1):109010.

Schaefer, H.-E. 2010. Nanoscience: The Science of the Small in Physics, Engineering, Chemistry, Biology and Medicine. Berlin, Heidelberg: Springer. https://doi.org/10.1007/978-3-642-10559-3.

Sengupta, D., S. Chen, A. Michael, C.Y. Kwok, S. Lim, Y. Pei, and A.G. Kottapalli Prakash. 2020. “Single and Bundled Carbon Nanofibers as Ultralightweight and Flexible Piezoresistive Sensors.” NPJ Flex Electron 4(1):9.

Shakoorioskooie, M., Y.Z. Menceloglu, S. Unal, and S.H. Soytas. 2018. “Rapid Microwave-Assisted Synthesis of Platinum Nanoparticles Immobilized in Electrospun Carbon Nanofibers for Electrochemical Catalysis.” ACS Applied Nano Materials 1(11):6236–6246.

Shen, Y., and A.C. Lua. 2015. “Synthesis of Ni and Ni–Cu Supported on Carbon Nanotubes for Hydrogen and Carbon Production by Catalytic Decomposition of Methane.” Applied Catalysis B: Environmental 164(March):61–69. https://doi.org/10.1016/j.apcatb.2014.08.038.

Sim, J., S.-H. Kim, J.-Y. Kim, K. Bong Lee, S.-C. Nam, and C.Y. Park. 2020. “Enhanced Carbon Dioxide Decomposition Using Activated SrFeO3−δ.” Catalysts 10(11):1278.

Snoeckx, R., and A. Bogaerts. 2017. “Plasma Technology—A Novel Solution for CO 2 Conversion?” Chemical Society Reviews 46(19):5805–5863.

Son, D.-H., D. Hwangbo, H. Suh, B.-I. Bae, S. Bae, and C.-S. Choi. 2023. “Mechanical Properties of Mortar and Concrete Incorporated with Concentrated Graphene Oxide, Functionalized Carbon Nanotube, Nano Silica Hybrid Aqueous Solution.” Case Studies in Construction Materials 18:e01603.

Spinner, N.S., J.A. Vega, and W.E. Mustain. 2012. “Recent Progress in the Electrochemical Conversion and Utilization of CO2.” Catalysis Science Technology 2(1):19–28. https://doi.org/10.1039/C1CY00314C.

Spradling, D.M., and R.A. Guth. 2003. “Carbon Foams.” Advanced Materials and Processes 161(11):29–31.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Strudwick, A.J., N.E. Weber, M.G. Schwab, M. Kettner, R.T. Weitz, J.R. Wünsch, K. Müllen, and H. Sachdev. 2015. “Chemical Vapor Deposition of High Quality Graphene Films from Carbon Dioxide Atmospheres.” ACS Nano 9(1):31–42.

Su, H., T. Wu, D. Cui, X. Lin, X. Luo, Y. Wang, and L. Han. 2020. “The Application of Graphene Derivatives in Perovskite Solar Cells.” Small Methods 4(10):2000507.

Surovtseva, D., E. Crossin, R. Pell, and L. Stamford. 2022. “Toward a Life Cycle Inventory for Graphite Production.” Journal of Industrial Ecology 26(3):964–979.

Tackett, B.M., E. Gomez, and J.G. Chen. 2019. “Net Reduction of CO2 via Its Thermocatalytic and Electrocatalytic Transformation Reactions in Standard and Hybrid Processes.” Nature Catalysis 2(5):381–386. https://doi.org/10.1038/s41929-019-0266-y.

Tamaura, Y., and M. Tahata. 1990. “Complete Reduction of Carbon Dioxide to Carbon Using Cation-Excess Magnetite.” Nature 346(6281):255–256.

Tang, J., J. Tang, M. Mayyas, M.B. Ghasemian, J. Sun, M.A. Rahim, J. Yang, et al. 2022. “Liquid-Metal-Enabled Mechanical-Energy-Induced CO2 Conversion.” Advanced Materials 34(1):2105789. https://doi.org/10.1002/adma.202105789.

Tao, Y., S.D. Ebbesen, W. Zhang, and M.B. Mogensen. 2014. “Carbon Nanotube Growth on Nanozirconia Under Strong Cathodic Polarization in Steam and Carbon Dioxide.” ChemCatChem 6(5):1220–1224. https://doi.org/10.1002/cctc.201300941.

Taylor, R., G.J. Langley, H.W. Kroto, and D.R.M. Walton. 1993. “Formation of C60 by Pyrolysis of Naphthalene.” Nature 366(6457):728–731.

Terrones, M., W.K. Hsu, A. Schilder, H. Terrones, N. Grobert, J.P. Hare, Y.Q. Zhu, et al. 1998. “Novel Nanotubes and Encapsulated Nanowires.” Applied Physics A: Materials Science and Processing 66(3).

Timmerberg, S., M. Kaltschmitt, and M. Finkbeiner. 2020. “Hydrogen and Hydrogen-Derived Fuels Through Methane Decomposition of Natural Gas—GHG Emissions and Costs.” Energy Conversion and Management: X 7:100043. https://doi.org/10.1016/j.ecmx.2020.100043.

Tomkute, V., A. Solheim, and E. Olsen. 2013. “Investigation of High-Temperature CO2 Capture by CaO in CaCl2 Molten Salt.” Energy and Fuels 27(9):5373–5379. https://doi.org/10.1021/ef4009899.

Tsang, W.M., S.J. Henley, V. Stolojan, and S.R.P. Silva. 2006. “Negative Differential Conductance Observed in Electron Field Emission from Band Gap Modulated Amorphous-Carbon Nanolayers.” Applied Physics Letters 89(19).

Tsuji, M., Y. Wada, T. Yamamoto, T. Sano, and Y. Tamaura. 1996a. “CO2 Decomposition by Metallic Phase on Oxygen-Deficient Ni (II)-Bearing Ferrite.” Journal of Materials Science Letters 15:156–158.

Tsuji, M., T. Kodama, T. Yoshida, Y. Kitayama, and Y. Tamaura. 1996b. “Preparation and CO2 Methanation Activity of an Ultrafine Ni (II) Ferrite Catalyst.” Journal of Catalysis 164(2):315–321.

Tucker, S.C. 1999. “Solvent Density Inhomogeneities in Supercritical Fluids.” Chemical Reviews 99(2):391–418.

ul Haque, S., A. Nasar, N. Duteanu, and S. Pandey. 2023. “Carbon Based-Nanomaterials Used in Biofuel Cells—A Review.” Fuel 331:125634.

Urade, A.R., I. Lahiri, and K.S. Suresh. 2023. “Graphene Properties, Synthesis and Applications: A Review.” JOM 75(3):614–630.

Van Tran, V., E. Wi, S.Y. Shin, D. Lee, Y.A. Kim, B.C. Ma, and M. Chang. 2022. “Microgels Based on 0D–3D Carbon Materials: Synthetic Techniques, Properties, Applications, and Challenges.” Chemosphere 307(Pt 3):135981.

Wang, R., L. Sun, X. Zhu, W. Ge, H. Li, Z. Li, H. Zhang et al. 2023. “Carbon Nanotube-Based Strain Sensors: Structures, Fabrication, and Applications.” Advanced Materials Technologies 8(1):2200855.

Wang, X., X. Liu, G. Licht, and S. Licht. 2020. “Calcium Metaborate Induced Thin Walled Carbon Nanotube Syntheses from CO2 by Molten Carbonate Electrolysis.” Scientific Reports 10(1):15146.

Wang, Z., Y. Zhang, E.C. Neyts, X. Cao, X. Zhang, B.W.-L. Jang, and C. Liu. 2018. “Catalyst Preparation with Plasmas: How Does It Work?” ACS Catalysis 8(3):2093–2110.

Wei, W., K. Sun, and Y.H. Hu. 2016. “Direct Conversion of CO2 to 3D Graphene and Its Excellent Performance for Dye-Sensitized Solar Cells with 10% Efficiency.” Journal of Materials Chemistry A 4(31):12054–12057.

Wei, S., H. Shi, X. Li, X. Hu, C. Xu, and X. Wang. 2022. “A Green and Efficient Method for Preparing Graphene Using CO2@ Mg In-Situ Reaction and Its Application in High-Performance Lithium-Ion Batteries.” Journal of Alloys and Compounds 902:163700.

Wexler, R.B., G.S. Gautam, E.B. Stechel, and E.A. Carter. 2021. “Factors Governing Oxygen Vacancy Formation in Oxide Perovskites,” Journal of the American Chemical Society 143(33):13212–13227. https://doi.org/10.1021/jacs.1c05570.

Wexler, R.B., G.S. Gautam, R. Bell, S. Shulda, N.A. Strange, J.A. Trindell, J.D. Sugar, E. Nygren, S. Sainio, A.H. McDaniel, D. Ginley, E.A. Carter, and E.B. Stechel. 2023a. “Multiple and Nonlocal Cation Redox in Ca–Ce–Ti–Mn Oxide Perovskites for Solar Thermochemical Applications.” Energy & Environmental Science 16:2550–2560. https://doi.org/10.1039/D3EE00234A.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Wexler, R.B., E.B. Stechel, and E.A. Carter. 2023b. “Materials Design Directions for Solar Thermochemical Water Splitting,” Pp. 3–64 in Solar Fuels, N.D. Sankir and M. Sankir, eds. Vol. 3. Wiley-Scrivener. https://doi.org/10.1002/9781119752097.ch1.

Windhorst, T., and G. Blount. 1997. “Carbon-Carbon Composites: A Summary of Recent Developments and Applications.” Materials and Design 18(1):11–15.

Wissler, M. 2006. “Graphite and Carbon Powders for Electrochemical Applications.” Journal of Power Sources 156(2):142–150. https://doi.org/10.1016/j.jpowsour.2006.02.064.

Wu, Z.-Y., C. Li, H.-W. Liang, J.-F. Chen, and S.-H. Yu. 2013. “Ultralight, Flexible, and Fire-Resistant Carbon Nanofiber Aerogels from Bacterial Cellulose.” Angewandte Chemie 125(10):2997–3001.

Xie, Z., E. Huang, S. Garg, S. Hwang, P. Liu, and J.G. Chen. 2024. “CO2 Fixation into Carbon Nanofibres Using Electrochemical–Thermochemical Tandem Catalysis.” Nature Catalysis 7(1):98–109. https://doi.org/10.1038/s41929-023-01085-1.

Yaashikaa, P.R., P.S. Kumar, S.J. Varjani, and A. Saravanan. 2019. “A Review on Photochemical, Biochemical and Electrochemical Transformation of CO2 into Value-Added Products.” Journal of CO2Utilization 33:131–147.

Yao, L., C. Sun, H. Lin, G. Li, Z. Lian, R. Song, S. Zhuang, and D. Zhang. 2023. “Electrospun Bi-Decorated BixTiyOz/TiO2 Flexible Carbon Nanofibers and Their Applications on Degradating of Organic Pollutants Under Solar Radiation.” Journal of Materials Science and Technology 150(July):114–123. https://doi.org/10.1016/j.jmst.2022.07.066.

Yap, Y.W., N. Mahmed, M.N. Norizan, S.Z.A. Rahim, M.N.A. Salimi, K.A. Razak, I.S. Mohamad, M.M. Al-Bakri Abdullah, and M.Y.M. Yunus. 2023. “Recent Advances in Synthesis of Graphite from Agricultural Bio-Waste Material: A Review.” Materials 16(9):3601.

Ye, L., N. Syed, D. Wang, J. Guo, J. Yang, J. Buston, R. Singh, M.S. Alivand, G.K. Li, and A. Zavabeti. 2023. “Low-Temperature CO2 Reduction Using Mg–Ga Liquid Metal Interface.” Advanced Materials Interfaces 10(3):2201625. https://doi.org/10.1002/admi.202201625.

Ye, R.-P., J. Ding, W. Gong, M.D. Argyle, Q. Zhong, Y. Wang, C.K. Russell, et al. 2019. “CO2 Hydrogenation to High-Value Products via Heterogeneous Catalysis.” Nature Communications 10(1):5698.

Yoshida, T., M. Tsuji, Y. Tamaura, T. Hurue, T. Hayashida, and K. Ogawa. 1997. “Carbon Recycling System Through Methanation of CO2 in Flue Gas in LNG Power Plant.” Energy Conversion and Management 38:S443–S448.

Yu, R., B. Deng, K. Du, D. Chen, M. Gao, and D. Wang. 2021. “Modulating Carbon Growth Kinetics Enables Electrosynthesis of Graphite Derived from CO2 via a Liquid–Solid–Solid Process.” Carbon 184:426–436.

Yu, W., L. Sisi, Y. Haiyan, and L. Jie. 2020. “Progress in the Functional Modification of Graphene/Graphene Oxide: A Review.” RSC Advances 10(26):15328–15345.

Yu, Z., S. Wang, Z. Xiao, F. Xu, C. Xiang, L. Sun, and Y. Zou. 2024. “Graphene-Supported Cu2O-CoO Heterojunctions for High Performance Supercapacitors.” Journal of Energy Storage 77:110009.

Zhai, S., J. Nam, G.S. Gautam, K. Lim, J. Rojas, M.F. Toney, E.A. Carter, I.-H. Jung, W.C. Chueh, and A. Majumdar. 2022. “Thermodynamic Guiding Principles of High-Capacity Phase Transformation Materials for Splitting H2O and CO2 by Thermochemical Looping,” Journal of Materials Chemistry A 10:3552–3561. https://doi.org/10.1039/D1TA10391A.

Zhang, B., F. Kang, J. Tarascon, and J. Kim. 2016. “Recent Advances in Electrospun Carbon Nanofibers and Their Application in Electrochemical Energy Storage.” Progress in Materials Science 76:319–380.

Zhang, J., T. Tian, Y. Chen, Y. Niu, J. Tang, and L.-C. Qin. 2014. “Synthesis of Graphene from Dry Ice in Flames and Its Application in Supercapacitors.” Chemical Physics Letters 591:78–81.

Zhang, P., J. Su, J. Guo, and S. Hu. 2023a. “Influence of Carbon Nanotube on Properties of Concrete: A Review.” Construction and Building Materials 369:130388.

Zhang, W., X. Qiu, C. Wang, L. Zhong, F. Fu, J. Zhu, Z. Zhang, Y. Qin, D. Yang, and C.C. Xu. 2022. “Lignin Derived Carbon Materials: Current Status and Future Trends.” Carbon Research 1(1):14. https://doi.org/10.1007/s44246-022-00009-1.

Zhang, X., R. Han, Y. Liu, H. Li, W. Shi, X. Yan, X. Zhao, Y. Li, and B. Liu. 2023b. “Porous and Graphitic Structure Optimization of Biomass-Based Carbon Materials from 0D to 3D for Supercapacitors: A Review.” Chemical Engineering Journal 460:141607.

Zhao, L., Y. Chang, S. Qiu, H. Liu, J. Zhao, and J. Gao. 2023. “High Mechanical Energy Storage Capacity of Ultranarrow Carbon Nanowires Bundles by Machine Learning Driving Predictions.” Advanced Energy and Sustainability Research 4(11):2300112. https://doi.org/10.1002/aesr.202300112.

Zhao, X., Y. Ando, Y. Liu, M. Jinno, and T. Suzuki. 2003. “Carbon Nanowire Made of a Long Linear Carbon Chain Inserted Inside a Multiwalled Carbon Nanotube.” Physical Review Letters 90(18):187401.

Zhou, D., C. Zheng, Y. Niu, D. Feng, H. Ren, Y. Zhang, and H. Yu. 2024. “Hydrogen Storage Property Improvement of Ball-Milled Mg2. 3Y0. 1Ni Alloy with Graphene.” International Journal of Hydrogen Energy 50(2024):123–135.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.

Zhou, T., and T. Zhang. 2021. “Recent Progress of Nanostructured Sensing Materials from 0D to 3D: Overview of Structure–Property–Application Relationship for Gas Sensors.” Small Methods 5(9):2100515.

Zhou, X., B. Liu, Y. Chen, L. Guo, and G. Wei. 2020. “Carbon Nanofiber-Based Three-Dimensional Nanomaterials for Energy and Environmental Applications.” Materials Advances 1(7):2163–2181.

Zoghi, M., M. Pourmadadi, F. Yazdian, M.N. Nigjeh, H. Rashedi, and R. Sahraeian. 2023. “Synthesis and Characterization of Chitosan/Carbon Quantum Dots/Fe2O3 Nanocomposite Comprising Curcumin for Targeted Drug Delivery in Breast Cancer Therapy.” International Journal of Biological Macromolecules 249:125788.

Zuraiqi, K., A. Zavabeti, J. Clarke-Hannaford, B.J. Murdoch, K. Shah, M.J.S. Spencer, C.F. McConville, T. Daeneke, and K. Chiang. 2022. “Direct Conversion of CO2 to Solid Carbon by Ga-Based Liquid Metals.” Energy and Environmental Science 15(2):595–600.

Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 203
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 204
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 205
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 206
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 207
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 208
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 209
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 210
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 211
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 212
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 213
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 214
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 215
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 216
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 217
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 218
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 219
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 220
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 221
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 222
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 223
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 224
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 225
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 226
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 227
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 228
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 229
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 230
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 231
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 232
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 233
Suggested Citation: "6 Chemical CO2 Conversion to Elemental Carbon Materials." National Academies of Sciences, Engineering, and Medicine. 2024. Carbon Utilization Infrastructure, Markets, and Research and Development: A Final Report. Washington, DC: The National Academies Press. doi: 10.17226/27732.
Page 234
Next Chapter: 7 Chemical CO2 Conversion to Fuels, Chemicals, and Polymers
Subscribe to Email from the National Academies
Keep up with all of the activities, publications, and events by subscribing to free updates by email.