Brought to you by:
Paper The following article is Open access

Computational modeling of CO2 conversion by a solar-enhanced microwave plasma reactor

, and

Published 29 June 2023 © 2023 The Author(s). Published by IOP Publishing Ltd
, , Citation Rasool Elahi et al 2023 Plasma Sources Sci. Technol. 32 065018 DOI 10.1088/1361-6595/acde08

0963-0252/32/6/065018

Abstract

The use of renewable energy to convert carbon dioxide (CO2) into higher-value products can help meet the demand for fuels and chemicals while reducing CO2 emissions. Solar-Enhanced Microwave Plasma (SEMP) CO2 conversion aims to combine the scalability and sustainability of solar thermochemical methods with the high efficiency and continuous operation of plasmachemical approaches. A computational study of a built SEMP reactor operating with up to 1250 W of microwave power together with up to 525 W of incident solar power at atmospheric pressure is presented. The study is based on a fully-coupled 2D computational model comprising the description of fluid flow, heat transfer, Ar-CO2 chemical kinetics, energy conservation for electrons and heavy-species, electrostatics, and radiative transport in participating media through the discharge tube, together with the description of the microwave electromagnetic field through the waveguide and the discharge tube. Numerical simulations reveal that the plasma is concentrated near the location of incident microwave energy, which is aligned with the radiation focal point, and that CO2 decomposition is highest in that region. The incident solar radiation flux leads to more uniform distributions of heavy-species temperature with moderately greater values throughout most of the discharge tube. Modeling results show that, at 700 W of electric power, conversion efficiency increases from 6.8% to 10.0% with increasing solar power from 0 to 525 W, in good agreement with the experimental findings of 6.4% to 9.2%. The enhanced process performance is a consequence of the greater power density of the microwave plasma due to the absorption of solar radiation.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 license. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Global energy demands have risen consistently over the previous few decades, with fossil fuels meeting most of the increase (International Energy Agency 2018). As a result, energy-related emissions of carbon dioxide (CO2), the primary greenhouse gas, have continually increased, exacerbating the need to adopt sustainable energy solutions to avoid the worst impacts of climate change. The sustainable use of CO2 for the synthesis of value-added products such as organic acids, esters, alcohols, etc, can play an important role in limiting CO2 emissions (Zhang et al 2017). Particularly, the synthesis of solar fuels, i.e. hydrocarbons from CO2, water, or other feedstock using solar energy, is seen as a sustainable way to meet the demand for fuels in difficult-to-decarbonize sectors, such as long-haul aviation and maritime transportation (Detz et al 2018). Solar fuels can be produced via direct or indirect processes. Direct processes harness the energy in solar radiation to produce fuel without an intermediary energy conversion step. Among direct processes, solar thermochemical ones, which make direct use of solar radiation to drive high-temperature endothermic chemical reactions, are very promising due to their potential for scalability and lower costs (Schäppi et al 2022). However, direct processes suffer from the intermittent nature of solar radiation, which limits their economic viability (Roy 2010, Abedini Najafabadi et al 2019). In contrast, indirect processes have solar energy converted to another form of energy, such as biomass or electricity, first, and then use that energy in fuel synthesis. By de-coupling the reception and conversion of solar energy, and moreover, by allowing the use of other forms of renewable energy such as wind or geothermal, indirect processes can operate continuously. Among indirect approaches, plasma-based CO2 conversion processes have demonstrated high conversion and energy efficiencies, and great potential for industrial-scale deployment (Liu et al 1999, Kozák and Bogaerts 2014, 2015, Lindon and Scime 2014, Snoeckx and Bogaerts 2017).

Solar thermochemical processes largely rely on the absorption of radiative energy by a catalytic solid medium to increase its temperature and promote chemical reactions over its surface, given that the most common feedstock gases (CO2, H2O, CH4, etc) are transparent to solar photons (Lapp et al 2012). These processes depend on high-temperature reactors and high concentration of solar radiation to achieve efficient operation (Marxer et al 2017). Typically, solar-driven thermochemical CO2 decomposition processes operate at temperatures above 1400 K (Buelens et al 2016). Operating temperatures above 2500 K are often needed to achieve reasonable degrees of CO2 decomposition in single-step reduction or direct thermolysis processes (Smestad and Steinfeld 2012). For instance, thermodynamic and chemical equilibrium considerations indicate that at temperatures above 2700 °C, nearly 50% of CO2 is decomposed in a single-step process (Abanades and Chambon 2010). In contrast, two-step CO2 reduction processes, commonly based on the use of metal oxides operating in reduction and oxidation cycles, can operate effectively at significantly lower temperatures (Al-Shankiti et al 2017). A review of the state-of-the-art solar thermochemical approaches for CO2 decomposition is given in (Pullar et al 2019).

Plasmachemical approaches are generally based on driving the feedstock gas to a plasma state using electrical discharges. Among the different types of plasma sources (arc, glow, corona discharges, inductively coupled, etc (Fridman 2008)), microwave plasma has potentially the highest energy efficiencies for molecular dissociation thanks to effective stepwise vibrational excitation, an effective channel for dissociation; as well as relatively high heavy-species temperature, which promotes thermal dissociation (van Rooij et al 2015, Bogaerts et al 2016, Bongers et al 2017, Silva et al 2017, 2021). Such characteristics make microwave plasma sources ideal for gas-phase chemical synthesis, including the decomposition of CO2. A significant amount of research has been performed on CO2 decomposition by microwave plasma. Among these, approaches based on (near) atmospheric pressure operation are particularly appealing for their potentially greater viability, given that the absence of vacuum systems can lead to less expensive installations and to greater compatibility with other unit operations. Mitsingas and collaborators investigated CO2 conversion by atmospheric pressure microwave plasma, achieving a maximum conversion of 9% with an energy efficiency of 50% at a Specific Energy Input (SEI, amount of energy per molecule) of 0.5 eV mol−1 (Mitsingas et al 2016). Their results showed that CO2 conversion increases significantly with decreasing flow rate and that the effect of electric power on conversion was not significant. Spencer and Gallimore investigated CO2 decomposition in an atmospheric pressure microwave plasma-catalyst system for a wide range of SEI, from 2 to 28 eV mol−1 (Spencer and Gallimore 2010, 2012). Their results showed that by increasing the SEI, conversion efficiency increases, but energy efficiency decreases. Their reported highest conversion efficiency of around 45% occurred at a SEI of 28 eV mol−1, corresponding to their lowest value of energy efficiency of around 5%. Bekerom and collaborators established the importance of thermal conversion in the dissociation of CO2 in microwave discharges (Bekerom et al 2019). The authors found that thermal processes dominate at higher power densities, favoring thermal dissociation over radial energy transport—a particularly relevant finding in atmospheric pressure processes and in processes that receive additional energy inputs, such as solar radiation in the present work.

Exploiting the advantages of solar thermochemical methods (i.e. direct use of solar radiation as a free form of sustainable energy and process scalability) and plasmachemical approaches (i.e. flexibility and continuous operation afforded by their reliance on electrical energy) into integrated solar-plasma processes could lead to more viable and/or sustainable CO2 conversion (Trelles 2022). Specifically, solar-plasma processes can utilize electricity to compensate for fluctuations in solar radiation (during daytime) or even substitute it (at nighttime) and therefore can operate continuously. Simultaneously, the direct use of solar radiation limits the sustainability penalties associated with the generation, transmission, and storage of electricity. Moreover, a previous study by the authors (Elahi et al 2020 Elahi and Trelles 2021a, 2021b) showed that CO2 in nonequilibrium plasma state depicts drastically greater solar radiation absorption than CO2 in thermodynamic equilibrium. The absorbed radiation can lead to new chemical pathways and enhanced CO2 conversion.

Solar-plasma processes can be divided between Plasma-Enhanced Solar thermochemical (PES) and Solar-Enhanced Plasmachemical (SEP), depending on the dominance of either incident solar power or electric input power, respectively (Trelles 2022). In PES processes, concentrated solar power is greater than the electric power used to sustain the plasma. Therefore, the role of plasma is to enhance solar thermochemistry. An example of a PES process is the decomposition of CO2 using the solar-gliding arc reactor by Nagassou et al (2019), (2020). The reactor was designed to operate with up to 240 W of electric power and up to 525 W of concentrated solar power, at atmospheric pressure conditions and with feedstock ranging from 100% CO2 to CO2 mixtures with water, nitrogen, and methane. In SEP processes, the amount of electric power to sustain the plasma is greater than the solar input power, whose role is to augment plasma-driven chemical reactions. A distinct implementation of a SEP process is the conversion of CO2 using the Solar-Enhanced Microwave Plasma (SEMP) reactor devised by Mohsenian et al (2019a), (2019b). The SEMP reactor was designed to operate with up to 1250 W of microwave power (MP) at 2.45 GHz and up to 525 W of concentrated solar power at atmospheric pressure conditions with argon-CO2 and nitrogen-CO2 mixtures.

The present work constitutes the first computational study of solar-plasma CO2 conversion in a SEMP reactor. Specifically, this study focuses on unveiling characteristic features of solar-plasma CO2 conversion, particularly the role of solar radiation, during operation of the SEMP reactor under the conditions experimentally investigated in Mohsenian et al (2019a) and (2019b). The computational solar-plasma reactor model couples a microwave plasma flow model, based on previously-reported models (i.e. Baeva et al 2018, Bekerom et al 2019, Baeva et al 2021) to a novel model of radiative transport in participating nonequilibrium plasma media to describe the effect of solar radiation on CO2 conversion. The model is validated with experimentally-determined outflow temperature and CO2 conversion efficiency as function of input electric power and input solar power. Future developments of the model will be aimed at integrating comprehensive descriptions of photon-driven and photon-mediated chemical kinetics and to guide reactor design and process improvements.

2. SEMP reactor

SEMP chemical conversion aims to combine the advantages of solar thermochemical and microwave plasma processes. The SEMP CO2 conversion system developed by Mohsenian et al (2019a) is schematically depicted in figure 1. The reactor was designed to operate with up to 1250 W of electric power from a 2.45 GHz magnetron and up to 525 W of concentrated solar power (effective incident power from a 6.5 kW high-flux solar simulator), using as feedstock CO2 diluted in argon or nitrogen.

Figure 1.

Figure 1. Solar-enhanced microwave plasma (SEMP) reactor. The reactor design aims to exploit the interaction between concentrated solar radiation and microwave plasma to enhance CO2 conversion.

Standard image High-resolution image

In the SEMP reactor, the input electromagnetic power propagates along the waveguide, which is tapered at the point of plasma formation. A concentrator is used to direct the flux of solar radiation from a high-flux solar simulator into the SEMP reactor. Radiation enters the reactor via an optical aperture at the base of a conical chamber. The conical chamber allows the uninterrupted transmission of the incoming concentrated solar radiation to its focal point inside the reactor, which by design coincides with the location of plasma formation. The feedstock gas enters the conical chamber tangentially near the aperture. The interaction between concentrated solar radiation and microwave plasma, starting at the concentrator's focal point inside the discharge tube and proceeding downstream, leads to the absorption of photons by plasma species. The experimental results obtained by Mohsenian et al showed that up to 20% of incident solar radiation is absorbed by the microwave plasma, and that the absorbed radiation led to an increase from ∼ 6.4% to 9.2% in CO2 conversion Mohsenian et al (2019a). Further studies focused on the effect of SEI and the use of argon-nitrogen mixtures on CO2 conversion (Mohsenian et al 2019b). The enhanced conversion seemed to be due to the greater power density of the microwave plasma, which, as indicated by the work of Bekerom et al (2019), favors CO2 dissociation. The use of solar radiation appears to be effective at depositing power into the plasma, circumventing the skin effect that typically limits electric power deposition in microwave discharges.

3. Mathematical model

The analysis of solar radiation—plasma interaction requires the concurrent description of physical-chemical-radiative phenomena due to the reception of electric power to sustain the plasma and radiative power from concentrated solar radiation. The SEMP reactor model is based on a fluid flow (continuum) approximation that encompasses the description of chemical kinetics, energy conservation of electrons and heavy-species (molecules, atoms, ions), fluid flow, electrostatics, and radiative transport in participating media through the discharge tube, together with the description of the microwave electromagnetic field through the waveguide and the discharge tube. The reactor operates at atmospheric pressure with an argon–carbon dioxide mixture (Ar:CO2) supplied in a 7:1 ratio by volume as gas feedstock.

The model considers the plasma in a state of chemical nonequilibrium and non-local thermodynamic equilibrium (NLTE), which includes thermal nonequilibrium. Under thermal nonequilibrium, free electrons and heavy-species have different Maxwellian velocity distributions characterized with an electron temperature Te and a heavy-species temperature Th, respectively, with Te > Th. It is to be noted that the relatively high values of Th encountered in microwave discharges at atmospheric (or higher) pressure often makes them to be considered as quasi-thermal (Trelles 2019). This characterization contrasts with thermal plasmas (e.g. high-intensity arcs and radio-frequency discharges at relatively high pressures), which depict large regions where the LTE assumption is valid, and with atmospheric pressure nonthermal plasmas (e.g. glow and corona discharges), which depict thermal nonequilibrium throughout their extent.

3.1. Electron transport

The rate of change of electron number density ne (m−3), considering flow advection, diffusive transport, and plasma reactions, is described by:

Equation (1)

where u (m s−1) is the bulk flow velocity, $\nabla $ is the gradient operator, ${{\boldsymbol{\Gamma }}_{\text{e}}}$ (m−2 s−1) the electron flux, and Re (m−3 s−1) the net reaction rate for electrons. The electron flux ${{\boldsymbol{\Gamma }}_{\text{e}}}$ is given by:

Equation (2)

where ${\mu _{\text{e}}} = \frac{e}{{{m_{\text{e}}}{v_{\text{e}}}}}$ (m2 V−1 s−1) is the electron mobility, Es (V m−1) is the electrostatic electric field, and ${D_{\text{e}}} = \frac{{{k_{\text{B}}}{T_{\text{e}}}}}{{{m_{\text{e}}}{v_{\text{e}}}}}$ (m2 s−1) the electron diffusion coefficient, with e (C), me (kg), kB (m2 kg s−2 K−1), and ve (s−1) as the electron charge, electron mass, Boltzmann constant, and electron collision frequency, respectively.

The evolution equation for the energy density of electrons, nepsilon (V m−3), is given by:

Equation (3)

where Repsilon (V m−3 s−1) represents the rate of energy change due to elastic and inelastic collisions, ${{\mathbf{E}}_{\mathbf{s}}}{ } \cdot { }{\boldsymbol{{\Gamma }}_{\mathbf{e}}}$ heating or cooling of electrons depending on whether their drift velocity is aligned or not with the electric field, ${\dot Q_{\text{h}}}$ (J m−3 s−1) describes the heating of electrons due to the microwaves, and ${\boldsymbol{{\Gamma }}_\varepsilon }$ (V m−2 s−1) is the electron energy flux, which is given by:

Equation (4)

where ${{{\mu }}_\varepsilon } = \frac{{\text{5}}}{{\text{3}}}{{{\mu }}_{\text{e}}}{ }$(m2 V−1 s−1) is the electron energy mobility, and ${D_\varepsilon } = \frac{{\text{5}}}{{\text{3}}}{\mu _{\text{e}}}{T_{\text{e}}}$ (m2 s−1) is the electron energy diffusivity. Provided that the electron energy distribution function is Maxwellian, the mean electron energy is ${n_\varepsilon }/{n_{\text{e}}} = \frac{{\text{3}}}{{\text{2}}}{k_{\text{B}}}{T_{\text{e}}}$.

The source terms Re and Repsilon follow from collision processes involving electrons. Considering that there are M reactions that contribute to the growth or decay of electron density and P inelastic electron-neutral collisions (in general, PM), the electron source term is given by:

Equation (5)

where xr is the mole fraction of the target species for reaction r, kr is the rate coefficient for reaction r (m3 s−1), and nn is the total number density of neutral species (m−3). Complementarily, the electron energy source term is obtained by summing the collisional energy change over all reactions, namely:

Equation (6)

where ${{\Delta }}{\varepsilon _r}$ (V) is the energy change due to reaction r.

The power transfer from the electromagnetic field to the electron in equation (3) is given by:

Equation (7)

where $\sigma $ is the electrical conductivity, E is the microwave field, * denotes the complex conjugate, and Re(·) represents the real component of a complex number.

3.2. Species transport

The Ar–CO2 working fluid is described as consisting of a total of 13 species, these are 7 neutral species, 2 positive ions, 3 negative ions, and 1 electronically-excited argon species, which are listed in table 1. The set of Ar species corresponds to that used by Baeva and collaborators in (Baeva et al 2018), the set of CO2 species were compiled from the reduced CO2 chemistry model presented by Aerts and collaborators in (Aerts et al 2015), together with the set used by Bekerom and colleagues (Bekerom et al 2019). Additionally, the carbon atom is included as a species in the model because of its role in Ar–CO2 interaction reactions and CO dissociation reaction.

Table 1. Species in Ar–CO2 solar-microwave plasma model.

Neutral ${\text{Ar}}$, ${\text{C}}{{\text{O}}_2}$, ${\text{CO}}$, ${{\text{O}}_3}$, ${{\text{O}}_2}$, ${\text{O}}$, ${\text{C}}$
Positive ions ${\text{A}}{{\text{r}}^ + }$, ${\text{CO}}_2^ + $
Negative ions ${{\text{e}}^ - }$, ${\text{O}}_2^ - $, ${{\text{O}}^ - }$
Excited states ${\text{A}}{{\text{r}}^*}$

For each ion and neutral species, a transport equation resembling the drift-diffusion equation for electrons is solved for the mass fraction of each species, yj , i.e.

Equation (8)

where the subscript j indicates the jth species, ρ (kg m−3) is the total mass density, Rj (kg m−3 s−1) is the net reaction rate, and J j (kg m−2 s−1) the mass flux describing mass transport due to migration from the ambipolar field by the electrostatic field Es and diffusion from concentration gradients.

The net reaction rate for species j is given by:

Equation (9)

where Mj (kg mol−1) is the molar mass of species j, ${\upsilon _{jr}}$ the stoichiometric coefficient for species j in reaction r, and ${r_r} = {k_r}\prod\nolimits_j {c_j}^{{\upsilon _{jr}}}$ the reaction rate for reaction r, where kr is the reaction rate coefficient and cj = $\rho {y_j}/{M_j}$ is the molar concentration of species j. The mass flux for species j is defined as:

Equation (10)

where ${{\mathbf{V}}_j}$ is the multicomponent diffusion velocity for species j described by a mixture-averaged mass diffusion model. The species mass fraction yj is related to its corresponding molar fraction xj by xjnn = ρ(yj /Mj )NA with NA as Avogadro's number.

The Ar–CO2 chemical kinetics model is composed of plasmachemical reactions consistent with the set of species in table 1. A total of 62 reactions are included in the Ar–CO2 plasma chemical kinetic model, accounting for 13 elastic and ionization electron impact, 10 electron attachment and recombination, 26 neutral, 4 dissociation, and 9 ion-molecule reactions. These reactions are a compilation of the set of reactions used by Baeva and collaborators (Baeva et al 2018) for argon plasma, the set used by Aerts and collaborators (Aerts et al 2015) for reduced CO2 plasma chemistry, that by Bekerom and colleagues (Bekerom et al 2019) for CO2 thermochemistry, and the set by Beuthe and Chang (1997) for Ar–CO2 interactions. The set of chemical reactions to evaluate Rj and their corresponding rate coefficients are listed in appendix. The only electron impact ionization reactions considered in the model are the ionization of the feedstock gases Ar and CO2 shown in table A.1, together with elastic electron impact reactions of the neutral species in the model and Ar electron excitation. The rate coefficient of all elastic and ionization electron impact reactions in table A.1 are obtained using cross-section data from the Biagi-v7.1 database for argon-related species and the Morgan database for oxygen and carbon-related species, both within LxCat (Carbone et al 2021, Pitchford et al 2017 and Pancheshnyi et al 2012). The reaction rate coefficients of electron attachment and recombination reactions are listed in table A.2, electron impact dissociation reactions following molecular electronic exictation in table A.3, ion-molecule reactions are presented in table A.4, and chemical reactions involving neutral species in table A.5.

While vibrational excitation has a key role in CO2 conversion in low-pressure microwave plasmas (Silva et al 2021) and in high temperature microwave discharges (Pietanza et al 2020), species describing vibrationally excited CO2-related species are not included in the model. Inclusion of vibrational kinetics would involve species describing the vibrationally-excited levels of CO2 (all 21 levels of the asymmetric stretch mode, together with the symmetric stretch and bending modes (Zhang et al 2020)), and vibrational modes of CO (63 levels (Kozák and Bogaerts 2015)) and of O2 (three levels (Zhang et al 2020)). Meritorious examples of the modeling of CO2 vibrational kinetics are given in (Berthelot and Bogaerts 2016) and (Wang et al 2016). Nevertheless, the work of Bekerom et al (2019) indicates that the role of vibrational kinetics is secondary compared to thermal effects at atmospheric pressure. Additionally, no photochemistry is included in the model, despite the important role of photons in plasma chemical kinetics (Capitelli et al 2017). This omission is due to the high complexity and computational cost associated with the description of radiative-collisional kinetics, especially for a broad-enough range of wavelengths suitable to describe solar radiation, and the reliance on the use of radiative properties to describe the effect of radiation on the plasma (section 3.7). Therefore, the solar-plasma kinetics model is able to describe potential CO2 conversion enhancements due to both thermal and ion interaction effects. The set of species and reactions provide a reasonable description of CO2 conversion by solar-microwave plasma, particularly of thermal effects, while making 2D fully-coupled computational simulations practicable.

3.3. Total mass and momentum conservation

The fluid flow through the discharge tube is described by the Navier–Stokes equations assuming laminar flow. These equations are constituted by the equation of total mass conservation, i.e.

Equation (11)

and the equation of mass-averaged momentum conservation, i.e.

Equation (12)

where p is the pressure, ${\boldsymbol{\unicode{x03C4}}}$ is the viscous stress tensor, and I is the identity tensor. In equation (12), the Lorentz force FL $ = \frac{{\text{1}}}{{\text{2}}}{\mu _{\text{0}}}{\text{Re(}}\sigma {\mathbf{E}} \times {{\mathbf{H}}^{\text{*}}}{\text{)}}$, where ${\mu _0}$ is the free space permeability and H the magnetostatic field, has been neglected.

3.4. Heavy-species energy conservation

The conservation of the energy of heavy-species is described by:

Equation (13)

where q is the heat conduction flux and Cp the specific heat at constant pressure for the mixture, i.e.

Equation (14)

where ${C_{p,j}}$ denotes the heat capacity of species j.

The source term ${ }{\dot Q_{el}}$ represents the energy gained due to elastic collisions between electrons and heavy-species, and is modeled as:

Equation (15)

where ${v_{{\text{e}}j}}$ is the collision frequency for elastic collisions with species j.

The source term ${\dot Q_{n - n}}$ describes the heat released due to non-electron collisions, and is given by:

Equation (16)

with $\upsilon _{jr}^{^{\prime}}$, ${\upsilon _{jr}}\prime \prime$, and hj representing the stoichiometric coefficients for forward and backward reactions, and the enthalpy of the products and reactants, respectively. Finally, the source term ${\dot Q_r}$ describes the net energy transported due to radiation, which is described below.

3.5. Electrostatic field

Poisson's equation is used to describe the evolution of the ambipolar electric field Es generated through the discharge tube, i.e.

Equation (17)

where V is the electric potential and ρv = e(n+ ne —n ) is the space charge density (C m−3), where n+ is the total number of positive ions, n the total number of negative ions, and epsilon0 the permittivity of free-space.

3.6. Microwave field

In contrast to the fluid flow-related variables in the model, whose equations are formulated in the time domain, the electromagnetic field evolution model is formulated in the frequency domain. Starting from Maxwell's equations and invoking the assumption of a harmonic time variation of the electric field E (V m−1) and the magnetic field H (A m−1) in a non-magnetic medium, leads to the equations:

Equation (18)

where i2 = −1 is the imaginary unit, μ0 (kg m s−2 A−2) is the free-space permeability, ω = 2πνf is the angular field frequency with νf = 2.45 GHz as the microwave excitation frequency, and epsilonpl is the relative electrical permittivity of the medium, which is given by the Lorentz formula:

Equation (19)

where ${v_{\text{m}}}$ is electron collision frequency for momentum transfer, and ${{\omega }}_p^{\text{2}} = \frac{{{n_{\text{e}}}{e^{\text{2}}}}}{{{\varepsilon _0}{m_{\text{e}}}}}$ is the plasma frequency. By introducing the electrical conductivity as:

Equation (20)

and combining the equations in equation (17) leads to the following equation for the microwave electric field (Baeva et al 2018, 2021):

Equation (21)

in which k0 = ω/c0, where c0 is the speed of light in free space.

As described in section 4.1, a rectangular waveguide is used to connect the plasma tube to the microwave source. The rectangular port is excited by a transverse electric (TE) wave, which is a wave that has no electric field component in the direction of propagation. At the microwave excitation frequency of 2.45 GHz, TE10 is the dominant mode, with the lowest cut-off frequency, depicting a half-cycle variation of the field across the width of the waveguide and no cycle variation of the field along the waveguide. As TE10 is the only propagating mode through the rectangular waveguide, electrons do not experience any change in the high-frequency electric field during the microwave time scale. This means that the phase coherence between the electrons and electromagnetic waves is only destroyed through collisions with the background gas. The loss of phase coherence between the electrons and high-frequency fields is what results in energy gain for the electrons. Therefore, the momentum collision frequency is set as the collision frequency between electrons and neutral species (i.e. ${v_{\text{m}}}$ $ = \,\,{{\text{v}}_{\text{e}}}$).

3.7. Radiative energy transport

Radiative transport in an absorbing (i.e. non-transparent) and emitting medium is described by the radiative transport equation (RTE) (Modest 2013), which for steady-state conditions and in the absence of scattering is given by:

Equation (22)

where I is the (gray-)radiation intensity, which depends on the spatial location x and propagation direction s; α is the total absorption coefficient, and epsilon represents total radiative emission. The description of radiative transport in non-gray media requires the solution of a set of RTEs, one for each wavelength (or frequency, wavenumber, etc) of interest, each equation with a corresponding spectral absorption coefficient αλ and an emission epsilonλ . Given the high computational cost of the solution of the RTE and consistent with the species transport model described in section 3.2, a gray-medium description of radiative transport (i.e. equation (22)) is adopted as a first step towards the analysis of solar radiation—plasma interaction in the SEMP reactor.

The radiative energy term in equation (13) is given by (Modest 2013, Sun et al 2016):

Equation (23)

where G is the incident radiation, which for a gray medium is G = $\int\nolimits_{{\text{4}}\pi } I{\text{d}\Omega }$, and Ω is the solid angle at the location given by the direction vector s and spanning the unitary sphere (solid angle 4π).

3.8. Radiative properties

Given that the microwave plasma is in a state of thermal nonequilibrium, the spectral radiative properties αλ and epsilonλ are functions of electron and heavy-species temperatures, Te and Th, in addition to pressure p and the set of species number densities nj (all needed to determine the thermodynamic state of the medium). The code SPARK (Simulation Platform for Aerodynamics Radiation and Kinetics) developed by Lino da Silva (2007) has been used to determine αλ and epsilonλ as a function of Te and Th, for p = 1 atm and composition corresponding to that of an Ar–CO2 (7:1 vol.) mixture in chemical equilibrium. SPARK is an adaptive line-by-line numerical code that calculates spectral emission and absorption of plasma in NLTE (including LTE, by setting Te = Th).

The calculations of αλ and epsilonλ considered 34 atomic and molecular transitions, namely: atomic discrete (bound–bound) transitions for C, C+, O, Ar+ and O+; atomic continuum transitions: C, C+, O, Ar+, and O+ photoionization, O Bremstrahlung, and C and O photo-detachment; vaccum ultraviolet (VUV) continuum transitions: CO2, C2, CO, and O2 photoionization; VUV-visible transitions: C2 Phillips, Mulliken, Deslandres–Azamb, Fox–Herzberg, Ballik–Ramsaw, and Swan transitions; CO Angstrom, Asundi, CO4+, CO3+, CO+ Comet tail, CO Triplet; O2 Shumann–Runge, and O2 Shumann–Runge Cont; and infrared transitions: CO2 and CO rovibrational, and O2 Bremstrahlung. Information about the databases used in the models for the calculation of discrete and continuum radiation is given in (Lino da Silva et al 2013). The calculations involved in the order of O(106 ) transitions.

Representative spectral absorption coefficient αλ of Ar–CO2 (7:1 vol.) is presented in figure 2(a) (left) for LTE (Th = Te= T) for different values of equilibrium temperature T = 500, 1000, and 1500 K and in figure 2(a) (right) for NLTE (T = Th and Te = 1 eV). The results show that, despite the high complexity of the absorption spectra, under both, LTE and NLTE conditions, αλ increases with increasing temperature. Importantly, αλ is significantly greater for NLTE compared to LTE. Nevertheless, epsilonλ is also drastically larger for NLTE than for LTE. Determining the net effect of radiative properties of a medium in NLTE compared to LTE requires the solution of the non-gray RTE. The study in (Elahi et al 2020) involved such analysis for a one-dimensional domain with constant composition (i.e. no chemical reactions). For the fully-coupled 2D SEMP reactor model in the present work, the solution of the spectral RTE with complete consideration of the dependency of αλ and epsilonλ on λ, Te, Th, p, and nj is largely impractical, if not unfeasible. Therefore, the model treats the plasma as a gray medium (i.e. solution of the RTE for a gray medium, equation (22)) within the range of visible wavelengths. The gray medium approximation uses total radiative properties, namely total absorption coefficient α and total emission epsilon, which are function of Te and Th only (given the assumption of chemical equilibrium at p = 1 atm). The total absorption coefficient α is obtained by integrating the spectral absorption coefficient αλ over the visible range, i.e.

Equation (24)

Figure 2.

Figure 2. Radiative properties. (a) Spectral absorption coefficient of Ar:CO2 (7:1 vol.) for different temperatures in LTE and NLTE. (b) Total absorption coefficient of the mixture as function of Th and Te.

Standard image High-resolution image

where $\lambda_{UV} = $380 nm and $\lambda_{IR} = $700 nm. The total emission epsilon is determined by using Planck's relation epsilon = αIb , where Ib is the blackbody radiation intensity. Figure 2(b) shows the resulting total absorption coefficient α as function of Te and Th. The surface representation of α(Te,Th) in figure 2(b) show that α is largely negligible under thermal equilibrium (TeTh) and, as expected given the results in figure 2(a), that α increases with increasing Th and rapidly increases with increasing Te.

4. Computational model

4.1. Domain geometry

A sectional view of the SEMP reactor presented in (Mohsenian et al 2019a) is shown in figure 3(a), depicting the regions of incidence of microwave power (MP) and solar power (SP), and the inflow and outflow regions together with the spatial domains for the microwave propagation model and the plasma flow model. To reduce the complexity and computational cost of simulating the complete three-dimensional (3D) reactor geometry, the cylindrical discharge tube is modelled using a 2D Cartesian domain (indicated by the regions enclosed by dashed lines in figure 3(a)). Therefore, the cylindrical discharge tube's circular cross-section in the reactor is approximated as a rectangular channel. Figure 3(b) shows the 2D computational domains (i.e. discharge tube and waveguide), with the notation used to identify the different boundaries.

Figure 3.

Figure 3. Computational domain of the SEMP reactor model. (a) Sectional view of the SEMP reactor depicting its main components and the spatial domains for the microwave and plasma flow models. (b) Computational domain encompassing the waveguide and the discharge tube, together with the demarcation of the domain boundaries. (c) Actual 3D domain of the waveguide (left) and geometric approximation for the 2D model (right) with the surface of microwave power deposition highlighted. (d) Actual 3D domain (left) and geometric approximation for the 2D model (right) with the region of incident solar radiation highlighted. Section A–A shows the incident solar radiation region in the 3D domain, and section B–B its counterpart in the 2D domain, together with the geometric parameters defining those regions.

Standard image High-resolution image

The origin of the coordinates system used to define the domains (x–y) is placed at the point of intersection between the discharge tube axis and the centerline along the waveguide. The origin of the coordinates system coincides with the focal point of the incident concentrated radiation (experimentally, the latter is determined by the relative placement of the high-flux solar simulator in front of the optical aperture). The boundary AB is constituted by a parabolic segment whose focal point coincides with the focal point of incident radiation. Therefore, the incident concentrated solar radiation entering the discharge tube can be imposed as a flux normal to the boundary AB. The boundary AB also acts as the inflow boundary, assuming that the swirl component of the flow is negligible (i.e. the flow is along the x-axis only). The boundary DC is used to define the outflow. Electromagnetic energy enters the domain through the boundary LE as a 2.45 GHz wave (i.e. microwave) in the TE10 mode with amplitude corresponding to the imposed input electric power. The cut-off tube, i.e. the metal tube surrounding the reaction chamber, is not included in the computational domain to reduce the complexity and size of the discrete model. The electromagnetic modeling reported in (Mohsenian et al 2019a) indicates that this approximation would result in less power absorbed by the plasma. Therefore, this approximation can be expected to lead to a lower temperature and/or CO2 conversion than that experimentally obtained. Nevertheless, given the shorter size of the domain and the omission of the outflow constriction (see figure 3(a)), simulation results can also be expected to under-estimate recombination events, and therefore over-estimate CO2 conversion. The net effect of these modeling approximations can only be determined through computational simulations.

The actual (3D) spatial domain of the waveguide together with the geometric approximation for the 2D model are depicted in figure 3(c), with the surface of MP deposition highlighted. Similarly, the incident solar radiation deposition surfaces are highlighted on the actual reactor and the geometric approximation for the 2D model in figure 3(d), with the surface of SP deposition highlighted. The power deposited into the 2D Cartesian domain is matched to the actual power deposited into the SEMP reactor by scaling up the incident SP and the MP with the ratio of the power deposition surface areas of the approximated and actual domains. Specifically, the geometric scaling factor (SF) for the incident SP and the MP are given by:

Equation (25)

where dopt is the out-of-plane thickness, which is used to associate the 2D domain of the model to the actual 3D domain of the reactor; a is radius the discharge tube; h is the height of the cap representing the surface of incidence of the influx of concentrated solar radiation; and θ is the angle spanning the cap by the radial distance R from the focal point of incident solar radiation. These geometric factors are depicted in the inserts in figure 3(d). Given the geometry of the reactor, as presented in (Mohsenian et al 2019a), the values of these geometric parameters are: dopt = 1 m (i.e. arbitrarily large length consistent with the 2D approximation), h = 2.16 mm, a = 13 mm, and $\theta $ = 37.9°. Therefore, the values of the SFs used in the simulations in section 5 are: SFSP = 48.45 and SFMP = 24.49.

Despite the drastic approximation of the SEMP reactor geometry and scaling of power inputs, the model captures the main characteristics of the solar-plasma CO2 conversion process, such as plasma formation, flow development, distribution of temperatures, electric field, radiative intensity, and, especially, chemical species, as shown in the results in section 5. A more appropriate description of the reactor requires full consideration of its 3D geometry, which will be the focus of future work.

4.2. Boundary conditions

The boundary conditions needed for solving the set of model equations for the variables: ne in equation (1), nepsilon in equation (3), set of yj in equation (8), V in equation (17), p and u in equation (11) and equation (12), Th in equation (13), E in equation (21), and I in equation (22) are summarized in table 2 for the microwave domain and in table 3 for the plasma domain. In these tables, Pmw is the power of the incident microwave, n is the (outer) normal to the boundary, ${v_{{\text{e,th}}}} = {\left( {{\text{8}}{k_{\text{B}}}{T_{\text{e}}}/\pi {m_{\text{e}}}} \right)^{\frac{{\text{1}}}{{\text{2}}}}}$ is the thermal velocity of electrons, n·J j is the normal mass flux of species j, uin (y) = uinmax (1—(y/a)2) is the inflow velocity component along the axis x, and p0 = 1 atm is the operating pressure. The gas temperature at the inlet (AB boundary) is set to the ambient temperature of T0 = 300 K, and hc is the convective heat transfer coefficient (corresponding to external natural convection in ambient air). Iin is the input radiation intensity (concentrated radiation from the solar simulator), and therefore the incident solar power Psolar = Iin Ss, where Ss = $\theta R{d_{{\text{opt}}}}$ is the area of solar power deposition, Ig is the gray radiation intensity calculated assuming a gray-body at the temperature of the corresponding boundary and with a surface emissivity of 0.01, and Ifix is the outlet (DC boundary) radiation intensity, which is set equal to 0.

Table 2. Boundary conditions for the microwave domain.

BoundaryVariable
  E
EL ${P_{{\text{mw}}}}$
IH, HG, FE, LK, JI, FD, DC, CG, JB, BA, AK ${\mathbf{n}} \times {\mathbf{E}}\, = {\boldsymbol{0}}$

Table 3. Boundary conditions for the plasma domain.

 Variable
Boundary ${n_{\text{e}}}$ ${n_\varepsilon }$ yj V u Th I
AB $ - {\mathbf{n}} {\boldsymbol{\cdot}} {{\boldsymbol{\Gamma }}_{\text{e}}} = \frac{{\text{1}}}{{\text{2}}}{v_{{\text{e,th}}}}{n_{\text{e}}}$ $ - {\mathbf{n}} {\boldsymbol{\cdot}} {{\boldsymbol{\Gamma }}_\varepsilon } = \frac{{\text{5}}}{{\text{6}}}{v_{{\text{e,th}}}}{n_{\text{e}}}$ ${\mathbf{n}} {\boldsymbol{\cdot}} {{\mathbf{J}}_{\mathbf{j}}} = {\text{ 0}}$ a ${x_{{\text{C}}{{\text{O}}_{\text{2}}}}} = {\text{ 0}}{\text{.13}}$ ${{\text{x}}_{\text{j}}} \approx {\text{ 0}}$ b ${\mathbf{n}} \times {\varepsilon _{\text{0}}}{{\mathbf{E}}_{\text{s}}} = {\text{ 0}}$ u = (uin,0) T0 Iin
DC $ - {\mathbf{n}} {\boldsymbol{\cdot}} {{\boldsymbol{\Gamma }}_{\text{e}}} = {\text{ 0}}$ $ - {\mathbf{n}} {\boldsymbol{\cdot}} {\boldsymbol{{\Gamma }}_\varepsilon } = {\text{ 0}}$ ${\mathbf{n}} {\boldsymbol{\cdot}} {{\mathbf{J}}_j} = {\text{ 0}}$ ${\mathbf{n}} \times {\varepsilon _{\text{0}}}{{\mathbf{E}}_{\text{s}}} = {\text{ 0}}$ p0 n $ \cdot $ q = 0 Ifix
AD, BC $ - {\mathbf{n}} {\boldsymbol{\cdot}} {\boldsymbol{{\Gamma }}_{\text{e}}} = \frac{{\text{1}}}{{\text{2}}}{v_{{\text{e,th}}}}{n_{\text{e}}}$ $ - {\mathbf{n}} {\boldsymbol{\cdot}} {\boldsymbol{{\Gamma }}_\varepsilon } = \frac{{\text{5}}}{{\text{6}}}{v_{{\text{e,th}}}}{n_{\text{e}}}$ ${\mathbf{n}} {\boldsymbol{\cdot}} {{\mathbf{J}}_j} = {\text{ 0}}$ V = 0 u = 0 n $ \cdot $ q = hc(T0 -Th) Ig

a All ion species. b All neutral species apart from Ar and CO2.

The simulations in section 5 correspond to an inflow of 8 slpm of an Ar:CO2 mixture in the ratio 7:1 by volume, leading to a maximum inflow velocity uinmax = 0.2 m s−1 and to molar fractions equal to 0.87 and 0.13 for Ar and CO2, respectively. Surface reactions of neutralization and de-excitation are defined over the sidewalls of the discharge tube, as well as zero mass flux for all the other species in the model.

4.3. Solution approach

The SEMP reactor model is implemented in Comsol Multiphysics, version 5.4 (Comsol 2022). Comsol uses a stabilized finite element method for the discretization of the set of coupled partial differential equations defining the model. The spatial domain is discretized using 6148 triangular elements. This relatively small number of elements was used due to the high computational cost of the simulations (in terms of both, CPU time and required memory), particularly due to the inclusion of radiative transport in participating media. To appropriately resolve the severe property gradients near the discharge tube walls, eight layers of boundary layer elements with a stretching factor of 1.2, with the first layer element size of 0.1 mm, are used near the AD and BC boundaries. The Debye length near the AD and BC boundaries is approximately 0.5 mm, and therefore the discretization mesh near the wall is capable to resolve the Debye length scale, as needed to resolve the charge separation implied by equation (17).

The Navier–Stokes equations that describe conservation of mass and momentum, the equation for conservation of heavy-species energy, the microwave field equation, the equations for electron and species transport, and the RTE equation are solved in the frameworks of laminar flow, heat transfer in fluids, RF, plasma, and radiation in participating media modules in Comsol, respectively.

Temporal advancement is done with a fully-implicit variable-order (1st or 2nd order) Backward Differences method with automatic step-size selection (between ∼10−16 and 10−5 s, with and average time-step size of ∼10−4 s), together with a highly nonlinear Newton method. A frequency-transient solution technique is used for the temporal evolution of the electromagnetic field in equation (21). Solution of the numerical model is accomplished using a segregated approach, in which the discrete linearized problem in each step is solved using the MUltifrontal Massively Parallel sparse direct Solver.

As initial conditions, uniform distribution of variables consistent with the boundary values are used, namely, p = p0, u(x, y) = (uin(y), 0), V = 0 V, E = 0 V m−1, Th = 300 K, nepsilon = 4 eV, ne= 1.0 ×1017 m−3, xCO2 = 0.13, xAr = 0.87, xj ≈ 0 for all other neural species, and nj = 1.0 ×1013 m−3 for all other ion species are set throughout the discharge tube. For the waveguide, E = 0 V m−1 and H = 0 A-m−1 are used. To initiate the solution process, the initial computational time-step $\Delta $ t is equal to 10−16 s. Such a small value is used to facilitate numerical convergence. Moreover, the highly nonlinear Newton method is configured with an initial damping of 10−4 and minimum damping of 10−8. Changes in the solution smaller than the relative tolerance of 0.01 define convergence in each time step.

The problem is initially solved without radiative transport (equation (22)), i.e. keeping the 12 radiative intensity variables fixed until plasma ignition is achieved (maximum value of ne reaches ∼1016 m−3). Despite the omission of the radiative transport model (RTE), the fully-coupled set of model equations is numerically stiff and, therefore, still very difficult to solve. Particularly, the resolution of discharge ignition involves strong electromagnetic field-plasma interaction. During the early stages of discharge evolution, the electric field distribution rapidly changes as the electron density grows, especially when the cut-off density nec = 7.6 ×1016 m−3 at 2.45 GHz is achieved. Moreover, in the plasma-microwave interaction region, steep gradients in the spatial distribution of properties such as electron number density, electron temperature, and electric potential are observed, necessitating smaller maximum timesteps (∼10−8 s).

Once plasma ignition is achieved, the solution of the RTE is included in the solution approach. Solution of the RTE requires the discretization angular domain (i.e. the unit sphere spanned by the vector s in equation (22)), which is accomplished using the discrete ordinates method (DOM) with the S4 discretization with level symmetric even quadrature set (i.e. 12 directions). This S4 method, involving 12 inter-dependent variables to discretely describe the radiative intensity I(x,s), has been shown to be sufficient for many applications (Comsol 2022). The DOM has significant advantages over other approaches (e.g. spherical harmonics), particularly regarding solution accuracy in spatial domains with arbitrary configurations and in problems with widely varying radiative properties. However, this method is computationally expensive, and the required memory for problems with complicated geometries can rapidly exceed the available memory capacity on typical workstations. Therefore, the solver divides these 12 directional intensity variables into three groups (each includes four intensity directions). Each group is computed in a single iterative step before moving on to the next one, which results in the required memory remaining relatively low.

With the above solver and simulation set-up, each simulation was run using eight shared-memory cores (single processor) with 64 GB of memory and required on the order of 10−5 time-steps to achieve steady-state (defined as a change of less than 0.05% in the average CO2 concentration in the outlet of the reactor).

5. Results and discussion

5.1. Operation with microwave power only

The SEMP reactor is first simulated operating with 700 W of microwave power only (i.e. Pmw = 700 W and Psolar= 0 W by setting Iin = 0 in table 3). Figure 4 shows the distribution of all the model variables through the microwave and plasma domains, i.e. magnitude of electric field ||E|| along the waveguide and discharge tube, magnitude of electrostatic field ||Es|| through the discharge tube, pressure p, velocity magnitude ||u||, heavy-species temperature Th, electron temperature Te, incident radiative intensity G (due to radiative heat transfer within the reactor chamber), electron number density ne, and mass fractions of all participating species yj with j = { CO2, CO, CO2 +, Ar, Ar*, Ar+, O2, O, O, O2 , O3, and C }.

Figure 4.

Figure 4. SEMP reactor operation with microwave power only (no solar power). Distribution fields of all the variables in the model for the SEMP reactor operating with 700 W of microwave power and no input solar power (i.e. Pmw = 700 W, Iin = 0 W m−2 hence Psolar = 0 W).

Standard image High-resolution image

The distribution of ||E|| shows the propagation of microwaves, from the boundary of incident MP (boundary EL, also see insert in figure 4), along the tapered waveguide, and their absorption within the discharge tube with small-amplitude waves leaving the waveguide domain (through boundary IH). The interaction of the microwave field E with the plasma within the discharge tube leads to a distribution of the electrostatic field Es. The distribution of Es directly affects the distribution of electron number density ne and especially, electron energy density ${n_\varepsilon }$. This dependency is seen in the distribution of Te in figure 4, resembling that of ||Es|| and depicting separate high-temperature zones indicative of standing-wave phenomena. Given the coupling between the heavy-species energy and the electron energy due to elastic collisions, the distribution of heavy-species temperature Th appears correlated to the distribution of Te.

The plasma region is elongated in the downstream direction due to advection by the inflow stream. The relatively high heavy-species temperature in the plasma leads to a significant reduction in mass density, and consequently, to significant acceleration of flow (i.e. the maximum velocity within the plasma is ∼ 0.84 m s−1, whereas the maximum inflow velocity is 0.2 m s−1) and variations in pressure.

The maximum values of Th (∼2200 K), Te (∼3.25 eV), and of ne (2.86 ×1019 m−3) occur near the region of incidence of microwave energy (KF boundary in figure 3(b)). The distribution of incident radiation G shows that the plasma acts as a radiative source that heats the inner walls of the discharge chamber. The high values of Th, Te, and ne lead to the highest degree of ionization (mass fraction of Ar+, CO2 +) and of CO2 conversion. Regarding the latter, the minimum value of CO2 mass fraction yCO2 ∼ 0.07 and maximum mass fraction of CO yCO ∼ 0.05 occur within the plasma core, near the region of incident MP, representing ∼6.8% conversion of CO2.

The results in figure 4 show that the distribution of negative molecular oxygen ion (O2 ) and ozone (O3) is highest in the region immediately upstream of the plasma core. This is attributed to the high reactivity of these species, which leads to their decomposition at high temperatures.

The maximum value of Th of ∼2200 K is relatively low for a microwave discharge operating at atmospheric pressure (Bekerom et al 2019). This is attributed to the geometric approximations used in the 2D model, as discussed in section 4.1. The effect of the scaling factor for MP (SFMP) used in the model is clearly appreciated in the results in figure 5. Figure 5 shows a sub-set of the results in figure 4 together with the corresponding results when no SF is applied (i.e. SFMP = 1).

Figure 5.

Figure 5. Effect of microwave power scaling factor. Modeling results of the SEMP reactor operating with 700 W of microwave power and no input solar power (a) with microwave power scaling factor (SFMP = 24.49) and (b) without scaling factor (SFMP = 1.0). Distribution of heavy-species temperature Th, velocity magnitude ||u||, electron temperature Te, electron number density ne, incident radiation G, mass fraction of CO2 yCO2, and mass fraction of CO yCO.

Standard image High-resolution image

The results in figure 5 show, as expected, that the use of the geometric SF has a dramatic effect on the results. Particularly, if no SF is used, the maximum heavy-species temperature is just above 900 K, the maximum electron temperature ∼1.35 eV, and the maximum number density ∼3.15 ×1018 m−3 (compared to ∼2200 K, 3.25 eV, and 2.86 ×1019 m−3, respectively when the scaling factor SFMP = 24.49 is used). These relatively low values of temperatures and electron number density lead to negligible CO2 conversion, as expected. These results show that the use of a direct 2D model (without a geometric SF) to describe the waveguide-discharge tube assembly results in poor approximation of the microwave plasma system. This result also emphasizes the need to embrace a full 3D description of the reactor geometry to describe the microwave plasma reactor operation more accurately.

5.2. Effect of solar input power

To determine the effects of solar input power, the SEMP reactor is simulated operating with 700 W of microwave power and the maximum solar power of 525 W. Representative results are presented in figure 6. These results show that the incident radiation leads to significant heating of the gas, with the maximum heavy-species temperature ∼2200 K without solar power compared to ∼3040 K with solar power. Such intense heating causes minor increases in electron temperature Te and number density ne. The incident radiation G near the optical aperture (boundary AB) is ∼1.4 MW m−2 when solar input power is used (and it is near 0 when it is not), which is comparable to the maximum incident radiation emitted by the plasma core (near the boundary KF). Moreover, part of the incoming solar radiation is absorbed by the discharge tube, resulting in additional heating of the walls.

Figure 6.

Figure 6. SEMP reactor operation with microwave and solar power. Modeling results of the SEMP reactor operating with 700 W of microwave power together with (a) 0 W or (b) 525 W of incident solar power. Distribution of heavy-species temperature Th, electron temperature Te, electron number density ne , incident radiation G, mass fraction of CO2 yCO2, and mass fraction of CO yCO.

Standard image High-resolution image

As observed in figure 2(b), the total radiative absorption coefficient α increases with increasing heavy-species temperature and/or electron temperature, leading to further increases in the absorption of solar radiation. This result suggests that the incorporation of solar radiation should lead to enhanced thermochemical reactions (i.e. among heavy-species), which as demonstrated by Bekerom and collaborators, plays a major role in CO2 decomposition (Bekerom et al 2019). This is corroborated by the results in figure 4, which show significantly lower minimum values of CO2 mass fraction yCO2 and higher maximum values of CO mass fraction yCO when input solar power is used than when it is not. The net effect of the incorporation of solar power on CO2 conversion is discussed in section 5.5.

To observe the effect of increasing solar power, figure 7 shows the distributions of heavy-species temperature Th and incident radiation G in the discharge tube for increasing values of solar input power, i.e. 0 (i.e. no solar power), 350, 440, and 525 W. As discussed in relation to the results in figure 6, the modeling results show that the discharge tube absorbs a significant part of incoming solar radiation, causing significant heating throughout the discharge tube, and given the increase in absorption coefficient with Th and/or Te, the plasma absorbs solar radiation directly as well (i.e. volumetric heating). This leads to additional heating of the plasma. The heating produced by the absorption of solar radiation is clearly observed in the results in figure 7, which show that Th reaches a maximum value of ∼2200, 2800, 2900, and 3000 K for input solar power of 0, 350, 440, and 525 W, respectively. In all cases, the maximum temperature is found near location of incident microwave power (boundary KF). The maximum incident radiation G is ∼3.8, 12.9, 15.6, and 17.9 MW m−2 for input solar power of 0, 350, 440, and 525 W, respectively. The location of maximum G occurs near the location of maximum Th, although for the case of maximum solar power (525 W) the influx of radiative energy (along the boundary AB) is comparable to the amount of radiative energy emitted by the plasma.

Figure 7.

Figure 7. Effect of increasing solar power. (a) Distribution of heavy-species temperatures Th and incident radiation G for 700 W of incident microwave power and solar radiation power of 0, 350, 440, and 525 W. (b) Experimentally-measured gas temperature at the gas outlet and Th at the center of the outflow boundary as function of solar input power for 700 W of microwave power.

Standard image High-resolution image

As an initial validation of the model, figure 7(b) shows the comparison of the experimentally-measured gas temperature at the flow outlet and the average heavy-species temperature obtained with the model at the outflow boundary (DC in figure 3(b)) as a function of solar input power and 700 W of MP (i.e. conditions used in the results in figure 7(a)). The experimental results correspond to those reported by Mohsenian et al in (2019a) of the exhaust gas temperature as measured by a K-type thermocouple. The experimental results showed an increase in exhaust temperature from 450 to 504 K with increasing solar power from 0 to 525 W. These results are contrasted against the simulation results, which show an increase in heavy-species temperature from ∼1150 to 1620 K with increasing solar power. The significantly higher temperatures obtained by the model can be a consequence of two main factors. The first one is the geometric approximation of the reactor model (i.e. 2D with geometric power SF versus 3D). And the second one is that, experimentally, the temperature is measured in the outflow region, which is not included in the computational domain and is far downstream the outflow boundary (boundary DC). Therefore, the modeling results under-predict the cooling of gas products downstream.

5.3. Effect of microwave power

Simulation results of the SEMP reactor operating with different amounts of microwave power, namely 500, 700, and 900 W, for constant input solar power of 320 W are presented in figure 8.

Figure 8.

Figure 8. Effect of increasing microwave power. Distribution of heavy-species temperatures Th and incident radiation G for representative results of the SEMP reactor model for 320 W of solar radiation power and incident microwave power of 500, 700, and 900 W.

Standard image High-resolution image

The distributions of heavy-species temperature Th and incident radiation G show that absorption of microwave power causes additional heating in the plasma region leading to higher maximum temperatures. Specifically, the heavy-species temperature Th reaches a maximum value of 2631, 2760, and 2894 K for incident microwave power of 500, 700, and 900 W, respectively. Correspondingly, the incident radiation reaches a maximum of 10.3, 11.8, and 14.2 MW m−2, respectively. As expected, given the strong temperature-dependence of radiative emission with temperature, the location of maximum G occurs near the location of maximum Th (i.e. near the boundary KF). Given that the radiative absorption coefficient increases with increasing Th and/or Te, the higher temperature within the plasma leads to more intense radiative energy exchange within the discharge tube. This inter-dependence limits further heating, partially explaining why, despite an 80% increase in microwave power (from 500 to 900 W) and almost 50% increase in total (microwave plus solar) power (from 820 to 1220 W), the maximum temperature within the discharge tube increases by only ∼10% (from 2631 to 2894 K).

5.4. Effect of solar power on species distributions

The changes in species number density along the main axis of the reactor (i.e. along the x-axis), for all the species considered in the model, are depicted in figure 9, for 700 W of microwave power together with either 0 W or 525 W of solar input power. It is to be noted that the focal point for the incident solar radiation is located at x = 0 cm.

Figure 9.

Figure 9. Effect of solar radiation on species number density along the SEMP reactor. Distribution of number density of all chemical species in the model along the central axis of the reactor for 700 W of microwave power together with either 0 or 525 W of solar input power. (a) Neutral species, (b) negatively-charged species, and (c) positively-charges species.

Standard image High-resolution image

The distributions of neutral species in figure 9(a) shows that the incorporation of solar input power leads to significantly more rapid conversion of Ar and CO2 species within the plasma core (region −0.008 < x < 0.008). The enhancement of CO2 conversion by the incorporation of plasma can be observed in the values of species number densities at the outflow boundary (x = 0.12 m). The fact that the CO2 number density nCO2 increases, while the CO number density nCO decreases, both from x ∼ 0.02 m to the end of the plasma domain (x = 0.12 m) is indicative of the significant role of recombination. This result suggests that rapid quenching of products, soon after the core plasma region, can significantly enhance CO2 conversion. Moreover, this result indicates that the SEMP modeling results are expected to over-predict CO2 conversion. This aspect of the model is quantitatively addressed in section 5.5.

The results in figure 9(a) also show that the enhanced heating due to the incorporation of solar power has a minor effect on excited argon species Ar*, but leads to significant increases in O, O3, and C species at the outlet, with the increase more pronounced for ozone. As already observed in the results in figure 4, ozone (as well as O2 ) present very high reactivity at the intermediate temperatures found in the upstream region immediately adjacent to the plasma code (i.e. −0.02 < x < 0). This leads to an abrupt and high increase in the number density of O3 in that region.

Figures 9(b) and (c) show the distributions of negatively-charged species and positively-charged species, respectively, along the main axis of the reactor. The results show that the maximum degree of ionization of the plasma (primarily given by the number densities of e and Ar+ species) remains relatively unaffected by the incorporation of input solar power. Nevertheless, it is interesting to note that the incorporation of solar power leads to higher densities of O and lower densities of CO2 +. This dependency is attributed to the role of heavy-species temperature (which increases with the incorporation of solar radiation) on the reaction rate constants associated with these species (e.g. reactions A1 and A10 in table A.1 in appendix). Additionally, the density of O2 is significantly larger when solar power is used, a result that is consistent with the increased population of O3, as discussed above.

5.5. CO2 conversion

The efficacy of CO2 conversion by the SEMP reactor is assessed by the conversion efficiency ηc, i.e.

Equation (26)

where $x_{{\text{C}}{{\text{O}}_{\text{2}}}}^{{\text{in}}}$ and $\overline {x_{{\text{C}}{{\text{O}}_{\text{2}}}}^{{\text{out}}}} $ are the input and average output molar fractions of CO2, respectively.

Figure 10 shows comparisons between the experimentally-obtained CO2 conversion efficiency and that obtained by the SEMP reactor model. Figure 10(a) shows ηc as a function of solar input power, i.e. from zero power to the maximum 525 W delivered by the high-flux solar simulator, while the microwave power is fixed to 700 W, and figure 10(b) shows ηc as a function of microwave power, from 500 to 900 W, for 320 W of solar power. In the experiments, $\overline {x_{{\text{C}}{{\text{O}}_{\text{2}}}}^{{\text{out}}}} $ is measured right after the constriction following the exhaust port of the discharge tube (see figure 3(a)) by gas chromatography. In the simulation results, $\overline {x_{{\text{C}}{{\text{O}}_{\text{2}}}}^{{\text{out}}}} $ corresponds to the average value xCO2 over the outflow boundary (DC in figure 3(b)).

Figure 10.

Figure 10. CO2 conversion as function of solar power and microwave power. CO2 conversion efficiency in the SEMP reactor (a) as function of input solar power and 700 W of microwave power and (b) as function of incident microwave power and 320 W of solar power.

Standard image High-resolution image

The experimental results from (Mohsenian et al 2019a) in figure 10(a) show an increase in ηc from 6.2% to 9.0% by adding 525 W of solar power to 700 W of microwave power. In contrast, the simulation results show an increase in conversion efficiency from 6.8% to 10.0% with 525 W of solar input power. These results can be contrasted against the results of the variation in outlet gas temperature in figure 7(b), which show the increase in temperature with input solar power and that the experimental temperatures are significantly lower than those obtained with the model. Together, these results indicate that the model over-predicts the effect solar radiation on CO2 conversion. This may be due to over-prediction of the absorption coefficient leading to relatively greater temperatures within the plasma but lower temperatures near the exhaust port, as well as by the geometric approximation implied in the 2D reactor model.

Results of the effect of microwave power on CO2 conversion efficiency in figure 10(b) show that the agreement between the experimental results in (Mohsenian et al 2019b) and the computational modeling results improves with increasing microwave power. Specifically, the experimental results show ηc increasing from ∼6% to ∼9% and the computational results ηc increasing from ∼9% to ∼10% with increasing microwave power from 500 to 900 W. The discrepancy between results is mainly attributed to the under-prediction of recombination reactions by the model. Specifically, given the limited extent of the plasma domain in the model, increasing microwave power leads to increased temperatures throughout the domain, including the near-outlet region, and consequently to lower rates of recombination reactions. Moreover, given the omission of photon-driven chemical kinetics in the model together with the use of a gray absorption coefficient, the model is only capable to describe thermal effects driven by the influx of solar radiation. The incorporation of photon-driven reactions, particularly given the high intensity of radiative fluxes, can have a significant effect in the prediction of CO2 decomposition by the model.

6. Conclusion

Solar-Enhanced Microwave Plasma (SEMP) conversion aims to the scalability and sustainability of solar thermochemical methods with the high efficiency and continuous operation of plasmachemical approaches. This paper presents the first computational study of a built and experimentally characterized SEMP reactor for the conversion CO2. The study is based on a fully-coupled 2D model of SEMP reactor developed and characterized by Mohsenian et al (2019a), (2019b). The model encompasses the description of fluid flow, heat transfer, energy conservation for electrons, energy conservation for heavy-species, electrostatics, and radiative transport in participating media through the discharge tube, together with the description of the microwave electromagnetic field through the waveguide and the discharge tube. The model is based on a 2D description of the actual reactor, together with the incorporation of geometric scaling factor to scale-up the incident microwave power and incident solar power. The model is used to simulate the operation of the SEMP reactor at atmospheric pressure conditions, with 8 slpm of an Ar–CO2 mixture (1:7 by volume), powered with 500–900 W of microwave power and from 0 to 525 W of solar power from a high-flux solar simulator used in the experiments.

Modeling results show that absorption of solar power leads to greater temperatures throughout the discharge tube. The maximum incident radiation inside the reactor is comparable with maximum radiation emitted by the plasma (at the plasma core, near the location of incident microwave power) when the reactor operates at the maximum solar power of 525 and 700 W of microwave power. The results indicate that, with 325 W of incident solar power, by increasing microwave power from 500 to 900 W, the conversion efficiency increases from ∼9% to ∼10%. Importantly, conversion efficiency increases from 6.8% to 10.0% with increasing solar power from 0 to 525 W, in good agreement with the experimental findings of 6.4% to 9.2%. The incorporation of solar radiation appears to be an effective means to increase the power density of the plasma circumventing the skin effect that typically limits the operation of microwave discharges at high power levels.

Given the absence of photon-driven kinetics, the model describes solar enhancement due to thermal effects only (i.e. heating due to interaction with solar radiation). Consistent with this simplification, the observed enhancement appears to be a consequence of the greater power density of the microwave plasma (i.e. greater heavy-species temperature) due to the absorption of solar radiation. Future efforts will be aimed at unveiling specific mechanisms of CO2 conversion enhancement, particularly the incorporation of photon-driven kinetics (collisional-radiative effects), and on adopting a more accurate geometric description of the SEMP reactor (i.e. 3D model geometry encompassing the rectangular tapered waveguide and the cylindrical discharge tube).

Acknowledgments

The authors acknowledge the financial support provided by the U.S. National Science Foundation through the Award CBET-1552037. The authors would also like to thank Dr Mario Lino da Silva at the Technical University of Lisbon for sharing the SPARK code.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Appendix: Chemical kinetics model

The Ar–CO2 chemical kinetics model used to describe the operation of the SEMP reactor is composed of plasmachemical reactions among the species listed in table 1. These reactions are grouped as: (X) elastic and ionization electron-impact reactions, (A) electron attachment and recombination reactions, (D) Electron-impact dissociation reactions, (I) ion-molecule, ion-ion reactions, and excited states reactions, and (N) reactions among neutral species. These sets of reactions are listed in the tables below.

Table A.1. Elastic and ionization electron-impact reactions.

IndexReactionRate coefficient
X1 ${{\text{e}}^ - }{ } + {\text{ Ar }} \to {\text{ Ar }} + {\text{ }}{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ a
X2 ${{\text{e}}^ - }{ } + {\text{ Ar }} \to {\text{ A}}{{\text{r}}^*}{ } + {\text{ }}{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ a
X3 ${{\text{e}}^ - }{ } + {\text{ Ar }} \to {\text{ A}}{{\text{r}}^ + } + {\text{ }}{{\text{e}}^ - }{\text{ }} + {\text{ }}{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ a
X4 ${{\text{e}}^ - }{ } + {\text{ A}}{{\text{r}}^*}\, \to {\text{ A}}{{\text{r}}^ + } + {\text{ }}{{\text{e}}^ - }\,\, + {\text{ }}{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ a
X5 ${{\text{e}}^ - }{ } + {\text{ A}}{{\text{r}}^*}\, \to {\text{ Ar }} + {\text{ }}{{\text{e}}^ - }\,$ ${f_M}\left( {CS} \right)$ a
X6 ${{\text{e}}^ - }{ } + {\text{ C}}{{\text{O}}_2} \to {\text{ CO}}_2^ + + {\text{ }}{{\text{e}}^ - }\,\, + {\text{ }}{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
X7 ${{\text{e}}^ - }{ } + {\text{ C}}{{\text{O}}_2} \to {\text{ C}}{{\text{O}}_2}{ } + {\text{ }}{{\text{e}}^ - }\,\,$ ${f_M}\left( {CS} \right)$ b
X8 ${{\text{e}}^ - }{ } + {\text{ CO }} \to {\text{ CO }} + {\text{ }}{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
X9 ${{\text{e}}^ - }{ } + {\text{ }}{{\text{O}}_3}{\text{ }} \to {\text{ }}{{\text{O}}_3}{\text{ }} + { }{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
X10 ${{\text{e}}^ - }{ } + {\text{ }}{{\text{O}}_2}{\text{ }} \to {\text{ }}{{\text{O}}_2}\,\,\, + \,{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
X11 ${{\text{e}}^ - }{ } + {\text{ O }} \to {\text{ O}}\,\,\,\,\, + \,{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
X12 ${{\text{e}}^ - }{ } + {\text{ }}{{\text{O}}^ - }{\text{ }} \to {\text{ }}{{\text{O}}^ - }\,\,\,\, + \,{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
X13 ${{\text{e}}^ - }{ } + {\text{ C }} \to {\text{ C}}\,\,\,\,\,\, + \,{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b

Note: ${f_M}\left( {CS} \right)$ indicates that the rate coefficients are calculated from cross sections assuming a Maxwellian electron energy distribution function. a Biagi-v7.1 database, www.lxcat.net, retrieved on 17 November 2022 (Pancheshnyi 2012). b Morgan database, www.lxcat.net, retrieved on 17 November 2022 (Pancheshnyi 2012).

Table A.2. Electron attachment and recombination reactions.

IndexReactionRate coefficient
A1 ${{\text{e}}^ - }{ } + {\text{ CO}}_2^ + { } \to {\text{ CO }} + {\text{ O}}$ $2.0 \times {10^{ - 11}}T_{\text{e}}^{ - 0.5}T_{\text{h}}^{ - 1}$ a
A2 ${{\text{e}}^ - }{ } + { }{{\text{e}}^ - }{ } + {\text{ A}}{{\text{r}}^ + }{ } \to {\text{ Ar }} + { }{{\text{e}}^ - }$ ${\text{8}}{{.75 \times 1}}{{\text{0}}^{{ - \text{39}}}}T_{\text{e}}^{ - 4.{\text{5}}}$ b
A3 ${{\text{e}}^ - }{ } + {\text{ C}}{{\text{O}}_2}\, \to {\text{ CO }} + { }{{\text{O}}^ - }$ ${f_M}\left( {CS} \right)$ c
A4 ${{\text{e}}^ - }{ } + { }{{\text{O}}_3} \to {\text{ O}}_2^ - $ ${f_M}\left( {CS} \right)$ c
A5 ${{\text{e}}^ - }{ } + { }{{\text{O}}_3} \to { }{{\text{O}}^ - }$ ${f_M}\left( {CS} \right)$ c
A6 ${{\text{e}}^ - }{ } + { }{{\text{O}}_2} \to {\text{ O }} + \,{{\text{O}}^ - }$ ${f_M}\left( {CS} \right)$ c
A7 ${{\text{e}}^ - }{ } + { }{{\text{O}}^ - }{ } \to {\text{ O}}\, + \,{{\text{e}}^ - }\, + \,{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ c
A8 ${{\text{e}}^ - }{ } + {\text{ CO }}\, \to {\text{ C}}\, + {{\text{O}}^ - }$ ${f_M}\left( {CS} \right)$ c
A9 ${{\text{e}}^ - }{ } + { }{{\text{O}}_2}{ } + { }{{\text{O}}_2}\, \to {\text{ O}}_2^ - \, + { }{{\text{O}}_2}$ $2.2 \times {10^{ - 41}}{\left( {300/{T_h}} \right)^{1.5}} \times {\text{exp}}\left( { - 600/{T_h}} \right)$ d
A10 ${{\text{e}}^ - }{ } + {\text{ CO}}_2^ + \, \to {\text{ C }} + { }{{\text{O}}_2}$ $3.94 \times {10^{ - 13}}{T_e}^{ - 0.4}$ d

Note: Rate coefficients are in (${{\text{m}}^3} \cdot {\text{s}}$) for the two-body reactions and in (${{\text{m}}^6} \cdot {\text{s}}$) for the three-body reactions. ${f_M}\left( {CS} \right)$ indicates that the rate coefficients are calculated from cross sections assuming a Maxwellian electron energy distribution function. In the rate constant expressions, ${T_{\text{h}}}$ is the heavy-species temperature in K and ${T_{\text{e}}}$ is the electron temperature in eV. a Reaction rate coefficient expression taken from (Hokazono and Fujimoto 1987). b Reaction rate coefficient expression taken from (Baeva et al 2018). c Morgan database, www.lxcat.net, retrieved on 17 November 2022 (Pancheshnyi 2012). d Reaction rate coefficient expression taken from (Beuthe and Chang 1997).

Table A.3. Electron-impact dissociation reactions.

IndexReactionRate coefficient
D1 ${{\text{e}}^ - }{ } + {\text{ C}}{{\text{O}}_2} \to {\text{ CO}} + {\text{ O }} + { }{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ a
D2 ${{\text{e}}^ - }{ } + {\text{ CO }} \to {\text{ C}} + {\text{ O }} + { }{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
D3 ${{\text{e}}^ - }{ } + { }{{\text{O}}_2}{\text{ }} \to {\text{ O }} + \,{\text{O }} + \,{ }{{\text{e}}^ - }$ ${f_M}\left( {CS} \right)$ b
D4 ${{\text{e}}^ - }{ } + { }{{\text{O}}_3}{\text{ }} \to {\text{O }} + { }{{\text{O}}_2} + {\text{ }}{{\text{e}}^ - }$ $9.0 \times {10^{ - 16}}$ c

Notes: Reaction rate expressions in (${{\text{m}}^3} \cdot {\text{s}}$). ${f_M}\left( {CS} \right)$ indicates that the rate coefficients are calculated from cross sections assuming a Maxwellian electron energy distribution function. a Polak and Slovetsky cross section (Polak and Slovetsky 1976) recommended by (Morillo-Candas et al 2020). b Morgan database, www.lxcat.net, retrieved on 17 November 2022 (Pancheshnyi 2012). c Reaction rate coefficient expression taken from (Beuthe and Chang 1997).

Table A.4. Ion-molecule, ion–ion reactions, and excited-state species reactions.

IndexReactionRate coefficient (${{\text{m}}^3} \cdot {\text{s}})$
I1 ${\text{A}}{{\text{r}}^{\text{*}}}{ } + {\text{ A}}{{\text{r}}^{\text{*}}}{\text{ }} \to {\text{ A}}{{\text{r}}^ + } + {\text{ Ar }} + { }{{\text{e}}^ - }$ $1.625 \times {10^{ - 16}}T_{\text{h}}^{0.5}$ a
I2 ${\text{A}}{{\text{r}}^*}{ } + {\text{ Ar }} \to {\text{ Ar }} + {\text{ Ar}}$ $3.0 \times {10^{ - 21}}$ a
I3 ${{\text{O}}^ - }{\text{ }} + {\text{ CO }} \to {\text{ C}}{{\text{O}}_2} + {\text{ }}{{\text{e}}^ - }$ $5.5 \times {10^{ - 16}}$ b
I4 ${{\text{O}}^ - }{\text{ }} + {\text{ }}{{\text{O}}_2}{\text{ }} \to { }{{\text{O}}_3}{\text{ }} + {\text{ }}{{\text{e}}^ - }$ $1.0 \times {10^{ - 18}}$ b
I5 ${\text{A}}{{\text{r}}^ + }{ } + {\text{ C}}{{\text{O}}_2}{\text{ }} \to {\text{ CO}}_2^ + + \,{\text{Ar}}$ $7.6 \times {10^{ - 16}}$ b
I6 ${\text{O}}_2^ - {\text{ }} + {\text{ Ar }} \to \,{{\text{O}}_2} + {\text{Ar}} + \,{{\text{e}}^ - }$ $2.7 \times {10^{ - 16}}{\left( {{T_{\text{h}}}/300} \right)^{0.5}}\exp \left( { - 5590/{T_{\text{h}}}} \right)$ b
I7 ${\text{O}}_2^ - {\text{ }} + {\text{ C}}{{\text{O}}_2}{\text{ }} \to \,{{\text{O}}_2} + {\text{C}}{{\text{O}}_2} + \,{{\text{e}}^ - }$ $2.7 \times {10^{ - 16}}{\left( {{T_{\text{h}}}/300} \right)^{0.5}}\exp \left( { - 5590/{T_{\text{h}}}} \right)$ b
I8 ${{\text{O}}^ - }{\text{ }} + {\text{ }}{{\text{O}}_3}{\text{ }} \to { }{{\text{O}}_2}{\text{ }} + {\text{ }}{{\text{O}}_2}{ } + { }{{\text{e}}^ - }$ $3.0 \times {10^{ - 16}}$ c
I9 ${\text{O}}_2^ - {\text{ }} + {\text{ CO}}_2^ + {\text{ }} \to {\text{ CO }} + \,{{\text{O}}_2}\, + \,{\text{O}}$ $6.0 \times {10^{ - 13}}$ d

a Reaction rate coefficient expression taken from (Baeva et al 2018). b Reaction rate coefficient expression taken from (Beuthe and Chang 1997). c Reaction rate coefficient expression taken from (Ionin et al 2007). d Reaction rate coefficient expression taken from (Hokazono et al 1998).

Table A.5. Chemical reactions involving neutral species.

IndexReactionRate coefficient
k0,i Ea,i (eV)
N1 ${\text{C}}{{\text{O}}_2}{ } + {\text{ C}}{{\text{O}}_2}\, \to \,{\text{CO }} + {\text{O }} + {\text{C}}{{\text{O}}_2}$ $4.38 \times {10^{ - 13}}$ 5.58
N2 ${\text{C}}{{\text{O}}_2}{ } + {\text{ CO }} \to \,{\text{CO }} + {\text{ O }} + {\text{CO}}$ ${\text{4}}{\text{.38}} \times {\text{1}}{{\text{0}}^{ - {\text{13}}}}$ 5.58
N3 ${\text{C}}{{\text{O}}_2}{ } + { }{{\text{O}}_2}{ } \to {\text{ CO }} + {\text{O}} + {{\text{O}}_2}$ ${\text{3}}{\text{.72}} \times {\text{1}}{{\text{0}}^{ - {\text{16}}}}$ 5.19
N4 ${\text{C}}{{\text{O}}_2}{ } + {\text{ O }} \to {\text{CO }} + {{\text{O}}_2}$ ${\text{7}}{\text{.77}} \times {\text{1}}{{\text{0}}^{ - {\text{18}}}}$ 1.57
N5 ${{\text{O}}_2} + {{\text{O}}_2}\,{ } \to {\text{ O }} + {\text{O}} + {\text{O}}2$ ${\text{8}}{\text{.14}} \times {\text{1}}{{\text{0}}^{ - {\text{15}}}}$ 5.14
N6 ${{\text{O}}_2} + {\text{O }} \to {\text{ O }} + {\text{ O }} + {\text{ O}}$ ${\text{1}}{\text{.99}} \times {\text{1}}{{\text{0}}^{ - {\text{14}}}}$ 4.98
N7 ${{\text{O}}_2} + {\text{CO }} \to {\text{ O }} + {\text{O }} + {\text{ CO}}$ ${\text{2}}{\text{.41}} \times {\text{1}}{{\text{0}}^{ - {\text{15}}}}$ 5.12
N8 ${{\text{O}}_2} + {\text{C}}{{\text{O}}_2}\, \to {\text{ O}} + {\text{ O }} + {\text{ C}}{{\text{O}}_2}$ ${\text{2}}{\text{.57}} \times {\text{1}}{{\text{0}}^{ - {\text{15}}}}$ 5.14
N9 ${\text{CO}} + {\text{O }} + {\text{C}}{{\text{O}}_2} \to \,{\text{C}}{{\text{O}}_2}{ } + {\text{ C}}{{\text{O}}_2}$ $6.54 \times {10^{ - {\text{45}}}}$ 0.19
N10 ${\text{CO }} + {\text{ O }} + {\text{CO }} \to \,{\text{C}}{{\text{O}}_2}{ } + {\text{ CO}}$ ${\text{6}}{\text{.54}} \times {\text{1}}{{\text{0}}^{ - {\text{45}}}}$ 0.19
N11 ${\text{CO }} + {\text{ O }} + {{\text{O}}_2}{ } \to {\text{ C}}{{\text{O}}_2}{ } + { }{{\text{O}}_2}$ ${\text{6}}{\text{.51}} \times {\text{1}}{{\text{0}}^{ - {\text{48}}}}$ −0.16
N12 ${\text{CO }} + { }{{\text{O}}_2}{ } \to {\text{ C}}{{\text{O}}_2}{ } + {\text{ O}}$ ${\text{1}}{\text{.23}} \times {\text{1}}{{\text{0}}^{ - {\text{18}}}}$ 1.32
N13 ${\text{O }} + {\text{ O }} + {{\text{O}}_2}{ } \to { }{{\text{O}}_2}{ } + { }{{\text{O}}_2}$ ${\text{6}}{\text{.81}} \times {\text{1}}{{\text{0}}^{ - {\text{46}}}}$ 0.00
N14 ${\text{O }} + {\text{ O }} + {\text{O }} \to \,{{\text{O}}_2}{ } + {\text{ O}}$ ${\text{2}}{\text{.19}} \times {\text{1}}{{\text{0}}^{ - {\text{45}}}}$ −0.20
N15 ${\text{O }} + {\text{ O }} + {\text{CO }} \to { }{{\text{O}}_2}{ } + {\text{ CO}}$ ${\text{2}}{\text{.76}} \times {\text{1}}{{\text{0}}^{ - {\text{46}}}}$ 0.00
N16 ${\text{O }} + {\text{ O }} + {\text{C}}{{\text{O}}_2}{ } \to { }{{\text{O}}_2}{ } + {\text{ C}}{{\text{O}}_2}$ ${\text{2}}{\text{.76}} \times {\text{1}}{{\text{0}}^{ - {\text{46}}}}$ 0.00
N17 ${\text{C}}{{\text{O}}_2}{ } + {\text{ Ar }} \to {\text{ CO }} + {\text{ O }} + {\text{ Ar}}$ $4.39 \times {10^{ - 13}}\exp \left( { - 65000/{T_{\text{h}}}} \right)$ a
N18 ${\text{CO }} + {\text{ O}} + {\text{ Ar }} \to {\text{ C}}{{\text{O}}_2}{ } + {\text{ Ar}}$ $8.2 \times {10^{ - 46}}\exp \left( { - 1510/{T_{\text{h}}}} \right)$ b
N19 ${\text{O }} + {\text{ O }} + {\text{ Ar }} \to { }{{\text{O}}_2}{ } + {\text{ Ar}}$ $1.27 \times {10^{ - 44}}{\left( {{T_{\text{h}}}/300} \right)^{ - 1}}\exp \left( { - 170/{T_{\text{h}}}} \right)$ c
N20 ${\text{O }} + {{\text{O}}_2} + {\text{Ar }} \to {{\text{O}}_3}\, + \,{\text{Ar}}$ $3.6 \times {10^{ - 46}}{\left( {{T_{\text{h}}}/300} \right)^{ - 1.93}}$ d
N21 ${{\text{O}}_3} + {\text{ Ar }} \to {{\text{O}}_2}\, + \,{\text{O }} + \,{\text{Ar}}$ $4.1175 \times {10^{ - 16}}\exp \left( { - 11430/{T_{\text{h}}}} \right)$ d
N22 ${\text{CO }} + {\text{ Ar }} \to {\text{C }} + \,{\text{O }} + \,{\text{Ar}}$ $1.46 \times {10^0}T_{\text{h}}^{ - 3.52}\exp \left( { - 128700/{T_{\text{h}}}} \right)$ d
N23 ${\text{O }} + {\text{ C }} + {\text{ Ar }} \to \,{\text{CO }} + \,{\text{Ar}}$ $2.14 \times {10^{ - 41}}{\left( {{T_{\text{h}}}/300} \right)^{ - 3.08}}\exp \left( { - 2114/{T_{\text{h}}}} \right)$ d
N24 ${\text{C}}{{\text{O}}_2}{ } + {\text{ C }} \to {\text{CO }} + \,{\text{CO}}$ $1.0 \times {10^{ - 21}}$ d
N25 ${\text{CO }} + { }{{\text{O}}_3}{ } \to {\text{C}}{{\text{O}}_2}{ } + \,{{\text{O}}_2}$ $4.06 \times {10^{ - 31}}$ d
N26 ${\text{O }} + {{\text{O}}_3}{ } \to {{\text{O}}_2}\, + \,{{\text{O}}_2}$ $8.0 \times {10^{ - 18}}\exp \left( { - 17.13/{T_{\text{h}}}} \right)$ e

Notes: N1–N17 reaction rates are of the form ${k_i}\left( {{T_{\text{h}}}} \right) = {k_{0,i}}\exp \left( { - {E_{a,i}}/{k_{\text{B}}}{T_{\text{h}}}} \right)$ where ${k_{0,i}}$ is the reaction rate coefficients and ${E_{a,i}}$ the activation energy. ${k_{\text{B}}}$ is the Boltzmann constant, and ${T_{\text{h}}}$ is the heavy-species temperature as expressed by Bekerom et al (2019). All other rate coefficients are in (${{\text{m}}^3} \cdot {\text{s}}$) for the two-body reactions and in (${{\text{m}}^6} \cdot {\text{s}}$) for the three-body reactions. a Reaction rate coefficient expression taken from (Fridman 2008). b Reaction rate coefficient expression taken from (Berthelot and Bogaerts 2016). c Reaction rate coefficient expression taken from (Hadj-Ziane et al 1992). d Reaction rate coefficient expression taken from (Beuthe and Chang 1997). e Reaction rate coefficient expression taken from (Aerts et al 2015).

Please wait… references are loading.
10.1088/1361-6595/acde08