ChemPlasKin: a general-purpose program for unified gas and plasma kinetics simulations

Xiao Shaoa,∗, Deanna A. Lacostea, Hong G. Ima
aCCRC, King Abdullah University of Science and Technolody (KAUST), Thuwal 23955-6900, Saudi Arabia
E-mail: [email protected]
Abstract

This work introduces ChemPlasKin, a freely accessible solver optimized for zero-dimensional (0D) simulations of chemical kinetics of neutral gas in non-equilibrium plasma environments. By integrating the electron Boltzmann equation solver, CppBOLOS, with the open-source combustion library, Cantera, at the source code level, ChemPlasKin computes time-resolved evolution of species concentration and gas temperature in a unified gas-plasma kinetics framework. The model allows high fidelity predictions of both chemical thermal effects and plasma-induced heating, including fast gas heating and slower vibrational-translational relaxation processes. Additionally, a new heat loss model is developed for nanosecond pulsed discharges, specifically within pin-pin electrode configurations. With its versatility, ChemPlasKin is well-suited for a wide range of applications, from plasma-assisted combustion (PAC) to fuel reforming. In this paper, the reliability, accuracy and efficiency of ChemPlasKin are validated through a number of test problems, demonstrating its utility in advancing gas-plasma kinetic studies.

Keywords: plasma-assisted combustion (PAC); ion chemistry; fuel reforming; electron-impact reactions; reaction kinetics

Source code: https://github.com/ShaoX96/ChemPlasKin

1 Introduction

Non-equilibrium plasma has gained an increasing interest within both the combustion and plasma research communities, owing to its potential to enhance combustion characteristics and fuel reforming [1, 2, 3]. The integration of plasma actuation into reacting flows, via kinetic, thermal, and hydrodynamic effects, presents a complex multi-physics challenge that requires a synergistic approach combining both experimental and computational methodologies to gain a comprehensive understanding. Due to the formidable computational complexity and expense associated with higher-dimensional models, detailed chemical kinetic analyses of systems that couple neutral gas with plasma have predominantly been confined to zero or one-dimensional simulations. Such 0D simulations, enriched with detailed gas-plasma kinetics, are not only foundational for the development of kinetic mechanisms but are also pivotal in identifying important pathways of plasma energy transfer. This insight is crucial for creating reduced-order plasma models that can be integrated efficiently into computational fluid dynamics (CFD) solvers [4, 5].

A combined kinetic solver requires the determination of species reaction rates for the gas-phase and plasma. For the latter, the electron energy distribution function (EEDF) is crucial for determining reaction rate coefficients of inelastic electron-impact collisions and calculating electron temperature (Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT). In non-equilibrium plasma, EEDF deviates from Maxwellian, and needs to be determined by solving the electron Boltzmann equation (EBE). Table 1 summarizes various modeling approaches for computing chemical kinetics in gas-phase coupled with non-equilibrium plasma available in the literature. Broadly, solver development methodologies fall into three categories: tabulation, operator-splitting, and full coupling. Solvers 1-4 employ either the Bolsig+ EBE solver [6] or the ZDPlasKin plasma kinetics solver [7] to pre-calculate electron-impact reaction rates to build look-up tables or polynomial fits as functions of reduced electric field (E/N𝐸𝑁E/Nitalic_E / italic_N) or Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, for use with CHEMKIN [8], the standard in gas-phase chemical kinetics software. While straightforward, tabulation may be subjected to large errors with significant gas composition changes. Solvers 5-13, following Lefkowitz et al. [9], integrate ZDPlasKin with chemical kinetics solvers like CHEMKIN or Cantera [10], alternating the integration of plasma and neutral gas kinetics with stepwise exchange of species concentration and temperature data. Solver 14 by Cheng et al. [11] is the only approach known to date that directly couples the BOLOS EBE solver [12] with Cantera, enabling unified gas-plasma kinetics simulations, although the details of its code implementation were not elaborated.

While CHEMKIN and Cantera are well-established open-source modules for combustion community, and ZDPlasKin is for the plasma community, the integration of the two schools for a unified version of gas-plasma kinetics solver is not straightforward, as evidenced by the fact that most of the tools shown in Table 1 are in-house codes and lack open availability. To address this deficiency, we introduce ChemPlasKin, an open-source code optimized for simulating neutral gas-phase chemical reactions in conjunction with non-thermal plasma chemistry within a unified code module framework. Similar to the CERFACS code [11], ChemPlasKin is developed with a particular interest in for nanosecond repetitively pulsed (NRP) plasma applications, which exhibits a large spectrum of time scales, offering new capabilities for plasma-assisted ignition (PAI) and fuel reforming applications.

The main content of this paper is divided into two sections. Section 2 outlines the methodologies, including the governing equations, the heat loss model, and code development. Section 3 presents extensive code validations that demonstrate the consistency and efficiency of ChemPlasKin in comparison with results from the literature. It also highlights the solver’s capability in predicting both fast and slow gas heating processes, and assesses the impact of the newly introduced heat loss model.

Table 1: Literature review of publications involving in-house 0D gas-plasma kinetics solvers
##\## Origin Method Topic Discharge Year
1 Ohio State Bolsig+ precalculation + ChemKin-Pro Air/H2subscriptH2\mathrm{H_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and air/hydrocarbon kinetics [13] NRP 2015
2 UC Berkeley ZDPlasKin precalculation + CHEMKIN CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT/air ignition [14] NRP 2016
3 UT Austin Bolsig+ precalculation + CHEMKIN CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT/air and C2H4subscriptC2subscriptH4\mathrm{C_{2}H_{4}}roman_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT/air ignition [15] NRP 2021
4 TU/e Bolsig+ precalculation + Sundials IDA CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT and H2subscriptH2\mathrm{H_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT oxidation in Ar [16] DBD 2023
5 Princeton ZDPlasKin + CHEMKIN CH4/O2/HesubscriptCH4subscriptO2He\mathrm{CH_{4}/O_{2}/He}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_He oxidization [9] NRP 2015
CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT/air ignition [17] NRP 2018
H2/O2/HesubscriptH2subscriptO2He\mathrm{H_{2}/O_{2}/He}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_He ignition [18] NRP/DC 2019
CH4/O2/HesubscriptCH4subscriptO2He\mathrm{CH_{4}/O_{2}/He}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_He ignition [19] NRP/DC 2019
H2subscriptH2\mathrm{H_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT/air ignition [20] NRP 2020
N-dodecane/O2/N2subscriptO2subscriptN2\mathrm{O_{2}/N_{2}}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT kinetics [21] NRP 2021
NH3/O2/HesubscriptNH3subscriptO2He\mathrm{NH_{3}/O_{2}/He}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_He ignition [22] NRP 2021
NH3subscriptNH3\mathrm{NH_{3}}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT/air ignition [23] NRP 2022
N2O/NOxsubscriptN2OsubscriptNOx\mathrm{N_{2}O/NO_{x}}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_O / roman_NO start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT mechanism in NH3subscriptNH3\mathrm{NH_{3}}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT oxidation [24] DBD 2023
NH3subscriptNH3\mathrm{NH_{3}}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT/air ignition [25] NRP 2024
n-pentane/air oxidation [26] NRP 2024
6 USC ZDPlasKin + Cantera DME/O2/ArsubscriptO2Ar\mathrm{O_{2}/Ar}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_Ar and C3H8/O2/ArsubscriptC3subscriptH8subscriptO2Ar\mathrm{C_{3}H_{8}/O_{2}/Ar}roman_C start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_Ar ignition [27] nanosecond 2017
7 Tsinghua ZDPlasKin + CHEMKIN DME oxidation kinetics [28] DBD 2021
8 UMN ZDPlasKin + CHEMKIN NH3subscriptNH3\mathrm{NH_{3}}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT pyrolysis and combustion [29] NRP 2021
NH3subscriptNH3\mathrm{NH_{3}}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT/air ignition and NOxsubscriptNOx\mathrm{NO_{x}}roman_NO start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT emission [30] NRP 2022
Plasma-based global pathway analysis [31] NRP 2023
9 KAUST ZDPlasKin + CHEMKIN Lean H2/O2subscriptH2subscriptO2\mathrm{H_{2}/O_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT kinetics [32] DBD 2022
Rich H2/O2subscriptH2subscriptO2\mathrm{H_{2}/O_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT kinetics [33] DBD 2022
NH3subscriptNH3\mathrm{NH_{3}}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT cracking [34] DBD 2023
O3subscriptO3\mathrm{O_{3}}roman_O start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT kinetics [35] DBD 2023
10 Birmingham ZDPlasKin + Cantera NH3/N2/O2/HesubscriptNH3subscriptN2subscriptO2He\mathrm{NH_{3}/N_{2}/O_{2}/He}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_He combustion [36] NRP 2022
11 MIT ZDPlasKin + Cantera 1D CH4/airsubscriptCH4air\mathrm{CH_{4}/air}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / roman_air flame [37] NRP-DBD 2023
12 XJTU ZDPlasKin + in-house chemistry solver NH3/N2/O2subscriptNH3subscriptN2subscriptO2\mathrm{NH_{3}/N_{2}/O_{2}}roman_NH start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ignition [38] NRP/DC 2023
13 SDU ZDPlasKin + CHEMKIN CH4/O2/HesubscriptCH4subscriptO2He\mathrm{CH_{4}/O_{2}/He}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_He ignition [39] NRP-SDBD 2024
14 CERFACS BOLOS + Cantera CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT/air mechanism [11] NRP 2022
CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT/air phenomenological model [5] NRP 2023

2 Methodology

2.1 Governing equations

The governing equations build upon the previous work by Cheng et al. [11] and are presented in a more generalized format. ChemPlasKin integrates the mass fraction Yksubscript𝑌𝑘Y_{k}italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and gas temperature Tgassubscript𝑇gasT_{\text{gas}}italic_T start_POSTSUBSCRIPT gas end_POSTSUBSCRIPT over time t𝑡titalic_t in a 0D neutral gas-plasma reactor for a total of N𝑁Nitalic_N species:

dYkdt=Wkρω˙k,𝑑subscript𝑌𝑘𝑑𝑡subscript𝑊𝑘𝜌subscript˙𝜔𝑘\frac{dY_{k}}{dt}=\frac{W_{k}}{\rho}\dot{\omega}_{k},divide start_ARG italic_d italic_Y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG = divide start_ARG italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ end_ARG over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , (1)
ρcvdTgasdt𝜌subscript𝑐𝑣𝑑subscript𝑇gas𝑑𝑡\displaystyle\rho c_{v}\frac{dT_{\text{gas}}}{dt}italic_ρ italic_c start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT divide start_ARG italic_d italic_T start_POSTSUBSCRIPT gas end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG =k=1Nω˙kukWk+E˙p+k(E˙vibk+R˙VTk),(constant volume)absentsuperscriptsubscript𝑘1𝑁subscript˙𝜔𝑘subscript𝑢𝑘subscript𝑊𝑘superscript˙𝐸𝑝subscript𝑘superscriptsubscript˙𝐸vib𝑘superscriptsubscript˙𝑅VT𝑘(constant volume)\displaystyle=-\sum_{k=1}^{N}\dot{\omega}_{k}u_{k}W_{k}+\dot{E}^{p}+\sum_{k}{(% -\dot{E}_{\text{vib}}^{k}+\dot{R}_{\text{VT}}^{k})},\quad\text{(constant % volume)}= - ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + over˙ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( - over˙ start_ARG italic_E end_ARG start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT + over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) , (constant volume) (2a)
ρcpdTgasdt𝜌subscript𝑐𝑝𝑑subscript𝑇gas𝑑𝑡\displaystyle\rho c_{p}\frac{dT_{\text{gas}}}{dt}italic_ρ italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT divide start_ARG italic_d italic_T start_POSTSUBSCRIPT gas end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG =k=1Nω˙khkWk+E˙p+k(E˙vibk+R˙VTk).(constant pressure)\displaystyle=-\sum_{k=1}^{N}\dot{\omega}_{k}h_{k}W_{k}+\dot{E}^{p}+\sum_{k}{(% -\dot{E}_{\text{vib}}^{k}+\dot{R}_{\text{VT}}^{k})}.\quad\text{(constant % pressure)}= - ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + over˙ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( - over˙ start_ARG italic_E end_ARG start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT + over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) . (constant pressure) (2b)

Here, Wksubscript𝑊𝑘W_{k}italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and ω˙ksubscript˙𝜔𝑘\dot{\omega}_{k}over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT denote the molecular weight and molar production rate of species k𝑘kitalic_k, respectively. ρ𝜌\rhoitalic_ρ represents the density. The terms cvsubscript𝑐𝑣c_{v}italic_c start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT and cpsubscript𝑐𝑝c_{p}italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT signify the mass heat capacities at constant volume and constant pressure, respectively. The term uksubscript𝑢𝑘u_{k}italic_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT refers to the internal energy of species k𝑘kitalic_k, and hksubscript𝑘h_{k}italic_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT represents the enthalpy of species k𝑘kitalic_k. The total plasma energy E˙psuperscript˙𝐸𝑝\dot{E}^{p}over˙ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT is derived from non-elastic electron collision processes 𝒫𝒫\mathcal{P}caligraphic_P:

E˙p=j𝒫εthjqj,superscript˙𝐸𝑝subscript𝑗𝒫superscriptsubscript𝜀th𝑗subscript𝑞𝑗\dot{E}^{p}=\sum_{j\in\mathcal{P}}\varepsilon_{\text{th}}^{j}q_{j},over˙ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_j ∈ caligraphic_P end_POSTSUBSCRIPT italic_ε start_POSTSUBSCRIPT th end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_q start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (3)

where εthjsuperscriptsubscript𝜀th𝑗\varepsilon_{\text{th}}^{j}italic_ε start_POSTSUBSCRIPT th end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT and qjsubscript𝑞𝑗q_{j}italic_q start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT are the threshold energy and net molar production rate for each process j𝑗jitalic_j, respectively.

A significant portion of plasma energy, not directly contributing to fast gas heating, is stored in the vibrational states of species k𝑘kitalic_k, denoted by E˙vibksuperscriptsubscript˙𝐸vib𝑘\dot{E}_{\text{vib}}^{k}over˙ start_ARG italic_E end_ARG start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT. This energy is released to facilitate slow gas heating, primarily through vibrational-translational (V-T) relaxation, at a rate R˙VTksuperscriptsubscript˙𝑅VT𝑘\dot{R}_{\text{VT}}^{k}over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT. Given the system’s non-equilibrium nature, a dedicated equation for the vibrational energy evibksuperscriptsubscript𝑒vib𝑘e_{\text{vib}}^{k}italic_e start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT is necessary:

devibkdt=E˙vibkR˙VTk.𝑑superscriptsubscript𝑒vib𝑘𝑑𝑡superscriptsubscript˙𝐸vib𝑘superscriptsubscript˙𝑅VT𝑘\frac{de_{\text{vib}}^{k}}{dt}=\dot{E}_{\text{vib}}^{k}-\dot{R}_{\text{VT}}^{k}.divide start_ARG italic_d italic_e start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_t end_ARG = over˙ start_ARG italic_E end_ARG start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT - over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT . (4)

R˙VTksuperscriptsubscript˙𝑅VT𝑘\dot{R}_{\text{VT}}^{k}over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT is modeled using the V-T relaxation timescale τVTksuperscriptsubscript𝜏VT𝑘\tau_{\text{VT}}^{k}italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT:

R˙VTk=evibkτVTk.superscriptsubscript˙𝑅VT𝑘subscript𝑒superscriptvib𝑘superscriptsubscript𝜏VT𝑘\dot{R}_{\text{VT}}^{k}=\frac{e_{\text{vib}^{k}}}{\tau_{\text{VT}}^{k}}.over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT = divide start_ARG italic_e start_POSTSUBSCRIPT vib start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG . (5)

Extending the two-component mixture relaxation formula in [40] to multiple species, we obtain:

(τVTk)1=m(Xm/τVTk,m),superscriptsuperscriptsubscript𝜏VT𝑘1subscript𝑚subscript𝑋𝑚superscriptsubscript𝜏VT𝑘𝑚\left(\tau_{\text{VT}}^{k}\right)^{-1}=\sum_{m}\left(X_{m}/\tau_{\text{VT}}^{k% ,m}\right),( italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_X start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_m end_POSTSUPERSCRIPT ) , (6)

where Xmsubscript𝑋𝑚X_{m}italic_X start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT represents the mole fraction of collider m𝑚mitalic_m and τVTk,msuperscriptsubscript𝜏VT𝑘𝑚\tau_{\text{VT}}^{k,m}italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_m end_POSTSUPERSCRIPT the relaxation time for oscillator k𝑘kitalic_k in a high dilution of m𝑚mitalic_m. The condition m=k𝑚𝑘m=kitalic_m = italic_k indicates relaxation in pure gas k𝑘kitalic_k. Experimental data fitting lines in [40] are expressed as:

p0τVTk,m=exp(a(T1/3b)18.42),[atms],subscript𝑝0superscriptsubscript𝜏VT𝑘𝑚exp𝑎superscript𝑇13𝑏18.42delimited-[]atmsp_{0}\tau_{\text{VT}}^{k,m}=\text{exp}\left(a(T^{-1/3}-b)-18.42\right),\ % \mathrm{[atm\cdot s]},italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_m end_POSTSUPERSCRIPT = exp ( italic_a ( italic_T start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT - italic_b ) - 18.42 ) , [ roman_atm ⋅ roman_s ] , (7)

with p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT denoting total pressure in atm, and (a,b)𝑎𝑏(a,b)( italic_a , italic_b ) representing the fitting parameter pair for a k-m𝑘-𝑚k\text{-}mitalic_k - italic_m mixture. Specifically, for air-dominated systems where vibrational energy is primarily stored in N2(v)subscriptN2𝑣\mathrm{N_{2}}(v)roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_v ) for E/N>50𝐸𝑁50E/N>50italic_E / italic_N > 50 Td, it is pertinent to consider N2,O2,subscriptN2subscriptO2\mathrm{N_{2},O_{2},}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , and OO\mathrm{O}roman_O as major collision partners. For m=N2𝑚subscriptN2m=\mathrm{N_{2}}italic_m = roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, (a,b)=(220,0.03)𝑎𝑏2200.03(a,b)=(220,0.03)( italic_a , italic_b ) = ( 220 , 0.03 ) are directly obtained from [40]. For m=O2𝑚subscriptO2m=\mathrm{O_{2}}italic_m = roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, (a,b)=(162,0.03)𝑎𝑏1620.03(a,b)=(162,0.03)( italic_a , italic_b ) = ( 162 , 0.03 ) are inferred from experimental results at 2500 K [41]. The most efficient V-T relaxation pathway involves reactions with O(3P)\mathrm{O(^{3}P)}roman_O ( start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_P ) atoms, characterized by a rate constant of 4.5×1015(T/300)2.1[cm3/s]4.5superscript1015superscript𝑇3002.1delimited-[]superscriptcm3s4.5\times 10^{-15}(T/300)^{2.1}\ [\mathrm{cm^{3}/s]}4.5 × 10 start_POSTSUPERSCRIPT - 15 end_POSTSUPERSCRIPT ( italic_T / 300 ) start_POSTSUPERSCRIPT 2.1 end_POSTSUPERSCRIPT [ roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / roman_s ] [42], setting τVTN2,Osuperscriptsubscript𝜏VTsubscriptN2O\tau_{\text{VT}}^{\mathrm{{N_{2},O}}}italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_O end_POSTSUPERSCRIPT to 488.5/(p0T1.1XO)488.5subscript𝑝0superscript𝑇1.1subscript𝑋O488.5/(p_{0}T^{1.1}X_{\mathrm{O}})488.5 / ( italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT 1.1 end_POSTSUPERSCRIPT italic_X start_POSTSUBSCRIPT roman_O end_POSTSUBSCRIPT ). Note that these τVTk,msuperscriptsubscript𝜏VT𝑘𝑚\tau_{\text{VT}}^{k,m}italic_τ start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k , italic_m end_POSTSUPERSCRIPT parameters, are different from those in the phenomenological model by Castela et al.[4], underscoring the need for validation and refinement for more precise V-T relaxation predictions.

The model’s approach to slow heating via V-T relaxation, encapsulated in equations (2a), (2b), (4) and (5), consolidates all vibrational states of species k𝑘kitalic_k into a singular variable, evibksuperscriptsubscript𝑒vib𝑘e_{\text{vib}}^{k}italic_e start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT. This aggregation significantly simplifies the species count, facilitating practicality in multi-dimensional simulations. However, users have the option to model each vibrational state explicitly along with their specific V-T relaxation reactions. In such scenarios, the slow gas heating term k(E˙vibk+R˙VTk)subscript𝑘superscriptsubscript˙𝐸vib𝑘superscriptsubscript˙𝑅VT𝑘\sum_{k}{(-\dot{E}_{\text{vib}}^{k}+\dot{R}_{\text{VT}}^{k})}∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( - over˙ start_ARG italic_E end_ARG start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT + over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) in equation (2a) and (2b) is unnecessary. Note that the operator-splitting solvers (5-13) listed in Table 1 utilize a model proposed by Flitti & Pancheshnyi [43] to describe gas heating. In this model, external power deposited into the system is distributed among gas heating, electron energy, and chemical energy. Similarly, ChemPlasKin incorporates this model option for gas temperature calculations, analogous to the formulas used in Lefkowitz et al. [9]. For constant volume configurations, the governing equation reads:

ρcvdTgasdt=k=1Nω˙kukWk+eNeveE32RTeω˙e,𝜌subscript𝑐𝑣𝑑subscript𝑇gas𝑑𝑡superscriptsubscript𝑘1𝑁subscript˙𝜔𝑘subscript𝑢𝑘subscript𝑊𝑘𝑒subscript𝑁𝑒subscript𝑣𝑒𝐸32𝑅subscript𝑇𝑒subscript˙𝜔𝑒\rho c_{v}\frac{dT_{\text{gas}}}{dt}=-\sum_{k=1}^{N}\dot{\omega}_{k}u_{k}W_{k}% +eN_{e}v_{e}E-\frac{3}{2}RT_{e}\dot{\omega}_{e},italic_ρ italic_c start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT divide start_ARG italic_d italic_T start_POSTSUBSCRIPT gas end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG = - ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_e italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_E - divide start_ARG 3 end_ARG start_ARG 2 end_ARG italic_R italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , (8)

where e𝑒eitalic_e represents the elementary charge, Nesubscript𝑁𝑒N_{e}italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT the electron number density, vesubscript𝑣𝑒v_{e}italic_v start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT the electron drift velocity under the electric field E𝐸Eitalic_E, R𝑅Ritalic_R the gas constant, and Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT the electron temperature. Employing this model requires the inclusion of reactions related to all vibrational states of species k𝑘kitalic_k, which significantly increases the number of equations that must be solved.

Accurately depicting a weakly ionized plasma necessitates accounting for the electron distribution function, fesubscript𝑓𝑒f_{e}italic_f start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, which adheres to the Boltzmann equation in a six-dimensional phase space [6]:

fet+𝐯efeem𝐄vfe=(fet)coll,subscript𝑓𝑒𝑡subscript𝐯𝑒subscript𝑓𝑒𝑒𝑚𝐄subscript𝑣subscript𝑓𝑒subscriptsubscript𝑓𝑒𝑡coll\frac{\partial f_{e}}{\partial t}+\mathbf{v}_{e}\cdot\nabla f_{e}-\frac{e}{m}% \mathbf{E}\cdot\nabla_{v}f_{e}=\left(\frac{\partial f_{e}}{\partial t}\right)_% {\text{coll}},divide start_ARG ∂ italic_f start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG + bold_v start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⋅ ∇ italic_f start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT - divide start_ARG italic_e end_ARG start_ARG italic_m end_ARG bold_E ⋅ ∇ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = ( divide start_ARG ∂ italic_f start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG ) start_POSTSUBSCRIPT coll end_POSTSUBSCRIPT , (9)

where 𝐯esubscript𝐯𝑒\mathbf{v}_{e}bold_v start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT represents electron velocity, e𝑒eitalic_e the elementary charge, m𝑚mitalic_m the electron mass, 𝐄𝐄\mathbf{E}bold_E the electric field, and vsubscript𝑣\nabla_{v}∇ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT the velocity gradient operator. The term on the right-hand side quantifies fesubscript𝑓𝑒f_{e}italic_f start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT’s collision-induced rate of change. Employing the widely used two-term approximation allows for the decomposition of equation (9), with components of fesubscript𝑓𝑒f_{e}italic_f start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT renormalized as probability distribution functions. The isotropic part, denoted F0subscript𝐹0F_{0}italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, serves as the electron energy EEDF [6, 44]. Reaction rate coefficients for collisions 𝒫𝒫\mathcal{P}caligraphic_P are derived from F0subscript𝐹0F_{0}italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and cross-section σjsubscript𝜎𝑗\sigma_{j}italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT by integrating over electron energy ϵitalic-ϵ\epsilonitalic_ϵ:

kj=2e/m0ϵσjF0𝑑ϵ,j𝒫.formulae-sequencesubscript𝑘𝑗2𝑒𝑚superscriptsubscript0italic-ϵsubscript𝜎𝑗subscript𝐹0differential-ditalic-ϵ𝑗𝒫k_{j}=\sqrt{2e/m}\int_{0}^{\infty}{\epsilon\sigma_{j}F_{0}}d\epsilon,\quad j% \in\mathcal{P}.italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = square-root start_ARG 2 italic_e / italic_m end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ϵ italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d italic_ϵ , italic_j ∈ caligraphic_P . (10)

Furthermore, F0subscript𝐹0F_{0}italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT facilitates the calculation of electron temperature:

Te=230ϵ3/2F0𝑑ϵ,subscript𝑇𝑒23superscriptsubscript0superscriptitalic-ϵ32subscript𝐹0differential-ditalic-ϵT_{e}=\frac{2}{3}\int_{0}^{\infty}{\epsilon^{3/2}F_{0}}d\epsilon,italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = divide start_ARG 2 end_ARG start_ARG 3 end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ϵ start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d italic_ϵ , (11)

which is commonly featured in the rate coefficient expressions for recombination reactions involving electrons.

2.2 Heat loss model for pin-pin NRP discharges

NRP discharges produced by pin-pin electrodes typically result in ultrafast gas heating within an approximately constant volume, generating a weak shock wave and thermal energy loss due to gas expansion [45, 46]. Such hydrodynamic phenomena cannot be directly resolved within the limitations of an 0D framework, leading to potential overestimations of temperature increase if heat loss mechanisms are neglected. This discrepancy may remain negligible for scenarios involving low plasma power; however, it becomes increasingly significant with higher energy deposition and a greater number of nanosecond pulses. To enhance ChemPlasKin’s ability to accurately model temperature dynamics, particularly for applications investigating plasma-assisted ignition delay times, we propose the following heat loss model.

During the nanosecond pulse, the gas within the discharge kernel undergoes a rise in temperature and pressure under constant volume conditions. Once the plasma energy input ceases, the discharge kernel experiences an isentropic expansion from state 1 to state 2, described by:

T2T1=(ρ2ρ1)γ1=(p2p1)γ1γ,subscript𝑇2subscript𝑇1superscriptsubscript𝜌2subscript𝜌1𝛾1superscriptsubscript𝑝2subscript𝑝1𝛾1𝛾\frac{T_{2}}{T_{1}}=\left(\frac{\rho_{2}}{\rho_{1}}\right)^{\gamma-1}=\left(% \frac{p_{2}}{p_{1}}\right)^{\frac{\gamma-1}{\gamma}},divide start_ARG italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG = ( divide start_ARG italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_γ - 1 end_POSTSUPERSCRIPT = ( divide start_ARG italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG italic_γ - 1 end_ARG start_ARG italic_γ end_ARG end_POSTSUPERSCRIPT , (12)

where γ𝛾\gammaitalic_γ represents the heat capacity ratio. The propagation of the resulting weak shock wave, occurring on the microsecond timescale, rapidly equalizes the kernel pressure back to the ambient pressure p2subscript𝑝2p_{2}italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. For practical code implementation, the temperature and density at the pulse’s conclusion are adjusted to reflect T2subscript𝑇2T_{2}italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and ρ2subscript𝜌2\rho_{2}italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT values within a single timestep, for given initial (p1subscript𝑝1p_{1}italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) and final (p2subscript𝑝2p_{2}italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) pressures. This model effectively modifies the gas temperature equation between constant volume and constant pressure conditions to mimic the isentropic process, a flexibility not typically offered by previous solvers listed in Table 1.

The heat loss model is augmented to account for radial thermal conduction from the hot discharge kernel to the ambient environment, described by:

q˙loss=λdTdrAdisVdisλC0TT0rdis2,subscript˙𝑞𝑙𝑜𝑠𝑠𝜆𝑑𝑇𝑑𝑟subscript𝐴dissubscript𝑉dis𝜆subscript𝐶0𝑇subscript𝑇0superscriptsubscript𝑟dis2\dot{q}_{loss}=-\lambda\frac{dT}{dr}\frac{A_{\text{dis}}}{V_{\text{dis}}}% \approx-\lambda C_{0}\frac{T-T_{0}}{r_{\text{dis}}^{2}},over˙ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_l italic_o italic_s italic_s end_POSTSUBSCRIPT = - italic_λ divide start_ARG italic_d italic_T end_ARG start_ARG italic_d italic_r end_ARG divide start_ARG italic_A start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT end_ARG start_ARG italic_V start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT end_ARG ≈ - italic_λ italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG italic_T - italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (13)

where q˙losssubscript˙𝑞𝑙𝑜𝑠𝑠\dot{q}_{loss}over˙ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_l italic_o italic_s italic_s end_POSTSUBSCRIPT represents the heat loss in W/m3Wsuperscriptm3\mathrm{W/m^{3}}roman_W / roman_m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, λ𝜆\lambdaitalic_λ the thermal conductivity, and rdissubscript𝑟disr_{\text{dis}}italic_r start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT the effective radius of the cylindrical discharge characterized by surface area Adissubscript𝐴disA_{\text{dis}}italic_A start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT and volume Vdissubscript𝑉disV_{\text{dis}}italic_V start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT. Here, C0subscript𝐶0C_{0}italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a dimensionless empirical coefficient, typically set to 1.0, and T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT denotes the ambient temperature. To simulate gas temperature dynamics during intervals without plasma energy input, the following equation is actually solved:

ρcpdTgasdt=k=1Nω˙khkWk+kR˙VTkλC0TT0rdis2.𝜌subscript𝑐𝑝𝑑subscript𝑇gas𝑑𝑡superscriptsubscript𝑘1𝑁subscript˙𝜔𝑘subscript𝑘subscript𝑊𝑘subscript𝑘superscriptsubscript˙𝑅VT𝑘𝜆subscript𝐶0𝑇subscript𝑇0superscriptsubscript𝑟dis2\rho c_{p}\frac{dT_{\text{gas}}}{dt}=-\sum_{k=1}^{N}\dot{\omega}_{k}h_{k}W_{k}% +\sum_{k}\dot{R}_{\text{VT}}^{k}-\lambda C_{0}\frac{T-T_{0}}{r_{\text{dis}}^{2% }}.italic_ρ italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT divide start_ARG italic_d italic_T start_POSTSUBSCRIPT gas end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG = - ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over˙ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over˙ start_ARG italic_R end_ARG start_POSTSUBSCRIPT VT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT - italic_λ italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG italic_T - italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (14)

2.3 Code development

2.3.1 Electron Boltzmann equation (EBE) solver

To enhance performance in simulations, ChemPlasKin incorporates an EBE solver, named CppBOLOS, directly integrated into the Cantera library at the source level. This integration represents a central feature of ChemPlasKin’s functionality. Among the available EBE solvers within the plasma community, none provides an open-source C++ implementation that employs the efficient two-term expansion method akin to Bolsig+ [6]. For instance, while MultiBolt [47] offers a C++ open-source version (v3.x), its multi-term model tends to be computationally intensive and not essential for a gas-plasma kinetics solver. Addressing this need, we developed a C++ version of BOLOS [12], an open-source Python solver that follows the algorithms specified in [6]. Now referred to as CppBOLOS, this solver’s development process was significantly assisted by GPT-4 [48]. It is seamlessly integrated with Cantera, enabling dynamic solutions of the electron Boltzmann equation and real-time updates of the EEDF based on the temporal evolution of gas temperature, mixture composition, and reduced electric field. CppBOLOS accepts input cross-section data in the popular LXCat format [49], which ensures broad compatibility and simplifies data integration. The validation of CppBOLOS is briefly presented in Section 3.1.

2.3.2 Code architecture

Refer to caption
Figure 1: Code architecture of ChemPlasKin.

Figure 1 illustrates the code architecture of ChemPlasKin, highlighting its streamlined approach to handling both plasma and neutral gas kinetics. Key to its design is the use of a unified YAML-type input file for all thermal properties and reactions. This choice facilitates ease of use and integration with existing datasets. ChemPlasKin extends Cantera’s reaction module to register inelastic electron-impact reactions classified as type of ‘Boltzmann’, as shown in the examples of Listing LABEL:lst:yaml_example_1:

1# Example 1: Vibrational excitation
2- equation: Electron + N2 => Electron + N2
3 type: Boltzmann
4 process: N2 -> N2(v1)
5 duplicate: true
6 energy_transfer: {evib_N2: 0.291_eV}
7
8# Example 2: Electronic excitation
9- equation: Electron + O2 => Electron + O2(a1)
10 type: Boltzmann
11 process: O2 -> O2(a1)
12 energy_transfer: {e_th: 0.977_eV}
Listing 1: Examples of YAML configurations for Boltzmann type reactions. Vibrational states of N2subscriptN2\mathrm{N_{2}}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are agrragated and the vibrational energy is stored in evib_N2.

To accommodate the diverse range of rate constant expressions encountered in plasma kinetics, ChemPlasKin employs muParser library [50] for parsing any complex mathematical expressions not natively supported by Cantera. Listing LABEL:lst:yaml_example_2 shows an example of e+N4+superscriptesuperscriptsubscriptN4\mathrm{e^{-}+N_{4}^{+}}roman_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT + roman_N start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT recombination reaction, where the reaction rate is calculated as a function of electron temperature. The energy_transfer entry values enforce zero gas heating and the corresponding energy (3.8 eV) is stored in evibN2superscriptsubscript𝑒vibsubscriptN2e_{\text{vib}}^{\mathrm{N_{2}}}italic_e start_POSTSUBSCRIPT vib end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT.

1# Example 3: Recombination
2- equation: Electron + N4+ => N2 + N2(C3)
3 type: PlasmaCustomExpr
4 energy_transfer: {e_th: -3.8_eV, evib_N2: 3.8_eV} # No gas heating
5 rateExpr: {A: 2.086e+19, Expr: Te^-0.5}
Listing 2: Example of YAML configurations for PlasmaCustomExpr type reactions.

For added convenience, an optional parser tool, parsePlasKin, is available to convert ZDPlaKin input mechanism files into the human-readable YAML format automatically. This feature is particularly beneficial given the prevalence of ZDPlaKin mechanisms published as supplementary materials in the field. Configuration parameters are set through the text files controlDict and chemPlasProperties, ensuring straightforward operation.

The core of ChemPlasKin’s simulation capabilities is encapsulated in plasmaReactor.h, where a single ordinary differential equation (ODE) system—encompassing equations (1, 2a, 2b 4)—is constructed and integrated within the main time loop of the master code. The ODEs are solved using the CVODE solver from the SUNDIALS suite [51, 52], which is efficiently utilized by Cantera for high-precision integration.

2.3.3 Code efficiency

ChemPlasKin’s unified ODE system offers distinct advantages over operator-splitting techniques, by eliminating splitting errors and enabling much larger timesteps in the main time loop. Although it is suggested that splitting the plasma kinetics could expedite integration due to the higher stiffness during nanosecond pulses, such benefits become marginal during the much longer pulse intervals where plasma stiffness diminishes. Moreover, the plasma kinetics component of ChemPlasKin, akin to that solved by ZDPlasKin, encompasses a broad spectrum of timescales beyond just ultrafast electron-impact reactions. Last but not least, switching between ZDPlasKin and CHEMKIN/Cantera requires reinitialization of their respective ODE solvers at every step, adding extra computational overhead to the solvers in the initial transient to find suitable integration timesteps. Consequently, ChemPlasKin’s unified approach not only simplifies the integration process but also enhances overall simulation efficiency compared to the operator-splitting method. To demonstrate this, we have also built a ZDPlasKin-Cantera solver consistent with previous counterparts and compared it with ChemPlasKin, as detailed in Section 3.2.

3 Code Validation

This section begins with a brief validation of CppBOLOS against Bolsig+, establishing the groundwork for subsequent evaluations. The subsequent five subsections are carefully structured to assess ChemPlasKin’s performance across diverse simulation scenarios. We present comparisons of simulation results between literature sources, our equivalent Cantera-ZDPlasKin solver, and ChemPlasKin in Section 3.2. The capability of ChemPlasKin to predict ultrafast gas heating and radical species production is assessed in Section 3.3. The slow gas heating model is validated through experimental data in Section 3.4, while a practical application case of fuel reforming is examined in Section 3.5. Finally, the efficacy of the heat loss model is tested in Section 3.6. For the simulations discussed in Sections 3.3, 3.4, and 3.6, we utilize a rigorously validated detailed PAC mechanism involving 100 species and 964 reactions for a methane-air mixture, as developed by [11].

3.1 CppBOLOS

Figure 2 illustrates the comparison between CppBOLOS and Bolsig+ for mean electron energy and ionization rate in pure N2subscriptN2\mathrm{N_{2}}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT across a wide range of reduced electric fields (E/N𝐸𝑁E/Nitalic_E / italic_N). Additionally, relative differences compared to Bolsig+ are plotted, demonstrating satisfactory consistency except at very low E/N𝐸𝑁E/Nitalic_E / italic_N values. The large relative differences in ionization rate at low E/N𝐸𝑁E/Nitalic_E / italic_N are considered acceptable because the infinitesimally small absolute values have a negligible impact on the accurate prediction of the plasma kinetics system’s evolution.

Refer to caption
Figure 2: Comparison between CppBOLOS and Bolsig+ for pure N2subscriptN2\mathrm{N_{2}}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT at 300 K, using cross-section data from Phelps [49].

3.2 Plasma assisted 𝐇𝟐/𝐎𝟐/𝐇𝐞subscript𝐇2subscript𝐎2𝐇𝐞\mathbf{H_{2}/O_{2}/He}bold_H start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT / bold_O start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT / bold_He ignition

Mao et al. [18] conducted numerical simulations of H2/O2/HesubscriptH2subscriptO2He\mathrm{H_{2}/O_{2}/He}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_He ignition assisted by hybrid NRP and DC discharges using the ZDPlasKin-CHEMKIN solver developed by Lefkowitz et al. [9]. As no openly available PAC solver exists for use as a benchmark, we integrated ZDPlasKin with Cantera to evaluate the computational efficiency of the popular operator-splitting technique, adhering to the strategy described in [9]. The Cantera-ZDPlasKin solver includes a Python wrapper of ZDPlasKin that facilitates efficient data communications between the two codes. The timestep in the main time loop is dynamically adjusted to accommodate the nanosecond timescale and exponential growth of electrons in NRP discharges, then gradually relaxed to 107superscript10710^{-7}10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT s during the pulse intervals, balancing accuracy with the computational cost of the operator-splitting method.

The mixture has an initial composition of H2:O2:He=0.1667:0.0883:0.75:subscriptH2subscriptO2:He0.1667:0.0883:0.75\mathrm{H_{2}:O_{2}:He=0.1667:0.0883:0.75}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT : roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT : roman_He = 0.1667 : 0.0883 : 0.75, and the adiabatic system is maintained at atmospheric pressure. The reduced electric field (E/N𝐸𝑁E/Nitalic_E / italic_N) for NRP discharges is set at 100 Td (1 Td = 1021Vm2superscript1021superscriptVm2\mathrm{10^{-21}Vm^{2}}10 start_POSTSUPERSCRIPT - 21 end_POSTSUPERSCRIPT roman_Vm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), with a frequency of 30 kHz. DC discharges at an E/N𝐸𝑁E/Nitalic_E / italic_N of 20 Td are applied between the nanosecond pulses. The deposited plasma energy is fixed at 0.1 mJ/cm3mJsuperscriptcm3\mathrm{mJ/cm^{3}}roman_mJ / roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT per pulse. To ensure consistency, ChemPlasKin operates with the Flitti & Pancheshnyi [43] model for gas temperature. The plasma kinetic mechanism used by [18], originally formatted for ZDPlasKin, has been automatically converted into YAML format compatible with ChemPlasKin.

Figure 3 compares these three solvers for their predictions of ignition delay times (IDT) for NRP and NRP/DC hybrid discharges assisted ignition across various initial temperatures. IDT is defined as the point of maximum temperature gradient during the pulse intervals in our simulations. Good agreement is achieved among the three solvers for both discharge types.

Refer to caption
Figure 3: Ignition delay times (IDT) for NRP discharge assisted ignition and NRP/DC hybrid discharge assisted ignition. Results from ChemPlasKin are compared against those from Mao et al. [18] and our Cantera-ZDPlasKin solver.

The predictive capabilities of the solvers for electron and O2(a1Δg)subscriptO2superscripta1subscriptΔg\mathrm{O_{2}(a^{1}\Delta_{g})}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( roman_a start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT ) production over the first ten pulses in the hybrid NRP/DC discharge scenario are displayed in Figure 4. These results underscore the performance consistency of our solvers, particularly Cantera-ZDPlasKin, with those reported by Mao et al. [18].

Refer to caption
Figure 4: Temporal evolution of electron and O2(a1Δg)subscriptO2superscripta1subscriptΔg\mathrm{O_{2}(a^{1}\Delta_{g})}roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( roman_a start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT ) mole fractions during the first ten pulses of a hybrid discharge at 400 K.

A comparison of computational efficiency between ChemPlasKin and Cantera-ZDPlasKin was carried out for NRP discharge assisted ignition cases, as shown in Figure 5. ChemPlasKin achieved a three-fold speed-up. Note that in Figure 5, ChemPlasKin employed the same dynamic timestep settings in the main loop to ensure a fair comparison. In actual applications, ChemPlasKin can utilize much larger outer timesteps, as it is not constrained by operator-splitting errors.

Refer to caption
Figure 5: Computational cost (clock time) comparison between ChemPlasKin and Cantera-ZDPlasKin corresponding to the NRP assisted ignition depicted in Figure 3. Single-thread execution at 2.3 GHz. Profiling results confirm that there is no data communication bottleneck in the Cantera-ZDPlasKin solver.

Note that many operator-splitting solvers listed in Table 1, which operate based on kinetic mechanisms developed in various studies, lack rigorous validation against experimental data at micro time scales, such as radical production and fast and slow gas heating during a single NRP discharge period. The following two subsections will present validation of ChemPlasKin that addresses this gap, similar to the approach taken by Cheng et al. [11].

3.3 Spark discharge in air

This test case aligns with Case A from [11], utilizing the experimental setup described in [53], where NRP spark discharges are generated between pin-pin electrodes. The gas temperature can be inferred from the rotational temperature of N2(C)subscriptN2C\mathrm{N_{2}(C)}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( roman_C ) for a duration of at least 17 ns [53]. The initial gas mixture temperature is set at 1500 K with a composition of N2:O2:O=77.4:18.6:4:subscriptN2subscriptO2:O77.4:18.6:4\mathrm{N_{2}:O_{2}:O=77.4:18.6:4}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT : roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT : roman_O = 77.4 : 18.6 : 4, accounting for the thermal and chemical effects of preceding pulses. Figure 6 presents a comparison of the gas temperature evolution and O(3P)\mathrm{O(^{3}P)}roman_O ( start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_P ) production as predicted by ChemPlasKin against experimental data.

Refer to caption
Figure 6: Comparison of ChemPlasKin predictions with experimental data on the temporal evolution of (a) O(3P)\mathrm{O(^{3}P)}roman_O ( start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_P ) production during and after a nanosecond pulse and (b) gas temperature, demonstrating good agreement. Experimental data from [53].

3.4 V-T relaxation in air

This validation case aligns with Case C from [11], utilizing the experimental setup detailed in [54]. The initial conditions of air are set at a temperature of 300 K and a pressure of 100 Torr. Between the pin-pin electrodes, the applied E/N𝐸𝑁E/Nitalic_E / italic_N is approximately 100 Td, with up to 50%percent5050\%50 % of the plasma energy contributing to the vibrational excitation of N2subscriptN2\mathrm{N_{2}}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. In this scenario, the model’s prediction of the rotational temperature, represented by the solid curve shows good agreement with the experimental data, as shown in Figure 7. Note that the simulation focuses solely on N2subscriptN2\mathrm{N_{2}}roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT for the vibrational energy equation, under the assumption of isobaric conditions due to the minimal impact of fast gas heating in this context. The exclusion of heat loss from thermal diffusion in the model accounts for the observed discrepancies between calculated results and experimental measurements on the millisecond timescale.

Refer to caption
Figure 7: ChemPlasKin predictions of gas temperature from VT-relaxation of N2(v)subscriptN2𝑣\mathrm{N_{2}}(v)roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_v ) in pure air, showcasing a comparison against experimental measurements from [54].

3.5 Hydrogen oxidation with DBD discharges

This validation case compares ChemPlasKin’s performance against both experimental and numerical results detailed in [32], which investigates the kinetics of plasma-assisted oxidation of H2subscriptH2\mathrm{H_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in an undiluted lean H2/O2subscriptH2subscriptO2\mathrm{H_{2}/O_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT system. In the referenced study, a preheated gas mixture is subjected to treatment in a dielectric barrier discharge (DBD) reactor under constant pressure and temperature conditions. To approximate the filamentary nature of the discharges, it is assumed that each fluid particle encounters a total of 288 rectangular pulses at a constant E/N𝐸𝑁E/Nitalic_E / italic_N.

Figure 8 presents a comparison of the H2subscriptH2\mathrm{H_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT oxidation rates derived from ChemPlasKin against the experimental and numerical results under varying gas temperatures. The agreement between ChemPlasKin’s predictions and the reference results is generally good, with a larger discrepancy at higher temperatures, potentially attributed to the omission of thermal cracking residence time in our model. ChemPlasKin completes each data point presented in Figure 8 within approximately 45 seconds of single-threaded execution on a 2.3 GHz core.

Refer to caption
Figure 8: Comparison of ChemPlasKin simulations with reference data on plasma-assisted oxidation rates of H2subscriptH2\mathrm{H_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT over a range of gas temperatures (Tgassubscript𝑇gasT_{\text{gas}}italic_T start_POSTSUBSCRIPT gas end_POSTSUBSCRIPT) for an H2/O2subscriptH2subscriptO2\mathrm{H_{2}/O_{2}}roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT mixture (ϕ=0.01italic-ϕ0.01\phi=0.01italic_ϕ = 0.01) at 10 W of plasma power. Reference data from [32].

3.6 Heat loss model

The proposed heat loss model, detailed in Section 3.6, is validated focusing on the fast gas-heating phenomenon, thereby omitting the need for detailed plasma kinetics and slow gas heating. Figure 9 illustrates a typical scenario for applying the heat loss model, aligned with the test case described in Section 3.3. Here, NRP discharges at 10 kHz are generated between pin-pin electrodes, with each pulse delivering an energy input of 670μJ670𝜇J670\ \mathrm{\mu J}670 italic_μ roman_J. According to the phenomenological plasma model by [4], 20%percent2020\%20 % of the pulse energy is allocated to fast gas-heating within a 10 ns duration. A horizontal 2D computational domain is defined, featuring the pulsed heating source at its center, modeled with an energy density spatial function =erfc(r2/6.9×108)2.5erfcsuperscriptsuperscript𝑟26.9superscript1082.5\mathcal{F}=\text{erfc}\left(r^{2}/{6.9\times 10^{-8}}\right)^{2.5}caligraphic_F = erfc ( italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 6.9 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT, where r𝑟ritalic_r denotes the radial distance from the discharge center and an effective discharge radius rdissubscript𝑟disr_{\text{dis}}italic_r start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT of 225μm225𝜇m225\ \mathrm{\mu m}225 italic_μ roman_m is specified. The domain employs a non-reflective boundary condition, ensuring the pressure field near the discharge zone recovers the ambient level within microseconds [55].

The averaged temperature within rdissubscript𝑟disr_{\text{dis}}italic_r start_POSTSUBSCRIPT dis end_POSTSUBSCRIPT in the 2D domain, solved using OpenFOAM [56], represents the discharge kernel temperature. In parallel, a 0D model, simply written as an independent Python code, utilizes the Cantera library to update thermal conductivity based on temperature changes. After applying ten heating pulses, Figure 10 displays the temperature comparison between the 2D and 0D models. The models demonstrate good overall agreement, with the 0D model’s assumption of instantaneous isentropic expansion capturing the sharp temperature drop at each pulse’s conclusion, as elaborated in the zoom-in subplot of Figure 10. Without incorporating the heat loss model, a conventional 0D chemistry-plasma model would likely predict a step-wise, monotonic temperature increase for NRP discharges, contrasting with the quasi-steady state indicated by the 2D model.

Refer to caption
Figure 9: Schematic of a 2D computational domain for simulating ultrafast heating effects from pin-pin generated NRP discharges using phenomenological model.
Refer to caption
Figure 10: Comparison of temperature evolution within the discharge kernel across 10 NRP discharges between 2D and 0D phenomenological models. V-T relaxation heating not included. The red bars highlight the highest temperature for each nanosecond pulse in the 2D model. C0=1.0subscript𝐶01.0C_{0}=1.0italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.0 is employed in the 0D model.

Following the standalone examination discussed previously, the impact of the heat loss model on gas temperature in ChemPlasKin, which incorporates detailed gas-plasma kinetics of a methane/air mixture [11], is presented in Figure 11. The discrepancy in temperature evolution with and without the heat loss model during the first ten pulses clearly demonstrates its potential influence on ignition delay time calculations. We acknowledge that further validation and improvement of this model are necessary, ideally through direct comparison with experimental data or high-fidelity multi-dimensional simulations.

Refer to caption
Figure 11: Temperature evolution in a 0.0499CH4/0.7505N2/0.1996O20.0499subscriptCH40.7505subscriptN20.1996subscriptO2\mathrm{0.0499CH_{4}/0.7505N_{2}/0.1996O_{2}}0.0499 roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / 0.7505 roman_N start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / 0.1996 roman_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT mixture with NRP discharge assisted ignition over the first ten pulses. Deposited plasma energy is 10mJ/cm310mJsuperscriptcm3\mathrm{10\ mJ/cm^{3}}10 roman_mJ / roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT per pulse, with C0=1.0subscript𝐶01.0C_{0}=1.0italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.0.

4 Conclusions

This study describes ChemPlasKin, a freeware tool developed for simulating gas-plasma kinetic processes. ChemPlasKin integrates the electron Boltzmann equation solver CppBOLOS with the Cantera library, enabling the solution of neutral gas and plasma kinetics within a unified ODE system. Additionally, a supplementary heat loss model is proposed to enhance the accuracy of temperature predictions for nanosecond repetitively pulsed (NRP) discharges in configurations utilizing pin-pin electrodes.

To evaluate the computational efficiency of ChemPlasKin, we constructed a Cantera-ZDPlasKin PAC solver using the widely used operator-splitting method. In the test cases, this configuration achieved a threefold speedup, and we anticipate even faster performance with a relaxed outer timestep.

The C++ solver has been validated against experimental results across various aspects and timescales, including ultrafast gas heating and radical production, slow gas heating from V-T relaxation and fuel reforming involving hundreds of pulses. ChemPlasKin shows its capability as a versatile tool, particularly useful for PAC and fuel reforming. It can also be used purely as a plasma kinetics solver, akin to ZDPlasKin. Its development provides a resource for researchers seeking to explore gas-plasma kinetics without the need to invest in custom coding efforts. Continued validation and enhancement by the user community are encouraged. We also anticipate ChemPlasKin to be incorporated into CFD codes to enable high-fidelity simulations of fully-coupled plasma-assisted reacting flows, acknowledging that computational costs remain a significant consideration.

Acknowledgments

This research was funded by King Abdullah University of Science and Technology (KAUST). The authors would like to thank Seunghwan Bang and Dr. Ramses Snoeckx for useful discussion. The first author also thanks friends Renston, Vijay, and Alessandro for the great tunes and good times that helped lighten the load.

References

  • [1] Yiguang Ju and Wenting Sun. Plasma assisted combustion: Dynamics and chemistry. Progress in Energy and Combustion Science, 48:21–83, June 2015.
  • [2] Deanna A. Lacoste. Flames with plasmas. Proceedings of the Combustion Institute, 39(4):5405–5428, January 2023.
  • [3] Yaolin Wang, Yanzhen Chen, Jonathan Harding, Hongyuan He, Annemie Bogaerts, and Xin Tu. Catalyst-free single-step plasma reforming of CH4 and CO2 to higher value oxygenates under ambient conditions. Chemical Engineering Journal, 450:137860, December 2022.
  • [4] Maria Luis Gracio Bilro Castela. Direct Numerical Simulations of plasma-assisted ignition in quiescent and turbulent flow conditions. phdthesis, Université Paris-Saclay, May 2016.
  • [5] Nicolas Barléon, Lionel Cheng, Bénédicte Cuenot, and Olivier Vermorel. A phenomenological model for plasma-assisted combustion with NRP discharges in methane-air mixtures: PACMIND. Combustion and Flame, 253:112794, July 2023.
  • [6] G. J. M. Hagelaar and L. C. Pitchford. Solving the Boltzmann equation to obtain electron transport coefficients and rate coefficients for fluid models. Plasma Sources Science and Technology, 14(4):722, October 2005.
  • [7] S Pancheshnyi, B Eismann, GJM Hagelaar, and LC Pitchford. Computer code ZDPlasKin. University of Toulouse, LAPLACE, CNRS-UPS-INP, Toulouse, France, 2008.
  • [8] R. J. Kee, F. M. Rupley, and J. A. Miller. Chemkin-II: A Fortran chemical kinetics package for the analysis of gas-phase chemical kinetics. Technical Report SAND-89-8009, Sandia National Lab. (SNL-CA), Livermore, CA (United States), September 1989.
  • [9] Joseph K Lefkowitz, Peng Guo, Aric Rousso, and Yiguang Ju. Species and temperature measurements of methane oxidation in a nanosecond repetitively pulsed discharge. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 373(2048):20140333, August 2015. Publisher: Royal Society.
  • [10] David G. Goodwin, Harry K. Moffat, Ingmar Schoegl, Raymond L. Speth, and Bryan W. Weber. Cantera: An Object-oriented Software Toolkit for Chemical Kinetics, Thermodynamics, and Transport Processes, 2022.
  • [11] L. Cheng, N. Barleon, B. Cuenot, O. Vermorel, and A. Bourdon. Plasma assisted combustion of methane-air mixtures: Validation and reduction. Combustion and Flame, 240:111990, June 2022.
  • [12] aluque. aluque/bolos, July 2023. original-date: 2014-06-10T08:17:00Z.
  • [13] Igor V. Adamovich, Ting Li, and Walter R. Lempert. Kinetic mechanism of molecular energy transfer and chemical reactions in low-temperature air-fuel plasmas. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 373(2048):20140336, August 2015. Publisher: Royal Society.
  • [14] Anthony C. DeFilippo and Jyh-Yuan Chen. Modeling plasma-assisted methane–air ignition using pre-calculated electron impact reaction rates. Combustion and Flame, 172:38–48, October 2016.
  • [15] Nicholas Deak, Aurélie Bellemans, and Fabrizio Bisetti. Plasma-assisted ignition of methane/air and ethylene/air mixtures: Efficiency at low and high pressures. Proceedings of the Combustion Institute, 38(4):6551–6558, January 2021.
  • [16] T. Hazenberg, J. van Dijk, and J. A. van Oijen. Chemical flux analysis of low-temperature plasma-enhanced oxidation of methane and hydrogen in argon. Combustion and Flame, 257:113037, November 2023.
  • [17] Xingqian Mao and Qi Chen. Effects of Vibrational Excitation on Nanosecond Discharge Enhanced Methane–Air Ignition. AIAA Journal, 56(11):4312–4320, 2018. Publisher: American Institute of Aeronautics and Astronautics _eprint: https://doi.org/10.2514/1.J057304.
  • [18] Xingqian Mao, Qi Chen, Aric C. Rousso, Timothy Y. Chen, and Yiguang Ju. Effects of controlled non-equilibrium excitation on H2/O2/He ignition using a hybrid repetitive nanosecond and DC discharge. Combustion and Flame, 206:522–535, August 2019.
  • [19] Xingqian Mao, Aric Rousso, Qi Chen, and Yiguang Ju. Numerical modeling of ignition enhancement of CH4/O2/He mixtures using a hybrid repetitive nanosecond and DC discharge. Proceedings of the Combustion Institute, 37(4):5545–5552, January 2019.
  • [20] Yuan Wang, Peng Guo, Haitao Chen, and Zheng Chen. Numerical modeling of ignition enhancement by repetitive nanosecond discharge in a hydrogen/air mixture II: forced ignition. Journal of Physics D: Applied Physics, 54(6):065502, November 2020. Publisher: IOP Publishing.
  • [21] Hongtao Zhong, Xingqian Mao, Aric C Rousso, Charles L Patrick, Chao Yan, Wenbin Xu, Qi Chen, Gerard Wysocki, and Yiguang Ju. Kinetic study of plasma-assisted n-dodecane/O2/N2 pyrolysis and oxidation in a nanosecond-pulsed discharge. Proceedings of the Combustion Institute, 38(4):6521–6531, January 2021.
  • [22] Galia Faingold and Joseph K. Lefkowitz. A numerical investigation of NH3/O2/He ignition limits in a non-thermal plasma. Proceedings of the Combustion Institute, 38(4):6661–6669, January 2021.
  • [23] Galia Faingold, Omer Kalitzky, and Joseph K. Lefkowitz. Plasma reforming for enhanced ammonia-air ignition: A numerical study. Fuel Communications, 12:100070, September 2022.
  • [24] Hongtao Zhong, Xingqian Mao, Ning Liu, Ziyu Wang, Timothy Ombrello, and Yiguang Ju. Understanding non-equilibrium N2O/NOx chemistry in plasma-assisted low-temperature NH3 oxidation. Combustion and Flame, 256:112948, October 2023.
  • [25] Xingqian Mao, Hongtao Zhong, Ning Liu, Ziyu Wang, and Yiguang Ju. Ignition enhancement and NOx formation of NH3/air mixtures by non-equilibrium plasma discharge. Combustion and Flame, 259:113140, January 2024.
  • [26] Nan Liu, Qi Chen, Xianwu Jiang, Jie Chen, and Lidong Zhang. Kinetic roles of excited state species in the oxidation of n-pentane/air assisted by nanosecond-pulsed discharge. Combustion and Flame, 259:113142, January 2024.
  • [27] Vyaas Gururajan and Fokion N. Egolfopoulos. Transient plasma effects on the autoignition of DME/O2/Ar and C3H8/O2/Ar mixtures. Proceedings of the Combustion Institute, 36(3):4165–4174, January 2017.
  • [28] Ruzheng Zhang, Handong Liao, Jiuzhong Yang, and Bin Yang. Exploring chemical kinetics of plasma assisted oxidation of dimethyl ether (DME). Combustion and Flame, 225:388–394, March 2021.
  • [29] Taaresh Sanjeev Taneja and Suo Yang. Comparing Low-Mach and Fully-Compressible CFD Solvers for Phenomenological Modeling of Nanosecond Pulsed Plasma Discharges with and without Turbulence. In AIAA SCITECH 2022 Forum. American Institute of Aeronautics and Astronautics, 2021. _eprint: https://arc.aiaa.org/doi/pdf/10.2514/6.2022-0976.
  • [30] Taaresh Sanjeev Taneja, Praise Noah Johnson, and Suo Yang. Nanosecond pulsed plasma assisted combustion of ammonia-air mixtures: Effects on ignition delays and NOx emission. Combustion and Flame, 245:112327, November 2022.
  • [31] Praise N. Johnson, Taaresh S. Taneja, and Suo Yang. Plasma-based global pathway analysis to understand the chemical kinetics of plasma-assisted combustion and fuel reforming. Combustion and Flame, 255:112927, September 2023.
  • [32] Ramses Snoeckx, Daeyoung Jun, Bok Jik Lee, and Min Suk Cha. Kinetic study of plasma assisted oxidation of H2 for an undiluted lean mixture. Combustion and Flame, 242:112205, August 2022.
  • [33] Ramses Snoeckx and Min Suk Cha. Kinetic study of plasma assisted oxidation of H2 for an undiluted rich mixture. Combustion and Flame, 250:112638, April 2023.
  • [34] Seunghwan Bang, Ramses Snoeckx, and Min Suk Cha. Kinetic Study for Plasma Assisted Cracking of NH3: Approaches and Challenges. The Journal of Physical Chemistry A, 127(5):1271–1282, February 2023. Publisher: American Chemical Society.
  • [35] Seunghwan Bang, Ramses Snoeckx, and Min Suk Cha. Temperature-Dependent Kinetics of Ozone Production in Oxygen Discharges. Plasma Chemistry and Plasma Processing, August 2023.
  • [36] Mohammad Shahsavari, Alexander A. Konnov, Agustin Valera-Medina, and Mehdi Jangi. On nanosecond plasma-assisted ammonia combustion: Effects of pulse and mixture properties. Combustion and Flame, 245:112368, November 2022.
  • [37] Colin A. Pavan and Carmen Guerra-Garcia. Modeling Flame Speed Modification by Nanosecond Pulsed Discharges to Inform Experimental Design. In AIAA SCITECH 2023 Forum, AIAA SciTech Forum. American Institute of Aeronautics and Astronautics, January 2023.
  • [38] Yifan Qiu, Yifei Zhu, Yun Wu, Ningqiu Zhao, Zhenyang Li, Mai Hao, Boya Zhang, and Di Pan. Numerical investigation of the hybrid pulse–DC plasma assisted ignition and NOx emission of NH3/N2/O2 mixture. Combustion and Flame, 258:113078, December 2023.
  • [39] Ziying Xin, Zhencao Zheng, Yong Hu, Ao Sun, Feiyang Zhao, and Wenbin Yu. Numerical modeling of plasma assisted ignition of CH4/O2/He mixture by the nanosecond repetitive pulsed surface dielectric barrier discharge. Fuel, 357:129975, February 2024.
  • [40] Roger C. Millikan and Donald R. White. Systematics of Vibrational Relaxation. The Journal of Chemical Physics, 39(12):3209–3213, December 1963.
  • [41] S. J. Colgan and B. P. Levitt. Vibrational relaxation of nitrogen by various collision partners at 2500°K. Transactions of the Faraday Society, 63(0):2898–2905, 1967. Publisher: Royal Society of Chemistry.
  • [42] N. A. Popov. Fast gas heating in a nitrogen–oxygen discharge plasma: I. Kinetic mechanism. Journal of Physics D: Applied Physics, 44(28):285201, June 2011.
  • [43] A. Flitti and S. Pancheshnyi. Gas heating in fast pulsed discharges in N2–O2 mixtures. The European Physical Journal Applied Physics, 45(2):21001, February 2009. Number: 2 Publisher: EDP Sciences.
  • [44] L. L. Alves, A. Bogaerts, V. Guerra, and M. M. Turner. Foundations of modelling of nonequilibrium low-temperature plasmas. Plasma Sources Science and Technology, 27(2):023002, February 2018. Publisher: IOP Publishing.
  • [45] Da A. Xu, Deanna A. Lacoste, and Christophe O. Laux. Schlieren Imaging of Shock-Wave Formation Induced by Ultrafast Heating of a Nanosecond Repetitively Pulsed Discharge in Air. IEEE Transactions on Plasma Science, 42(10):2350–2351, October 2014. Conference Name: IEEE Transactions on Plasma Science.
  • [46] Fabien Tholin and Anne Bourdon. Simulation of the hydrodynamic expansion following a nanosecond pulsed spark discharge in air at atmospheric pressure. Journal of Physics D: Applied Physics, 46(36):365205, August 2013. Publisher: IOP Publishing.
  • [47] J. Stephens. A multi-term Boltzmann equation benchmark of electron-argon cross-sections for use in low temperature plasma models. Journal of Physics D: Applied Physics, 51(12):125203, March 2018. Publisher: IOP Publishing.
  • [48] GPT-4.
  • [49] S. Pancheshnyi, S. Biagi, M. Bordage, G. J. M. Hagelaar, W. Morgan, Arthur Phelps, and L. Pitchford. The LXCat project: Electron scattering cross sections and swarm parameters for low temperature plasma modeling. Chemical Physics, 398:148–153, April 2012.
  • [50] muparser - fast math parser library.
  • [51] Alan C. Hindmarsh, Peter N. Brown, Keith E. Grant, Steven L. Lee, Radu Serban, Dan E. Shumaker, and Carol S. Woodward. SUNDIALS: Suite of nonlinear and differential/algebraic equation solvers. ACM Transactions on Mathematical Software, 31(3):363–396, September 2005.
  • [52] David J. Gardner, Daniel R. Reynolds, Carol S. Woodward, and Cody J. Balos. Enabling New Flexibility in the SUNDIALS Suite of Nonlinear and Differential/Algebraic Equation Solvers. ACM Transactions on Mathematical Software, 48(3):31:1–31:24, September 2022.
  • [53] D. L. Rusterholtz, D. A. Lacoste, G. D. Stancu, D. Z. Pai, and C. O. Laux. Ultrafast heating and oxygen dissociation in atmospheric pressure air by nanosecond repetitively pulsed discharges. Journal of Physics D: Applied Physics, 46(46):464010, October 2013. Publisher: IOP Publishing.
  • [54] A. Montello, Z. Yin, D. Burnette, I. V. Adamovich, and W. R. Lempert. Picosecond CARS measurements of nitrogen vibrational loading and rotational/translational temperature in non-equilibrium discharges. Journal of Physics D: Applied Physics, 46(46):464002, October 2013. Publisher: IOP Publishing.
  • [55] Xiao Shao, Narjisse Kabbaj, Deanna A. Lacoste, and Hong G. Im. A computational study of a laminar methane–air flame assisted by nanosecond repetitively pulsed discharges. Journal of Physics D: Applied Physics, 57(20):205201, February 2024. Publisher: IOP Publishing.
  • [56] H. G. Weller, G. Tabor, H. Jasak, and C. Fureby. A tensorial approach to computational continuum mechanics using object-oriented techniques. Computer in Physics, 12(6):620–631, November 1998.