Hostname: page-component-745bb68f8f-s22k5 Total loading time: 0 Render date: 2025-02-12T01:58:04.713Z Has data issue: false hasContentIssue false

Theoretical and numerical study of the binary scaling law for electron distribution in thermochemical non-equilibrium flows under extremely high Mach number

Published online by Cambridge University Press:  05 April 2022

You Wu
Affiliation:
School of Astronautics, Beihang University, Beijing102206, PR China
Xu Xu*
Affiliation:
School of Astronautics, Beihang University, Beijing102206, PR China
Bing Chen
Affiliation:
School of Astronautics, Beihang University, Beijing102206, PR China
Qingchun Yang
Affiliation:
School of Astronautics, Beihang University, Beijing102206, PR China
*
Email address for correspondence: xuxu_1521@126.com

Abstract

The binary scaling law is a classical similarity law used in analysing hypersonic flow fields. The objective of this study is to investigate the applicability of the binary scaling law in thermochemical non-equilibrium airflow. Dimensional analysis of vibrational and electron–electronic energy conservation equations was employed to explore the theoretical reasons for the failure of the binary scaling law. Numerical simulation based on a multi-temperature model (translational–rotational temperature T, electron–electronic excitation temperature ${T_e}$ and the vibrational temperatures of ${\textrm{O}_2}$ and ${\textrm{N}_2}$, $\; {T_{{v_{{\textrm{O}_2}}}}}$and ${T_{{v_{{\textrm{N}_2}}}}}$) with two chemical models (the Gupta model and the Park model) was adopted to study the accuracy of the binary scaling law for electron distribution at high altitude with extremely high Mach number. The results of theoretical analysis indicate that the three-body collision reactions and the translation–electron energy exchange from collisions between electrons and ions, ${Q_{t - e\_ions}}$, can cause the failure of the binary scaling law. The results of numerical simulation show that the electron-impact ionization reactions are the main reasons for the invalidation of the binary scaling law for electron distribution at high altitude with high Mach number. With an increase of free-stream Mach number, the negative effect on the binary scaling law caused by ${Q_{t - e\_ions}}$ cannot be ignored.

Type
JFM Papers
Copyright
© The Author(s), 2022. Published by Cambridge University Press

1. Introduction

Small-scale models are usually used in hypersonic testing because the complete replication of real flight conditions is not always possible. However, scale effect exists in hypersonic non-equilibrium flow and brings difficulty to experimental simulation. Similarity law plays an essential role in scaled model tests, which can ensure that the experimental results are similar to the data in actual hypersonic flight. The binary scaling law is a classical similarity law for hypersonic and non-equilibrium flow. The binary scaling law states (Birkhoff Reference Birkhoff1960; Birkhoff & Eckerman Reference Birkhoff and Eckerman1963; Shen Reference Shen2005) that under the same inflow static temperature ${T_\infty }$ and Mach number $M{a_\infty }$, the flow fields around two scaled models are similar if the product of the incoming flow density and the characteristic size of the object ${\rho _\infty }L$ is equal. The binary scaling law provides enormous flexibility for high-altitude flight simulation. For instance, nozzle expansion used in creating hypersonic incoming flow in some wind tunnels can result in pressure lower than that in real flight. In order to ameliorate this problem, binary scaling can be employed to compensate for lack of pressure by using a model size larger than full scale (Hall, Eschenroeder & Marrone Reference Hall, Eschenroeder and Marrone1962). In addition, a reduced-scale model is usually used in order to reduce costs and pressure higher than that in flight can be employed to compensate for reduction in size according to the binary scaling law.

The binary scaling law was first proposed by Birkhoff (Reference Birkhoff1960) and was called the binary collision modelling law. Two-body binary collision processes such as dissociation are characterized by ${\rho _\infty }L$ scaling, so the ${\rho _\infty }L$ scaling is termed binary scaling (Birkhoff Reference Birkhoff1960; Hornung Reference Hornung1972; Candler Reference Candler2019). Three-body processes are characterized by $\rho _\infty ^2L$ scaling, so the fundamental reason for the failure of the binary scaling law is that three-body processes cannot be ignored (Hall et al. Reference Hall, Eschenroeder and Marrone1962; Gibson & Marrone Reference Gibson and Marrone1964). Hall et al. (Reference Hall, Eschenroeder and Marrone1962) demonstrated the practicality of the binary scaling law for a range of altitude and showed its application in hypersonic blunt-nose flows. Chemical kinetics are dominated by two-body dissociation reactions in some regimes while three-body collision processes are relatively unimportant (Hall et al. Reference Hall, Eschenroeder and Marrone1962; Hornung Reference Hornung1988), which is the reason why the binary scaling law is well suited for high-altitude airflow with coupled chemical non-equilibrium. Ellington (Reference Ellington1967) examined the limitations of binary scaling for hypersonic flight and suggested that there exists a class of low-density hypersonic flows for which the binary scaling law is still applicable. Lee & Baker (Reference Lee and Baker1969) identified the controlling parameter for non-equilibrium air dissociation and ionization using a simplified set of boundary-layer equations and demonstrated binary scaling by comparing with axisymmetric solutions. To determine the peak electron density in the boundary layer, they provided a simple engineering means by which air ionization solutions for the axisymmetric case can be scaled to that for the angle-of-attack case. The validity of the binary scaling law was also demonstrated through analysis of a cone with constant free-stream velocity at three different altitudes (Thoenes Reference Thoenes1969). Zhang (Reference Zhang1990) derived the law of similarity for high-temperature real air through the analysis of the Navier–Stokes equations and explored the scaling parameters for frozen, equilibrium and non-equilibrium flows. Zeng et al. (Reference Zeng, Lin, Liu and Qu2009) considered that the characteristic time of a dissociation reaction (two-body collision) is inversely proportional to free-stream density $({\tau _D} \propto 1/{\rho _\infty })$ and the characteristic time of a recombination reaction (three-body collision) is inversely proportional to the square of free-stream density $({\tau _R} \propto 1/\rho _\infty ^2)$. When the incoming flow density ${\rho _\infty }$ decreases to a certain degree, ${\tau _R}$ is far greater than ${\tau _D}$. If the characteristic time of flow ${\tau _F}$ reaches the same order of magnitude as ${\tau _D}$, the three-body recombination reaction is frozen relative to the flow. Zeng (Reference Zeng2007) also proved that scale effect does not exist in inviscid frozen airflow. Therefore, the similarity condition in this circumstance is the similarity condition for a dissociation reaction, ${\rho _\infty }L$. Candler (Reference Candler2019) reviewed the non-dimensionalization and scaling used in hypersonic flow and compared the scaling of gas–surface and gas-phase reactions.

Accurate prediction of peak heat flux is crucial to the development of hypersonic vehicles (Sivolella Reference Sivolella2014; Zhu et al. Reference Zhu, Lee, Chen, Wu, Chen and Gad-el-Hak2018). The applicability of the binary scaling law in predicting heat flux was verified by wind tunnel test data of the ELECTRE vehicle under $M{a_\infty } \approx 13$ (Muylaert et al. Reference Muylaert, Walpot, Häuser, Sagnier, Devezeaux, Papirnyk and Lourme1992). Camm et al. (Reference Camm, Kivel, Taylor and Teare1961) discussed the scaling of non-equilibrium radiation and suggested that correct scaling of radiative heat transfer requires electronic non-equilibrium to be governed by two-body collision processes. Teare, Georgiev & Allen (Reference Teare, Georgiev and Allen1962) also used binary scaling to correlate the non-equilibrium radiation from a normal shock and pointed out that radiation loss can invalidate the binary scaling concept.

The distribution and number density of electrons around hypersonic vehicles are closely related to the ‘blackout’ phenomenon and extra attention is paid to these conditions in entry and re-entry missions (Farbar, Boyd & Martin Reference Farbar, Boyd and Martin2013; Ramjatan et al. Reference Ramjatan, Lani, Boccelli, Hove, Karatekin, Magin and Thoemel2020). Gibson & Marrone (Reference Gibson and Marrone1964) investigated the scaling of electron and nitric oxide concentrations using Newtonian theory and the problem of binary scaling for bluff-body flows was reduced to the simpler one of binary scaling of normal shocks. They also gave a summary of key points concerned with the range of validity of binary scaling for blunt-body flows. The prediction of peak electron number density and distribution using the binary scaling law was proved to be feasible for the ELECTRE vehicle and its scaled model under a free-stream velocity of 5530 m s−1 (Zeng et al. Reference Zeng, Lin, Liu and Qu2009). Zeng et al. Reference Zeng, Lin, Liu and Qu2009 proposed that the binary scaling law can be applied to all flow parameters if the incoming flow speed is in the range of 7000–11 000 m s−1 and ${\rho _\infty }L$ is less than $3 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$.

Although many scholars have undertaken research on the binary scaling law through theoretical analysis and numerical and experimental simulation, there are still some deficiencies in these works. Firstly, most theoretical analyses of the binary scaling law are limited to either kinetic theory of gases with only binary collisions considered or the mass, momentum and total energy conservation equations of the Navier–Stokes equations (Birkhoff Reference Birkhoff1960; Zhang Reference Zhang1990; Zeng Reference Zeng2007). Most previous studies did not consider thermal non-equilibrium with vibrational energy conservation equations and electron–electronic energy conservation equations. For example, translational, rotational, vibrational and electronic degrees of freedom were assumed to be fully equilibrated in the analysis of Gibson & Marrone (Reference Gibson and Marrone1964). This means that only the effects of chemical reactions (chemical non-equilibrium) on the binary scaling law were taken into account and the effects of complex energy exchange and relaxation between different energy modes (thermal non-equilibrium) were ignored. However, the effects of thermal non-equilibrium play a crucial role in hypersonic flow at high altitude. Therefore, it calls into question if there exist any other factors besides chemical three-body collisions, perhaps from energy exchange, that could theoretically cause binary scaling to fail. If such factors do exist, the extent of their contribution to the failure of the similarity law in practice must also be questioned. Secondly, although Houwing et al. (Reference Houwing, Nonaka, Mizuno and Takayama2000) explored the effect of vibrational non-equilibrium and modified the theoretical analysis predicting shock standoff distance, chemical dissociation and ionization effects were negligible in that study. Spacecraft enter the atmosphere at extremely high Mach number (Ma = 30–40) and ionization reactions are violent at extraordinarily high gas temperatures. This begs the question as to whether the binary scaling law is applicable to electron distribution under such conditions. Of all the elementary reactions in air chemistry, those which cause the binary scaling law to fail for electron distribution have yet to be identified. All of these issues mentioned above are discussed in this article.

This paper is organized as follows. Section 2 describes the physical models used in this work. Section 3 is devoted to the theoretical analysis of the binary scaling law in thermochemical non-equilibrium flows. Numerical analysis including numerical method and detailed results is presented in § 4. Conclusions are given in § 5.

2. Physical models

2.1. Fluid governing equations

Multi-species hypersonic reacting flows are governed by the Navier–Stokes equations with thermochemical non-equilibrium processes described by a multi-temperature model.

The continuity equation consists of a total mass equation and $(ns - 1)$ species equations:

(2.1)\begin{gather}\frac{{\partial \rho }}{{\partial t}} + \boldsymbol{\nabla }\boldsymbol{\cdot }(\rho \boldsymbol{V}) = 0,\end{gather}
(2.2)\begin{gather}\frac{{\partial \rho {y_s}}}{{\partial t}} + \boldsymbol{\nabla }\boldsymbol{\cdot }(\rho {y_s}\boldsymbol{V}) = \boldsymbol{\nabla }\boldsymbol{\cdot }(\rho {D_s}\boldsymbol{\nabla }{y_s}) + {\omega _s},\quad s = 1,2, \ldots ,ns - 1,\end{gather}

where the index $ns$ represents the total number of species, $\; t$ is the time, $\rho $ is the mixture density and ${y_s}$ is the mass fraction of species s. Parameter $\; \boldsymbol{V}$ denotes the velocity and ${\omega _s}$ represents the source term of production of species s. Coefficient ${D_s}$ is the effective diffusion coefficient of species s.

The momentum conservation equation is

(2.3)\begin{equation}\frac{{\partial \rho \boldsymbol{V}}}{{\partial t}} + \boldsymbol{\nabla }\boldsymbol{\cdot }(\rho \boldsymbol{VV}) + \boldsymbol{\nabla }p = \boldsymbol{\nabla }\boldsymbol{\cdot }\bar{\boldsymbol{\tau}},\end{equation}

where p is the pressure of the mixture and $\bar{\boldsymbol{\tau}}$ denotes the viscous stress tensor (Garicano-Mena, Lani & Degrez Reference Garicano-Mena, Lani and Degrez2018).

The total energy conservation equation is

(2.4)\begin{align}\frac{{\partial \rho E}}{{\partial t}} + \boldsymbol{\nabla }\boldsymbol{\cdot }(\boldsymbol{V}\rho H) = \boldsymbol{\nabla }\boldsymbol{\cdot }\left( {{k_{tr}}\boldsymbol{\nabla }T + \sum\limits_{m = 1}^{nmol} {{k_{{v_m}}}\boldsymbol{\nabla }{T_{{v_m}}}} + {k_e}\boldsymbol{\nabla }{T_e} + \bar{\boldsymbol{\tau}}\boldsymbol{\cdot }\boldsymbol{V} + \sum\limits_{s = 1}^{ns} {{H_s}\rho {D_s}\boldsymbol{\nabla }{y_s}} } \right),\end{align}

where E and H represent the total energy and total enthalpy per unit mass of mixture, respectively, and ${H_s}$ is enthalpy per unit mass of species s. Index $nmol$ denotes the total number of molecular species. Parameters ${k_{tr}}$, ${k_{{v_m}}}$ and ${k_e}$ represent translational–rotational thermal conductivity of mixture, vibrational thermal conductivity of molecular species m and electron–electronic thermal conductivity of mixture, respectively. It is assumed that the translational temperature is equal to the rotational temperature and T denotes translational–rotational temperature. Here ${T_{{v_m}}}$ is vibrational temperature of molecular species m. In this study, the vibrational temperature of ${\textrm{O}_2}$, ${T_{{v_{{\textrm{O}_2}}}}}$, and the vibrational temperature of ${\textrm{N}_2}$, ${T_{{v_{{\textrm{N}_2}}}}}$, are considered. The electronic–free electron modes are described by electron–electronic excitation temperature ${T_e}$.

The vibrational energy conservation equation for a non-equilibrium molecule of species m is

(2.5)\begin{equation}\frac{{\partial \rho {y_m}{E_{{v_m}}}}}{{\partial t}} + \boldsymbol{\nabla }\boldsymbol{\cdot }(\boldsymbol{V}\rho {y_m}{E_{{v_m}}}) = \boldsymbol{\nabla }\boldsymbol{\cdot }({k_{{v_m}}}\boldsymbol{\nabla }{T_{{v_m}}} + {E_{{v_m}}}\rho {D_m}\boldsymbol{\nabla }{y_m}) + {\omega _{{v_m}}},\end{equation}

where ${E_{{v_m}}}$ is vibrational energy per unit mass of molecular species $m$ and ${\omega _{{v_m}}}$ represents the source term of vibrational energy of species m.

The electron–electronic energy conservation equation is

(2.6)\begin{equation}\frac{{\partial \rho {E_e}}}{{\partial t}} + \boldsymbol{\nabla }\boldsymbol{\cdot }(\boldsymbol{V}\rho {E_e}) = \boldsymbol{\nabla }\boldsymbol{\cdot }\left( {{k_e}\boldsymbol{\nabla }{T_e} + \sum\limits_{s = 1}^{ns} {{H_{{e_s}}}\rho {D_s}\boldsymbol{\nabla }{y_s}} } \right) + {\omega _e},\end{equation}

where ${E_e}$ and ${H_{{e_s}}}$ denote electron–electronic energy per unit mass of mixture and electron–electronic enthalpy per unit mass of species s, respectively, and $\; {\omega _e}$ represents the source term of electron–electronic energy of mixture.

2.2. Thermodynamic relations

The pressure of mixture is the sum of the partial pressures of each species:

(2.7)\begin{equation}p = \sum\limits_{s = 1}^{ns} {{p_s}} = \sum\limits_{s \ne {\textrm{e}^ - }} {\frac{{{\rho _s}{R_u}T}}{{{M_s}}} + \frac{{{\rho _e}{R_u}{T_e}}}{{{M_e}}}} ,\end{equation}

where ${p_s}$ and ${M_s}$ are the partial pressure and molecular weight of species $s$, respectively, ${M_e}$ is the electron molecular weight and ${R_u}$ is the universal gas constant.

The internal energy per unit mass of species $s$, i.e.

(2.8)\begin{equation}{E_s} = {E_{t{r_s}}} + {E_{ro{t_s}}} + {E_{{v_s}}} + {E_{{e_s}}} + \Delta {h_s},\end{equation}

is split between the translational, rotational, vibrational and electron–electronic contributions and the latent chemical energy of species s.

The translational energy and rotational energy per unit mass of species s are given by

(2.9a,b)\begin{equation}{E_{t{r_s}}} = \left\{ {\begin{array}{@{}ll} {\dfrac{{3{R_u}T}}{{2{M_s}}}}&{s \ne {\textrm{e}^ - }}\\ 0&{s = {\textrm{e}^ - }} \end{array}} \right.\quad \textrm{and}\quad {E_{ro{t_s}}} = \left\{ {\begin{array}{@{}ll} {\dfrac{{{R_u}T}}{{{M_s}}}}&{s = dia}\\ 0&{s \ne dia,} \end{array}} \right.\end{equation}

where index $dia$ represents diatomic molecule.

The vibrational energy and electron–electronic energy per unit mass of species s are

(2.10a,b)\begin{align}{E_{{v_s}}} = \left\{ {\begin{array}{@{}ll} {\dfrac{{{R_u}{\theta_{{v_s}}}/{M_s}}}{{{\textrm{e}^{{\theta_{{v_s}}}/{T_{{v_s}}}}} - 1}}}&{s = dia}\\ 0&{s \ne dia} \end{array}} \right.\quad \textrm{and}\quad {E_{{e_s}}} = \left\{ {\begin{array}{@{}ll} {\dfrac{{\left( {\dfrac{{{g_{{1_s}}}}}{{{g_{{0_s}}}}}} \right)\,{\textrm{e}^{ - {\theta_{{e_s}}}/{T_e}}}}}{{1 + \left( {\dfrac{{{g_{{1_s}}}}}{{{g_{{0_s}}}}}} \right)\,{\textrm{e}^{ - {\theta_{{e_s}}}/{T_e}}}}}\dfrac{{{R_u}{\theta_{{e_s}}}}}{{{M_s}}}}&{s \ne {\textrm{e}^ - }}\\ {\dfrac{{3{R_u}{T_e}}}{{2{M_e}}}}&{s = {\textrm{e}^ - },} \end{array}} \right.\end{align}

where ${\theta _{{v_s}}}$ and ${\theta _{{e_s}}}$ denote the characteristic vibrational temperature and the characteristic electronic temperature of species s, respectively, and ${g_{{0_s}}}$ and ${g_{{1_s}}}$ are the degeneracy of the ground and the first electronic level of species s, respectively.

2.3. Chemistry model

This study selects the Gupta chemical reaction model (Gupta et al. Reference Gupta, Yos, Thompson and Lee1990) and the Park model (Park Reference Park1993) which are applied to 11 air species $\textrm{(}{\textrm{O}_2},{\textrm{N}_2},\,\textrm{N},\,\textrm{O},\textrm{ NO},\, {\textrm{N}^ + },\, {\textrm{O}^ + }, \textrm{N}_2^ + ,\; \; \textrm{O}_2^ + ,\textrm{ N}{\textrm{O}^ + },\, {\textrm{e}^ - })$. The mass production rate ${\omega _s}$ is calculated by

(2.11)\begin{equation}{\omega _s} = {M_s}\sum\limits_{z = 1}^{nr} {(\nu _{z,s}^b - \nu _{z,s}^f)R_z^{fb}} = {M_s}\sum\limits_{z = 1}^{nr} {(\nu _{z,s}^b - \nu _{z,s}^f)(R_z^f - R_z^b)} ,\end{equation}

where $\nu _{z,s}^f$ and $\nu _{z,s}^b$ are the stoichiometric coefficients of species s for the reactants and products in reaction z, respectively, and $R_z^{fb}$ denotes the total reaction rate of reaction z. The forward and backward reaction rates can be written as

(2.12a,b)\begin{equation}R_z^f = {\varGamma _z}k_z^f\prod\limits_{l = 1}^{ns} {{{\left( {\frac{{{\rho_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^f}}} ,\quad R_z^b = {\varGamma _z}k_z^b\prod\limits_{l = 1}^{ns} {{{\left( {\frac{{{\rho_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^b}}} ,\end{equation}

where ${\varGamma _z}$ is given as follows:

(2.13)\begin{equation}{\varGamma _z} = {\left[ {\sum\limits_{l = 1}^{ns} {{\alpha_{z,l}}} \left( {\frac{{{\rho_l}}}{{{M_l}}}} \right)} \right]^{{L_z}}},\end{equation}

where ${\alpha _{z,l}}$ denotes the third-body efficiency for species l in reaction z. Here ${L_z}$ equals 1 when a third body M is needed in reaction z. Otherwise, ${L_z}$ equals 0.

When the Gupta model is employed, the forward and backward reaction rate coefficients, $k_z^f$ and $k_z^b$, are expressed in modified Arrhenius form as

(2.14a,b)\begin{equation}k_z^f = {A_{f,z}}T_f^{{B_{f,z}}}\,{\textrm{e}^{ - {C_{f,z}}/{T_f}}},\quad k_z^b = {A_{b,z}}T_b^{{B_{b,z}}}\,{\textrm{e}^{ - {C_{b,z}}/{T_b}}},\end{equation}

where ${T_f}$ and ${T_b}$ represent the controlling temperature of the forward and backward reactions, respectively. When the Park model is applied, $k_z^f$ is still calculated by (2.14a,b) and $k_z^b$ is obtained by the equilibrium constant (Park Reference Park1985):

(2.15)\begin{equation}k_z^b = \frac{{k_z^f({T_b})}}{{{K_{eq,z}}({T_b})}},\end{equation}

where the equilibrium constant ${K_{eq,z}}$ is given as the following curve-fitting expression (Park Reference Park1990):

(2.16)\begin{equation}{K_{eq,z}}({T_b}) = exp \left( {\frac{{A_1^z}}{S} + A_2^z + A_3^z\ln S + A_4^zS + A_5^z{S^2}} \right),\end{equation}

where $S = 10\;000/{T_b}$ and the constants $A_1^z,A_2^z,A_3^z,A_4^z,A_5^z$ are given by Park (Reference Park1990). Those five constants for ${\textrm{N}_2} + {\textrm{N}^ + }\rightleftharpoons \textrm{N} + \textrm{N}_2^ + $ are not included in Park (Reference Park1990), so the equilibrium constant ${K_{eq,z}}$ for ${\textrm{N}_2} + {\textrm{N}^ + }\rightleftharpoons \textrm{N} + \textrm{N}_2^ + $ is calculated using the similar method proposed by Park (Reference Park1985) and Ghezali, Haoui & Chpoun (Reference Ghezali, Haoui and Chpoun2019). Radiative recombination reactions in the Park model are not considered, because radiation is ignored in this paper.

2.4. Energy exchange model

The vibrational energy source term of molecular species m can be decomposed as

(2.17)\begin{equation}{\omega _{{v_m}}} = {Q_{t - {v_m}}} + {Q_{v - {v_m}}} + {Q_{e - {v_m}}} + {Q_{r - {v_m}}},\end{equation}

where ${Q_{t - {v_m}}}$ denotes the energy exchange between translation and vibration energies. Energy exchange ${Q_{t - {v_m}}}$ is evaluated using Landau–Teller theory (Landau & Teller Reference Landau and Teller1936) as

(2.18)\begin{equation}{Q_{t - {v_m}}} = \rho {y_m}\frac{{{E_{{v_m}}}(T) - {E_{{v_m}}}({T_{{v_m}}})}}{{{\tau _{vt,m}}}},\end{equation}

and the relaxation time of vibration–translation exchanges ${\tau _{vt,m}}$ modelled by the Millikan–White empirical formula (Millikan & White Reference Millikan and White1963; Miller et al. Reference Miller, Tannehill, Lawrence and Edwards1995) is expressed as

(2.19)\begin{gather}\tau _{vt,m}^{MW} = {{\sum\limits_{s \ne {\textrm{e}^ - }} {{X_s}} } / {\sum\limits_{s \ne {\textrm{e}^ - }} {\frac{{{X_s}}}{{{\tau _{vt,m,s}}}}} }},\end{gather}
(2.20)\begin{gather}{\tau _{vt,m,s}} = \frac{{101\;325}}{p}exp [{a_{m,s}}({T^{ - 1/3}} - {b_{m,s}}) - 18.42],\end{gather}

where ${X_s}$ is the mole fraction of heavy particle s. Here ${a_{m,s}}$ and ${b_{m,s}}$ are given by

(2.21)\begin{gather}{a_{m,s}} = 0.00116\mu _{m,s}^{1/2}\theta _{{v_m}}^{4/3},\end{gather}
(2.22)\begin{gather}{b_{m,s}} = 0.015\mu _{m,s}^{1/4},\end{gather}

where the reduced mass ${\mu _{m,s}}$ is defined as

(2.23)\begin{equation}{\mu _{m,s}} = \frac{{1000{M_m}{M_s}}}{{{M_m} + {M_s}}}.\end{equation}

When the temperature is higher than 8000 K, Park (Reference Park1990) suggests that the correction term should be added on the basis of $\tau _{vt,m}^{MW}$ and the total relaxation time ${\tau _{vt,m}}$ can be expressed as

(2.24)\begin{equation}{\tau _{vt,m}} = \tau _{vt,m}^{MW} + {({\sigma _m}{\bar{C}_m}{n_m})^{ - 1}},\end{equation}

where the average molecular velocity ${\bar{C}_m}$, the effective cross-section ${\sigma _m}$ and number density ${n_m}$ of species m are given by

(2.25)\begin{gather}{\bar{C}_m} = {\left( {\frac{{8{R_u}T}}{{{\rm \pi} {M_m}}}} \right)^{1/2}},\end{gather}
(2.26)\begin{gather}{\sigma _m} = {10^{ - 21}}{\left( {\frac{{50\;000}}{T}} \right)^2},\end{gather}
(2.27)\begin{gather}{n_m} = \frac{{\rho {y_m}{N_0}}}{{{M_m}}},\end{gather}

where ${N_0}$ is Avogadro's constant. Time ${\tau _{vt,m}}$ can be expressed as

(2.28)\begin{equation}{\tau _{vt,m}} = {f_{{\tau _{vt,m}}}}(\rho ,T,{y_1}, \ldots ,{y_{ns}}) = \frac{1}{\rho }{f^{\prime}_{{\tau _{vt,m}}}}(T,{T_e},{y_1}, \ldots ,{y_{ns}}),\end{equation}

where ${f_{{\tau _{vt,m}}}}$ stand for the functions by which ${\tau _{vt,m}}$ can be calculated. The term ${f^{\prime}_{{\tau _{vt,m}}}}$ does not involve the density $\rho $. Thus, the term ${Q_{t - {v_m}}}$ can be written as

(2.29)\begin{equation}{Q_{t - {v_m}}} = {\rho ^2}{f^{\prime}_{{Q_{t - {v_m}}}}}(T,{T_{{v_m}}},{T_e},{y_1}, \ldots ,{y_{ns}}),\end{equation}

where ${f^{\prime}_{{Q_{t - {v_m}}}}}$ is independent of $\rho $.

The vibrational energy exchange between the different molecules ${Q_{v - {v_m}}}$ is described according to Roberts (Reference Roberts1994):

(2.30)\begin{align}{Q_{v - {v_m}}} &= \frac{{\rho {y_m}}}{{{M_m}}}\mathop \sum \limits_{s = 1,s \ne m}^{mol} \left\{ \frac{{{X_s}}}{{{\tau_{vv,m,s}}}}\left[ {\frac{{1 - exp \left( { - \frac{{{\theta_{{v_m}}}}}{T}} \right)}}{{1 - exp \left( { - \frac{{{\theta_{{v_s}}}}}{T}} \right)}}} \right]\frac{{{e_{{v_s}}}({T_{{v_s}}})}}{{{e_{{v_s}}}(T)}}[{e_{{v_m}}}(T) - {e_{{v_m}}}({T_{{v_m}}})]\right.\nonumber\\ &\quad - \left.\frac{{{X_s}}}{{{\tau_{vv,m,s}}}}\frac{{{e_{{v_m}}}({T_{{v_m}}})}}{{{e_{{v_s}}}(T)}}[{e_{{v_s}}}(T) - {e_{{v_s}}}({T_{{v_s}}})] \vphantom{\left\{ \frac{{{X_s}}}{{{\tau_{vv,m,s}}}}\left[ {\frac{{1 - exp \left( { - \frac{{{\theta_{{v_m}}}}}{T}} \right)}}{{1 - exp \left( { - \frac{{{\theta_{{v_s}}}}}{T}} \right)}}} \right]\frac{{{e_{{v_s}}}({T_{{v_s}}})}}{{{e_{{v_s}}}(T)}}[{e_{{v_m}}}(T) - {e_{{v_m}}}({T_{{v_m}}})]\right.}\right\},\end{align}

where ${e_{{v_m}}}$ denotes the vibrational energy per mole of species m. Note that species m can only be ${\textrm{N}_2}$ in (2.30). For other species, the transfer of vibrational energy can be obtained according to the assumption of conservation of vibrational energy (Miller et al. Reference Miller, Tannehill, Lawrence and Edwards1995). The relaxation time of vibration–vibration exchanges ${\tau _{vv,m,s}}$ can be written as

(2.31)\begin{equation}{\tau _{vv,m,s}} = \frac{1}{{{P_{m,s}}{{(8{\rm \pi} KT{N_0}/{\mu _{m,s}})}^{1/2}}\sigma _{m,s}^2{n_s}}},\end{equation}

where K and ${\sigma _{m,s}}$ represent Boltzmann's constant and the collision cross-sections (Miller et al. Reference Miller, Tannehill, Lawrence and Edwards1995), respectively. The probability of energy transfer during each collision ${P_{m,s}}$ is provided by Park (Reference Park1993) as

(2.32)\begin{gather}{P_{{\textrm{N}_2},\textrm{NO}}} = 5.5 \times {10^{ - 5}} {\left( {\frac{T}{{1000}}} \right)^{2.32}},\end{gather}
(2.33)\begin{gather}{P_{{\textrm{N}_2},{\textrm{O}_2}}} = 3.0 \times {10^{ - 6}} {\left( {\frac{T}{{1000}}} \right)^{2.87}}.\end{gather}

In this paper, only collision between ${\textrm{N}_2}$ and ${\textrm{O}_2}$ is taken into consideration. Like ${Q_{t - {v_m}}}$, the term ${Q_{v - {v_m}}}$ can be expressed as

(2.34)\begin{equation}{Q_{v - {v_m}}} = {\rho ^2}{f^{\prime}_{{Q_{v - {v_m}}}}}(T,{T_{{v_m}}},{y_m},{y_{s1}}, \ldots ,{y_{smol}}),\end{equation}

where ${y_{s1}}, \ldots ,{y_{smol}}$ represent the molecular species except species m.

The energy exchange between vibrational and electronic modes ${Q_{e - {v_m}}}$ is modelled as (Candler & Maccormack Reference Candler and Maccormack1991)

(2.35)\begin{equation}{Q_{e - {v_m}}} = \rho {y_m}\frac{{{E_{{v_m}}}({T_e}) - {E_{{v_m}}}({T_{{v_m}}})}}{{{\tau _{e,m}}}},\end{equation}

where ${\tau _{e,m}}$ denotes the relaxation time of electronic–vibrational exchanges for species m. Parameter $\; {\tau _{e,{\textrm{N}_2}}}$ is given based on research of Lee (Reference Lee1993) and Candler (Reference Candler1988) by

(2.36)\begin{gather}{\log _{10}}\left( {\frac{{{p_e}{\tau_{e,{\textrm{N}_2}}}}}{{101\;325}}} \right) = 3.91\log _{10}^2({T_e}) - 30.36{\log _{10}}({T_e}) + 48.90,\quad 1000\;\textrm{K} \le {T_e} \le 7000\;\textrm{K},\end{gather}
(2.37)\begin{gather}{\log _{10}}\left( {\frac{{{p_e}{\tau_{e,{\textrm{N}_2}}}}}{{101\;325}}} \right) = 1.30\log _{10}^2({T_e}) - 9.09{\log _{10}}({T_e}) + 5.58,\quad 7000 < {T_e} \le 50\;000\;\textrm{K},\end{gather}

where ${p_e}$ is the electron partial pressure.

Parameter ${\tau _{e,{\textrm{O}_2}}}$ is obtained from the formula of Park & Lee (Reference Park and Lee1995) based on ${\tau _{e,{\textrm{N}_2}}}$:

(2.38)\begin{equation}{\tau _{e,{\textrm{O}_2}}} = 300 \times {\tau _{e,{\textrm{N}_2}}}(1.492{T_e}).\end{equation}

The term ${Q_{e - {v_m}}}$ can also be written as

(2.39)\begin{equation}{Q_{e - {v_m}}} = {\rho ^2}{f^{\prime}_{{Q_{e - {v_m}}}}}(T,{T_e},{y_m},{y_{{\textrm{e}^ - }}}),\end{equation}

where ${f^{\prime}_{{Q_{e - {v_m}}}}}$ does not depend on $\rho $.

The change in vibrational energy due to chemical reactions ${Q_{r - {v_m}}}$ can be expressed as (Miller et al. Reference Miller, Tannehill, Lawrence and Edwards1995)

(2.40)\begin{equation}{Q_{r - {v_m}}} = {\omega _m}{E_{{v_m}}}.\end{equation}

The electron–electronic energy source term can be decomposed as

(2.41)\begin{equation}{\omega _e} = {Q_{t - e}} - \sum\limits_{m = mol} {{Q_{e - {v_m}}}} + {Q_{r - e}} + {Q_{i - e}} - {p_e}\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{V}.\end{equation}

The energy exchange between translation and electron modes ${Q_{t - e}}$ can be written as (Miller et al. Reference Miller, Tannehill, Lawrence and Edwards1995; Roberts Reference Roberts1996)

(2.42)\begin{equation}{Q_{t - e}} = 3{\rho _e}{R_u}(T - {T_e})\sum\limits_{s \ne {\textrm{e}^ - }} {\frac{{{v_{e,s}}}}{{{M_s}}}} ,\end{equation}

where ${v_{e,s}}$ denotes the collision frequency between electrons and species s. For ions and electrons, it follows Coulomb collisions and can be expressed as (Appleton & Bray Reference Appleton and Bray1964)

(2.43)\begin{equation}{\nu _{e,s}} = \frac{8}{3}{\left( {\frac{{\rm \pi} }{{{m_e}}}} \right)^{1/2}}{n_s}e_c^4\frac{1}{{{{(2K{T_e})}^{3/2}}}}\ln \left( {\frac{{{K^3}T_e^3}}{{{\rm \pi} {n_e}e_c^6}}} \right),\end{equation}

where ${m_e}$ is the electron mass, ${n_e}$ denotes number density of electrons and ${e_c}$ is the magnitude of the electronic charge. For collisions between neutral species and electrons, the collision frequency can be expressed as (Appleton & Bray Reference Appleton and Bray1964)

(2.44)\begin{equation}{\nu _{e,s}} = {n_s}{\sigma _{e,s}}{\left( {\frac{{8K{T_e}}}{{{\rm \pi} {m_e}}}} \right)^{1/2}},\end{equation}

where the effective cross-section ${\sigma _{e,s}}$ is evaluated by curve fits (Gnoffo, Gupta & Shinn Reference Gnoffo, Gupta and Shinn1989) as

(2.45)\begin{equation}{\sigma _{e,s}} = {a_s} + {b_s}{T_e} + {c_s}T_e^2.\end{equation}

The collision frequency ${v_{e,s}}$ and the effective cross-section ${\sigma _{e,s}}$ can also be calculated by similar formulae obtained by Petschek & Byron (Reference Petschek and Byron1957) or Matsuzaki (Reference Matsuzaki1988) and the simulation in § 4 uses Matsuzaki's model. Regardless of the model mentioned, the term ${Q_{t - e}}$ can be written as

(2.46)\begin{align}\begin{aligned}{Q_{t - e}} & = {Q_{t - e\_neu}} + {Q_{t - e\_ions}}\\ & = {\rho ^2}\textrm{[}f_{t - e}^{neu}(T,{T_e},{y_{{\textrm{e}^ - }}},{y_{neu1}}, \ldots ,{y_{neun}}) + f_{t - e}^{ion}(\rho ,T,{T_e},{y_{{\textrm{e}^ - }}},{y_{ion1}}, \ldots ,{y_{ionn}})\textrm{] ,} \end{aligned}\end{align}

where ${Q_{t - e\_neu}}$ denotes the translation–electron energy exchange from collisions between electrons and neutral particles and ${Q_{t - e\_ions}}$ denotes that from collisions between electrons and ions. The term $f_{t - e}^{neu}$ represents the contribution of the electron–neutral particle collisions and is not correlated with $\rho $; $f_{t - e}^{ion}$ stands for the contribution of the electron–ion collisions and depends on $\rho $.

The added or removed electron energy due to chemical reactions ${Q_{r - e}}$ can be specified as (Hao, Wang & Lee Reference Hao, Wang and Lee2016)

(2.47)\begin{equation}{Q_{r - e}} = \sum\limits_{s = 1}^{ns} {{\omega _s}{E_{{e_s}}}} .\end{equation}

The term ${Q_{i - e}}$ represents the energy loss of free electrons by the impact ionization reactions (Scalabrin Reference Scalabrin2007), which gives

(2.48)\begin{equation}{Q_{i - e}} ={-} {\textstyle{1 \over 3}}({I_\textrm{O}}{\omega _{{\textrm{O}^ + }}} + {I_\textrm{N}}{\omega _{{\textrm{N}^ + }}}),\end{equation}

where ${I_\textrm{O}}$ and ${I_\textrm{N}}$ are the first energy of ionization of $\textrm{O}$ and $\textrm{N}$; ${\omega _{{\textrm{O}^ + }}}$ and ${\omega _{{\textrm{N}^ + }}}$ are the formation rate of ${\textrm{O}^ + }$ and ${\textrm{N}^ + }$ in the electron-impact ionization reactions.

The term $- {p_e}\boldsymbol{\nabla }\boldsymbol{\cdot }\boldsymbol{V}$ denotes the contribution of electron partial pressure in electron–electronic energy change (Gnoffo et al. Reference Gnoffo, Gupta and Shinn1989; Lee Reference Lee1984).

2.5. Transport properties

The dynamic viscosity coefficient of each species is obtained by curve fits taken from Gupta et al. (Reference Gupta, Yos, Thompson and Lee1990) in the temperature range (1000–30 000 K) and Lennard-Jones potential (Anderson Reference Anderson2006) in the other temperature range. The translational–rotational, vibrational and electron–electronic thermal conductivities of each species are given by Eucken's relation (Anderson Reference Anderson2006). The dynamic viscosity $\mu $ and thermal conductivities of gas mixture are derived from Wilke's rule (Wilke Reference Wilke1950). So $\mu $, ${k_{tr}}$, ${k_{{v_m}}}$ and ${k_e}$ can be expressed as

(2.49)\begin{gather}\mu = {f_\mu }(T,{T_e},{y_1}, \ldots ,{y_{ns}}),\end{gather}
(2.50)\begin{gather}{k_{tr}} = {f_{{k_{tr}}}}(T,{y_1}, \ldots ,{y_{ns}}),\end{gather}
(2.51)\begin{gather}{k_{{v_m}}} = {f_{{k_{{v_m}}}}}(T,{T_{{v_m}}},{y_1}, \ldots ,{y_{ns}}),\end{gather}
(2.52)\begin{gather}{k_e} = {f_{{k_e}}}(T,{T_e},{y_1}, \ldots ,{y_{ns}}),\end{gather}

where ${f_\mu }$, ${f_{{k_{tr}}}}$, ${f_{{k_{{v_m}}}}}$ and ${f_{{k_e}}}$ stand for the functions by which $\mu $, ${k_{tr}}$, ${k_{{v_m}}}$ and ${k_e}$ can be calculated, respectively. One can observe that $\mu $, ${k_{tr}}$, ${k_{{v_m}}}$ and ${k_e}$ depend on temperatures and mass fraction of each species.

It is assumed that the binary diffusion coefficients between each pair of components ${D_{sj}}$ are the same and the electric current of the flow field is zero. The multi-component diffusion coefficient of each species ${D_s}$ is given as follows:

(2.53)\begin{align}{D_s} = \left\{ {\begin{array}{@{}ll} {\dfrac{{(1 - {y_s})\mu }}{{\rho (1 - {X_s})Sc}}}&{s \ne {\textrm{e}^ - },}\\ {\dfrac{{\sum\limits_{s = 1}^{nion} {{y_s}{D_s}} }}{{\sum\limits_{s = 1}^{nion} {{y_s}} }}}&{s = {\textrm{e}^ - },} \end{array}} \right.\end{align}

where $Sc$ is the Schmidt number and index $nion$ denotes the total number of ion species. For neutral particles, $Sc = 0.5$. For ions, $Sc = 0.25$. In a similar way, ${D_s}$ can also be expressed as

(2.54)\begin{equation}{D_s} = {f_{{D_s}}}(T,{T_e},\rho ,{y_1}, \ldots ,{y_{ns}}) = {f^{\prime}_{{D_s}}}(T,{T_e},{y_1}, \ldots ,{y_{ns}})/\rho ,\end{equation}

where the functions ${f_{{D_s}}}$ and ${f^{\prime}_{{D_s}}}$ show that ${D_s}$ depends on density, temperature and mass fraction of each species.

3. The binary scaling law in thermochemical non-equilibrium

The binary scaling law requires the same ${\rho _\infty }L$, ${T_\infty }$, $\; {U_\infty }$ for two different flows and the mass fractions of free streams ${y_{1,\infty }}, \ldots ,{y_{ns,\infty }}$ also need to be consistent. Flow parameters need to be transformed to the dimensionless form for similarity analysis (see Appendix A). By analysing the dimensionless continuity, momentum conservation and total energy conservation equation, the similarity law for real gas flow is given by Zhang (Reference Zhang1990). In this paper, the similarity analysis based on the dimensionless vibrational and electron–electronic energy conservation equation is discussed in detail.

Radiation and ablation might become important at large Mach number. Teare et al. (Reference Teare, Georgiev and Allen1962) considered that, in a certain range of altitude and free-stream velocity, the intensity of radiation is proportional to the particle density when two-body collision processes prevail sufficiently. If the condition of the binary scaling law $({\rho _\infty }R = \textrm{const}\textrm{.)}$ is satisfied, the integrated radiation flux emitted from the non-equilibrium region is independent of the particle density and binary scaling is applicable to radiation (Camm et al. Reference Camm, Kivel, Taylor and Teare1961; Teare et al. Reference Teare, Georgiev and Allen1962). Zeng et al. (Reference Zeng, Lin, Liu and Qu2009) investigated the radiation of re-entry flows at ${U_\infty } = 7650,\;11\;360\;\textrm{m}\;{\textrm{s}^{ - 1}}$ and the results revealed that non-equilibrium radiation abides by the binary scaling law. However, the binary scaling law is invalid for analysing radiation in some extreme cases (Teare et al. Reference Teare, Georgiev and Allen1962; Zeng et al. Reference Zeng, Lin, Liu and Qu2009). Ablation involves various gas–surface reactions. The scaling of the gas–surface reactions is not as clear as that of the gas-phase reactions (Candler Reference Candler2019). The rates of gas–surface reactions depend on the fraction of active sites in a certain phase (Marschall & MacLean Reference Marschall and Maclean2011). For example, the catalytic recombination of nitrogen $\textrm{N} + \textrm{N}(\textrm{s})\rightleftharpoons {\textrm{N}_2}$ is one of the basic gas–surface reactions. Here, $\textrm{N(s)}$ denotes a nitrogen atom that is bonded to an open surface site. The rate of this reaction depends on both the gas-phase $\textrm{N}$ atom concentration and the number of $\textrm{N}$ atoms bonded to the surface sites (Candler Reference Candler2019). Thus, both radiation and ablation might inhibit the binary scaling. However, this article aims to investigate the effects of energy exchange and air reactions on the binary scaling law. To avoid interference, radiation and ablation are not considered in this paper. Thus, the simplest wall boundary condition, i.e. an isothermal and fully catalytic wall boundary, is employed in similarity analysis.

3.1. Similarity analysis for continuity equation

The analysis based on the dimensionless continuity equation (Zhang Reference Zhang1990) is reviewed first:

(3.1)\begin{equation}\frac{{\partial \bar{\rho }{y_s}}}{{\partial \bar{t}}} + \frac{{\partial \bar{\rho }{y_s}{{\bar{u}}_j}}}{{\partial {{\bar{x}}_j}}} = \frac{1}{{R{e_\infty }}}\frac{{\partial \left( {\frac{{\bar{\rho }{{\bar{D}}_s}\partial {y_s}}}{{\partial {{\bar{x}}_j}}}} \right)}}{{\partial {{\bar{x}}_j}}} + \frac{L}{{{\rho _\infty }{U_\infty }}}{\omega _s},\end{equation}

where the overbar is used to denote non-dimensional variables. Parameter $\; {\bar{D}_s}$ can be expressed in the following form (see (2.49), (2.54) and (A4), (A5) in Appendix A):

(3.2)\begin{align}{\bar{D}_s} = \frac{{{{f^{\prime}}_{{D_s}}}(\bar{T} {T_\infty },{{\bar{T}}_e} {T_\infty },{y_1}, \ldots ,{y_{ns}})/\rho }}{{{f_\mu }( {T_\infty },{T_{e,\infty }},{y_{1,\infty }}, \ldots ,{y_{ns,\infty }})/{\rho _\infty }}} = \frac{{{{f^{\prime}}_{{D_s},{T_\infty }}}(\bar{T},{{\bar{T}}_e},{y_1}, \ldots ,{y_{ns}})}}{{{f_{\mu ,{T_\infty }}}(1,{T_{e,\infty }}/{T_\infty },{y_{1,\infty }}, \ldots ,{y_{ns,\infty }}) \bar{\rho }}},\end{align}

where the index ${T_\infty }$ represents that the functions ${f_{\mu ,{T_\infty }}}$ and ${f^{\prime}_{{D_s},{T_\infty }}}$ are relevant to the temperature of free stream ${T_\infty }$. Formula (3.2) shows that ${\bar{D}_s}$ only depends on dimensional parameter ${T_\infty }$ in addition to non-dimensional parameters.

The second item on the right-hand side of (3.1) can be written as

(3.3)\begin{align}\begin{aligned} rh{s_{c2}} & = \dfrac{L}{{{\rho _\infty }{U_\infty }}}{M_s}\sum\limits_{z = 1}^{nr} {(\nu _{z,s}^b - \nu _{z,s}^f)} {\left[ {\sum\limits_{l = 1}^{ns} {{\alpha_{z,l}}} \left( {\dfrac{{{\rho_l}}}{{{M_l}}}} \right)} \right]^{{L_z}}}\left[ {k_z^f\prod\limits_{l = 1}^{ns} {{{\left( {\dfrac{{{\rho_l}}}{{{M_l}}}} \right)}^{\nu_{z,l}^f}}} - k_z^b\prod\limits_{l = 1}^{ns} {{{\left( {\dfrac{{{\rho_l}}}{{{M_l}}}} \right)}^{\nu_{z,l}^b}}} } \right]\\ & = \dfrac{{{\rho _\infty }L}}{{{U_\infty }}}{M_s}\sum\limits_{z = 1}^{nr} {(\nu _{z,s}^b - \nu _{z,s}^f)} {\left[ {\sum\limits_{l = 1}^{ns} {{\alpha_{z,l}}} \left( {\dfrac{{{y_l}}}{{{M_l}}}} \right)} \right]^{{L_z}}}\left[ {k_z^f{{\bar{\rho }}^{{L_z} + \sum\nolimits_{l = 1}^{ns} {\nu_{z,l}^f} }} \rho_\infty^{{L_z} + \sum\nolimits_{l = 1}^{ns} {\nu_{z,l}^f - 2} }\prod\limits_{l = 1}^{ns} {{{\left( {\dfrac{{{y_l}}}{{{M_l}}}} \right)}^{\nu_{z,l}^f}}} } \right.\\ & \left. \quad -\, k_z^b {{\bar{\rho }}^{{L_z} + \sum\nolimits_{l = 1}^{ns} {\nu_{z,l}^b} }} \rho_\infty^{{L_z} + \sum\nolimits_{l = 1}^{ns} {\nu_{z,l}^f - 2} }\prod\limits_{l = 1}^{ns} {{{\left( {\dfrac{{{y_l}}}{{{M_l}}}} \right)}^{\nu_{z,l}^b}}} \right]\\ &= \dfrac{{{\rho _\infty }L}}{{{U_\infty }}}{M_s}\sum\limits_{z = 1}^{nr} {(\nu _{z,s}^b - \nu _{z,s}^f)} {\left[ {\sum\limits_{l = 1}^{ns} {{\alpha_{z,l}}} \left( {\dfrac{{{y_l}}}{{{M_l}}}} \right)} \right]^{{L_z}}}{{R^{\prime}}_z}, \end{aligned}\end{align}

where ${\rho _\infty }L$ and ${U_\infty }$ are each required to be the same between two different flows according to the binary scaling law. Terms $k_z^f$ and $k_z^b$ can be expressed as a function of dimensionless parameters like ${\bar{D}_s}$:

(3.4)\begin{equation}k_z^{f/b} = {f_{k_z^{f/b}}}(T,{T_e},{T_{{v_1}}}, \ldots ,{T_{{v_{mol}}}}) = {f_{k_z^{f/b},{T_\infty }}}(\bar{T},{\bar{T}_e},{\bar{T}_{{v_1}}}, \ldots ,{\bar{T}_{{v_{mol}}}}).\end{equation}

As far as elementary reactions of high-temperature air are concerned (Gupta et al. Reference Gupta, Yos, Thompson and Lee1990; Park Reference Park1993), the forward reactions can be only second-order reactions and the backward reactions can be either second-order or third-order reactions. When the reverse of reaction z is a second-order reaction, ${L_z} + \sum\nolimits_{l = 1}^{ns} {\nu _{z,l}^f} = 2$ and ${L_z} + \sum\nolimits_{l = 1}^{ns} {\nu _{z,l}^b} = 2$ in (3.3). Thus, ${R^{\prime}_z}$ in (3.3) can be written as (refer to (3.4))

(3.5)\begin{align}{R^{\prime}_z} &= k_z^f {\bar{\rho }^2}\prod\limits_{l = 1}^{ns} {{{\left( {\frac{{{y_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^f}}} - k_z^b {\bar{\rho }^2}\prod\limits_{l = 1}^{ns} {{{\left( {\frac{{{y_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^b}}}\nonumber\\ &\quad = {f_{{{R^{\prime}}_z},{T_\infty }}}(\bar{\rho },\bar{T},{\bar{T}_e},{\bar{T}_{{v_1}}}, \ldots ,{\bar{T}_{{v_{mol}}}},{y_1}, \ldots ,{y_{ns}}),\end{align}

where ${f_{{{R^{\prime}}_z},{T_\infty }}}$ only depends on the dimensional parameter ${T_\infty }$ in addition to dimensionless parameters.

When the reverse of reaction z is a third-order reaction, ${L_z} + \sum\nolimits_{l = 1}^{ns} {\nu _{z,l}^f} = 2$ and ${L_z} + \sum\nolimits_{l = 1}^{ns} {\nu _{z,l}^b} = 3$. Therefore, ${R^{\prime}_z}$ can be expressed as

(3.6)\begin{align}{R^{\prime}_z} &= k_z^f {\bar{\rho }^2}\prod\limits_{l = 1}^{ns} {{{\left( {\frac{{{y_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^f}}} - k_z^b {\bar{\rho }^3}{\rho _\infty }\prod\limits_{l = 1}^{ns} {{{\left( {\frac{{{y_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^b}}}\nonumber\\ &\quad = {f_{{{R^{\prime}}_z},{T_\infty },{\rho _\infty }}}(\bar{\rho },\bar{T},{\bar{T}_e},{\bar{T}_{{v_1}}}, \ldots ,{\bar{T}_{{v_{mol}}}},{y_1}, \ldots ,{y_{ns}}),\end{align}

where ${f_{{{R^{\prime}}_z},{T_\infty },{\rho _\infty }}}$ depends on both ${T_\infty }$ and ${\rho _\infty }$ in addition to dimensionless parameters. In this case, the similarity between two different hypersonic airflows not only needs the same ${\rho _\infty }L$ but also needs the same ${\rho _\infty }$. In other words, the scale of the flow field L needs to be equal everywhere and the two flows are actually the same.

Generally, the density $\rho $ of the flow field around hypersonic vehicles is relatively low because of the low density of the free stream ${\rho _\infty }$ at high altitude. In this case, the backward third-order reaction rate is far less than the forward second-order reaction rate for reaction z in most areas of the flow field. Thus, for high-altitude flow, formula (3.6) can be written as follows:

(3.7)\begin{align}\begin{aligned} {{R^{\prime}}_z} & = k_z^f {{\bar{\rho }}^2}\prod\limits_{l = 1}^{ns} {{{\left( {\dfrac{{{y_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^f}}} - \rho \left[ {k_z^b {{\bar{\rho }}^2}\prod\limits_{l = 1}^{ns} {{{\left( {\dfrac{{{y_l}}}{{{M_l}}}} \right)}^{\nu_{z,l}^b}}} } \right] \approx k_z^f {{\bar{\rho }}^2}\prod\limits_{l = 1}^{ns} {{{\left( {\dfrac{{{y_l}}}{{{M_l}}}} \right)}^{\nu _{z,l}^f}}} \\ & = {{f^{\prime}}_{{{R^{\prime}}_z},{T_\infty }}}(\bar{\rho },\bar{T},{{\bar{T}}_e},{{\bar{T}}_{{v_1}}}, \ldots ,{{\bar{T}}_{{v_{mol}}}},{y_1}, \ldots ,{y_{ns}}), \end{aligned}\end{align}

where ${f^{\prime}_{{{R^{\prime}}_z},{T_\infty }}}$ is unrelated to ${\rho _\infty }$. In this case, should ${\rho _\infty }L$, ${T_\infty }$, $\; {U_\infty }$, $\; {y_{1,\infty }}, \ldots ,{y_{ns,\infty }}$ be identical, the airflows over geometrically similar objects will be similar and the binary scaling law can be applied to the continuity equation.

3.2. Similarity analysis for vibrational energy conservation equation

The dimensionless vibrational energy conservation equation is written as

(3.8)\begin{equation}\frac{{\partial \bar{\rho }{y_m}{{\bar{E}}_{{v_m}}}}}{{\partial \bar{t}}} + \frac{{\partial \bar{\rho }{y_m}{{\bar{E}}_{{v_m}}}{{\bar{u}}_j}}}{{\partial {{\bar{x}}_j}}} = \frac{1}{{R{e_\infty }}}\frac{{\partial \left( {{{\bar{k}}_{{v_m}}}\frac{{\partial {{\bar{T}}_{{v_m}}}}}{{\partial {{\bar{x}}_j}}} + \bar{\rho }{{\bar{D}}_m}{{\bar{E}}_{{v_m}}}\frac{{\partial {y_m}}}{{\partial {{\bar{x}}_j}}}} \right)}}{{\partial {{\bar{x}}_j}}} + \frac{L}{{{\rho _\infty }U_\infty ^3}}{\omega _{{v_m}}},\end{equation}

where ${\bar{k}_{{v_m}}}$ and ${\bar{E}_{{v_m}}}$ can be expressed as follows (see (2.49), (2.51) and (A3), (A6) in Appendix A):

(3.9)\begin{gather}{\bar{k}_{{v_m}}} = \frac{{{f_{{k_{{v_m}}}}}(T,{T_{{v_m}}},{y_1}, \ldots ,{y_{ns}})}}{{{\mu _\infty }U_\infty ^2/{T_\infty }}} = \frac{{{f_{{k_{{v_m}}},{T_\infty }}}(\bar{T},{{\bar{T}}_{{v_m}}},{y_1}, \ldots ,{y_{ns}})}}{{{f_{\mu ,{T_\infty }}}(1,{T_{e,\infty }}/{T_\infty },{y_{1,\infty }}, \ldots ,{y_{ns,\infty }})U_\infty ^2/{T_\infty }}},\end{gather}
(3.10)\begin{gather}{\bar{E}_{{v_m}}} = \frac{{{f_{{E_{{v_m}}},{T_\infty }}}({{\bar{T}}_{{v_m}}})}}{{U_\infty ^2}}.\end{gather}

Formulae (3.9) and (3.10) indicate that ${T_\infty }$ and ${U_\infty }$ are the only dimensional parameters on which ${\bar{k}_{{v_m}}}$ and ${\bar{E}_{{v_m}}}$ depend. The second item on the right-hand side of (3.8) can be written as (refer to (2.29), (2.34), (2.39) and (2.40))

(3.11)\begin{align}rh{s_{v2}} &= \frac{{{\rho _\infty }L}}{{\rho _\infty ^2U_\infty ^3}}({Q_{t - {v_m}}} + {Q_{v - {v_m}}} + {Q_{e - {v_m}}} + {Q_{r - {v_m}}}) = \frac{{{\rho _\infty }L}}{{U_\infty ^3}}({\bar{\rho }^2}{f^{\prime}_{{Q_{t - {v_m}}}}} + {\bar{\rho }^2}{f^{\prime}_{{Q_{v - {v_m}}}}}\nonumber\\ &\quad + {\bar{\rho }^2}{f^{\prime}_{{Q_{e - {v_m}}}}}) + \frac{L}{{{\rho _\infty }U_\infty ^3}}{\omega _m}{E_{{v_m}}},\end{align}

where ${f^{\prime}_{{Q_{t - {v_m}}}}}$ can be written as a function of dimensionless parameters by (2.29):

(3.12)\begin{equation}{f^{\prime}_{{Q_{t - {v_m}}}}} = {f^{\prime}_{{Q_{t - {v_m},{T_\infty }}}}}(\bar{T},{\bar{T}_{{v_m}}},{\bar{T}_e},{y_1}, \cdots ,{y_{ns}}),\end{equation}

where ${f^{\prime}_{{Q_{t - {v_m},{T_\infty }}}}}$ only depends on dimensional parameter ${T_\infty }$ besides dimensionless parameters. Similarly, ${f^{\prime}_{{Q_{v - {v_m}}}}}$ and ${f^{\prime}_{{Q_{e - {v_m}}}}}$ can also be expressed as functions ${f^{\prime}_{{Q_{v - {v_m},{T_\infty }}}}}(\bar{T},{\bar{T}_{{v_m}}},{y_m},{y_{s1}}, \ldots ,{y_{smol}})$ and ${f^{\prime}_{{Q_{e - {v_m},{T_\infty }}}}}(\bar{T},{\bar{T}_e},{y_m},{y_{{\textrm{e}^ - }}})$, which are only related to dimensional parameter ${T_\infty }$. These three terms can ensure the similarity of the vibrational energy conservation equation when the preconditions of the binary scaling law are satisfied. The term $(L/{\rho _\infty }U_\infty ^3){\omega _m}{E_{{v_m}}}$ can be analogous to the term $(L/{\rho _\infty }{U_\infty }){\omega _s}$ in continuity equation (3.1) and depends on both ${T_\infty }$ and ${\rho _\infty }$. Thus, the term $(L/{\rho _\infty }U_\infty ^3){\omega _m}{E_{{v_m}}}$ can cause dissimilarity in two hypersonic airflows which have different scales but meet the conditions of the binary scaling law. The fundamental cause of this dissimilarity also lies in the three-body reactions of the reverse reaction. Just as in the continuity equation (3.1), the binary scaling law can be applied to the vibrational energy conservation equation as long as the incoming flow density ${\rho _\infty }$ is low enough.

Thus, only ${Q_{r - {v_m}}}$ in the vibrational energy conservation equation can theoretically result in the failure of the binary scaling law and the root cause is still three-body reactions.

3.3. Similarity analysis for electron–electronic energy conservation equation

The dimensionless electron–electronic energy conservation equation is written as

(3.13)\begin{equation}\frac{{\partial \bar{\rho }{{\bar{E}}_e}}}{{\partial \bar{t}}} + \frac{{\partial \bar{\rho }{{\bar{E}}_e}{{\bar{u}}_j}}}{{\partial {{\bar{x}}_j}}} = \frac{1}{{R{e_\infty }}}\frac{{\partial \left( {{{\bar{k}}_e}\frac{{\partial {{\bar{T}}_e}}}{{\partial {{\bar{x}}_j}}} + \bar{\rho }\sum\limits_{s = 1}^{ns} {{{\bar{D}}_s}{{\bar{H}}_{{e_s}}}} \frac{{\partial {y_s}}}{{\partial {{\bar{x}}_j}}}} \right)}}{{\partial {{\bar{x}}_j}}} + \frac{L}{{{\rho _\infty }U_\infty ^3}}{\omega _e},\end{equation}

where ${\bar{k}_e}$, ${\bar{E}_e}$ and ${\bar{H}_{{e_s}}}$ can be expressed as follows (see (2.49), (2.52) and (A3), (A6) in Appendix A):

(3.14)\begin{gather}{\bar{k}_e} = \frac{{{f_{{k_e}}}(T,{T_e},{y_1}, \ldots ,{y_{ns}})}}{{{\mu _\infty }U_\infty ^2/{T_\infty }}} = \frac{{{f_{{k_e},{T_\infty }}}(\bar{T},{{\bar{T}}_e},{y_1}, \ldots ,{y_{ns}})}}{{{f_{\mu ,{T_\infty }}}(1,{T_{e,\infty }}/{T_\infty },{y_{1,\infty }}, \ldots ,{y_{ns,\infty }})U_\infty ^2/{T_\infty }}},\end{gather}
(3.15)\begin{gather}{\bar{E}_e} = \frac{{{f_{{E_e},{T_\infty }}}({{\bar{T}}_e},{y_1}, \ldots ,{y_{ns}})}}{{U_\infty ^2}},\end{gather}
(3.16)\begin{gather}{\bar{H}_{{e_s}}} = \frac{{{f_{{H_{{e_s}}},{T_\infty }}}({{\bar{T}}_e})}}{{U_\infty ^2}}.\end{gather}

Similar to ${\bar{k}_{{v_m}}}$ and ${\bar{E}_{{v_m}}}$ in § 3.2, the dimensional parameters on which ${\bar{k}_e}$, ${\bar{E}_e}$ and ${\bar{H}_{{e_s}}}$ depend are only ${T_\infty }$ and ${U_\infty }$. The second item on the right-hand side of (3.13) can be written as (refer to (2.39), (2.46), (2.47) and (2.48))

(3.17)\begin{align} rh{s_{e2}} & = \dfrac{{{\rho _\infty }L}}{{\rho _\infty ^2U_\infty ^3}}\left( {{Q_{t - e}} - \mathop \sum \limits_{m = mol} {Q_{e - {v_m}}} + {Q_{r - e}} + {Q_{i - e}} - {p_e}\dfrac{{\partial {u_j}}}{{\partial {x_j}}}} \right)\nonumber\\ &\quad = \dfrac{{{\rho _\infty }L}}{{U_\infty ^3}}\left( {{{\bar{\rho }}^2}f_{t - e}^{ion} + {{\bar{\rho }}^2}f_{t - e}^{neu} - {{\bar{\rho }}^2}\mathop \sum \limits_{m = mol} {{f^{\prime}}_{{Q_{e - {v_m}}}}}} \right)\nonumber\\ & \quad + \dfrac{L}{{{\rho _\infty }U_\infty ^3}}\left( {\mathop \sum \limits_{s = 1}^{ns} {\omega_s}{E_{{e_s}}} - \dfrac{1}{3}({I_\textrm{O}}{\omega_{{\textrm{O}^ + }}} + {I_\textrm{N}}{\omega_{{\textrm{N}^ + }}})} \right) - \dfrac{{L{p_e}}}{{{\rho _\infty }U_\infty ^3}}\dfrac{{\partial {u_j}}}{{\partial {x_j}}}, \end{align}

where $f_{t - e}^{ion}$ and $f_{t - e}^{neu}$ can be written as functions $f_{t - e,{T_{\infty ,}}{\rho _\infty }}^{ion}(\bar{\rho },\bar{T},{\bar{T}_e},{y_{{\textrm{e}^ - }}},{y_{ion1}}, \ldots ,{y_{ionn}})$ and $f_{t - e,{T_\infty }}^{neu}(\bar{T},{\bar{T}_e},{y_{{\textrm{e}^ - }}},{y_{neu1}}, \ldots ,{y_{neun}})$, respectively, based on (2.46), similar to $\; {f^{\prime}_{{Q_{e - {v_m},{T_\infty }}}}}(\bar{T},{\bar{T}_e},{y_m},{y_{{\textrm{e}^ - }}})$ in § 3.2. One can notice that $f_{t - e,{T_\infty }}^{neu}$ only depends on dimensional variable ${T_\infty }$ like ${f^{\prime}_{{Q_{e - {v_m}}},{T_\infty }}}$, whereas $f_{t - e,{T_{\infty ,}}{\rho _\infty }}^{ion}$ not only depends on ${T_\infty }$ but also is dependent on ${\rho _\infty }$. Therefore, it is observed that the energy exchange between translation and electron mode caused by electron–ion collisions, ${Q_{t - e\_ions}}$, can invalidate the binary scaling law. If the energy exchange caused by electron–ion collisions is small enough compared with electron–electronic energy change caused by other means, the binary scaling law will be approximately applicable.

Due to the preceding three-body reactions in the reverse reaction, the electron–electronic energy change resulting from chemical reactions (including $\sum\nolimits_{s = 1}^{ns} {{\omega _s}{E_{{e_s}}}} $, ${I_\textrm{O}}{\omega _{{\textrm{O}^ + }}}$ and ${I_\textrm{N}}{\omega _{{\textrm{N}^ + }}}$) can make the binary scaling law invalid. Similar to the analysis for term $(L/{\rho _\infty }U_\infty ^3){\omega _m}{E_{{v_m}}}$ in (3.11), the binary scaling law can be applied to the electron–electronic energy conservation equation if the incoming flow density ${\rho _\infty }$ is sufficiently low.

The term $- (L{p_e}/{\rho _\infty }U_\infty ^3)(\partial {u_j}/\partial {x_j})$ can be expressed as

(3.18)\begin{equation}- \frac{{L{p_e}}}{{{\rho _\infty }U_\infty ^3}}\frac{{\partial {u_j}}}{{\partial {x_j}}} ={-} \frac{{{T_\infty }{R_u}{y_{{e^ - }}}}}{{U_\infty ^2{M_{{e^ - }}}}}\frac{{\bar{\rho }{{\bar{T}}_e}\partial {{\bar{u}}_j}}}{{\partial {{\bar{x}}_j}}}.\end{equation}

One can observe from (3.18) that the term $- (L{p_e}/{\rho _\infty }U_\infty ^3)(\partial {u_j}/\partial {x_j})$ only depends on dimensional parameters ${T_\infty }$, ${U_\infty }$ and does not affect the feasibility of the binary scaling law.

Thus, ${Q_{t - e}}$, ${Q_{r - e}}$ and ${Q_{i - e}}$ in the electron–electronic energy conservation equation are the terms that could potentially invalidate the binary scaling law. For ${Q_{r - e}}$ and ${Q_{i - e}}$, the primary cause is still three-body reactions. For ${Q_{t - e}}$, the primary cause is the translation–electron energy exchange resulting from electron–ion collisions, ${Q_{t - e\_ions}}$.

4. Numerical analysis

The governing equations with the multi-temperature model are discretized by cell-centred finite volume method based on unstructured scheme. The inviscid fluxes are calculated using a modified Harten–Lax–van Leer contact scheme (Kim et al. Reference Kim, Lee, Lee and Jeung2009) with second-order multi-dimensional limiting process (MLP) (Park, Yoon & Kim Reference Park, Yoon and Kim2010) reconstruction and MLP pressure-weighted limiter (Zhang, Liu & Chen Reference Zhang, Liu and Chen2018). The second-order central difference scheme is applied for viscous flux calculation. The lower–upper symmetric Gauss–Seidel implicit algorithm (Blazek Reference Blazek2015) is employed for time integration. The Gupta and the Park chemical reaction models with 11-species and four-temperature model with vibrational temperature of ${\textrm{N}_2}$ and ${\textrm{O}_2}$ are employed.

In this section, all the meshes of different cases have high quality near the wall. The wall mesh Reynolds number based on local parameters next to the wall is defined as follows:

(4.1)\begin{equation}R{e_{cell}} = \frac{{{\rho _w}{c_w}\Delta x}}{{{\mu _w}}},\end{equation}

where c is speed of sound and $\Delta x$ denotes the normal height of the first grid at the wall. The subscript w denotes the parameters near the wall. The accuracy of heat flux calculation is guaranteed when the value of $R{e_{cell}}$ is less than or equal to 3 (Men'shov & Nakamura Reference Men'shov and Nakamura2000). In this paper, the maximum wall mesh Reynolds numbers $R{e_{cell,max}}$ of all the cases are located at their corresponding stagnation point and are less than 3.

4.1. Code validation

The flight experimental data of RAM-C II vehicle (Jones & Cross Reference Jones and Cross1972) are used for code validation. The geometry of RAM-C II vehicle and grid of the computation field are illustrated in figure 1. The detailed computational conditions are listed in table 1.

Figure 1. Geometry and grid of RAM-C II vehicle (top: grid; bottom: geometry of vehicle).

Table 1. Computational conditions for RAM-C II vehicle.

A comparison of peak electron number density along the RAM-C II vehicle surface at 71 km between CFD predicted result and flight data is plotted in figure 2. It is observed that the distribution of maximum electron number densities calculated by the Gupta model presents a good agreement with the experimental data. Although the distribution of peak electron number density predicted by the Park model results in a poorer agreement with the flight data compared with that predicted by the Gupta model, the result obtained by the Park model is no more than an order of magnitude different from the experimental data on the whole and has the same trend as the experimental data.

Figure 2. Peak electron number density along the surface of RAM-C II vehicle at 71 km.

4.2. The failure of the binary scaling law for electron distribution

The applicability of the binary scaling similarity law for thermochemical non-equilibrium air was examined using the classical case of flow around cylinders. The original cylinder which had a radius of 0.15 m and its 10-times-smaller model were used for numerical simulation. The original cylinder was subjected to a free-stream condition of 70 km and Mach 27 while the scaled-down cylinder had its free-stream density magnified 10 times to ensure that the condition of the binary scaling law $({\rho _\infty }R = \textrm{const}\textrm{.)}$ is satisfied. The detailed computational conditions are listed in table 2.

Table 2. Computational conditions for original (R = 0.15 m) and 10-times-scaled (R = 0.015 m) cylinders.

Grid independence studies of the original and scaled-down cylinders were implemented with three different grids: A, B and C. The grid quality increases from A to C by increasing both the circumferential and radial grid numbers. Detailed information of the grids is given in table 3. Grids A, B and C are constructed with the same normal size of the first cell next to surfaces. All the meshes are refined near the bow shock and Grid B with medium quality for two cylinders is illustrated in figure 3. The distributions of electron number density along the stagnation line for both the original and the scaled-down cylinders using Grids A, B and C are shown in figure 4. The results with different grid resolutions apparently demonstrate good agreement. According to the results with Grids B and C, the flow parameters no longer rely on the quality of the grid. Therefore, only the results calculated with Grid B are used for the later analysis.

Figure 3. Grid B of cylinders at 70 km and $M{a_\infty } = 27$. (a) Original cylinder of R = 0.15 m. (b) Scaled cylinder of R = 0.015 m.

Figure 4. Comparisons of distributions of electron number density along the stagnation line using different grids at $M{a_\infty } = 27$ with the Gupta model. (a) Original cylinder. (b) Scaled cylinder.

Table 3. Mesh information for cylinders at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$.

Figure 5 shows the Mach number and ${\textrm{e}^ - }$ mass fraction contours of the original and scaled cylinders at $M{a_\infty } = 27$ using the Gupta model. Although there is a good similarity in the bow-shock shape between the two cylinders, the distributions of free electrons are significantly different between the two cylinders. Figure 6 shows the distributions of T, ${T_e}$, ${T_{{v_{{\textrm{N}_2}}}}}$ and ${T_{{v_{{\textrm{O}_2}}}}}$ along the stagnation streamline at $M{a_\infty } = 27$ using the Gupta model. It is observed that the temperatures of the two flows maintain a good similarity on the whole despite the relatively small differences in local regions. The mass fractions of free electrons and ions along the stagnation line are plotted in figure 7. As shown figure 7(a,b), the gaps of maximum mass fraction between original and scaled cylinders for ${\textrm{N}^ + }$, $\textrm{N}{\textrm{O}^ + }$ and $\textrm{N}_2^ + $ are all no more than 6 %. Nevertheless, figure 7(c,d) indicates that there are tremendous differences between the original and scaled cylinders for mass fractions of ${\textrm{e}^ - }$ and ${\textrm{O}^ + }$. The difference in peak mass fraction on the stagnation line between the two cylinders reaches 50.4 % for ${\textrm{e}^ - }$ and 59.0 % for ${\textrm{O}^ + }$. As such, the binary scaling law fails completely for distribution of ${\textrm{e}^ - }$ at $M{a_\infty } = 27$ with the Gupta model.

Figure 5. Contours of (a) Mach number and (b) mass fraction of ${\textrm{e}^ - }$ at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model (top: original cylinder; bottom: scaled cylinder).

Figure 6. Distributions of temperatures along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. (a) Translational–rotational and electron–electronic temperatures and (b) ${\textrm{N}_2}$ and ${\textrm{O}_2}$ vibrational temperatures.

Figure 7. Distributions of electrons and ions along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. Mass fractions of (a) $\textrm{N}{\textrm{O}^ + }$ and $\textrm{O}_2^ + $, (b) $\textrm{N}_2^ + $ and ${\textrm{N}^ + }$, (c) ${\textrm{e}^ - }$ and (d) ${\textrm{O}^ + }$.

Figure 8 shows the distributions of ${\textrm{e}^ - }$, ${\textrm{O}^ + }$ and temperatures along the stagnation line at $M{a_\infty } = 27$ using the Park model. It is observed that the binary scaling law holds well for the distributions of ${\textrm{e}^ - }$ and ${\textrm{O}^ + }$ when the Park model is employed. Furthermore, the similarity in temperatures between the two cylinders using the Park model is greater than that using the Gupta model. However, the binary scaling law was found to be invalid for modelling electron distribution using the Park model when $M{a_\infty }$ increases to 39 (other boundary conditions remain unchanged). The distributions of ${\textrm{e}^ - }$ and ${\textrm{N}^ + }$ at $M{a_\infty } = 39$ using the Park model are illustrated in figure 9 (the detailed information of grids is listed in table 4). It is observed that the peak mass fraction of ${\textrm{e}^ - }$ and ${\textrm{N}^ + }$ in the original cylinder is obviously higher that in the scaled-down cylinder at $M{a_\infty } = 39$ using the Park model. Zeng et al. (Reference Zeng, Lin, Liu and Qu2009) also found that, at ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}$ and ${U_\infty } = 7650\;\textrm{m}\;{\textrm{s}^{ - 1}}$, the binary scaling law failed for distribution of ${\textrm{O}^ + }$ by numerical simulation of RAM-C II and its scaled model using the Gupta model.

Figure 8. Distributions of ${\textrm{e}^ - }$, ${\textrm{O}^ + }$ and temperatures along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model. Mass fractions of (a) ${\textrm{e}^ - }$ and (b) ${\textrm{O}^ + }$. (c) Translational–rotational and electron–electronic temperatures and (d) ${\textrm{N}_2}$ and ${\textrm{O}_2}$ vibrational temperatures.

Figure 9. Distributions of ${\textrm{e}^ - }$ and ${\textrm{N}^ + }$ at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model. (a) Contours of mass fraction of ${\textrm{e}^ - }$, (b) contours of mass fraction of N+, (c) mass fraction of ${\textrm{e}^ - }$ along the stagnation line and (d) mass fraction of ${\textrm{N}^ + }$ along the stagnation line.

Table 4. Mesh information for cylinders at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$.

4.3. Effect of the electron-impact ionization reactions

According to the analysis in § 3.1, the three-body reaction is one of the main reasons for the invalidation of the binary scaling law. We try to find out which three-body reactions caused the failure of the binary scaling law for the electron distribution at high altitude with extremely high Mach number. The formation of free electrons is governed by seven elementary reactions in the Gupta air-chemistry model: (i) associative ionization reactions ($\textrm{N} + \textrm{O}\rightleftharpoons \textrm{N}{\textrm{O}^ + } + {\textrm{e}^ - }$, $\textrm{O} + \textrm{O}\rightleftharpoons \textrm{O}_2^ +{+} {\textrm{e}^ - }$ and $\textrm{N} + \textrm{N}\rightleftharpoons \textrm{N}_2^ +{+} {\textrm{e}^ - }$), (ii) electron-impact ionization reactions ($\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ and $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$) and (iii) other reactions (${\textrm{O}_2} + {\textrm{N}_2}\rightleftharpoons \textrm{N}{\textrm{O}^ + } + \textrm{NO} + {\textrm{e}^ - }$ and $\textrm{NO} + \textrm{M}\rightleftharpoons \textrm{N}{\textrm{O}^ + } + {\textrm{e}^ - } + \textrm{M}$). Noticing that the shape of the distribution curve of ${\textrm{e}^ - }$ is very similar to that of ${\textrm{O}^ + }$ from figure 7(c,d), it is speculated that the dissimilarity of electron distribution might be related to $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ in the Gupta model. The dimensionless reaction rates can be defined as follows:

(4.2)\begin{equation}\overline {R_z^{fb}} ,\overline {R_z^f} ,\overline {R_z^b} = \frac{{R_z^{fb},R_z^f,R_z^b}}{{{\rho _\infty }{U_\infty }/(R{M_{mix,\infty }})}},\end{equation}

where ${M_{mix,\infty }}$ is the average molar mass of free stream and R is the radius of the cylinder. The dimensionless total reaction rates of the foregoing seven elementary reactions are plotted in figure 10. It is obvious that the reaction $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ plays the most dominant role in formation of free electrons. Figure 11 shows the dimensionless reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ on the stagnation line. As shown in figure 11(a), the dimensionless total reaction rates of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ for the original and scaled models are dissimilar and the difference of maximum reaction rates between the two reaches 26.3 %. As described in § 3.1, the precondition for the binary scaling law is that the two-body forward reaction rate is far larger than the three-body backward reaction rate, whereas both the original and scaled cylinders demonstrate a backward reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ of the same order of magnitude as the forward reaction rate, as illustrated in figure 11(b).

Figure 10. The dimensionless total rates of reactions related to generation of ${\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model.

Figure 11. The dimensionless reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. (a) Total reaction rate. (b) Forward and backward reaction rate.

In light of the fact that the binary scaling law did not fail at $M{a_\infty } = 27$ using the Park model, reaction rate coefficients of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ in the Gupta model were replaced with those in the Park model to verify the preceding assumption. Leaving the computational conditions listed in table 2 unchanged, the Gupta model was still adopted for coefficients of other reactions and the results are illustrated in figure 12. It can be observed that the binary scaling law works well for distributions of electrons, ions and temperatures. As a consequence, it can be inferred that the electron-impact ionization reaction $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ is the main reason causing the dissimilarity in flow field using the Gupta model.

Figure 12. Distributions of electrons, ions and temperatures along the stagnation line at $M{a_\infty } = 27$ with only the reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ calculated using the Park model (the Gupta model is employed for other reactions). (a) Mass fraction of ${\textrm{e}^ - }$, (b) mass fraction of ${\textrm{O}^ + }$, (c) mass fraction of $\textrm{N}_2^ + $ and ${\textrm{N}^ + }$, (d) mass fraction of $\textrm{N}{\textrm{O}^ + }$ and $\textrm{O}_2^ + $, (e) translational–rotational and electron–electronic temperatures and (f) ${\textrm{N}_2}$ and ${\textrm{O}_2}$ vibrational temperatures.

Figure 13 compares the dimensionless reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ along the stagnation streamline between the two cylinders with only reaction rate coefficients of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ calculated using the Park model. Although there are obvious differences between the two cylinders for the three-body reverse reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ from figure 13(a), the two-body forward reaction rates of that between the two cylinders have good similarity and the three-body reverse reaction rate can be negligible compared to the forward reaction rate from figure 13(b). Thus, $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ satisfies the necessary conditions of the binary scaling law in this case.

Figure 13. The dimensionless reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 27$ with only reaction rate coefficients of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ calculated using the Park model (the Gupta model is employed for other reactions). (a) Backward reaction rate. (b) Forward and backward reaction rate.

The ratio of the forward reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ to the reverse reaction rate ${\gamma _{\textrm{O} + {\textrm{e}^ - }}}$ is defined as follows:

(4.3)\begin{equation}{\gamma _{\textrm{O} + {\textrm{e}^ - }}} = \frac{{\frac{{{\rho _\textrm{O}}}}{{{M_\textrm{O}}}}\frac{{{\rho _{{\textrm{e}^ - }}}}}{{{M_{{\textrm{e}^ - }}}}}k_{\textrm{O} + {\textrm{e}^ - }}^f({T_e})}}{{\frac{{{\rho _{{\textrm{O}^ + }}}}}{{{M_{{\textrm{O}^ + }}}}}\frac{{{\rho _{{\textrm{e}^ - }}}}}{{{M_{{\textrm{e}^ - }}}}}\frac{{{\rho _{{\textrm{e}^ - }}}}}{{{M_{{\textrm{e}^ - }}}}}k_{\textrm{O} + {\textrm{e}^ - }}^b({T_e})}} = \frac{{{M_{{\textrm{O}^ + }}}{M_{{\textrm{e}^ - }}}}}{{{M_\textrm{O}}\rho }}\frac{{{y_\textrm{O}}}}{{{y_{{\textrm{O}^ + }}}{y_{{\textrm{e}^ - }}}}}{K_{eq,\textrm{O} + {\textrm{e}^ - }}}({T_e}).\end{equation}

The reverse reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ calculated by the Gupta model is of the same order of magnitude as the forward reaction rate $({\gamma _{\textrm{O} + {\textrm{e}^ - }}}\sim 1)$ and the precondition for the binary scaling law $({\gamma _{\textrm{O} + {\textrm{e}^ - }}} \gg 1)$ is no longer satisfied. Here ${T_e}$ is the controlling temperature of the forward and backward reaction rates for the electron-impact ionization reactions. At $M{a_\infty } = 27$, the maximum ${T_e}$ in cases using the Gupta model or the Park model for $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ are close from figures 6(a) and 12(e). One can notice that the equilibrium constant of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ in the Gupta model is close to that in the Park model from figure 14(a). Combined with (4.3), figures 7(c,d) and 12(a,b), it can be seen that higher mass fraction of ${\textrm{O}^ + }$ and ${\textrm{e}^ - }$ is the reason that ${\gamma _{\textrm{O} + {\textrm{e}^ - }}}$ with the Gupta model is far less than that with the Park model. Figure 14(b) compares the forward reaction rate coefficient of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ from Gupta and Park models as a function of the controlling temperature according to (2.14a,b). It is clear that the forward reaction rate coefficient of the Gupta model is more than one order of magnitude larger than that of the Park model at the corresponding peak ${T_e}$. Higher forward reaction rate with the Gupta model caused the reaction to move in the positive direction and produce more ${\textrm{O}^ + }$ and ${\textrm{e}^ - }$, which increased the backward reaction rate and reduced ${\gamma _{\textrm{O} + {\textrm{e}^ - }}}$ to around 1.

Figure 14. The equilibrium constants and forward reaction rate coefficients of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ in the Gupta and Park models. (a) Equilibrium constants (particle number density equals 1.0 × 1015 cm3 in the Park model). (b) Forward reaction rate coefficients (the arrows represent the maximum electron–electronic excitation temperatures in figures 6a and 12e).

The formation of free electrons is governed by five elementary reactions in the Park model: (i) associative ionization reactions ($\textrm{N} + \textrm{O}\rightleftharpoons \textrm{N}{\textrm{O}^ + } + {\textrm{e}^ - }$, $\textrm{O} + \textrm{O}\rightleftharpoons \textrm{O}_2^ +{+} {\textrm{e}^ - }$ and $\textrm{N} + \textrm{N}\rightleftharpoons \textrm{N}_2^ +{+} {\textrm{e}^ - }$) and (ii) electron-impact ionization reactions ($\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ and $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$). The dimensionless total reaction rates of these five elementary reactions along the stagnation line at $M{a_\infty } = 39$ using the Park model are depicted in figure 15. It can be seen that the electron-impact ionization reaction $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ is dominant in the formation of ${\textrm{e}^ - }$. The other electron-impact ionization reaction $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ also plays an important role in electron formation near the location where a significant difference in electron mass fraction began to appear (x/R = −0.09; see figure 9c). Figure 16 shows the dimensionless reaction rates of the two electron-impact ionization reactions along the stagnation streamline at $M{a_\infty } = 39$ using the Park model. It is noticed that the backward reaction rate is of the same order of magnitude as the forward reaction rate for both reactions, which leads to the significant differences in total reaction rate between the original and scaled cylinders in figure 16(b,d).

Figure 15. The dimensionless total rates of reactions related to generation of ${\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 39$ using the Park model.

Figure 16. The dimensionless reaction rate of two electron-impact ionization reactions along the stagnation line at $M{a_\infty } = 39$ using the Park model. (a) Forward and backward reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$. (b) Total reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$. (c) Forward and backward reaction rate of $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$. (d) Total reaction rate of $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$.

Figure 17 compares the forward reaction rate coefficient of two electron-impact ionization reactions in the Gupta and Park models. Considering the foregoing results at $M{a_\infty } = 27$ and 39 with the Gupta and Park models, one can see that the binary scaling law is invalid for electron distribution when the forward reaction rate coefficient ${k^f}$ of electron-impact ionization reactions rises to a certain degree with an increase of $M{a_\infty }$. In order to further verify this, $M{a_\infty }$ is reduced to 17 (other boundary conditions remain unchanged and the detailed information of grids is listed in table 5). Figure 18 shows the distributions of ${\textrm{e}^ - }$ and ${\textrm{O}^ + }$ along the stagnation line at $M{a_\infty } = 17$ using the Gupta model. One can observe that there is good agreement in the distributions of ${\textrm{e}^ - }$ and ${\textrm{O}^ + }$ between the original and scaled cylinders (see figure 18) with a decrease in ${k^f}$ of electron-impact ionization reactions at $M{a_\infty } = 17$ using the Gupta model (see figure 17).

Figure 17. Forward reaction rate coefficient of two electron-impact ionization reactions in the Gupta and Park models (the red arrow represents the maximum electron temperature of the flow field around the cylinder under the corresponding $M{a_\infty }$ using the Gupta model and the green arrow represents that using the Park model).

Figure 18. Distributions of ${\textrm{e}^ - }$ and ${\textrm{O}^ + }$ along the stagnation line at $M{a_\infty } = 17$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. Mass fractions of (a) ${\textrm{e}^ - }$ and (b) ${\textrm{O}^ + }$.

Table 5. Mesh information for cylinders at ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ with $M{a_\infty } = 17$.

Due to effects of the electron avalanche, it is hard to measure the forward reaction rates of electron-impact ionization reactions (Park Reference Park1990; Hao et al. Reference Hao, Wang and Lee2016). However, the reverse reactions of electron-impact ionization are very nearly in quasi-steady-state distribution and the forward reaction rates can be calculated using the reverse reaction rates and equilibrium constants (Park Reference Park1990). Park measured the ionic recombination rate of $\textrm{N}$ atom ${\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - } \to \textrm{N} + {\textrm{e}^ - }$ by observing the rate of decrease of electron density in an expanding nozzle (Park Reference Park1990, Reference Park1968, Reference Park1973). Although the rates of ${\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - } \to \textrm{O} + {\textrm{e}^ - }$ have never been measured, the ionic recombination rate coefficients of various atoms (such as $\textrm{C}$, $\textrm{Ar}$, $\textrm{He}$) are fairly similar (Park Reference Park1990). Thus, Park assumed the recombination rates of ${\textrm{O}^ + }$ to be the same as that of ${\textrm{N}^ + }$ and obtained the rate coefficient of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$. This assumption is also embodied in the Gupta model. The rate coefficients of electron-impact ionization reactions in the Gupta model are taken from the model of Dunn and Kang (Dunn & Kang Reference Dunn and Kang1973; Gupta et al. Reference Gupta, Yos, Thompson and Lee1990). Only the three-body ionic recombination rate coefficient for ${\textrm{C}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - } \to \textrm{C} + {\textrm{e}^ - }$ was measured by Dunn (Reference Dunn1971). Hao et al. (Reference Hao, Wang and Lee2016) found an abnormal non-equilibrium relaxation process in the FIRE II case at $M{a_\infty } = 39$ simulated using the Gupta model, which proved to be a consequence of electron-impact ionization reaction rates calculated using the Gupta model. The problem was solved using modified coefficients of electron-impact ionization reactions based on the Park model. Hao considered that parameters for electron-impact ionization reactions in the Park model might be more accurate because Park's data were taken from more recent experiments. In addition, a more recent chemical kinetic model was developed by Ozawa, Zhong & Levin (Reference Ozawa, Zhong and Levin2008) and the forward rate coefficient of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ at high temperature calculated using the Ozawa model is closer to that calculated using the Park model (Niu et al. Reference Niu, Yuan, Dong and Tan2018). Although the Gupta and Park models have their own advantages in predictions of species formations (Niu et al. Reference Niu, Yuan, Dong and Tan2018), based on the points discussed above, the result at $M{a_\infty } = 27$ obtained using the Park model might be more reliable. The rates of electron-impact ionization reactions require to be further studied experimentally and theoretically because they strongly affect the applicable range of the binary scaling law for modelling electron distribution. Although the differences in forward electron-impact ionization reaction rates between the Gupta and Park models can reach as much as an order of magnitude or more (Niu et al. Reference Niu, Yuan, Dong and Tan2018), the electron-impact ionization reactions can lead to the failure of the binary scaling law in electron distribution at extremely high Mach number no matter what chemical model is employed, which reflects that it is probably a general conclusion in nature.

4.4. Discussion of electron–ion collisions

In § 3.3, it is found that the translation–electron energy exchange from collisions between electrons and ions ${Q_{t - e\_ions}}$ can theoretically cause the failure of the binary scaling law. The dimensionless energy exchange source term can be defined as follows:

(4.4)\begin{equation}\overline {{Q_x}} = \frac{{{Q_x}}}{{{\rho _\infty }{U_\infty }^3/R}},\end{equation}

where ${Q_x}$ represents a certain energy exchange term in the vibrational or electron–electronic energy conservation equation and R is the radius of cylinder.

The dimensionless constituents of electron–electronic energy source term ${\omega _e}$ at $M{a_\infty } = 17$ with the Gupta model are depicted in figure 19. It is noticed that ${Q_{t - e\_ions}}$ is so low relative to the dominant terms $({Q_{e - v\_total}},\; \; {Q_{i - e}})$ that ${Q_{t - e\_ions}}$ can be ignored at $M{a_\infty } = 17$. In figure 20(a), ${\bar{Q}_{t - e\_ions}}$ of the original cylinder is dissimilar to that of the scaled cylinder, which is consistent with theoretical analysis. Nevertheless, ${\bar{\omega }_e}$ of two cylinders still have good similarity from figure 20(b) because the contribution of ${\bar{Q}_{t - e\_ions}}$ is negligible.

Figure 19. Dimensionless energy exchange between electrons and other mode along the stagnation line at $M{a_\infty } = 17$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model.

Figure 20. Dimensionless energy exchange (a) ${\bar{Q}_{t - e\_ions}}$ and (b) ${\bar{\omega }_e}$ along the stagnation line at $M{a_\infty } = 17$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model.

Figure 21 shows the dissimilarity property of ${Q_{t - e\_ions}}$ at $M{a_\infty } = 27$. Similar to the situation at $M{a_\infty } = 17$, ${Q_{t - e\_ions}}$ is still too small to affect the total source term ${\omega _e}$ from figure 22.

Figure 21. Dimensionless energy exchange ${\bar{Q}_{t - e\_ions}}$ along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$. (a) Gupta model. (b) Park model.

Figure 22. Dimensionless energy exchange between electrons and other mode along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$. (a) Gupta model. (b) Park model.

As the number of ions and electrons increases at $M{a_\infty } = 39$, it can be found from figure 23 that ${Q_{t - e\_ions}}$ and ${Q_{i - e}}$ play the dominant role in the electron–electronic energy source term. Figure 24(a) shows the obvious dissimilarity of ${Q_{t - e\_ions}}$ and ${Q_{i - e}}$ between the two cylinders at $M{a_\infty } = 39$. It is clear that the trend of ${Q_{t - e\_ions}}$ is opposite to that of ${Q_{i - e}}$ and the negative effects of the two on the similarity of the source term can be offset in part to some extent. The effect of dissimilarity of electron–electronic energy source term is directly embodied by the electron–electronic temperature ${T_e}$. In figure 24(b), there are relatively large differences in electron temperature between the two cylinders at some regions. This means that the influence of ${Q_{t - e\_ions}}$ on the binary scaling law cannot be neglected at extremely high Mach number. Temperature ${T_e}$ is the controlling temperature of some reactions (for example, the electron-impact ionization reactions), which affects the formation of electrons and ions and the production of electrons and ions is directly related to ${Q_{t - e\_ions}}$. As a result, ${Q_{t - e\_ions}}$ and production of ${\textrm{e}^ - }$ are coupled, which means that the quantification of the effect on the dissimilarity of electron distribution caused by ${Q_{t - e\_ions}}$ is complicated.

Figure 23. Dimensionless energy exchange between electrons and other mode along the stagnation line at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model.

Figure 24. (a) Dimensionless energy exchange ${\bar{Q}_{t - e\_ions}}$, ${\bar{Q}_{i - e}}$ and ${\bar{\omega }_e}$ and (b) temperature ${T_e}$ along the stagnation line at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model.

5. Conclusions

The applicability of the binary scaling law to electron distribution in thermochemical non-equilibrium flows at high altitude with high Mach number is investigated through theoretical analysis and numerical simulation in the present study. The main results show the following.

The binary scaling law is no longer applicable to electron distribution at extremely high Mach number even though the density of free stream is very low. In addition to the three-body reactions, the translation–electron energy exchange from collisions between electrons and ions ${Q_{t - e\_ions}}$ can theoretically lead to the invalidation of the binary scaling law. The vibrational energy change caused by ${Q_{t - {v_m}}}$, ${Q_{v - {v_m}}}$ and ${Q_{e - {v_m}}}$ cannot theoretically result in the failure of the binary scaling law.

As the Mach number increases, electron-impact ionization reactions can result in the failure of the binary scaling law. If the Gupta model is employed, $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ is the main factor which causes the invalidation of the binary scaling law for electron distribution at $M{a_\infty } = 27$. If the Park model is used, $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ becomes the main factor for the failure of the binary scaling law at $M{a_\infty } = 39$. The root cause is that the forward reaction rate coefficients of electron-impact ionization reactions increase rapidly due to an increase of electron–electronic excitation temperature, which further leads to an increase in concentration of electrons and ions. The increase in the concentration of electrons and ions causes the reverse reaction rates of electron-impact ionization reactions to increase to the same order of magnitude as those of the forward reaction, which leads to the invalidation of the binary scaling law for distribution of ${\textrm{e}^ - }$. Based on previous researches, the parameters for electron-impact ionization reactions of the Park model might be more reliable than those of the Gupta model. The rates of electron-impact ionization reactions need further study because they can strongly affect the applicable range of the binary scaling law for modelling electron distribution.

The effect of dissimilarity caused by ${Q_{t - e\_ions}}$ can be negligible though free-stream velocity is at a high level $(M{a_\infty } = 17,\; 27)$. However, as the Mach number increases further $(M{a_\infty } = 39)$, ${Q_{t - e\_ions}}$ and ${Q_{i - e}}$ are dominant in the electron–electronic energy source term, which means that the negative effect on the binary scaling law caused by ${Q_{t - e\_ions}}$ cannot be ignored. The trend of ${Q_{t - e\_ions}}$ is opposite to that of ${Q_{i - e}}$ at $M{a_\infty } = 39$ and the negative effects of the two on the similarity of source term can be offset in part to some extent. On the basis of understanding the mechanism for the failure of the binary scaling law in electron distribution, future work will focus on developing a method to quickly estimate the difference in peak electron number density between the scaled and original models. Future work will also involve quantifying the effects of ${Q_{t - e\_ions}}$ on the dissimilarity of electron distribution.

Acknowledgements

The authors would like to thank L. Leheng for his help with English writing and revision of this paper. We would also like to thank the referees for their helpful comments and suggestions.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Method of normalization

The non-dimensional variables are expressed as follows:

(A1ac)\begin{gather}{\bar{x}_j} = \frac{{{x_j}}}{L},\quad \bar{t} = \frac{t}{{L/{U_\infty }}},\quad {\bar{u}_i} = \frac{{{u_i}}}{{{U_\infty }}},\end{gather}
(A2ac)\begin{gather}\bar{\rho } = \frac{\rho }{{{\rho _\infty }}},\quad \bar{T},{\bar{T}_{{v_m}}},{\bar{T}_e} = \frac{{T,{T_{{v_m}}},{T_e}}}{{{T_\infty }}},\quad \bar{p} = \frac{p}{{{\rho _\infty }U_\infty ^2}},\end{gather}
(A3a,b)\begin{gather}\bar{E},{\bar{E}_{{v_m}}},{\bar{E}_e} = \frac{{E,{E_{{v_m}}},{E_e}}}{{U_\infty ^2}},\quad \bar{H},{\bar{H}_s},{\bar{H}_e} = \frac{{H,{H_s},{H_e}}}{{U_\infty ^2}},\end{gather}
(A4)\begin{gather}\bar{\mu } = \frac{\mu }{{{\mu _\infty }}} = \frac{{{f_\mu }(T,{T_e},{y_1}, \ldots ,{y_{ns}})}}{{{\mu _\infty }}}, \end{gather}
(A5)\begin{gather}{\bar{D}_s} = \frac{{{D_s}}}{{{\mu _\infty }/{\rho _\infty }}} = \frac{{{{f^{\prime}}_{{D_s}}}(T,{T_e},{y_1}, \ldots ,{y_{ns}})/\rho }}{{{\mu _\infty }/{\rho _\infty }}},\end{gather}
(A6)\begin{gather}\left. {\begin{array}{*{20}{l}} {{{\bar{k}}_{tr}} = \dfrac{{{k_{tr}}}}{{{\mu_\infty }U_\infty^2/{T_\infty }}} = \dfrac{{{f_{{k_{tr}}}}(T,{y_1}, \ldots ,{y_{ns}})}}{{{\mu_\infty }U_\infty^2/{T_\infty }}}}\\ {{{\bar{k}}_{{v_m}}} = \dfrac{{{k_{{v_m}}}}}{{{\mu_\infty }U_\infty^2/{T_\infty }}} = \dfrac{{{f_{{k_{{v_m}}}}}(T,{T_{{v_m}}},{y_1}, \ldots ,{y_{ns}})}}{{{\mu_\infty }U_\infty^2/{T_\infty }}}}\\ {{{\bar{k}}_e} = \dfrac{{{k_e}}}{{{\mu_\infty }U_\infty^2/{T_\infty }}} = \dfrac{{{f_{{k_e}}}(T,{T_e},{y_1}, \ldots ,{y_{ns}})}}{{{\mu_\infty }U_\infty^2/{T_\infty }}}} \end{array}} \right\},\end{gather}

where the functions ${f_\mu }$, ${f^{\prime}_{{D_s}}}$, ${f_{{k_{tr}}}}$, ${f_{{k_{{v_m}}}}}$ and ${f_{{k_e}}}$ are described in § 2.5.

References

REFERENCES

Anderson, J.D.J.R. 2006 Hypersonic and High-Temperature Gas Dynamics, 2nd edn. American Institute of Aeronautics and Astronautics.CrossRefGoogle Scholar
Appleton, J.P. & Bray, K.N.C. 1964 The conservation equations for a non-equilibrium plasma. J. Fluid Mech. 20 (4), 659672.CrossRefGoogle Scholar
Birkhoff, G. 1960 Hydrodynamics: A Study in Logic, Fact and Similitude, 2nd edn. Princeton University Press.Google Scholar
Birkhoff, G. & Eckerman, J. 1963 Binary collision modeling. J. Math. Mech. 12 (4), 543556.Google Scholar
Blazek, J. 2015 Computational Fluid Dynamics: Principles and Applications. Butterworth-Heinemann.Google Scholar
Camm, J.C., Kivel, B., Taylor, R.L. & Teare, J.D. 1961 Absolute intensity of non-equilibrium radiation in air and stagnation heating at high altitudes. J. Quant. Spectrosc. Radiat. Transfer 1 (1), 5375.CrossRefGoogle Scholar
Candler, G.V. 1988 The computation of weakly ionized hypersonic flows in thermo-chemical nonequilibrium. PhD dissertation, Stanford University.Google Scholar
Candler, G.V. 2019 Rate effects in hypersonic flows. Annu. Rev. Fluid Mech. 51, 379402.CrossRefGoogle Scholar
Candler, G.V. & Maccormack, R.W. 1991 Computation of weakly ionized hypersonic flows in thermochemical nonequilibrium. J. Thermophys. Heat Transfer 5 (3), 266273.CrossRefGoogle Scholar
Dunn, M.G. 1971 Measurement of C+ + e + e and CO+ + e recombination in carbon monoxide flows. AIAA J. 9 (11), 21842191.CrossRefGoogle Scholar
Dunn, M.G. & Kang, S.W. 1973 Theoretical and Experimental Studies of Reentry Plasmas, vol. 2232. National Aeronautics and Space Administration.Google Scholar
Ellington, D. 1967 Binary scaling limits for hypersonic flight. AIAA J. 5 (9), 17051706.CrossRefGoogle Scholar
Farbar, E., Boyd, I.D. & Martin, A. 2013 Numerical prediction of hypersonic flowfields including effects of electron translational nonequilibrium. J. Thermophys. Heat Transfer 27 (4), 593606.CrossRefGoogle Scholar
Garicano-Mena, J., Lani, A. & Degrez, G. 2018 An entropy-variables-based formulation of residual distribution schemes for non-equilibrium flows. J. Comput. Phys. 362, 163189.CrossRefGoogle Scholar
Ghezali, Y., Haoui, R. & Chpoun, A. 2019 Study of physico-chemical phenomena in a non-equilibrium hypersonic air flow behind a strong shock wave. Thermophys. Aeromech. 26 (5), 693710.CrossRefGoogle Scholar
Gibson, W.E. & Marrone, P.V. 1964 A similitude for non-equilibrium phenomena in hypersonic flight. In AGARDograph, vol. 68, pp. 105–131. Elsevier.CrossRefGoogle Scholar
Gnoffo, P.A., Gupta, R.N. & Shinn, J.L. 1989 Conservation equations and physical models for hypersonic air flows in thermal and chemical nonequilibrium. NASA Tech. Paper 2867.Google Scholar
Gupta, R.N., Yos, J.M., Thompson, R.A. & Lee, K.P. 1990 A review of reaction and thermodynamic and transport properties for an 11-species air model for chemical and thermal nonequilibrium calculations to 30000K. NASA Tech. Rep. 1232.Google Scholar
Hall, J.G., Eschenroeder, A.Q. & Marrone, P.V. 1962 Blunt-nose inviscid airflows with coupled nonequilibrium processes. J. Aerosp. Sci. 29 (9), 10381051.CrossRefGoogle Scholar
Hao, J., Wang, J. & Lee, C. 2016 Numerical study of hypersonic flows over reentry configurations with different chemical nonequilibrium models. Acta Astronaut. 126, 110.CrossRefGoogle Scholar
Hornung, H.G. 1972 Non-equilibrium dissociating nitrogen flow over spheres and circular cylinders. J. Fluid Mech. 53 (1), 149176.CrossRefGoogle Scholar
Hornung, H.G. 1988 28th Lanchester memorial lecture—experimental real-gas hypersonics. Aeronaut. J. 92 (920), 379389.CrossRefGoogle Scholar
Houwing, A.F.P., Nonaka, S., Mizuno, H. & Takayama, K. 2000 Effects of vibrational relaxation on bow shock standoff distance for nonequilibrium flows. AIAA J. 38 (9), 17601763.CrossRefGoogle Scholar
Jones, W.L. & Cross, A.E. 1972 Electrostatic-probe measurements of plasma parameters for two reentry flight experiments at 25000 feet per second. Rep. no. 6617. National Aeronautics and Space Administration.Google Scholar
Kim, S.D., Lee, B.J., Lee, H.J. & Jeung, I.S. 2009 Robust HLLC Riemann solver with weighted average flux scheme for strong shock. J. Comput. Phys. 228 (20), 76347642.CrossRefGoogle Scholar
Landau, L. & Teller, E. 1936 Theory of sound dispersion. Phys. Z. Sowjetunion 10, 3443.Google Scholar
Lee, J.H. 1984 Basic governing equations for the flight regimes of aeroassisted orbital transfer vehicles. In 19th Thermophysics Conference, Snowmass, Colorado, AIAA Paper 84-1729.Google Scholar
Lee, J.H. 1993 Electron-impact vibrational relaxation in high-temperature nitrogen. J. Thermophys. Heat Transfer 7 (3), 399405.CrossRefGoogle Scholar
Lee, R.H. & Baker, R.L. 1969 Binary scaling of air ionization on slender cones at small yaw. AIAA J. 7 (2), 355358.CrossRefGoogle Scholar
Marschall, J. & Maclean, M. 2011 Finite-rate surface chemistry model, I: formulation and reaction system examples. In 42nd AIAA Thermophysics Conference, Honolulu, Hawaii, San Diego, CA, AIAA Paper 2011-3783.Google Scholar
Matsuzaki, R. 1988 Quasi-one-dimensional aerodynamics with chemical, vibrational and thermal nonequilibrium. Trans. Japan Soc. Aeronaut. Space Sci. 30, 243258.Google Scholar
Men'shov, I.S. & Nakamura, Y. 2000 Numerical simulations and experimental comparisons for high-speed nonequilibrium air flows. Fluid Dyn. Res. 27 (5), 305334.CrossRefGoogle Scholar
Miller, J., Tannehill, J., Lawrence, S. & Edwards, T. 1995 Development of an upwind pns code for thermo-chemical nonequilibrium flows. In 30th AIAA Thermophysics Conference, AIAA Paper 95-2009.Google Scholar
Millikan, R.C. & White, D.R. 1963 Systematics of vibrational relaxation. J. Chem. Phys. 39 (12), 32093213.CrossRefGoogle Scholar
Muylaert, J., Walpot, L., Häuser, J., Sagnier, P., Devezeaux, D., Papirnyk, O. & Lourme, D. 1992 Standard model testing in the European high enthalpy facility F4 and extrapolation to flight. In 17th Aerospace Ground Testing Conference, Nashville, TN, AIAA Paper 92-3905.Google Scholar
Niu, Q., Yuan, Z., Dong, S. & Tan, H. 2018 Assessment of nonequilibrium air-chemistry models on species formation in hypersonic shock layer. Intl J. Heat Mass Transfer 127, 703716.CrossRefGoogle Scholar
Ozawa, T., Zhong, J. & Levin, D.A. 2008 Development of kinetic-based energy exchange models for noncontinuum, ionized hypersonic flows. Phys. Fluids 20 (4), 442–165.CrossRefGoogle Scholar
Park, C. 1968 Measurement of ionic recombination rate of nitrogen. AIAA J. 6 (11), 20902094.CrossRefGoogle Scholar
Park, C. 1973 Comparison of electron and electronic temperatures in recombining nozzle flow of ionized nitrogen—hydrogen mixture. Part 1. Theory. J. Plasma Phys. 9 (2), 187215.CrossRefGoogle Scholar
Park, C. 1985 Convergence of computations of chemical reacting flows. Prog. Astronaut. Aeronaut. 103, 478513.Google Scholar
Park, C. 1990 Nonequilibrium Hypersonic Aerothermodynamics. Wiley.Google Scholar
Park, C. 1993 Review of chemical-kinetic problems of future NASA missions. I-earth entries. J. Thermophys. Heat transfer 7 (3), 385398.CrossRefGoogle Scholar
Park, C. & Lee, S.H. 1995 Validation of multitemperature nozzle flow code. J. Thermophys. Heat Transfer 9 (1), 916.CrossRefGoogle Scholar
Park, J.S., Yoon, S.H. & Kim, C. 2010 Multi-dimensional limiting process for hyperbolic conservation laws on unstructured grids. J. Comput. Phys. 229 (3), 788812.CrossRefGoogle Scholar
Petschek, H. & Byron, S. 1957 Approach to equilibrium ionization behind strong shock waves in argon. Ann. Phys. 1, 270315.CrossRefGoogle Scholar
Ramjatan, S., Lani, A., Boccelli, S., Hove, B.V., Karatekin, Ö, Magin, T. & Thoemel, J. 2020 Blackout analysis of Mars entry missions. J. Fluid Mech. 904, A26.CrossRefGoogle Scholar
Roberts, T. 1994 The development of a multi-vibrational model and incorporation into a chemical reacting Navier-Stokes solver. In 6th AIAA/ASME Joint Thermophysics and Heat Transfer Conference, Colorado Springs, CO AIAA Paper 94-1988.Google Scholar
Roberts, T. 1996 Implementation into TINA modelling for electron/electronic energy equation. In 31st AIAA Thermophysics Conference, New Orleans, LA, AIAA Paper 96-1851.Google Scholar
Scalabrin, L.C. 2007 Numerical simulation of weakly ionized hypersonic flow over reentry capsules. PhD dissertation, University of Michigan.CrossRefGoogle Scholar
Shen, C. 2005 Rarefied Gas Dynamics: Fundamentals, Simulations and Micro Flows. Springer Science & Business Media.CrossRefGoogle Scholar
Sivolella, D. 2014 To Orbit and Back Again: How the Space Shuttle Flew in Space. Springer.CrossRefGoogle Scholar
Teare, J.D., Georgiev, S. & Allen, R.A. 1962 Radiation from the non-equilibrium shock front. In Hypersonic Flow Research, pp. 281–317. Elsevier.CrossRefGoogle Scholar
Thoenes, J. 1969 Inviscid hypersonic flow of chemically relaxing air about pointed circular cones. UARI Res. Rep. 66, Alabama Univ, Huntsville, AL.Google Scholar
Wilke, C.R. 1950 A viscosity equation for gas mixtures. J. Chem. Phys. 18 (4), 517519.CrossRefGoogle Scholar
Zeng, M. 2007 Numerical rebuilding of free-stream measurement and analysis of nonequilibrium effects in high-enthalpy tunnel. PhD dissertation, Institute of Mechanics, CAS (in Chinese).Google Scholar
Zeng, M., Lin, Z., Liu, J. & Qu, Z. 2009 Numerical analysis of the validity of binary scaling parameter ρL in nonequilibrium flow. Chin. J. Theor. Appl. Mech. 41 (2), 177184 (in Chinese).Google Scholar
Zhang, H.X. 1990 The similarity law for real gas flow. Acta Aerodyn. Sin. 8 (1), 18 (in Chinese).Google Scholar
Zhang, F., Liu, J. & Chen, B. 2018 Modified multi-dimensional limiting process with enhanced shock stability on unstructured grids. Comput. Fluids 161, 171188.CrossRefGoogle Scholar
Zhu, Y., Lee, C., Chen, X., Wu, J., Chen, S. & Gad-el-Hak, M. 2018 Newly identified principle for aerodynamic heating in hypersonic flows. J. Fluid Mech. 855, 152180.CrossRefGoogle Scholar
Figure 0

Figure 1. Geometry and grid of RAM-C II vehicle (top: grid; bottom: geometry of vehicle).

Figure 1

Table 1. Computational conditions for RAM-C II vehicle.

Figure 2

Figure 2. Peak electron number density along the surface of RAM-C II vehicle at 71 km.

Figure 3

Table 2. Computational conditions for original (R = 0.15 m) and 10-times-scaled (R = 0.015 m) cylinders.

Figure 4

Figure 3. Grid B of cylinders at 70 km and $M{a_\infty } = 27$. (a) Original cylinder of R = 0.15 m. (b) Scaled cylinder of R = 0.015 m.

Figure 5

Figure 4. Comparisons of distributions of electron number density along the stagnation line using different grids at $M{a_\infty } = 27$ with the Gupta model. (a) Original cylinder. (b) Scaled cylinder.

Figure 6

Table 3. Mesh information for cylinders at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$.

Figure 7

Figure 5. Contours of (a) Mach number and (b) mass fraction of ${\textrm{e}^ - }$ at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model (top: original cylinder; bottom: scaled cylinder).

Figure 8

Figure 6. Distributions of temperatures along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. (a) Translational–rotational and electron–electronic temperatures and (b) ${\textrm{N}_2}$ and ${\textrm{O}_2}$ vibrational temperatures.

Figure 9

Figure 7. Distributions of electrons and ions along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. Mass fractions of (a) $\textrm{N}{\textrm{O}^ + }$ and $\textrm{O}_2^ + $, (b) $\textrm{N}_2^ + $ and ${\textrm{N}^ + }$, (c) ${\textrm{e}^ - }$ and (d) ${\textrm{O}^ + }$.

Figure 10

Figure 8. Distributions of ${\textrm{e}^ - }$, ${\textrm{O}^ + }$ and temperatures along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model. Mass fractions of (a) ${\textrm{e}^ - }$ and (b) ${\textrm{O}^ + }$. (c) Translational–rotational and electron–electronic temperatures and (d) ${\textrm{N}_2}$ and ${\textrm{O}_2}$ vibrational temperatures.

Figure 11

Figure 9. Distributions of ${\textrm{e}^ - }$ and ${\textrm{N}^ + }$ at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model. (a) Contours of mass fraction of ${\textrm{e}^ - }$, (b) contours of mass fraction of N+, (c) mass fraction of ${\textrm{e}^ - }$ along the stagnation line and (d) mass fraction of ${\textrm{N}^ + }$ along the stagnation line.

Figure 12

Table 4. Mesh information for cylinders at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$.

Figure 13

Figure 10. The dimensionless total rates of reactions related to generation of ${\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model.

Figure 14

Figure 11. The dimensionless reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. (a) Total reaction rate. (b) Forward and backward reaction rate.

Figure 15

Figure 12. Distributions of electrons, ions and temperatures along the stagnation line at $M{a_\infty } = 27$ with only the reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ calculated using the Park model (the Gupta model is employed for other reactions). (a) Mass fraction of ${\textrm{e}^ - }$, (b) mass fraction of ${\textrm{O}^ + }$, (c) mass fraction of $\textrm{N}_2^ + $ and ${\textrm{N}^ + }$, (d) mass fraction of $\textrm{N}{\textrm{O}^ + }$ and $\textrm{O}_2^ + $, (e) translational–rotational and electron–electronic temperatures and (f) ${\textrm{N}_2}$ and ${\textrm{O}_2}$ vibrational temperatures.

Figure 16

Figure 13. The dimensionless reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 27$ with only reaction rate coefficients of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ calculated using the Park model (the Gupta model is employed for other reactions). (a) Backward reaction rate. (b) Forward and backward reaction rate.

Figure 17

Figure 14. The equilibrium constants and forward reaction rate coefficients of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$ in the Gupta and Park models. (a) Equilibrium constants (particle number density equals 1.0 × 1015 cm3 in the Park model). (b) Forward reaction rate coefficients (the arrows represent the maximum electron–electronic excitation temperatures in figures 6a and 12e).

Figure 18

Figure 15. The dimensionless total rates of reactions related to generation of ${\textrm{e}^ - }$ along the stagnation line at $M{a_\infty } = 39$ using the Park model.

Figure 19

Figure 16. The dimensionless reaction rate of two electron-impact ionization reactions along the stagnation line at $M{a_\infty } = 39$ using the Park model. (a) Forward and backward reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$. (b) Total reaction rate of $\textrm{O} + {\textrm{e}^ - }\rightleftharpoons {\textrm{O}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$. (c) Forward and backward reaction rate of $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$. (d) Total reaction rate of $\textrm{N} + {\textrm{e}^ - }\rightleftharpoons {\textrm{N}^ + } + {\textrm{e}^ - } + {\textrm{e}^ - }$.

Figure 20

Figure 17. Forward reaction rate coefficient of two electron-impact ionization reactions in the Gupta and Park models (the red arrow represents the maximum electron temperature of the flow field around the cylinder under the corresponding $M{a_\infty }$ using the Gupta model and the green arrow represents that using the Park model).

Figure 21

Figure 18. Distributions of ${\textrm{e}^ - }$ and ${\textrm{O}^ + }$ along the stagnation line at $M{a_\infty } = 17$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model. Mass fractions of (a) ${\textrm{e}^ - }$ and (b) ${\textrm{O}^ + }$.

Figure 22

Table 5. Mesh information for cylinders at ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ with $M{a_\infty } = 17$.

Figure 23

Figure 19. Dimensionless energy exchange between electrons and other mode along the stagnation line at $M{a_\infty } = 17$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model.

Figure 24

Figure 20. Dimensionless energy exchange (a) ${\bar{Q}_{t - e\_ions}}$ and (b) ${\bar{\omega }_e}$ along the stagnation line at $M{a_\infty } = 17$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Gupta model.

Figure 25

Figure 21. Dimensionless energy exchange ${\bar{Q}_{t - e\_ions}}$ along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$. (a) Gupta model. (b) Park model.

Figure 26

Figure 22. Dimensionless energy exchange between electrons and other mode along the stagnation line at $M{a_\infty } = 27$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$. (a) Gupta model. (b) Park model.

Figure 27

Figure 23. Dimensionless energy exchange between electrons and other mode along the stagnation line at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model.

Figure 28

Figure 24. (a) Dimensionless energy exchange ${\bar{Q}_{t - e\_ions}}$, ${\bar{Q}_{i - e}}$ and ${\bar{\omega }_e}$ and (b) temperature ${T_e}$ along the stagnation line at $M{a_\infty } = 39$ with ${\rho _\infty }R = 1.3131 \times {10^{ - 5}}\;\textrm{kg}\;{\textrm{m}^{ - 2}}$ using the Park model.