This article introduces a novel model for describing the electronic excited states in the direct simulation Monte Carlo (DSMC) technique. The model involves the coupling the vibrational and electronic modes of molecular species, enabling each electronic excited state to excite its unique vibrational quantum levels. Numerical techniques are developed for equilibrium and post-collision sampling, as well as for measuring the internal temperature. The DSMC results demonstrate excellent agreement with theoretical predictions, providing verification of the successful implementation in a DSMC solver. For important thermophysical properties of molecular oxygen, such as the specific heat capacity, it is shown that the new model provides a better prediction than a compilation of past studies in comparison to the standard uncoupled approach in DSMC. The model is then applied to simulate a canonical nonreactive oxygen hypersonic flow past a cylindrical body. The population distribution of electronic excited states exhibit significant deviation from the standard approach typically used in the coupling between DSMC and radiation transport solvers.

In recent decades, the direct simulation Monte Carlo (DSMC) method has become one of the standard methods for simulating rarefied non-equilibrium flows.1,2 One reason for this is the relatively simple implementation of the method itself, along with the various physical and chemical models that can be implemented. Thus, with DSMC, it is possible to model very complex non-equilibrium effects in a flow, such as non-equilibrium in the internal degrees of freedom of molecules, i.e., rotation and vibrational energies, or non-equilibrium chemical processes, depending on the chosen physical models. It has long been known that excitation of the internal states strongly influences chemical processes, which in the case of dissociation and vibrational energy has already been modeled using the vibrationally favored dissociation (VDF) model.3 The link between the vibrational energy and the dissociation probability became even more pronounced with the introduction of the quantum kinetic (QK) chemistry models4—and its numerous extensions.5–7 In contrast, the electronic excitation of chemical species is sometimes disregarded, as some applications occur within temperature ranges in which the electronic excitation can still be neglected.8–10 

The assumption of neglecting the electronic mode depends on the chemical species involved in the flow. For instance, quasi-classical trajectory (QCT) calculations for the dissociation of oxygen conducted by Jo et al.11 reveal that dissociation reactions not only occur from the electronic ground state but also from electronic excited states. These QCT calculations demonstrate the importance of introducing a correction factor to account for the contribution of electronic excited states to reproduce the most representative experimental measurements. Similarly, the low electronic excitation energies observed in iron have substantial implications for flow physics, even at low temperatures, such as during meteorite reentry.12 For applications involving typical hypersonic reentry, e.g., Hayabusa,13 Stardust,14,15 and Fire II,16,17 ionization reactions and radiation are predominantly governed by the excitation of the electronic states.13–21 In these applications, the additional degrees of freedom of the electronic excitation modify the thermal conductivity, hence influencing the Prandtl number of the flow and the specific heats of the chemical species. Likewise, in applications involving plasma, e.g., electric propulsion systems or laser–plasma interactions, electronic excitation should not be disregarded, as it plays a significant role in the transport properties, chemical composition, and radiation processes.22–26 

In the DSMC community, four methods have been developed to model electronic excitation in flows. The first method has been introduced by Liechty and Lewis5,27 and treats the electronic excitation in a similar manner to the vibrational energies. In this method, each chemical species is initialized with an electronic quantum number sampled from the Boltzmann distribution. During collisions, energy exchange with the translational degrees of freedom occurs through the Larsen–Borgnakke (LB) model.28 The primary advantage of this technique lies in the storage of only one additional energy value per particle, while many aspects of vibrational modeling can be incorporated. However, a drawback of this method is that it exhibits significant noise, especially at low temperatures, requiring a large number of particles to compensate for the noise.

The second method, initially introduced by Bird,29 improved by Carlson et al.,30,31 and later expanded by Burt and Josyula,32,33 partially addresses this problem by equipping each particle with a complete distribution function of electronic excitation instead of an electronic quantum number. However, this leads to a significant increase in storage requirements.

The third method developed by Li et al.21 treats each transition between two electronic states as a chemical process during collisions, where each electronic excited state is regarded as a distinct species. However, to maintain a reasonable level of complexity, assumptions are made on the total number of electronic excited states involved. The merits of this method are that the transition rates can be tuned to reproduce any experimental measurements or high-fidelity calculations; albeit, noise remains a problem.

The fourth method has been designed by Gallis and Harvey.34,35 While the preceding methods rely on the assumption of an equilibrium distribution of the electronic state after a collision, this method computes the probability of an electronic excitation resulting from a collision with the aid of electronic excitation cross sections. To achieve a realistic description of the electronic excitation, accurate cross sections are required. However, the experimental measurement of accurate cross sections is extremely challenging and covers only a fraction of the electronic transitions of interest. Consequently, electronic excitation cross sections are typically either calculated from theoretical considerations or approximated from other known cross sections for similar interactions.

A common aspect of all methods is that each internal mode, i.e., rotational, vibrational, and electronic, is regarded separately. However, real-gas effects involve coupling between all internal modes, such as rotational–vibrational coupling and rotational–vibrational–electronic coupling.36 This becomes particularly significant in the analysis of molecular radiation, where achieving the correct distribution of vibrational and electronic excitation energy is crucial.

In recent years, coupling of DSMC methods with radiation transport solvers, e.g., PARADE,37 NEQAIR,38 or Specair,39 has emerged for such investigations.19,22,24,25,40,41 In contrast to the DSMC solvers, these radiation solvers adopt a fully coupled approach in which all types of interactions between the internal modes are incorporated. To perform such coupling, previous coupled radiation-DSMC simulations introduced additional assumptions, i.e., Boltzmann distribution19,41 or quasi-steady-state (QSS),14,21,40,42,43 to transfer uncoupled information from DSMC into the coupled models of radiation solvers. With an ever-increasing number of missions involving spectral studies of flows during in-flight or experimental measurements, a detailed description of the physics of the flow is fundamental to allow for spectral comparison. However, a limitation arises due to the assumption of decoupling each internal mode of the chemical species in DSMC simulations, preventing the achievement of this level of detail.

The novelty of the present article lies in the development of a model for coupling the electronic and vibrational modes of molecular species in DSMC. The model involves the development of new numerical techniques for the initialization of particles in the domain, the redistribution of the internal energy after a collision, and the measurement of the corresponding internal temperature. Considering the challenge of performing experiments to measure chemical processes involving electronic excited states, the new model is verified against an extensive compilation of theoretical studies for the reproduction of thermophysical properties of molecular oxygen. Finally, the model is applied for a canonical nonreactive oxygen hypersonic flow past an infinite cylindrical body at an altitude of 85 km in Earth's atmosphere.

The rest of the paper is organized as follows: Section II describes the new model for the coupling of the electronic and vibrational modes. In Sec. III, the numerical techniques for the implementation of the new model in a DSMC solver are developed. Section IV provides a comprehensive verification of the model and demonstrates its application under typical hypersonic flow conditions. The final section provides conclusions and perspectives for future work.

The derivation of the internal energies of a molecular system, i.e., rotational, vibrational, and electronic, involves resolving the appropriate Schrödinger equation for a given total energy, which is represented by the Hamiltonian operator
(1)
where T and V are the kinetic and potential energy operators, respectively.
The resolution of the Schrödinger equation is challenging with analytical solutions for the vibrational excitation only possible for the traditional harmonic oscillator (HO) model, or the anharmonic oscillator (aHO) model calculated with the Morse potential energy function.44 To simplify the derivation of the time-independent Schrödinger equation, the Born–Oppenheimer approximation is typically employed.45 This assumption allows each mode to be treated independently, assuming no interaction between internal modes. It leads to the total mean energy of the system being calculated as the summation of independent contributions
(2)
where subscripts (.)t,(.)r,(.)v, and (.)e refer to the translational, rotational, vibrational, and electronic modes of a molecular system, respectively.

This approximation is traditionally applied in DSMC simulations, i.e., each mode of a molecular system is regarded as insensitive to any other excitation. In the current section, the traditional approach for DSMC simulations is reviewed, and a new model involving the coupling between the vibrational and electronic modes is presented.

The traditional approach imposes that each mode of the molecule is individually treated assuming no interactions with one another. This approach takes into account the translational, rotational, vibrational, and electronic modes.

For the description of the rotational energy, the common technique46–48 is to consider the rigid-rotor model in which the rotational energy yields
(3)
where k is the Boltzmann constant, J is the rotational quantum level, and θr is the characteristic rotational temperature
(4)
where is the reduced Planck' constant, μ is the reduced mass, req is the inter-nuclei equilibrium distance, and J is the rotational quantum number. In addition, each rotational quantum level J is degenerated gJ=2J+1 times.

For the most dominant chemical species in Earth's atmosphere, the characteristic rotational temperatures are typically very low, e.g., θr,O2=2.064 K, θr,N2=2.869 K, and θr,NO=2.470 K.49 Therefore, the classical limit Tθr is applicable in most DSMC simulations, allowing for continuous treatment of the rotational mode.

For the vibrational mode, the traditional approach1,48,50 consists of modeling the vibrational excitation with the harmonic oscillator (HO) model in which the vibrational energy is formulated as
(5)
where h is the Planck constant, c the speed of light in vacuum, and ωe,0 is the harmonic oscillator spectroscopic constant of the ground electronic state.
A more detailed approach51–53 consists of describing the vibrational excitation of molecular systems with an anharmonic oscillator (aHO) model that is, for the Morse-aHO model
(6)
where ωe,0χe,0 is the first anharmonic oscillator spectroscopic constant of the ground electronic state. Note that for both vibrational models, each vibrational quantum level is non-degenerate, i.e., gv = 1.

The maximum vibrational quantum number, vmax, allowed by Eqs. (5) and (6), is calculated as the vibrational quantum number immediately before the vibrational energy reaches the dissociation limit. In this approach, it is important to emphasize that each electronic excited state can only excite the same set of vibrational levels as the ground electronic state.

Additionally, the energies of the electronic states are calculated as
(7)
where κe,i is the electronic transition of state i. Note that all electronic states have a distinct degeneracy corresponding to their electronic configuration.45,49
For a given diatomic molecule under thermal equilibrium, statistical mechanics theory54 demonstrates that, under the Born–Oppenheimer approximation, the internal energy is distributed according to the Boltzmann distribution, that is,
(8)
where T is the temperature, subscripts i, j, k, and l refer to the translational, rotational, vibrational, and electronic modes, respectively, and gi, gj, gk, and gl are the degeneracy of the translational, rotational, vibrational, and electronic modes, respectively.
Following this approach, the total partition function is defined as
(9)
where Qt, Qr, Qv, and Qe are the partition function of the translational, rotational, vibrational, and electronic modes, respectively.
It follows that the contribution of each mode to the total mean energy, Eq. (2), of such a system is expressed as
(10)
Finally, the specific heat capacity is directly connected to the partition function as
(11)
where R is the specific gas constant of the molecular species.

The evaluation of the thermodynamic properties of such a system becomes a matter of addressing each mode individually; hence, it is denominated the uncoupled approach in the current work. Note that each internal mode, i.e., rotational, vibrational, and electronic, is treated separately resulting in three one-dimensional distribution functions. It is important to emphasize that the Boltzmann distribution is limited to discrete energy levels. In the present article, the translational and rotational modes are treated continuously, therefore, their distribution functions are described by the Hinshelwood function.55 

The second approach assumes a coupling between the vibrational and electronic modes which carries significant implications. Specifically, each electronic excited state has a unique set of vibrational levels described by
(12)
where κe,i,ωe,i and ωe,iχe,i are the spectroscopic constants of electronic state i.

Considering an electronic state i, the maximum vibrational quantum level permissible is determined as the vibrational quantum immediately before the dissociation energy of electronic state i, i.e., εve(i)<εd(i). For some electronic configurations, all the vibrational quantum levels lie below the corresponding dissociation energy. In this specific scenario, the maximum vibrational quantum level is determined as the last vibrational quantum level before the gradient εve(i)j changes sign.

Since the electronic excited states allow for vibrational excitation, the corresponding distribution function for the vibrational-electronic mode consists of a coupled Boltzmann distribution such that
(13)
where Ni,j is the number of particles in electronic excited state i and vibrational quantum level j, and Qve is the partition function of the vibrational-electronic mode.
Additionally, adjusting the distribution function of the internal quantum numbers has implications for the thermophysical properties. As shown in Eq. (13), the partition function becomes
(14)
with the mean internal energy formulated as
(15)
and the isochoric specific heat capacity
(16)

As illustrated by Eqs. (8) and (13), the two approaches yield distinct distribution functions for the vibrational and electronic modes. The uncoupled approach uses a single one-dimensional distribution function for each mode, whereas the coupled approach introduces a two-dimensional distribution function that accounts for the coupling between the vibrational and electronic modes. This fundamental difference between the two approaches leads to significant consequences. Specifically, in the new coupled approach, a unique partition function, mean internal energy, and specific heat capacity can be defined. From a numerical perspective, it imposes the definition of a unique mean internal temperature, i.e., Tve. This contrasts with the traditional uncoupled approach which requires separate internal temperatures for each mode, i.e., Tv and Te.

As highlighted in Sec. II, the coupled approach has a two-dimensional distribution function that requires sampling of both the electronic and the vibrational quantum numbers simultaneously. The coupled approach is implemented in dsmcFoam+ based on the procedure reported by Liechty and Lewis.5 Three procedures are revisited to accommodate the vibrational excitation of the electronic excited states. These functions are the equilibrium sampling, the post-collision sampling, and the measurement of the corresponding mean temperature.

To initially populate the numerical mesh, each particle is assigned a set of properties, including its quantum numbers. The distribution of these quantum numbers follows the Boltzmann distribution, Eq. (13). However, a set of quantum numbers, i.e., (i, j), cannot be directly sampled from Eq. (13); hence, an acceptance–rejection scheme is employed. This scheme involves selecting a pair of quantum numbers from the normalized distribution
(17)
where fmax is the normalization factor, and (is, js) are the electronic and vibrational quantum numbers, respectively, for which Eq. (13) has a maximum. Note that the degeneracy differs from one electronic excited state to another, see Table I. As a result, the maximum of Eq. (13) cannot be determined beforehand but must be searched for in each iteration.
TABLE I.

Spectroscopy constants of the electronic excited states for molecular oxygen.

States κe (eV) g ωe (cm−1) ωeχe (cm−1) θv (K) θr (K) εD (eV) jmax
X3Σg  0.0  1688.17  16.43  2423.91  2.06  5.2  47 
a1Δg  0.98  1483.5  12.9  2130.05  2.05  5.2  33 
b1Σg+  1.63  1432.8  14.0  2057.25  2.01  5.2  29 
c1Σu  4.09  794.2  12.73  1140.33  1.31  5.2  16 
A3Δu  4.29  850.0  20.0  1220.45  1.38  5.2  13 
A3Σu+  4.38  799.07  12.16  1147.32  1.31  5.2  10 
B3Σu  6.16  709.3  10.65  1018.43  1.18  7.16  15 
States κe (eV) g ωe (cm−1) ωeχe (cm−1) θv (K) θr (K) εD (eV) jmax
X3Σg  0.0  1688.17  16.43  2423.91  2.06  5.2  47 
a1Δg  0.98  1483.5  12.9  2130.05  2.05  5.2  33 
b1Σg+  1.63  1432.8  14.0  2057.25  2.01  5.2  29 
c1Σu  4.09  794.2  12.73  1140.33  1.31  5.2  16 
A3Δu  4.29  850.0  20.0  1220.45  1.38  5.2  13 
A3Σu+  4.38  799.07  12.16  1147.32  1.31  5.2  10 
B3Σu  6.16  709.3  10.65  1018.43  1.18  7.16  15 

Figure 1 illustrates the general procedure to assign the initial electronic and vibrational quantum numbers. The first step consists of searching for the electronic and vibrational quantum numbers, i.e., (is, js), for which Eq. (13) has a maximum. Then, a pair of electronic and vibrational quantum numbers, i.e., (i,j), uniformly distributed between 0 and imax1 and 0 and jmax1, respectively, are independently chosen randomly. Finally, an acceptance–rejection scheme is used to select a pair of electronic and vibrational quantum numbers, i.e., (i,j), from the distribution, Eq. (17) that satisfies f>R(0,1).

FIG. 1.

Flowchart of the equilibrium sampling function implemented in dsmcFoam+.

FIG. 1.

Flowchart of the equilibrium sampling function implemented in dsmcFoam+.

Close modal
In the DSMC method,1 the internal energies are commonly redistributed through a serial application of the quantum Larsen–Borgnakke (LB) method.28 The LB method samples the post-collision quantum numbers, i.e., i and j, from a combined distribution of the translational and the vibrational-electronic mode of the colliding particle. The translational distribution function inherently depends on the inter-molecular model employed as it defines the collision probability. In the present article, the base collision scheme is the variable hard sphere (VHS) collision model developed by Bird.1 For the VHS model, the probability of distribution of translational energy during a collision is
(18)
where Γ(x) is the ordinary gamma function, and ωA,B is the mean viscosity exponent between colliding partners A and B.
Following the approach of Bergemann and Boyd,28 a Dirac delta function is applied to the Boltzmann distribution, Eq. (13), to define the following continuous distribution function:
(19)
Using Eqs. (19) and (18), the combined distribution for sampling post-collision quantum vibrational and electronic levels from the collision energy, εc=εt+εi,j=εt+εi,j, is
(20)
The LB scheme assumes that local thermodynamic equilibrium prevails and the collision energy remains constant across the redistribution process. Under these assumptions, the sampling of the post-collision quantum numbers can therefore be performed on a simplified distribution, i.e.,
(21)
For similar reasons to that of the equilibrium sampling, an acceptance–rejection scheme is applied. The normalization of the distribution is similarly obtained by searching for the pair of quantum numbers for which Eq. (21) has a maximum. The acceptance–rejection scheme is therefore conducted on
(22)
where hmax is the normalization factor that corresponds to the maximum of Eq. (21).
Figure 2 presents the general procedure for assigning post-collision quantum numbers within the dsmcFoam+ framework. Two inelastic mechanisms are distinguished depending on whether the particle experiences a vibrational excitation/de-excitation, Reaction 1
(Reaction 1)
or electronic and vibrational excitation/de-excitation, Reaction 2
(Reaction 2)
FIG. 2.

Flowchart of the post-collision sampling function implemented in dsmcFoam+.

FIG. 2.

Flowchart of the post-collision sampling function implemented in dsmcFoam+.

Close modal

In Reaction 1, the energy exchange occurs from a change in the vibrational quantum number. In Reaction 2, the energy exchange involves an electronic transition, see Table I, which, in turn, also necessitates a modification of the vibrational quantum number.

Typically, in DSMC simulations, a relaxation probability of Pv=0.02 is allowed to result in vibrational excitation/de-excitation while a relaxation probability of Pe=0.002 is allowed for electronic energy exchange.29 For the coupled approach to reproduce the two separate redistributions of internal energy in the uncoupled approach, two relaxation mechanisms, Reaction 1 and Reaction 2, are implemented in dsmcFoam+. Each mechanism operates with a relaxation probability, namely, P1 for Reaction 1 and P2 for Reaction 2. Note that the relaxation probability P2 is a conditional probability on P1 to be true, see Fig. 2. Therefore, to reproduce the two separate redistributions of internal energy, P1 must be set to 0.02 and P2 to 0.1. Note that these relaxation probabilities are user-defined; hence, they can be modified to reproduce any baseline relaxation model.

The energy exchange involving vibrational energy, i.e., Reaction 1, follows the quantum LB approach. A detailed explanation for the vibrational excitation/de-excitation during an inelastic collision under the assumption of an anharmonic oscillator model can be found in Civrais et al.51 For brevity, the present section focuses on the redistribution of internal energy involving an electronic and vibrational excitation/de-excitation.

If an inelastic collision is accepted for Reaction 2 energy exchange, the first step is to search the maximum allowed vibrational quantum number for each electronic excited state, i.e., (is, js), that satisfies εc<εis,js and to determine a pair of quantum numbers for which the distribution function, Eq. (22), is maximum. Then, an acceptance–rejection procedure is performed to sample (i,j) from the distribution, Eq. (22). Finally, a pair (i,j) is accepted if Eq. (22) satisfies h>R; otherwise, the procedure is repeated until values for (i,j) are obtained.

Finally, a point should be made about the calculation of the vibrational-electronic temperature. In the uncoupled approach, the vibrational and electronic modes, i.e., Tv and Te, are treated separately leading to the one mean temperature for each mode. To maintain mathematical consistency with the two-dimensional distribution function, Eq. (13), a unique temperature, Tve, must be defined for the vibrational-electronic mode. The difference between the two approaches will be assessed in Secs. IV A–IV C.

The vibrational-electronic temperature ultimately represents a mean excitation of the vibrational-electronic mode. Since the electronic mode admits a set of vibrational levels, the vibrational-electronic temperature is unable to reflect the vibrational excitation of each electronic excited state but, rather, a mean excitation of the electronic excited states and their corresponding vibrational quantum levels. Despite its minimal physical insights, the vibrational-electronic temperature serves the purpose of retaining the mean energy of a molecular system, making it valuable for verification exercise.

Furthermore, in the derivation, Eqs. (13)–(16), no assumption has been made regarding thermal equilibrium between the vibrational and electronic modes. Thus, these modes can exhibit thermal non-equilibrium with each other. These considerations contrast with the definition of Tve employed in the vast majority of multi-temperature extended Navier–Stokes equations solvers based on Gnoffo;56 albeit, there exist high-speed and high-temperature computational fluid dynamics (CFD) methods that do not impose thermal equilibrium between modes.57–59 

In the uncoupled approach, where vibrational excitation is modeled with an infinite harmonic oscillator, the mean vibrational degrees of freedom can be computed using the equipartition theorem, and the vibrational temperature derived analytically from the mean vibrational quantum number. However, with an anharmonic oscillator model, the vibrational temperature cannot be derived analytically.51 Specifically, it must be determined by solving an implicit equation. Similarly, to account for the vibrational excitation of the electronic excited states, the vibrational-electronic temperature cannot be directly calculated using the equipartition theorem. A similar approach to that described by Civrais et al.51 is herein adopted.

Consider a volume filled with chemical species s and a corresponding mean vibrational-electronic energy denoted Es. Calculating the corresponding vibrational-electronic temperature involves resolving
(23)

This implicit equation, Eq. (23), is numerically resolved with a Newton iterative approach. It involves initially evaluating Eq. (15) at a guessed temperature and comparing it to the mean vibrational-electronic energy in a cell. The procedure is repeated until a user-defined tolerance factor is reached.

A series of adiabatic reactor simulations is conducted to verify the derivation and implementation of the coupled approach. A single cubic cell with edge length 1.88×104 m filled with 1 × 106 DSMC simulator particles and periodic boundaries is used for this purpose. A fixed time step size of 1×109 s is adopted. The working gas is molecular oxygen with the electronic excited states summarized in Table I. Collision partners are selected according to the No Time Counter (NTC) model.1 The inter-molecular collisions are processed with the variable hard sphere (VHS) model1 with the properties at a reference temperature Tref of 273 K, i.e., m=48.12×1027 kg, d=4.07×1010 m, and ω=0.77. The probabilities for a particle to experience Reaction 1 or Reaction 2 are those presented in Sec. III unless stated otherwise. The simulations are performed in serial on a machine equipped with 12 Intel® Core™ i7-10850H central processing unit (CPU) cores (base: 2.70 GHz, max: 5.1 GHz).

In the present article, the seven lowest electronic excited states are considered.60–62 Note that additional electronic excited states are reported by Liu et al.;60 however, the decision has been made to be consistent with the number of electronic excited states engaged in previous DSMC studies.63 The spectroscopy constants are extracted from the National Institute for Standards and Technology (NIST) database49 with the exception of the ground state, i.e., X3Σg, which is extracted from Civrais et al.7 The spectroscopic constants of the electronic excited states of O2 are summarized in Table I. In addition, it is assumed that the electronic excited states conserve the collision properties of the ground state; hence, the mass, m, diameter, d and viscosity exponent, ω, of the electronic excited states are those of the ground state.

1. Vibrational-electronic temperature

The first verification involves the measurement of the vibrational-electronic temperature. The adiabatic reactor is initialized in thermal equilibrium conditions for temperatures ranging between 2000 and 20 000 K. The vibrational-electronic temperature is monitored and sampled for 105 iterations.

Table II shows that the measurement technique implemented for the calculation of the vibrational-electronic temperature is accurate across the temperature range considered. The numerical scatter of the measured temperature is inversely proportional to the square root of the sampling size;2 hence, for a sampling size of about 106, the numerical scatter is about 0.1%. Table II shows that the difference between the initialized and measured temperatures reduces as expected. This demonstrates the correct implementation of both the initialization algorithm and the new measurement technique.

TABLE II.

Comparison of the measured vibrational-electronic temperature against the initialized temperature.

Initialized temperature (K) Measured temperature (K) Error (%)
2 000  1 999.02  0.05 
4 000  4 000.89  −0.02 
6 000  6 002.21  −0.04 
8 000  7 998.08  0.02 
10 000  10 005.26  −0.05 
12 000  12 003.56  −0.03 
14 000  14 002.66  −0.02 
16 000  16 001.78  −0.01 
18 000  18 008.9  −0.05 
20 000  20 003.35  −0.02 
Initialized temperature (K) Measured temperature (K) Error (%)
2 000  1 999.02  0.05 
4 000  4 000.89  −0.02 
6 000  6 002.21  −0.04 
8 000  7 998.08  0.02 
10 000  10 005.26  −0.05 
12 000  12 003.56  −0.03 
14 000  14 002.66  −0.02 
16 000  16 001.78  −0.01 
18 000  18 008.9  −0.05 
20 000  20 003.35  −0.02 

2. Adiabatic relaxation

The next verification case assesses the capabilities of the model to reproduce the theoretical thermal equilibrium of a system initially in a state of thermal non-equilibrium. Two scenarios are herein considered. The adiabatic reactor is initialized with a vibrational-electronic temperature greater or lower than the translational and rotational temperatures such that excitation or de-excitation of the electronic excited states and their vibrational quantum levels occurs as a result of the relaxation process. The two scenarios are of major interest in hypersonic flow conditions as the former is encountered in post-shock flow conditions (Tt > Tve) and the latter is typically encountered in nozzle flow conditions (Tt < Tve). For the excitation scenario, the adiabatic reactor is initialized in the following thermal non-equilibrium conditions: Tt=Tr=20000 K and Tve = 5000 K, and for the de-excitation scenario with Tt=Tr=5000 K and Tve=20000 K. Under such conditions, a relaxation process toward a thermal equilibrium state is driven by inelastic collisions between particles and the corresponding exchange of internal energies. After a number of collisions have occurred, kinetic theory demonstrates that a system relaxes toward an equilibrium temperature that is defined as
(24)
where ξt, ξr, and ξve are the translational, rotational, vibrational and electronic mean degrees of freedom, respectively, and the subscripts (.)i and (.)f refer to the initial and final times.

Figure 3 demonstrates that the DSMC results achieve excellent reproduction of the theoretical predictions for both the excitation and de-excitation scenarios. Quantitatively, for the excitation scenario, the theoretical equilibrium temperature is Teq,extheo=13740 K and the equilibrium temperature measured in DSMC is Teq,exDSMC=13741 K. For the de-excitation scenario, the theoretical equilibrium temperature is Teqtheo=10521 K, and the equilibrium temperature measured in DSMC is TeqDSMC=10518 K. Both scenarios show good reproduction of the theoretical prediction that demonstrates the successful implementation of the post-collision sampling technique.

FIG. 3.

Thermal relaxation of an adiabatic reactor to equilibrium for (a) electronic and vibrational excitation (Tt=Tr=20000 K and Tve = 5000 K) and (b) electronic and vibrational de-excitation (Tt=Tr=5000 K and Tve=20000 K).

FIG. 3.

Thermal relaxation of an adiabatic reactor to equilibrium for (a) electronic and vibrational excitation (Tt=Tr=20000 K and Tve = 5000 K) and (b) electronic and vibrational de-excitation (Tt=Tr=5000 K and Tve=20000 K).

Close modal

3. Population of the electronic excited states

Figure 4 shows further verification of the model for the population of the electronic excited states and their corresponding vibrational quantum numbers. The adiabatic reactor is initialized in thermal equilibrium conditions, i.e., Tt=Tr=Tve=10000 K, and the population of each quantum number is sampled for 105 time steps to reduce the scatter. The theoretical and DSMC two-dimensional distribution of energy across all electronic excited states and their corresponding vibrational quantum levels are depicted in Figs. 4(a) and 4(b), respectively. Additionally, the population of the electronic excited states and vibrational quantum levels are, respectively, depicted in Figs. 4(c) and 4(d).

FIG. 4.

Comparison of the population of the electronic excited states for molecular oxygen in thermal equilibrium (Tt=Tr=Tve=10000 K).

FIG. 4.

Comparison of the population of the electronic excited states for molecular oxygen in thermal equilibrium (Tt=Tr=Tve=10000 K).

Close modal

Figures 4(a) and 4(b) show that the DSMC achieves a good reproduction of the two-dimensional theoretical distribution for all quantum levels. Figure 4(c) demonstrates an excellent agreement between the population of the electronic excited states and statistical mechanics theory. The consistency between DSMC results and the theoretical calculations, Eq. (13), serves as a further verification of the implementation of the model in the DSMC code. Similarly, Fig. 4(d) also confirms the accuracy of the model for the population of the vibrational quantum numbers in all seven electronic excited states. Some scatter is evident at the tails of the distributions. This is expected in a DSMC simulation, because the probability of finding a molecule in these higher vibrational levels is relatively small, resulting in a low signal-to-noise ratio.

4. Thermophysical properties

The uncoupled and coupled approaches are examined for the calculation of the isochoric-specific heat capacity. The adiabatic reactor is initialized in thermal equilibrium for temperatures ranging from 2000 to 20 000 K, and the thermophysical properties are sampled for 105 time steps to reduce the scatter. These results are compared against four numerical studies, i.e., Capitelli et al.,64 Jaffe,65 Qin et al.,66 and McBride et al.67 This represents a selection of many works available in the literature. Additionally, the specific heat capacity results presented in Civrais et al.53 are included for reference.

Figure 5 compares the uncoupled and coupled approaches for the calculation of the isochoric specific heat capacity of molecular oxygen. The DSMC results are in excellent agreement with the theoretical calculations, Eq. (16), for the entire range of temperatures considered which serves as a verification of the implementation of the model in the DSMC code. Figure 5 illustrates the benefits of the coupled approach. This approach provides a more accurate description of the specific heat capacity compared to its counterparts. Figure 5 emphasizes the importance of including the electronic mode of molecular systems and modeling their corresponding vibrational excitation to yield accurate predictions in low to moderate temperatures.

FIG. 5.

Comparison of the normalized isochoric specific heat capacity of molecular oxygen.

FIG. 5.

Comparison of the normalized isochoric specific heat capacity of molecular oxygen.

Close modal

For temperatures in excess of 7000 K, the coupled approach deviates from the prediction of past studies.64–67 In the derivation of the coupled approach, the rotational–vibrational coupling effects of the electronic excited modes have been disregarded. While this assumption holds for low-to-moderate temperatures, it becomes evident that as the temperature increases the coupled approach deviates from the compilation of past studies due to the omission of rotational–vibrational coupling effects. To improve the accuracy of the coupled approach, further investigation should be directed toward incorporating the contribution of rotational-vibrational coupling effects.68–71 

The present section aims to compare the uncoupled and coupled approaches for the simulation of a canonical nonreactive oxygen hypersonic flow past a cylindrical body entering Earth's atmosphere at an altitude of about 85 km. The free stream conditions are extracted from the U.S. Standard Atmosphere72 and summarized in Table III.

TABLE III.

Free stream parameters.

Parameter Values
Altitude (km)  85 
Temperature (K)  186.95 
Number density (m−3 1.45×1020 
Pressure (Pa)  0.374 
Speed (km s−1 7.50 
Mach number  28.76 
Knudsen number  0.085 
Parameter Values
Altitude (km)  85 
Temperature (K)  186.95 
Number density (m−3 1.45×1020 
Pressure (Pa)  0.374 
Speed (km s−1 7.50 
Mach number  28.76 
Knudsen number  0.085 

The geometry represents a two-dimensional slice of a cylindrical body with diameter d = 0.1 m, with a domain length equal to two and a half radii upstream of the stagnation point. The mesh is refined near the stagnation point to ensure that the cell size, Δx, remains around one-quarter of the local mean free path, λ, throughout, resulting in a total of 32 400 cells. The time step is carefully chosen to be an order of magnitude smaller than both the mean collision time and the cell residence time, the latter of which corresponds to the time for a DSMC particle to cross a cell length under the freestream conditions. To maintain a minimal number of 100 particles per cell throughout, the numerical mesh is populated with a total number of 8.59×106 DSMC particles at steady state. Similar to the adiabatic reactor simulations, the inter-molecular collisions are computed with the VHS model1 for a reference temperature of Tref = 273 K. Collision partners are selected according to the NTC model.73 To reproduce the typical relaxation behavior of the uncoupled approach, a constant probability of P1=0.02 and P2=0.1 for a particle to undergo Reaction 1 and Reaction 2, respectively, is applied. The gas–surface interactions are fully diffusive with an isothermal wall at temperature Tw = 1000 K. The simulations are performed in parallel on 8 cores on the aforementioned machine and over 105 samples are taken after steady-state to reduce the numerical scatter. To illustrate the structure of the flow, the internal temperatures and velocity magnitude along the stagnation streamline are presented in Fig. 6. The distance to the stagnation point is normalized by the diameter of the cylinder. A value of 0 refers to the inlet boundary conditions and 2.5 refers to the stagnation point.

FIG. 6.

Flow properties along the stagnation streamline.

FIG. 6.

Flow properties along the stagnation streamline.

Close modal

As mentioned previously, the standard approach to couple DSMC and radiation transport solvers consists of transferring the uncoupled internal temperatures, i.e., rotational, vibrational, and electronic, calculated by the DSMC solver into a radiation solver. Note that these internal temperatures are calculated by assuming a Boltzmann distribution and resolving an implicit equation for both vibrational and electronic temperatures.51 However, radiation solvers, e.g., NEQAIR,38 PARADE/PICLas,19,37 and Specair,39 integrate a fully coupled approach. Specifically, these solvers compute the radiative properties in a cell from the distribution function of the internal quantum levels. To compute these radiative properties, an assumption on the distribution function is made where it is assumed to follow Boltzmann statistics19,41 or the non-equilibrium quasi-steady-state method.21,40

The main objective of this section is twofold: to compare the two approaches for the distribution of the internal quantum levels measured by the DSMC method and to assess the assumption that the internal quantum levels are distributed according to the Boltzmann distribution. For both approaches, the vibrational and electronic density functions are calculated with the local internal temperatures, namely, f=fi(Tv)fj(Te) for the uncoupled approach and f=fi,j(Tve) for the coupled approach.

Figure 7 compares the DSMC results for the population of the internal quantum levels and the corresponding Boltzmann distribution along the stagnation streamline. The DSMC results are denoted by color markers, where solid symbols refer to the uncoupled approach and open symbols to the coupled approach, and the Boltzmann distributions are denoted by color lines where solid lines refer to the uncoupled approach and dashed lines refer to the coupled approach. Similar to Fig. 6, the distance to the stagnation point is normalized by the diameter of the cylinder. A value of 0 refers to the inlet boundary conditions and 2.5 refers to the stagnation point.

FIG. 7.

Comparison of the DSMC results against Boltzmann distribution along the stagnation streamline.

FIG. 7.

Comparison of the DSMC results against Boltzmann distribution along the stagnation streamline.

Close modal

Downstream of the shock wave, Fig. 7 shows that the highest vibrational and electronic quantum levels exhibit large deviations from the assumption of a Boltzmann distribution. These high-lying quantum levels are of primary importance for the modeling of the radiative properties of molecular species. Furthermore, a noticeable discrepancy arises between assuming a Boltzmann distribution corresponding to the cell temperature and density, and the actual distribution observed in DSMC simulations, which represents a source of error in typical flow-radiation coupling.

Figure 7 also indicates a significant population of the high-lying quantum levels in the upstream region. This non-equilibrium effect manifests itself in Knudsen number regimes in which the degree of rarefaction is sufficient for a particle to collide with the surface of the body and travel backwards without undergoing sufficient collisions with any other particles to de-excite the quantum levels to return to an equilibrium distribution; hence, carrying post-shock information upstream of the bow-shock. This non-equilibrium effect is characterized by a strong deviation from a Maxwell velocity distribution to the extent of quasi-Maxwellian or bi-modal distributions.21 Note that this phenomenon has also been observed previously.12,18,20,41

The deviation from the Boltzmann distribution is further investigated at four locations throughout the shock wave denoted by vertical red lines in Fig. 6 with thermal conditions summarized in Table IV. A series of adiabatic reactor simulations initialized with the thermal conditions summarized in Table IV is performed. Figure 8 shows the Boltzmann distribution and DSMC measurements of the coupled approach for the first three electronic states of O2 obtained along the stagnation line. Contrastingly, Fig. 9 shows the Boltzmann distribution and DSMC measurements obtained from the series of adiabatic reactor simulations. The DSMC results are denoted by color markers and the Boltzmann distributions by color lines.

TABLE IV.

Thermal flow conditions in four cells along the stagnation streamline.

xD Tt (K) Tr (K) Tve (K)
0.500  1 902.70  423.40  550.06 
1.000  5 813.31  1059.91  795.58 
1.500  17 168.40  3273.69  1311.87 
2.000  31 031.00  9390.44  2518.74 
xD Tt (K) Tr (K) Tve (K)
0.500  1 902.70  423.40  550.06 
1.000  5 813.31  1059.91  795.58 
1.500  17 168.40  3273.69  1311.87 
2.000  31 031.00  9390.44  2518.74 
FIG. 8.

Vibrational density function of the first three electronic excited states at four locations along the stagnation streamline.

FIG. 8.

Vibrational density function of the first three electronic excited states at four locations along the stagnation streamline.

Close modal
FIG. 9.

Vibrational density function of the first three electronic excited states measured in a series of adiabatic reactor simulations with thermal conditions corresponding to four locations along the stagnation streamline.

FIG. 9.

Vibrational density function of the first three electronic excited states measured in a series of adiabatic reactor simulations with thermal conditions corresponding to four locations along the stagnation streamline.

Close modal

Figure 8 shows that the Boltzmann distributions at all four locations are significantly different to the DSMC measurements along the stagnation streamline, whereas Fig. 9 indicates that the DSMC simulations precisely reproduce the theoretical vibrational density function for the same thermal conditions in an adiabatic reactor. This demonstrates that the different distributions in the hypersonic case are due to non-equilibrium effects that a radiation solver would not be capturing using solely the temperatures.

So far, the comparison of the two approaches was limited to the population of the quantum levels along the stagnation streamline. In Fig. 10, the total number density of the first six electronic states of O2, i.e., ni=jni,j, measured in the DSMC simulations are regarded for the flow field.

FIG. 10.

Population of the first four electronic states of O2. Uncoupled approach (top) and coupled approach (bottom).

FIG. 10.

Population of the first four electronic states of O2. Uncoupled approach (top) and coupled approach (bottom).

Close modal

As anticipated, Fig. 10 indicates that the ground electronic state is the most populated. Both approaches offer a similar description of the distribution of the ground state throughout the domain. However, the two approaches differ in the distribution for all the electronic excited states herein considered. Specifically, the coupled approach predicts lower populations of the electronic excited states in comparison to the uncoupled approach. This discrepancy arises from the difference in the modeling of the vibrational energies of the electronic excited states in each approach. The coupled approach implies that each electronic state can excite a unique set of vibrational levels, see Fig. 4. In contrast, the uncoupled approach assumes that each electronic state can only excite the same set of vibrational energies as the ground electronic state. In the context of a flow-radiation coupling, such a difference between the two approaches in terms of the distribution of the electronic excited states throughout the domain could lead to a significant difference in the radiation properties.

While the coupled approach has demonstrated good reproduction of theoretical calculations, see Figs. 3–5, for hypersonic flow past a cylindrical body, see Figs. 6–10, the approach has limitations that merit additional discussions.

The first limitation concerns the omission of the rotational–vibrational coupling. Figure 5 has evidenced the importance of the rotational–vibrational coupling effects for the reproduction of the most representative database.64–71 There exist various approaches to leverage this limitation. One approach for modeling the rotational–vibrational coupling consists of a continuous treatment of the rotational mode and a quantized treatment of the vibrational and electronic modes.74 This approach models the elongation of the bond length due to vibrational excitation in the rotational partition function. Specifically, it assumes that each vibrational quantum level has its own rotational and dissociation energies. The benefits of this approach lie in its capacity to provide a rotational–vibrational coupling without increasing the number of quantum levels considered. Another approach assumes a quantized treatment of the internal modes, i.e., rotational, vibrational, and electronic modes.75 This approach models triplets of quantum levels that lie in the bounded and quasi-bounded areas.64,65 While this approach offers a detailed representation of the dynamics of the internal quantum levels, it requires the addition of numerous quantum levels to the molecular systems, resulting in a significant increase in computational expense. For a typical air mixture simulation modeled with the five most dominant species, i.e., O2, N2, NO, O, and N, this approach requires ∼100 000 internal quantum numbers.

The second limitation concerns the chemical activity of the electronic excited states. In the current work, the chemical activity of the electronic excited states has consciously been disregarded. The inherent lack of experimental measurements of chemical processes involving electronic excited states of molecular species for temperatures relevant for hypersonic flow conditions does not allow the development of more sophisticated models. Such an issue would ideally be leveraged by QCT calculations on non-adiabatic potential energy surfaces, incorporating electronic excited state transitions. However, state-of-the-art QCT calculations of relevant species for Earth reentry are, as yet, limited to ground electronic configurations.11 Therefore, non-adiabatic QCT calculations are required to allow for model development.

A novel model for modeling the vibrational excitation of the electronic excited states is developed and implemented in a DSMC solver. The model is verified against an extensive compilation of theoretical studies for both thermal equilibrium and non-equilibrium conditions. Additionally, it is evaluated against the traditional uncoupled approach utilized in DSMC for a canonical hypersonic non-reactive oxygen gas flow past a cylindrical body at an altitude of about 85 km.

The series of adiabatic reactor simulations have shown excellent agreement between theoretical predictions and DSMC results, thereby demonstrating the successful implementation in a DSMC solver. Furthermore, the coupled approach has been shown to improve the reproduction of the specific heat capacity of molecular oxygen compared to the uncoupled approach.

The hypersonic flow scenario has revealed that the assumption of the quantum levels being distributed according to the Boltzmann distribution is invalid. In high Knudsen number regimes, non-equilibrium effects lead to significant deviations between the local distribution of quantum levels and that predicted by the Boltzmann distribution calculated from internal temperatures. Consequently, one should benefit from the capability of DSMC solvers to keep track of the local distribution of energy across the internal quantum levels and solely rely on these distributions for the determination of the radiative properties. Furthermore, the coupled approach has demonstrated a lower population of electronic excited states compared to the uncoupled approach. This discrepancy is attributed to the possibility for each electronic excited state to allow vibrational excitation, hence providing additional channels for redistributing internal energy during collisions. Such a difference in the distribution of electronic excited states between the two approaches is expected to result in modifications in the radiation properties.

The limitations of the model have been discussed with the results that both rotational–vibrational–electronic coupling and chemical reactions involving electronic excited states have been disregarded. Further investigations will focus on incorporating rotational–vibrational–electronic coupling into the mathematical description of the model and ultimately in a DSMC solver. Additionally, computational chemistry studies11 have evidenced that adding the contribution of electronic excited states to the global dissociation rates significantly improves the reproduction of experimental measurements. Therefore, future research will involve the modeling of the chemical activity of the electronic excited states in the context of hypersonic flows.

This work was initiated during a secondment at the Institute of Space Systems at the University of Stuttgart. The lead author gratefully acknowledges scholarship funding from the James Watt School of Engineering through the School of Engineering Scholarship 2020 and the Jim Gatheral Travel Grant 2023.

The authors have no conflicts to disclose.

C. H. B. Civrais: Conceptualization (equal); Data curation (lead); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (lead); Software (lead); Supervision (equal); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). M. Pfeiffer: Conceptualization (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal). C. White: Supervision (supporting); Writing – review & editing (supporting). R. Steijl: Supervision (equal); Writing – review & editing (supporting).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
G. A.
Bird
,
Molecular Gas Dynamics and the Direct Simulation of Gas Flows
(
Oxford University Press, Inc.
,
1994
).
2.
G. A.
Bird
, “
Sophisticated DSMC
,” in
Notes Prepared for a Short Course at the DSMC07 Meeting
(
Sandia National Laboratories
,
2007
).
3.
B. L.
Haas
and
I. D.
Boyd
, “
Models for direct Monte Carlo simulation of coupled vibration–dissociation
,”
Phys. Fluids A
5
,
478
489
(
1993
).
4.
G. A.
Bird
, “
The QK model for gas-phase chemical reaction rates
,”
Phys. Fluids
23
,
106101
(
2011
).
5.
D. S.
Liechty
and
M. J.
Lewis
, “
Extension of the quantum-kinetic model to lunar and Mars return physics
,”
Phys. Fluids
26
,
027106
(
2014
).
6.
C. H. B.
Civrais
,
C.
White
, and
R.
Steijl
, “
Evaluation of a kinetic-theory approach for chemical reaction rates in upper-atmosphere hypersonic flows
,” AIAA Paper No. 2023-3809,
2023
.
7.
C. H. B.
Civrais
,
C.
White
, and
R.
Steijl
, “
Quantum-kinetic chemistry model with an anharmonic oscillator model: Model derivation and limitations
,”
Phys. Fluids
(to be published).
8.
K. J.
Higdon
,
B. A.
Cruden
,
A. M.
Brandis
,
D. S.
Liechty
,
D. B.
Goldstein
, and
P. L.
Varghese
, “
Direct simulation Monte Carlo shock simulation of Saturn entry probe conditions
,”
J. Thermophys. Heat Transfer
32
,
680
690
(
2018
).
9.
K. J.
Higdon
,
D. B.
Goldstein
, and
P. L.
Varghese
, “
Sensitivity analysis of direct simulation Monte Carlo parameters for ionizing hypersonic flows
,”
J. Thermophys. Heat Transfer
32
,
90
102
(
2018
).
10.
M.
Schouler
,
Y.
Prévereaud
, and
L.
Mieussens
, “
IXV post-flight reconstruction and analysis of the aerothermodynamic measurements along the rarefied portion of the reentry trajectory
,”
Int. J. Heat Mass Transfer
178
,
121582
(
2021
).
11.
S. M.
Jo
,
S.
Venturi
,
J. G.
Kim
, and
M.
Panesi
, “
Rovibrational internal energy transfer and dissociation of high-temperature oxygen mixture
,”
J. Chem. Phys.
158
,
064305
(
2023
).
12.
M.
Pfeiffer
,
J.
Beyer
,
J.
Vaubaillon
,
P.
Matlovič
,
J.
Tóth
,
S.
Fasoulas
, and
S.
Löhle
, “
Numerical simulation of an iron meteoroid entering into Earth's atmosphere using DSMC and a radiation solver with comparison to ground testing data
,”
Icarus
407
,
115768
(
2024
).
13.
T.
Ozawa
,
T.
Suzuki
,
H.
Takayanagi
, and
K.
Fujita
, “
Analysis of non-continuum hypersonic flows for the Hayabusa reentry
,” AIAA Paper No. 2011-3311,
2011
.
14.
I.
Sohn
,
Z.
Li
,
D. A.
Levin
, and
M. F.
Modest
, “
Coupled DSMC-PMC radiation simulations of a hypersonic reentry
,”
J. Thermophys. Heat Transfer
26
,
22
35
(
2012
).
15.
I. D.
Boyd
and
P.
Jenniskens
, “
Modeling of stardust entry at high altitude, Part 2: Radiation analysis
,”
J. Spacecr. Rockets
47
,
901
909
(
2010
).
16.
E.
Farbar
and
I.
Boyd
, “
Simulation of Fire II reentry flow using the direct simulation Monte Carlo method
,” AIAA Paper No. 2008-4103,
2008
.
17.
E. D.
Farbar
and
I. D.
Boyd
, “
Modeling of the plasma generated in a rarefied hypersonic shock layer
,”
Phys. Fluids
22
,
106101
(
2010
).
18.
D.
Liechty
,
C.
Johnston
, and
M.
Lewis
, “
Comparison of DSMC and CFD solutions of Fire II including radiative heating
,” AIAA Paper No. 2011-3494,
2011
.
19.
J.
Beyer
,
M.
Pfeiffer
, and
S.
Fasoulas
, “
Non-equilibrium radiation modeling in a gas kinetic simulation code
,”
J. Quant. Spectrosc. Radiat. Transfer
280
,
108083
(
2022
).
20.
T.
Ozawa
,
Z.
Li
,
I.
Sohn
, and
D.
Levin
, “
Modeling of electronic excitation and radiation for hypersonic re-entry flows in DSMC
,” AIAA Paper No. 2010-987,
2010
.
21.
Z.
Li
,
T.
Ozawa
,
I.
Sohn
, and
D. A.
Levin
, “
Modeling of electronic excitation and radiation in non-continuum hypersonic reentry flows
,”
Phys. Fluids
23
,
066102
(
2011
).
22.
R.
Jambunathan
and
D. A.
Levin
, “
Kinetic, 3-D, PIC-DSMC simulations of ion thruster plumes and the backflow region
,”
IEEE Trans. Plasma Sci.
48
,
2017
2034
(
2020
).
23.
S.
Copplestone
,
P.
Ortwein
,
C.-D.
Munz
,
T.
Binder
,
A.
Mirza
,
P.
Nizenkov
,
M.
Pfeiffer
, and
S.
Fasoulas
, “
Coupled PIC-DSMC simulations of a laser-driven plasma expansion
,” in
High Performance Computing in Science and Engineering'15: Transactions of the High Performance Computing Center, Stuttgart (HLRS) 2015
(
Springer
,
2016
), pp.
689
701
.
24.
T.
Binder
,
M.
Pfeiffer
, and
S.
Fasoulas
, “
Validation of grid current simulations using the particle-in-cell method for a miniaturized ion thruster
,”
AIP Conf. Proc.
2132
,
040003
(
2019
).
25.
C. J.
Capon
,
M.
Brown
,
C.
White
,
T.
Scanlon
, and
R. R.
Boyce
, “
pdFOAM: A PIC-DSMC code for near-Earth plasma-body interactions
,”
Comput. Fluids
149
,
160
171
(
2017
).
26.
T.
Ozawa
,
T.
Suzuki
,
H.
Osawa
,
H.
Takayanagi
, and
K.
Fujita
, “
Electronic excitation modeling for inductively coupled plasma flows
,” AIAA Paper No. 2011-142,
2011
.
27.
D.
Liechty
and
M.
Lewis
, “
Treatment of electronic energy level transition and ionization following the particle-based chemistry model
,” AIAA Paper No. 2010-449,
2010
.
28.
F.
Bergemann
and
I. D.
Boyd
, “
New discrete vibrational energy model for the direct simulation Monte Carlo method
,” in
Rarefied Gas Dynamics: Experimental Techniques and Physical Systems
(
American Institute of Aeronautics and Astronautics
,
1994
), pp.
174
183
.
29.
G. A.
Bird
, “
Nonequilibrium radiation during re-entry at 10 km/s
,” AIAA Paper No. 1987-1543,
1987
.
30.
A. B.
Carlson
and
H. A.
Hassan
, “
Radiation modeling with direct simulation Monte Carlo
,”
J. Thermophys. Heat Transfer
6
,
631
636
(
1992
).
31.
J. C.
Taylor
,
A. B.
Carlson
, and
H. A.
Hassan
, “
Monte Carlo simulation of radiating re-entry flows
,”
J. Thermophys. Heat Transfer
8
,
478
485
(
1994
).
32.
J. M.
Burt
and
E.
Josyula
, “
DSMC modeling of nonequilibrium electronic excitation and emission for hypersonic sensor applications
,” AIAA Paper No. 2015-2511,
2015
.
33.
J. M.
Burt
and
E.
Josyula
, “
Direct simulation Monte Carlo modeling of gas electronic excitation for hypersonic sensing
,”
J. Thermophys. Heat Transfer
31
,
858
870
(
2017
).
34.
M. A.
Gallis
and
J. K.
Harvey
, “
Nonequilibrium thermal radiation from air shock layers modeled with direct simulation Monte Carlo
,”
J. Thermophys. Heat Transfer
8
,
765
772
(
1994
).
35.
M. A.
Gallis
and
J. K.
Harvey
, “
Atomic species radiation from sir shock layers modelled with DSMC
,” AIAA Paper No. 1994-2018,
1994
.
36.
P. A.
Gnoffo
, “
Planetary–Entry Gas Dynamics
,”
Annu. Rev. Fluid Mech.
31
,
459
494
(
1999
).
37.
A. J.
Smith
, “
ESA PlasmA RAdiation DatabasE (PARADE). Development history, status, current developments and future prospects
,” in
Radiation of High Temperature Gases in Atmospheric Entry
(
ESA Special Publication
,
Noordwijk, The Netherlands
,
2003
), Vol.
533
, pp.
75
83
.
38.
A. M.
Brandis
and
B. A.
Cruden
, “
NEQAIR V15.0 release notes: Nonequilibrium and equilibrium radiative transport and spectra program
,”
Technical Report No. ARC-E-DAA-TN72963
,
National Aeronautics and Space Administration
,
2019
.
39.
C. O.
Laux
, “
Optical diagnostics and radiative emission of air plasmas
,” Ph.D. thesis (
Stanford University
,
1993
).
40.
I. T.
Karpuzcu
,
M. P.
Jouffray
, and
D. A.
Levin
, “
Collisional radiative modeling of electronically excited states in a hypersonic flow
,”
J. Thermophys. Heat Transfer
36
,
982
1002
(
2022
).
41.
J.
Beyer
,
P.
Nizenkov
,
S.
Fasoulas
, and
M.
Pfeiffer
, “
Simulation of radiating non-equilibrium flows around a capsule entering Titan's atmosphere
,”
AIP Conf. Proc.
2996
,
200002
(
2024
).
42.
C.
Park
,
Nonequilibrium Hypersonic Aerothermodynamics
, 1st ed. (
John Wiley & Sons
,
1990
).
43.
T.
Ozawa
,
D. A.
Levin
,
A.
Wang
, and
M.
Modest
, “
Development of coupled particle hypersonic flowfield–photon Monte Carlo radiation methods
,”
J. Thermophys. Heat Transfer
24
,
612
622
(
2010
).
44.
P. M.
Morse
, “
Diatomic molecules according to the wave mechanics. II. Vibrational levels
,”
Phys. Rev.
34
,
57
(
1929
).
45.
P. W.
Atkins
and
R. S.
Friedman
,
Molecular Quantum Mechanics
, 5th ed. (
Oxford University Press, Inc
.,
2011
).
46.
I. D.
Boyd
, “
Rotational–translational energy transfer in rarefied nonequilibrium flows
,”
Phys. Fluids A
2
,
447
452
(
1990
).
47.
I. D.
Boyd
, “
Relaxation of discrete rotational energy distributions using a Monte Carlo method
,”
Phys. Fluids A
5
,
2278
2286
(
1993
).
48.
M.
Pfeiffer
,
P.
Nizenkov
,
A.
Mirza
, and
S.
Fasoulas
, “
Direct simulation Monte Carlo modeling of relaxation processes in polyatomic gases
,”
Phys. Fluids
28
,
027103
(
2016
).
49.
P. J.
Linstrom
and
W. G.
Mallard
,
NIST Chemistry WebBook, NIST Standard Reference Database Number 69
(
National Institute of Standards and Technology
,
2024
).
50.
M.
Pfeiffer
, “
Extending the particle ellipsoidal statistical Bhatnagar-Gross-Krook method to diatomic molecules including quantized vibrational energies
,”
Phys. Fluids
30
,
116103
(
2018
).
51.
C. H. B.
Civrais
,
C.
White
, and
R.
Steijl
, “
Vibrational modeling with an anharmonic oscillator model in direct simulation Monte Carlo
,”
J. Thermophys. Heat Transfer
37
,
534
548
(
2023
).
52.
C. H. B.
Civrais
,
C.
White
, and
R.
Steijl
, “
Influence of anharmonic oscillator model for flows over a cylindrical body
,”
AIP Conf. Proc.
2996
,
080008
(
2024
).
53.
C. H. B.
Civrais
,
C.
White
, and
R.
Steijl
, “
Extension of the normal shock wave relations for calorically imperfect gas
,”
Shock Waves
33
,
533
551
(
2023
).
54.
W. G.
Vincenti
and
C. H.
Kruger
,
Introduction to Physical Gas Dynamics
(
John Wiley & Sons
,
1965
).
55.
C. N.
Hinshelwood
,
The Kinetics of Chemical Change
(
Clarendon Press
,
1940
).
56.
P. A.
Gnoffo
, “
Conservation equations and physical models for hypersonic air flows in thermal and chemical nonequilibrium
,”
Technical Report No. TP-2867
,
National Aeronautics and Space Administration
,
1989
.
57.
G.
Colonna
,
I.
Armenise
,
D.
Bruno
, and
M.
Capitelli
, “
Reduction of state-to-state kinetics to macroscopic models in hypersonic flows
,”
J. Thermophys. Heat Transfer
20
,
477
486
(
2006
).
58.
M.
Panesi
,
T. E.
Magin
,
A.
Bourdon
,
A.
Bultel
, and
O.
Chazot
, “
Electronic excitation of atoms and molecules for the FIRE II flight experiment
,”
J. Thermophys. Heat Transfer
25
,
361
374
(
2011
).
59.
A.
Munafò
,
N. N.
Mansour
, and
M.
Panesi
, “
A reduced-order NLTE kinetic model for radiating plasmas of outer envelopes of stellar atmospheres
,”
Astrophys. J.
838
,
126
(
2017
).
60.
H.
Liu
,
D.
Shi
,
J.
Sun
,
Z.
Zhu
, and
Z.
Shulin
, “
Accurate calculations on the 22 electronic states and 54 spin-orbit states of the O2 molecule: Potential energy curves, spectroscopic parameters and spin-orbit coupling
,”
Spectrochim. Acta, Part A
124
,
216
229
(
2014
).
61.
W.
Zhao
,
Q.
Hong
,
C.
Yang
,
Q.
Sun
, and
Y.
Hu
, “
Collision integrals of electronically excited atoms in air plasmas. I. N–N and O–O interactions
,”
Plasma Sources Sci. Technol.
32
,
125002
(
2023
).
62.
M.
Buchowiecki
and
P.
Szabó
, “
Transport cross sections and collision integrals for O(3P)–O(3P) interaction
,”
Plasma Chem. Plasma Process.
44
,
755
763
(
2024
).
63.
D. S.
Liechty
and
M.
Lewis
, “
Electronic energy level transition and ionization following the quantum-kinetic chemistry model
,”
J. Spacecr. Rockets
48
,
283
290
(
2011
).
64.
M.
Capitelli
,
G.
Colonna
,
D.
Giordano
,
L.
Maraffa
,
A.
Casavola
,
P.
Minelli
,
D.
Pagano
,
L.
Pietanza
, and
F.
Taccogna
, “
Tables of internal partition functions and thermodynamic properties of high-temperature Mars-atmosphere species from 50 K to 50,000 K
,”
Technical Report No. STR-246
,
European Space Agency Publications Division
,
2005
.
65.
R.
Jaffe
, “
The calculation of high-temperature equilibrium and nonequilibrium specific heat data for N2, O2 and NO
,” AIAA Paper No. 1987-1633,
1987
.
66.
Z.
Qin
,
J. M.
Zhao
, and
L. H.
Liu
, “
High-temperature partition functions, specific heats and spectral radiative properties of diatomic molecules with an improved calculation of energy levels
,”
J. Quant. Spectrosc. Radiat. Transfer
210
,
1
18
(
2018
).
67.
B. J.
McBride
,
M. J.
Zehe
, and
S.
Gordon
, “
NASA Glenn coefficients for calculating thermodynamic properties of individual species
,”
Technical Report No. TP-2002–211556
,
National Aeronautics and Space Administration
,
2002
.
68.
R.
Jaffe
,
D.
Schwenke
, and
G.
Chaban
, “
Theoretical analysis of N2 collisional dissociation and rotation-vibration energy transfer
,” AIAA Paper No. 2009-1569,
2009
.
69.
T. E.
Magin
,
M.
Panesi
,
A.
Bourdon
,
R. L.
Jaffe
, and
D. W.
Schwenke
, “
Coarse-grain model for internal energy excitation and dissociation of molecular nitrogen
,”
Chem. Phys.
398
,
90
95
(
2012
).
70.
J. G.
Kim
and
I. D.
Boyd
, “
Monte Carlo simulation of nitrogen dissociation based on state-resolved cross sections
,”
Phys. Fluids
26
,
012006
(
2014
).
71.
M.
Buchowiecki
, “
Ro-vibrational coupling in high temperature thermochemistry of the BBr molecule
,”
Chem. Phys. Lett.
692
,
236
241
(
2018
).
72.
National Oceanic and Atmospheric Administration,
U.S. standard atmosphere
,”
Technical Report No. TM-X-74335
,
National Aeronautics and Space Administration
,
1976
.
73.
G. A.
Bird
, “
Perception of numerical methods in rarefied gasdynamics
,” in
Rarefied Gas Dynamics: Theoretical and Computational Techniques
(
American Institute of Aeronautics and Astronautics
,
1989
), Vol. 118, pp. 211–226.
74.
M.
Buchowiecki
, “
Vibrational partition function for the multitemperature theories of high-temperature flows of gases and plasmas
,”
J. Phys. Chem. A
124
,
4048
4052
(
2020
).
75.
Y.
Babou
,
P.
Rivière
,
M.-Y.
Perrin
, and
A.
Soufiani
, “
High-temperature and nonequilibrium partition function and thermodynamic data of diatomic molecules
,”
Int. J. Thermophys.
30
,
416
438
(
2009
).
Published open access through an agreement withJISC Collections