License: CC BY 4.0
arXiv:2403.13586v1 [physics.plasm-ph] 20 Mar 2024

Non-Equilibrium and Self-Organization Evolution in Hot-Spot Ignition Processes

X.-Y. Fu    Z.-Y. Guo    Q.-H. Wang    R.-C. Wang    D. Wu [email protected] Key Laboratory for Laser Plasmas and Department of Physics and Astronomy, Collaborative Innovation Center of IFSA (CICIFSA), Shanghai Jiao Tong University, Shanghai 200240, People’s Republic of China Zhiyuan College, Shanghai Jiao Tong University, Shanghai 200240, People’s Republic of China    J. Zhang [email protected] Key Laboratory for Laser Plasmas and Department of Physics and Astronomy, Collaborative Innovation Center of IFSA (CICIFSA), Shanghai Jiao Tong University, Shanghai 200240, People’s Republic of China Zhiyuan College, Shanghai Jiao Tong University, Shanghai 200240, People’s Republic of China Institute of Physics, Chinese Academy of Sciences, Bei**g 100190, People’s Republic of China
(March 20, 2024)
Abstract

Due to disparate formation mechanisms, as for central hot-spot ignition and fast ignition, the initial temperatures of electron and ions usually differs from each other in the hot spot. Considering the percipient dependence of fusion cross-section and energy losses on temperature, this difference manifests the inadequacy of the equilibrium theoretical model in accurately depicting the ignition condition and evolution of the hot-spot. In this work, we studied a non-equilibrium model and extended this model to both isobaric and isochoric scenarios, characterized by varying hot-spot densities, temperatures and expansion velocities. In both cases, a spontaneous self-organization evolution was observed, manifesting as the bifurcation of ion and electron temperatures. Notably, the ion temperature is particularly prominent during the ignition process. This inevitability can be traced to the preponderant deposition rates of alpha-particles into D-T ions and the decreasing rate of energy exchange between electrons and D-T ions at elevated temperatures. The inherent structure, characterized by higher ion temperature and lower electron temperature during ignition, directly contributes to the augmentation of D-T reactions and mitigates energy losses through electron conduction and bremsstrahlung, thereby naturally facilitating nuclear fusions.

I Introduction

In practical researches, there exist two distinct implosion design strategies, i.e., central hot-spot ignition and fast ignition, characterized by unique processes for the formation of hot-spots Atzeni . In the case of central hot-spot ignition, which is powered by X-rays or lasers, the D-T gas, along with the surrounding high density D-T fuel, is compressed inward, leading to the formation of a high-temperature hot spot at the core of the fuel. Conventionally, the hot spot achieved through this design maintains nearly constant pressure compared to the surrounding cold fuels Clavin (2017). During the implosion process, the majority of the shock wave energy is deposited into ions, resulting in elevated ion temperatures, as experimentally verified Rygg et al. (2009). In contrast, the fast ignition scheme separates the compression and hot-spot formation processes Tabak et al. (2006); Atzeni ; Ghasemi et al. (2014). During the compression, the density remains uniform, necessitating the employment of an isochoric design Xu et al. (2023); Clark and Tabak (2007); Farahbod et al. (2014). After the compression process, a beam of fast electrons is injected and deposited, resulting in the formation of a localized hot-spot with electron temperature significantly higher than that of the ions Ghasemi et al. (2014).

The central hot-spot design has remained the prevailing strategy and has attained several milestones in recent years, e.g. burning plasma state Zylstra et al. (2022), ignition Acree et al. (2022) and “scientific breakeven” Acree et al. (2024). Despite the achievement achieved, there are still unresolved facets of new physics within the burning plasmas and ignition processes, as detailed in Zylstra et al. (2022); Hurricane et al. (2014); Lindl et al. (2014). These include kinetic effects and the energy transfer mechanisms of α𝛼\alphaitalic_α-particles during the self-burning processes, as evidenced in Hartouni et al. (2023).

Concerning the unresolved ignition queries, previous theoretical efforts have primarily resorted to the simplified equilibrium model Chang et al. (2010); Döppner et al. (2015); Gopalaswamy et al. (2024); Churnetski et al. (2024); Atzeni and Meyer-ter Vehn . For the sake of simplicity, it has become a common assumption to presume nearly identical ion and electron temperatures throughout the entire ignition process Hurricane et al. (2023); Daughton et al. (2023). However, this approach seems to overlook the likelihood of non-equilibrium ion and electron temperatures. We hypothesize that, at high temperatures, the equilibrium condition might be disrupted due to different heating mechanisms acting on ions and electrons, resulting in distinct evolution.

In Z. F. Fan’s work, the ion-electron non-equilibrium ignition was theoretically studied through the introduction of a two-temperature mode Fan et al. (2016). They consider separate thermal equilibrium of ions and electrons at temperatures Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT (TiTesubscript𝑇𝑖subscript𝑇𝑒T_{i}\neq T_{e}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≠ italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT). Consequently, these two components evolve independently with energy exchange adjusting between them due to ion-electron collisions. According to their findings, an initially higher hot-spot ion temperature compared to electrons enhances nuclear reactions and reduces energy loss of electron conduction and bremsstrahlung, thereby facilitating ignition. Building upon their model, we further develop an isochoric model for fast ignition, depict the ignition condition in different cases, and provide a detailed analysis of different heating stages during both hot-spot ignition and fast ignition processes. In the context of fast ignition, the disparity between ion and electron temperatures exerts a significant influence on the heating process due to extreme density, offering valuable insights for optimizing the hot-spot formation process. Additionally, disparate deposition rates of alpha-particle heating to ions and electrons lead to the bifurcated evolution of ion and electron temperatures, ultimately and naturally benefiting ignition.

The organization of this paper is as follows. In Section II, we introduce the equations that form the basis of our non-equilibrium model. Subsequently, Section III offers an overview of the data sources employed and presents our simulation outcomes for both the isobaric and isochoric models. Furthermore, Section IV delves into a theoretical analysis of the distinct temporal stages observed in the heating process. Lastly, Section V concludes the paper by summarizing our key findings and discussing their implications.

II Non-Equilibrium Model

The hot-spot, characterized by its radius Rhsubscript𝑅R_{h}italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and total density ρhsubscript𝜌\rho_{h}italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, can be simplified as comprising two components: D-T ions and electrons. Assuming ion temperatures and pressures are denoted as Ti,Pisubscript𝑇𝑖subscript𝑃𝑖T_{i},~{}P_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and electron temperatures and pressures as Te,Pisubscript𝑇𝑒subscript𝑃𝑖T_{e},~{}P_{i}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, we define the hot-spot equilibrium temperature as Th(Ti+Te)/2subscript𝑇subscript𝑇𝑖subscript𝑇𝑒2T_{h}\equiv(T_{i}+T_{e})/2italic_T start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ≡ ( italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) / 2 and introduce a non-equilibrium factor, f=Ti/Th𝑓subscript𝑇𝑖subscript𝑇f=T_{i}/T_{h}italic_f = italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT.

During ignition, the alpha particles produced by D-T reactions, primarily concerned with the ion temperature Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, deposit energy into the hot spot at a total power of Wα=Aαρh2σvfαsubscript𝑊𝛼subscript𝐴𝛼superscriptsubscript𝜌2delimited-⟨⟩𝜎𝑣subscript𝑓𝛼W_{\alpha}=A_{\alpha}\rho_{h}^{2}\langle\sigma v\rangle f_{\alpha}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = italic_A start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟨ italic_σ italic_v ⟩ italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, with Aα=8×1040erg/g2subscript𝐴𝛼8superscript1040ergsuperscriptg2A_{\alpha}=8\times 10^{40}~{}{\rm erg/g^{2}}italic_A start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = 8 × 10 start_POSTSUPERSCRIPT 40 end_POSTSUPERSCRIPT roman_erg / roman_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Here, σvdelimited-⟨⟩𝜎𝑣\langle\sigma v\rangle⟨ italic_σ italic_v ⟩ is the cross-section of D-T fusion, expressed as a function of Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT Atzeni . The deposition rate of α𝛼\alphaitalic_α in hot spots fαsubscript𝑓𝛼f_{\alpha}italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT has a form of

fα={(3/2)τα(4/5)τα2,τα1/21(1/4)τα1+(1/160)τα3,τα1/2,subscript𝑓𝛼cases32subscript𝜏𝛼45superscriptsubscript𝜏𝛼2subscript𝜏𝛼12114superscriptsubscript𝜏𝛼11160superscriptsubscript𝜏𝛼3subscript𝜏𝛼12f_{\alpha}=\begin{cases}({3}/{2})\tau_{\alpha}-({4}/{5})\tau_{\alpha}^{2},&% \tau_{\alpha}\leq 1/2\\ 1-({1/4}){\tau_{\alpha}}^{-1}+({1}/{160}){\tau_{\alpha}^{-3}},&\tau_{\alpha}% \geq 1/2\end{cases},italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = { start_ROW start_CELL ( 3 / 2 ) italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT - ( 4 / 5 ) italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , end_CELL start_CELL italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≤ 1 / 2 end_CELL end_ROW start_ROW start_CELL 1 - ( 1 / 4 ) italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + ( 1 / 160 ) italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , end_CELL start_CELL italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≥ 1 / 2 end_CELL end_ROW , (1)

where τα9×lnΛρhRh/Th3/2similar-to-or-equalssubscript𝜏𝛼9Λsubscript𝜌subscript𝑅hsuperscriptsubscript𝑇h32\tau_{\alpha}\simeq 9\times{\ln\Lambda}{\rho_{h}R_{\mathrm{h}}}/{T_{\mathrm{h}% }^{3/2}}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≃ 9 × roman_ln roman_Λ italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT. This energy deposition heats the ions and electrons in the hot-spot by fractions of fαi=Te/(32+Te)subscript𝑓𝛼𝑖subscript𝑇𝑒32subscript𝑇𝑒f_{\alpha i}=T_{e}/(32+T_{e})italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / ( 32 + italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) and fαe=1fαisubscript𝑓𝛼𝑒1subscript𝑓𝛼𝑖f_{\alpha e}=1-f_{\alpha i}italic_f start_POSTSUBSCRIPT italic_α italic_e end_POSTSUBSCRIPT = 1 - italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT, respectively, where Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is given in keV Fraley et al. (1974). The energy losses in the system occur through electron thermal conduction with the surrounding main fuel, given by We=3AeTe7/2/Rh2/lnΛsubscript𝑊𝑒3subscript𝐴𝑒superscriptsubscript𝑇𝑒72superscriptsubscript𝑅2ΛW_{e}=3A_{e}T_{e}^{7/2}/R_{h}^{2}/\ln\Lambdaitalic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 3 italic_A start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 7 / 2 end_POSTSUPERSCRIPT / italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / roman_ln roman_Λ with Ae=9.5×1019ergs1cm1keV7/2subscript𝐴𝑒9.5superscript1019ergsuperscripts1superscriptcm1superscriptkeV72A_{e}=9.5\times 10^{19}~{}{\rm{erg~{}s^{-1}cm^{-1}keV^{-7/2}}}italic_A start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 9.5 × 10 start_POSTSUPERSCRIPT 19 end_POSTSUPERSCRIPT roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_keV start_POSTSUPERSCRIPT - 7 / 2 end_POSTSUPERSCRIPT, and through electron bremsstrahlung Wr=Arρh2Te1/2subscript𝑊𝑟subscript𝐴𝑟superscriptsubscript𝜌2superscriptsubscript𝑇𝑒12W_{r}=A_{r}\rho_{h}^{2}T_{e}^{1/2}italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_A start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT with Ar=3.05×1023ergcm3g2s1keV1/2subscript𝐴𝑟3.05superscript1023ergsuperscriptcm3superscriptg2superscripts1superscriptkeV12A_{r}=3.05\times 10^{23}~{}{\rm erg~{}cm^{3}g^{-2}s^{-1}keV^{-1/2}}italic_A start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 3.05 × 10 start_POSTSUPERSCRIPT 23 end_POSTSUPERSCRIPT roman_erg roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_keV start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT. The Coulumb logarithm lnΛΛ\ln\Lambdaroman_ln roman_Λ can be expressed as 0.5×ln[1+(bmin/bmax)2]0.51superscriptsubscript𝑏subscript𝑏20.5\times\ln\left[1+({b_{\min}}/{b_{\max}})^{2}\right]0.5 × roman_ln [ 1 + ( italic_b start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT / italic_b start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] Wu et al. (2018) where bminsubscript𝑏b_{\min}italic_b start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT is the maximum between the classical impact parameter and the De Broglie wavelength, and bmaxsubscript𝑏b_{\max}italic_b start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT is the Debye length. Additionally, the expansion of the ions and electrons within the hot spot, together with speed uhsubscript𝑢u_{h}italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, contributes to energy loss with powers of Wm,i=3Piuh/Rhsubscript𝑊𝑚𝑖3subscript𝑃𝑖subscript𝑢subscript𝑅W_{m,i}=3P_{i}u_{h}/R_{h}italic_W start_POSTSUBSCRIPT italic_m , italic_i end_POSTSUBSCRIPT = 3 italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and Wm,e=3Peuh/Rhsubscript𝑊𝑚𝑒3subscript𝑃𝑒subscript𝑢subscript𝑅W_{m,e}=3P_{e}u_{h}/R_{h}italic_W start_POSTSUBSCRIPT italic_m , italic_e end_POSTSUBSCRIPT = 3 italic_P start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, respectively. Furthermore, energy exchange between ions and electrons occurs through collisions, contributing to a power of Wie=Aω,eiρh2lnΛ(TiTe)/Te3/2subscript𝑊𝑖𝑒subscript𝐴𝜔𝑒𝑖superscriptsubscript𝜌2Λsubscript𝑇𝑖subscript𝑇𝑒superscriptsubscript𝑇𝑒32W_{ie}=A_{\omega,ei}\rho_{h}^{2}\ln\Lambda(T_{i}-T_{e})/T_{e}^{3/2}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT = italic_A start_POSTSUBSCRIPT italic_ω , italic_e italic_i end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_ln roman_Λ ( italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) / italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT, where Aω,ei=5.6×1014kJcm3g2s1keV1/2subscript𝐴𝜔𝑒𝑖5.6superscript1014kJsuperscriptcm3superscriptg2superscripts1superscriptkeV12A_{\omega,ei}=5.6\times 10^{14}~{}{\rm kJ~{}cm^{3}g^{-2}s^{-1}keV^{1/2}}italic_A start_POSTSUBSCRIPT italic_ω , italic_e italic_i end_POSTSUBSCRIPT = 5.6 × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT roman_kJ roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_keV start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT.

The temporal evolution for the temperatures of ion and electron energies, accounting for both the energy gains and losses in the hot spot, are expressed as follows Fan et al. (2016):

CV,iρidTidtsubscript𝐶𝑉𝑖subscript𝜌𝑖dsubscript𝑇𝑖d𝑡\displaystyle C_{V,i}\rho_{i}\frac{\mathrm{d}T_{i}}{\mathrm{d}t}italic_C start_POSTSUBSCRIPT italic_V , italic_i end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG roman_d italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG =WαfαiWieWm,i,absentsubscript𝑊𝛼subscript𝑓𝛼𝑖subscript𝑊𝑖𝑒subscript𝑊𝑚𝑖\displaystyle=W_{\alpha}f_{\alpha i}-W_{ie}-W_{m,i},= italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT - italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT - italic_W start_POSTSUBSCRIPT italic_m , italic_i end_POSTSUBSCRIPT , (2)
CV,eρedTedtsubscript𝐶𝑉𝑒subscript𝜌𝑒dsubscript𝑇𝑒d𝑡\displaystyle C_{V,e}\rho_{e}\frac{\mathrm{d}T_{e}}{\mathrm{d}t}italic_C start_POSTSUBSCRIPT italic_V , italic_e end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT divide start_ARG roman_d italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG =Wαfαe+WieWm,eWrWe,absentsubscript𝑊𝛼subscript𝑓𝛼𝑒subscript𝑊𝑖𝑒subscript𝑊𝑚𝑒subscript𝑊𝑟subscript𝑊𝑒\displaystyle=W_{\alpha}f_{\alpha e}+W_{ie}-W_{m,e}-W_{r}-W_{e},= italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_α italic_e end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT - italic_W start_POSTSUBSCRIPT italic_m , italic_e end_POSTSUBSCRIPT - italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , (3)

where CV,isubscript𝐶𝑉𝑖C_{V,i}italic_C start_POSTSUBSCRIPT italic_V , italic_i end_POSTSUBSCRIPT and CV,esubscript𝐶𝑉𝑒C_{V,e}italic_C start_POSTSUBSCRIPT italic_V , italic_e end_POSTSUBSCRIPT are the specific heat of ions and electrons respectively; ρisubscript𝜌𝑖\rho_{i}italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and ρesubscript𝜌𝑒\rho_{e}italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT are the density of ions and electrons.

Energy spreading outside the hot spot, include electron bremsstrahlung power Wrsubscript𝑊𝑟W_{r}italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and electron thermal conduction Wesubscript𝑊𝑒W_{e}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. The latter contributes to main fuel ablation, ultimately increasing the hot spot mass, which is expressed as Hurricane et al. (2023); Daughton et al. (2023); Spears et al. (2008):

dρhVhdt=[Wr+We+Wα(1fα)/fα]VhCv,iTi+Cv,eTe.dsubscript𝜌subscript𝑉d𝑡delimited-[]subscript𝑊𝑟subscript𝑊𝑒subscript𝑊𝛼1subscript𝑓𝛼subscript𝑓𝛼subscript𝑉subscript𝐶𝑣𝑖subscript𝑇𝑖subscript𝐶𝑣𝑒subscript𝑇𝑒\frac{\mathrm{d}\rho_{h}V_{h}}{\mathrm{d}t}=\frac{[W_{r}+W_{e}+W_{\alpha}(1-f_% {\alpha})/f_{\alpha}]V_{h}}{C_{v,i}T_{i}+C_{v,e}T_{e}}.divide start_ARG roman_d italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = divide start_ARG [ italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 1 - italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) / italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ] italic_V start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG italic_C start_POSTSUBSCRIPT italic_v , italic_i end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_C start_POSTSUBSCRIPT italic_v , italic_e end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG . (4)

Our primary interest lies in the density variation, therefore, we formulate the contribution of volume Vhsubscript𝑉V_{h}italic_V start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and rewrite Eq. (4) as

dρhdt=Wr+We+Wα(1fα)/fαCv,iTi+Cv,eTe4πRh2ρhuhVh,dsubscript𝜌d𝑡subscript𝑊𝑟subscript𝑊𝑒subscript𝑊𝛼1subscript𝑓𝛼subscript𝑓𝛼subscript𝐶𝑣𝑖subscript𝑇𝑖subscript𝐶𝑣𝑒subscript𝑇𝑒4𝜋superscriptsubscript𝑅2subscript𝜌subscript𝑢subscript𝑉\frac{\mathrm{d}\rho_{h}}{\mathrm{d}t}=\frac{W_{r}+W_{e}+W_{\alpha}(1-f_{% \alpha})/f_{\alpha}}{C_{v,i}T_{i}+C_{v,e}T_{e}}-\frac{4\pi R_{h}^{2}\rho_{h}u_% {h}}{V_{h}},divide start_ARG roman_d italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = divide start_ARG italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 1 - italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) / italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG start_ARG italic_C start_POSTSUBSCRIPT italic_v , italic_i end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_C start_POSTSUBSCRIPT italic_v , italic_e end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG - divide start_ARG 4 italic_π italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG , (5)

considering the hot-spot expansion rate given by

dRhdt=uh.dsubscript𝑅d𝑡subscript𝑢\frac{\mathrm{d}R_{h}}{\mathrm{d}t}=u_{h}.divide start_ARG roman_d italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT . (6)
Refer to caption
Figure 1: Schematic description of isobaric and isochoric hot-spot ignition models.

In our research, significant attention is focused on the stagnation moment, which lies between the end of compression and the decomposition of the hot spot. This transient yet pivotal moment exhibits distinct behaviors in response to various ignition approaches. In the case of central hot-spot ignition, the hot spot follows the isobaric description, where its pressure approximates that of the surrounding cold fuel, resulting in minimal alterations to its radius Daughton et al. (2023); Atzeni and Caruso . Conversely, the fast ignition scheme features a hot spot density that closely mirrors the cold fuel, utilizing the isochoric model Xu et al. (2023). Consequently, hot spots expand due to pressure differentials. The intricate relationship between these models is graphically represented in Fig. 1. The respective expansion rates for these models are mathematically described in Atzeni ,

uh={0,Isobaric Model(34ΓBThρhρc)1/2,Isochoric Modelsubscript𝑢casessimilar-toabsent0Isobaric Modelsimilar-toabsentsuperscript34subscriptΓ𝐵subscript𝑇subscript𝜌subscript𝜌𝑐12Isochoric Model\displaystyle u_{h}=\begin{cases}\sim 0,&\text{Isobaric Model}\\ \sim\displaystyle{\left(\frac{3}{4}\Gamma_{B}T_{h}\frac{\rho_{h}}{\rho_{c}}% \right)^{1/2}},&\text{Isochoric Model}\end{cases}italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = { start_ROW start_CELL ∼ 0 , end_CELL start_CELL Isobaric Model end_CELL end_ROW start_ROW start_CELL ∼ ( divide start_ARG 3 end_ARG start_ARG 4 end_ARG roman_Γ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT , end_CELL start_CELL Isochoric Model end_CELL end_ROW (7)

where ΓBsubscriptΓ𝐵\Gamma_{B}roman_Γ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is related to the Boltzmann constant, for D-T fuel, ΓB=4kB/(2me+5mp)=7.66×1014erg/(gkeV)subscriptΓ𝐵4subscript𝑘𝐵2subscript𝑚𝑒5subscript𝑚𝑝7.66superscript1014erggkeV\displaystyle{\Gamma_{B}=4k_{B}/(2m_{e}+5m_{p})=7.66\times 10^{14}~{}\mathrm{% erg/(g~{}keV)}}roman_Γ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = 4 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT / ( 2 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT + 5 italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) = 7.66 × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT roman_erg / ( roman_g roman_keV ); ρhsubscript𝜌\rho_{h}italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and ρcsubscript𝜌𝑐\rho_{c}italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are respectively the density of the hot spot and the main fuel.

By solving the aforementioned set of four differential equations, we obtain valuable insights into the ignition processes, which are elucidated in Section III.

III Numerical results

In the previous section, we have listed a set of differential equations for temporal evolution of physical quantities. In this sections, this set of equations are numerically solved and displayed. In Table 1, the initial parameters are displayed, and especially, the areal density is kept the same in both models for a consistent comparison between them.

Table 1: Numerical configurations of the initial state, including the radius, density and temperature of the hot spot, for both isobaric and isochoric models.
Scheme Rhsubscript𝑅R_{h}italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (μm𝜇m\mu\mathrm{m}italic_μ roman_m) ρhsubscript𝜌\rho_{h}italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (g/cm3gsuperscriptcm3\mathrm{g}\mathrm{/}\mathrm{c}\mathrm{m}^{3}roman_g / roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT) Thsubscript𝑇T_{h}italic_T start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (keV)$\mathrm{k}\mathrm{e}\mathrm{V}$)roman_keV )
Isobaric Model 100 100 8
Isochoric Model 10 1000 8

To investigate the non-equilibrium dynamics comprehensively, we conduct numerical analyses with initial states of various conditions for both isochoric and isobaric models. Our study involves analyzing the temporal evolution of electron temperature and ion temperature. Additionally, we explore the distribution of initial values leading to successful ignition within the phase space (areal density and temperature) over a finite time. These analyses aim to delineate the ignition thresholds associated with distinct equilibrium factors and models.

III.1 Isobaric Model

Figure 2 is the temperature evolution for the isobaric model. From this evolution curve, we can find that electrons and ions first undergo energy exchange with each other, similar to a relaxation process Daligault and Simoni (2019). Then the temperature of two species rise steadily until it reaches a certain value. During the entire process, a remarkable phenomenon occurs: after reaching a specific value, the temperatures of ions and electrons spontaneously diverge, ultimately reaching their saturated states. This heating process can be therefore broadly categorized into four distinct stages, and in the following we will elucidate each of them.

Refer to caption
Figure 2: In the isobaric model, with areal density of ρhRh=10g/cm2subscript𝜌subscript𝑅10gsuperscriptcm2\rho_{h}R_{h}=10\mathrm{g/cm^{2}}italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = 10 roman_g / roman_cm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, the ion and electron temperature rise as a function of time for different initial non-equilibrium factors f=0.8𝑓0.8f=0.8italic_f = 0.8, 1111 and 1.21.21.21.2. These two temperatures would finally become bifurcated and reach saturated values.

Under the influence of the non-equilibrium factor, we can find that the ignition is more prominent when the non-equilibrium factor is high, which means the initial temperature of ions is higher than that of electrons.

III.2 Isochoric Model

Refer to caption
Figure 3: In the isochoric model with an areal density ρhRh=10g/cm2subscript𝜌subscript𝑅10gsuperscriptcm2\rho_{h}R_{h}=10\ \mathrm{g/cm^{2}}italic_ρ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = 10 roman_g / roman_cm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, the ion and electron temperature rise with time at different initial non-equilibrium factors f=0.8𝑓0.8f=0.8italic_f = 0.8, 1111, and 1.21.21.21.2. These two temperatures will become bifurcated and eventually decreased due to expansion.

Figure 3 reveals that the temporal evolution of temperatures for the isochoric model is quite similar with that of the isobaric model. The influence of the non-equilibrium factor remains consistent with that observed in the isobaric model. Notably, the temperature of ions plays a pivotal role in the heating process. However, there still exits two main differences. Firstly, if the time duration is significantly extended, the temperature of both ions and electrons will decrease as a result of the expansion of the hot-spot, and this trend is more significant in the isochoric model. Secondly, by comparing the first 5 ps in Fig. 2 and Fig. 3, we observe significant differences in between the isobaric and isochoric models. Notably, in the isochoric model, DT hot-spot heats up considerably faster than the isobaric model, resulting in a much higher peak temperature.

III.3 Ignition Condition

The ignition threshold curve establishes the boundaries within the initial phase space (areal density and temperature) that permit successful ignition within a finite time span. Specifically, an excessively low areal density is precluded as it hinders electron heat conduction at elevated temperatures, as discussed in Temporal et al. (2012). Conversely, a high areal density enables the temperature to approach closely the critical value of 4.3 keV, which is essential for self-heating.

Refer to caption
Figure 4: The ignition condition for both isobaric and isochoric models. When the initial condition of the hot spot is above the curve, the ignition would take place.

In comparison, the isobaric model exhibits less stringent ignition requirements than the isochoric model. Regardless of the model chosen, an increase in the non-equilibrium factor plays a pivotal role in expanding the ignition area. This underscores the positive impact of enhancing the non-equilibrium factor, which effectively lowers ignition thresholds.

IV Analysis of Ignition Evolution

Our focus will primarily be on analyzing the isobaric model. As mentioned in Section III.2, the isochoric model exhibits similarities to the isobaric model in terms of its physical characteristics. However, a key distinction lies in the final process, which is influenced by the dominance of expansion power.

In the ignition process, depicted in Fig. 5, intriguing phenomena emerge, indicating that the heating process can be segmented into four distinct stages: thermal equilibrium (Stage A), co-heating (Stage B), bifurcated heating (Stage C), and attaining saturated temperatures (Stage D). We aim to delve into each of these processes and provide a theoretical analysis. Fig. 5 also illustrates the power per volume throughout the ignition process. Notably, the expansion work in the isobaric model is ignored. The exchange energy between ions and electrons, however, is influenced by both the temperature of the ions and electrons, necessitating the utilization of an absolute value for this particular energy component.

Refer to caption
Figure 5: The ignition process for an isobaric model, with Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, Wαsubscript𝑊𝛼W_{\alpha}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT, Wrsubscript𝑊𝑟W_{r}italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and Wesubscript𝑊𝑒W_{e}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT evolving as a function of time.

IV.1 Reaching a Dynamic Equilibrium

At the beginning, the average temperature of ions and electrons does not increase significantly, but the temperature difference gradually decreases, as presented in the Stage A of Fig. 5. We therefore subtract Eq. (2) and Eq. (3), to derive the temperature differences between ions and electrons. With relations CV,i=3kB/(2mi)subscript𝐶𝑉𝑖3subscript𝑘𝐵2subscript𝑚𝑖C_{V,i}=3k_{B}/(2m_{i})italic_C start_POSTSUBSCRIPT italic_V , italic_i end_POSTSUBSCRIPT = 3 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT / ( 2 italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ), CV,e=3kB/(2me)subscript𝐶𝑉𝑒3subscript𝑘𝐵2subscript𝑚𝑒C_{V,e}=3k_{B}/(2m_{e})italic_C start_POSTSUBSCRIPT italic_V , italic_e end_POSTSUBSCRIPT = 3 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT / ( 2 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ), and CV,iρi=CV,eρesubscript𝐶𝑉𝑖subscript𝜌𝑖subscript𝐶𝑉𝑒subscript𝜌𝑒C_{V,i}\rho_{i}=C_{V,e}\rho_{e}italic_C start_POSTSUBSCRIPT italic_V , italic_i end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT italic_V , italic_e end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, we can get

CV,iρidΔTdt=subscript𝐶𝑉𝑖subscript𝜌𝑖dΔ𝑇d𝑡absent\displaystyle C_{V,i}\rho_{i}\frac{\mathrm{d}\Delta T}{\mathrm{d}t}=italic_C start_POSTSUBSCRIPT italic_V , italic_i end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG roman_d roman_Δ italic_T end_ARG start_ARG roman_d italic_t end_ARG = WαTe32keVTe+32keV2Wiesubscript𝑊𝛼subscript𝑇𝑒32keVsubscript𝑇𝑒32keV2subscript𝑊𝑖𝑒\displaystyle W_{\alpha}\frac{T_{e}-32\text{keV}}{T_{e}+32\text{keV}}-2W_{ie}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT divide start_ARG italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT - 32 keV end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT + 32 keV end_ARG - 2 italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT (8)
Wm,i+Wm,e+Wr+Wesubscript𝑊𝑚𝑖subscript𝑊𝑚𝑒subscript𝑊𝑟subscript𝑊𝑒\displaystyle-W_{m,i}+W_{m,e}+W_{r}+W_{e}- italic_W start_POSTSUBSCRIPT italic_m , italic_i end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_m , italic_e end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT

where ΔT=TiTeΔ𝑇subscript𝑇𝑖subscript𝑇𝑒\Delta T=T_{i}-T_{e}roman_Δ italic_T = italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. In the isobaric model, uh0similar-tosubscript𝑢0u_{h}\sim 0italic_u start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ∼ 0, meaning that the expansion of ions and electrons does not contribute to energy loss. Consequently, both Wm,isubscript𝑊𝑚𝑖W_{m,i}italic_W start_POSTSUBSCRIPT italic_m , italic_i end_POSTSUBSCRIPT and Wm,esubscript𝑊𝑚𝑒W_{m,e}italic_W start_POSTSUBSCRIPT italic_m , italic_e end_POSTSUBSCRIPT are zero. At the beginning of the ignition process, when the electron and ion temperature is sufficiently low, Wασvproportional-tosubscript𝑊𝛼delimited-⟨⟩𝜎𝑣W_{\alpha}\propto\langle\sigma v\rangleitalic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∝ ⟨ italic_σ italic_v ⟩, which highly depends on the ion temperature Atzeni , WrTe1/2proportional-tosubscript𝑊𝑟superscriptsubscript𝑇𝑒12W_{r}\propto T_{e}^{1/2}italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ∝ italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT and WeTe7/2proportional-tosubscript𝑊𝑒superscriptsubscript𝑇𝑒72W_{e}\propto T_{e}^{7/2}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∝ italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 7 / 2 end_POSTSUPERSCRIPT. Instead, WieΔTTe3/2proportional-tosubscript𝑊𝑖𝑒Δ𝑇superscriptsubscript𝑇𝑒32W_{ie}\propto\Delta T\cdot T_{e}^{-3/2}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT ∝ roman_Δ italic_T ⋅ italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 3 / 2 end_POSTSUPERSCRIPT, which means that at low temperatures, energy exchange between the ions and electrons dominates, as demonstrated in Fig. 5. Also, due to the ΔTΔ𝑇\Delta Troman_Δ italic_T dependence and the large coefficient Aω,eisubscript𝐴𝜔𝑒𝑖A_{\omega,ei}italic_A start_POSTSUBSCRIPT italic_ω , italic_e italic_i end_POSTSUBSCRIPT in the ion-electron energy exchange power Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT, it will dominate for a while before the temperature reaches a high level. Consequently, the temperature difference between ion and electron starts to shrink, reaching a dynamic equilibrium. However, when the initial temperature difference is minimal (i.e., f𝑓fitalic_f is close to 1), the process becomes less distinct, as it can be interpreted as having reached a state of dynamic equilibrium.

IV.2 Co-Heating as an Equilibrium Model

As the ion and electron temperatures converge, they can be heated simultaneously under the influence of an increasing α𝛼\alphaitalic_α-particle heating power Wαsubscript𝑊𝛼W_{\alpha}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, as shown in B stage of Fig. 5. During this process, the ion-electron collisions maintain their close temperatures dynamically, by balancing the terms, e.g., Wα(fαifαe)subscript𝑊𝛼subscript𝑓𝛼𝑖subscript𝑓𝛼𝑒W_{\alpha}(f_{\alpha i}-f_{\alpha e})italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_α italic_e end_POSTSUBSCRIPT ), 2Wie2subscript𝑊𝑖𝑒2W_{ie}2 italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT, Wrsubscript𝑊𝑟W_{r}italic_W start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and Wesubscript𝑊𝑒W_{e}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, in Eq. (8). During the B stage depicted in Fig. 5, the α𝛼\alphaitalic_α-heating power significantly surpasses other energy loss powers, resulting in a co-heating process. However, as the electron temperature Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT rises, the electron conduction power experiences a remarkable increase due to its Te7/2superscriptsubscript𝑇𝑒72T_{e}^{7/2}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 7 / 2 end_POSTSUPERSCRIPT dependency. When the electron temperature approaches approximately 32eVsimilar-toabsent32eV\sim 32\ \text{eV}∼ 32 eV, Wα(fαifαe)subscript𝑊𝛼subscript𝑓𝛼𝑖subscript𝑓𝛼𝑒W_{\alpha}(f_{\alpha i}-f_{\alpha e})italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_α italic_e end_POSTSUBSCRIPT ) undergoes a sign change, indicating that the difference between the α𝛼\alphaitalic_α-particle heating contributions to the ion and electron temperatures is poised to reverse. These observations suggest that the balance maintained by the ion-electron collisions in Eq. (8) may be disrupted.

IV.3 Bifurcated Heating

In Stage C depicted in Fig. 5, the evolution of the ion and electron temperatures, Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, diverges significantly. This divergence can be attributed to the weakened ion-electron energy exchange power, WieΔT/(Te)3/2proportional-tosubscript𝑊𝑖𝑒Δ𝑇superscriptsubscript𝑇𝑒32W_{ie}\propto\Delta T/(T_{e})^{3/2}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT ∝ roman_Δ italic_T / ( italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT at high electron temperatures, as well as the differing energy loss and heating mechanisms between ions and electrons.

The heating effect on electron temperature is significantly suppressed, while ion temperature heating remains robust during Stage C. For electrons, the significant enhancement in electron conduction loss, WeTe7/21028proportional-tosubscript𝑊𝑒superscriptsubscript𝑇𝑒72similar-tosuperscript1028W_{e}\propto T_{e}^{7/2}\sim 10^{28}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∝ italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 7 / 2 end_POSTSUPERSCRIPT ∼ 10 start_POSTSUPERSCRIPT 28 end_POSTSUPERSCRIPT strongly constrains the growth of the electron temperature, Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. As evident in Eq. (3), the α𝛼\alphaitalic_α-particle heating power, Wα1028similar-tosubscript𝑊𝛼superscript1028W_{\alpha}\sim 10^{28}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 28 end_POSTSUPERSCRIPT is effectively suppressed by Wesubscript𝑊𝑒W_{e}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. Furthermore, the electron temperature Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT increases at a rate proportional to WieΔT/(Te)3/2proportional-tosubscript𝑊𝑖𝑒Δ𝑇superscriptsubscript𝑇𝑒32W_{ie}\propto\Delta T/(T_{e})^{3/2}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT ∝ roman_Δ italic_T / ( italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT, which is insufficient to keep up with the increasing temperatures. This trend is clearly illustrated in Fig. 5. Consequently, the electron temperature exhibits a slow growth (ΔTe5similar-toΔsubscript𝑇𝑒5\Delta T_{e}\sim 5roman_Δ italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∼ 5 keV) that can be disregarded, resulting in a gradual increase of Wesubscript𝑊𝑒W_{e}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT towards a maximum value. Meanwhile fαisubscript𝑓𝛼𝑖f_{\alpha i}italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT and fαesubscript𝑓𝛼𝑒f_{\alpha e}italic_f start_POSTSUBSCRIPT italic_α italic_e end_POSTSUBSCRIPT remain virtually unchanged, and Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT exhibits a nearly linear increase with ΔTproportional-toabsentΔ𝑇\propto\Delta T∝ roman_Δ italic_T. For ions, we delve into Eq. (2) and discover that they lack a significant energy loss mechanism besides transferring energy to electrons through collisions. Given the slow but steady increase in the relatively low energy exchange power Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT, the heating effect from α𝛼\alphaitalic_α-particles, represented by Wαi=Wαfαisubscript𝑊𝛼𝑖subscript𝑊𝛼subscript𝑓𝛼𝑖W_{\alpha i}=W_{\alpha}f_{\alpha i}italic_W start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT = italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT dominates the ion temperature growth, resulting in a significant increase. Consequently, Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT rises extremely rapidly, eventually pushing the ion-electron energy exchange power Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT to a notably high level. Meanwhile, the decreasing value of fαsubscript𝑓𝛼f_{\alpha}italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is accompanied by a weakening of Wαsubscript𝑊𝛼W_{\alpha}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT. To conclude, we refer to Eq. (8), which reveals that the difference in α𝛼\alphaitalic_α-particle heating, Wα(fαifαe)1027similar-tosubscript𝑊𝛼subscript𝑓𝛼𝑖subscript𝑓𝛼𝑒superscript1027W_{\alpha}(f_{\alpha i}-f_{\alpha e})\sim 10^{27}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_f start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_α italic_e end_POSTSUBSCRIPT ) ∼ 10 start_POSTSUPERSCRIPT 27 end_POSTSUPERSCRIPT, is significant. Additionally, the ion-electron energy exchange power, which is slowly increasing, contributes 2Wie102627similar-to2subscript𝑊𝑖𝑒superscript1026272W_{ie}\sim 10^{26-27}2 italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 26 - 27 end_POSTSUPERSCRIPT to the overall heating process. However, the dominant factor is the electron conduction loss, estimated to be We1028similar-tosubscript𝑊𝑒superscript1028W_{e}\sim 10^{28}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 28 end_POSTSUPERSCRIPT. Consequently, the ion and electron temperatures diverge significantly as they evolve.

IV.4 Reaching Saturated Temperatures

During the stage D depicted in Fig. 5, a saturation phenomenon becomes evident in both ion and electron temperatures. We have mentioned the slowly growing electron temperature Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, electron conduction loss Wesubscript𝑊𝑒W_{e}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, the increasing ion temperature Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, ion-electron energy exchange Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT and the decreasing α𝛼\alphaitalic_α-particle heating Wαsubscript𝑊𝛼W_{\alpha}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT in stage C. With an increasing temperature gradient, the energy exchange Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT between ions and electrons significantly intensifies. This enhancement allows the energy exchange to be comparable with both the electron thermal conduction loss Wesubscript𝑊𝑒W_{e}italic_W start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and the α𝛼\alphaitalic_α-particle heating Wαfαesubscript𝑊𝛼subscript𝑓𝛼𝑒W_{\alpha}f_{\alpha e}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_α italic_e end_POSTSUBSCRIPT, ultimately, a balance is achieved between the energy loss and gain for electrons. Subsequently, the electron temperature nearly attains a state of saturation. Later, with the increase of Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, the α𝛼\alphaitalic_α-particle heating power decreases, while the ion-electron energy exchange Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT continues to adjust, maintaining the saturation of the electron temperature Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. Eventually, the heating power from α𝛼\alphaitalic_α-particles, represented by Wα,Wαisubscript𝑊𝛼subscript𝑊𝛼𝑖W_{\alpha},W_{\alpha i}italic_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , italic_W start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT will decline to a certain low level, reaching a balance with the ion-electron conduction Wiesubscript𝑊𝑖𝑒W_{ie}italic_W start_POSTSUBSCRIPT italic_i italic_e end_POSTSUBSCRIPT as expressed in Eq. (2). This balance contributes to the saturation of the ion temperature Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

The intricate internal plasma environment poses significant non-linear constraints, stemming from the diverse energy gain and loss mechanisms at play. However, the ion-electron energy exchange, functioning as a self-regulating quantity, dynamically adjusts in response to the varying energy gain and loss mechanisms of ions and electrons, ultimately leading to the saturation phenomenon. This phenomenon serves as a vivid illustration of the complex non-linear constraints inherent in fusion plasmas. Furthermore, our results underscore a non-isothermal phenomenon that plays a pivotal role in enhancing the efficient burning of hot-spots. Notably, the primary energy loss mechanism in hot-spots is closely linked to the electron temperature, whereas the fusion heating mechanism exhibits a positive correlation with ion temperature. This dynamic imbalance gives rise to a non-equilibrium hot-spot model, which emerges as an inevitable outcome and represents one of our most valuable predictions.

V Conclusion and discussions

In our research, we delve into a non-equilibrium model, extending it to consider both isobaric and isochoric conditions. These conditions exhibit variations in the densities, temperatures and expansion velocities of the hot spot. Our results reveal intriguing self-organization phenomena in ion and electron temperatures during the ignition process. Specifically, we find that ion temperature dominates over the electron temperature in this process. This phenomenon arises due to the significant heating effect caused by alpha particles, as well as the distinct deposition rates of alpha particle heating at high temperatures. Additionally, the reduced rate of energy exchange between electrons and D-T ions contributes to the observed bifurcation. During ignition, the inherent structure of higher ion temperature and lower electron temperature directly promotes the enhancement of the D-T reaction and reduces energy loss through electron conduction. Consequently, our ion-electron non-equilibrium model holds promise for improving inertial fusion ignition performed at current mega-joule laser facilities.

Acknowledgements.
This work is supported by the Strategic Priority Research Program of Chinese Academy of Sciences (Grant Nos. XDA25010100 and XDA250050500), National Natural Science Foundation of China (Grants No. 12075204), and Shanghai Municipal Science and Technology Key Project (No. 22JC1401500). Dong Wu thanks the sponsorship from Yangyang Development Fund. The junior undergraduate students, X.-Y. Fu, Z.-Y. Guo, Q.-H. Wang, and R.-C. Wang, all contributed equally to this work.

References

  • (1) S. Atzeni, in Laser-Plasma Interactions and Applications, edited by P. McKenna, D. Neely, R. Bingham,  and D. Jaroszynski (Springer International Publishing) pp. 243–277.
  • Clavin (2017) P. Clavin, Combustion and Flame Special Issue in Honor of Norbert Peters, 175, 80 (2017).
  • Rygg et al. (2009) J. R. Rygg, J. A. Frenje, C. K. Li, F. H. Séguin, R. D. Petrasso, D. D. Meyerhofer,  and C. Stoeckl, Physical Review E 80, 026403 (2009).
  • Tabak et al. (2006) M. Tabak, D. Hinkel, S. Atzeni, E. M. Campbell,  and K. Tanaka, Fusion Science and Technology 49, 254 (2006), publisher: Taylor & Francis _eprint: https://doi.org/10.13182/FST49-3-254.
  • Ghasemi et al. (2014) S. A. Ghasemi, A. H. Farahbod,  and S. Sobhanian, AIP Advances 4, 077130 (2014).
  • Xu et al. (2023) Z. Xu, F. Wu, B. Jiang, S. Kawata,  and J. Zhang, Nuclear Fusion 63, 126062 (2023), publisher: IOP Publishing.
  • Clark and Tabak (2007) D. S. Clark and M. Tabak, Nuclear Fusion 47, 1147 (2007).
  • Farahbod et al. (2014) A. H. Farahbod, S. A. Ghasemi, M. J. Jafari, S. Rezaei,  and S. Sobhanian, The European Physical Journal D 68, 314 (2014).
  • Zylstra et al. (2022) A. B. Zylstra, O. A. Hurricane, D. A. Callahan, A. L. Kritcher, J. E. Ralph, H. F. Robey, J. S. Ross, C. V. Young, K. L. Baker, D. T. Casey, T. Döppner, L. Divol, M. Hohenberger, S. Le Pape, A. Pak, P. K. Patel, R. Tommasini, S. J. Ali, P. A. Amendt, L. J. Atherton, B. Bachmann, D. Bailey, L. R. Benedetti, L. Berzak Hopkins, R. Betti, S. D. Bhandarkar, J. Biener, R. M. Bionta, N. W. Birge, E. J. Bond, D. K. Bradley, T. Braun, T. M. Briggs, M. W. Bruhn, P. M. Celliers, B. Chang, T. Chapman, H. Chen, C. Choate, A. R. Christopherson, D. S. Clark, J. W. Crippen, E. L. Dewald, T. R. Dittrich, M. J. Edwards, W. A. Farmer, J. E. Field, D. Fittinghoff, J. Frenje, J. Gaffney, M. Gatu Johnson, S. H. Glenzer, G. P. Grim, S. Haan, K. D. Hahn, G. N. Hall, B. A. Hammel, J. Harte, E. Hartouni, J. E. Heebner, V. J. Hernandez, H. Herrmann, M. C. Herrmann, D. E. Hinkel, D. D. Ho, J. P. Holder, W. W. Hsing, H. Huang, K. D. Humbird, N. Izumi, L. C. Jarrott, J. Jeet, O. Jones, G. D. Kerbel, S. M. Kerr, S. F. Khan, J. Kilkenny, Y. Kim, H. Geppert Kleinrath, V. Geppert Kleinrath, C. Kong, J. M. Koning, J. J. Kroll, M. K. G. Kruse, B. Kustowski, O. L. Landen, S. Langer, D. Larson, N. C. Lemos, J. D. Lindl, T. Ma, M. J. MacDonald, B. J. MacGowan, A. J. Mackinnon, S. A. MacLaren, A. G. MacPhee, M. M. Marinak, D. A. Mariscal, E. V. Marley, L. Masse, K. Meaney, N. B. Meezan, P. A. Michel, M. Millot, J. L. Milovich, J. D. Moody, A. S. Moore, J. W. Morton, T. Murphy, K. Newman, J.-M. G. Di Nicola, A. Nikroo, R. Nora, M. V. Patel, L. J. Pelz, J. L. Peterson, Y. **, B. B. Pollock, M. Ratledge, N. G. Rice, H. Rinderknecht, M. Rosen, M. S. Rubery, J. D. Salmonson, J. Sater, S. Schiaffino, D. J. Schlossberg, M. B. Schneider, C. R. Schroeder, H. A. Scott, S. M. Sepke, K. Sequoia, M. W. Sherlock, S. Shin, V. A. Smalyuk, B. K. Spears, P. T. Springer, M. Stadermann, S. Stoupin, D. J. Strozzi, L. J. Suter, C. A. Thomas, R. P. J. Town, E. R. Tubman, C. Trosseille, P. L. Volegov, C. R. Weber, K. Widmann, C. Wild, C. H. Wilde, B. M. Van Wonterghem, D. T. Woods, B. N. Woodworth, M. Yamaguchi, S. T. Yang,  and G. B. Zimmerman, Nature 601, 542 (2022), number: 7894 Publisher: Nature Publishing Group.
  • Acree et al. (2022) R. Acree, H. Abu-Shawareb,  and et al, Physical Review Letters 129, 075001 (2022).
  • Acree et al. (2024) R. Acree, H. Abu-Shawareb,  and et al, Physical Review Letters 132, 065102 (2024).
  • Hurricane et al. (2014) O. A. Hurricane, D. A. Callahan, D. T. Casey, P. M. Celliers, C. Cerjan, E. L. Dewald, T. R. Dittrich, T. Döppner, D. E. Hinkel, L. F. B. Hopkins, J. L. Kline, S. Le Pape, T. Ma, A. G. MacPhee, J. L. Milovich, A. Pak, H.-S. Park, P. K. Patel, B. A. Remington, J. D. Salmonson, P. T. Springer,  and R. Tommasini, Nature 506, 343 (2014), publisher: Nature Publishing Group.
  • Lindl et al. (2014) J. Lindl, O. Landen, J. Edwards, E. Moses,  and NIC Team, Physics of Plasmas 21, 020501 (2014).
  • Hartouni et al. (2023) E. P. Hartouni, A. S. Moore, A. J. Crilly, B. D. Appelbe, P. A. Amendt, K. L. Baker, D. T. Casey, D. S. Clark, T. Döppner, M. J. Eckart, J. E. Field, M. Gatu-Johnson, G. P. Grim, R. Hatarik, J. Jeet, S. M. Kerr, J. Kilkenny, A. L. Kritcher, K. D. Meaney, J. L. Milovich, D. H. Munro, R. C. Nora, A. E. Pak, J. E. Ralph, H. F. Robey, J. S. Ross, D. J. Schlossberg, S. M. Sepke, B. K. Spears, C. V. Young,  and A. B. Zylstra, Nature Physics 19, 72 (2023), number: 1 Publisher: Nature Publishing Group.
  • Chang et al. (2010) P. Chang, R. Betti, B. K. Spears, K. S. Anderson, J. Edwards, M. Fatenejad, J. D. Lindl, R. L. McCrory, R. Nora,  and D. Shvarts, Physical Review Letters 104, 135002 (2010).
  • Döppner et al. (2015) T. Döppner, D. Callahan, O. Hurricane, D. Hinkel, T. Ma, H.-S. Park, L. Berzak Hopkins, D. Casey, P. Celliers, E. Dewald, T. Dittrich, S. Haan, A. Kritcher, A. MacPhee, S. Le Pape, A. Pak, P. Patel, P. Springer, J. Salmonson, R. Tommasini, L. Benedetti, E. Bond, D. Bradley, J. Caggiano, J. Church, S. Dixit, D. Edgell, M. Edwards, D. Fittinghoff, J. Frenje, M. Gatu Johnson, G. Grim, R. Hatarik, M. Havre, H. Herrmann, N. Izumi, S. Khan, J. Kline, J. Knauer, G. Kyrala, O. Landen, F. Merrill, J. Moody, A. Moore, A. Nikroo, J. Ralph, B. Remington, H. Robey, D. Sayre, M. Schneider, H. Streckert, R. Town, D. Turnbull, P. Volegov, A. Wan, K. Widmann, C. Wilde,  and C. Yeamans, Physical Review Letters 115, 055001 (2015).
  • Gopalaswamy et al. (2024) V. Gopalaswamy, C. A. Williams, R. Betti, D. Patel, J. P. Knauer, A. Lees, D. Cao, E. M. Campbell, P. Farmakis, R. Ejaz, K. S. Anderson, R. Epstein, J. Carroll-Nellenbeck, I. V. Igumenshchev, J. A. Marozas, P. B. Radha, A. A. Solodov, C. A. Thomas, K. M. Woo, T. J. B. Collins, S. X. Hu, W. Scullin, D. Turnbull, V. N. Goncharov, K. Churnetski, C. J. Forrest, V. Y. Glebov, P. V. Heuer, H. McClow, R. C. Shah, C. Stoeckl, W. Theobald, D. H. Edgell, S. Ivancic, M. J. Rosenberg, S. P. Regan, D. Bredesen, C. Fella, M. Koch, R. T. Janezic, M. J. Bonino, D. R. Harding, K. A. Bauer, S. Sampat, L. J. Waxer, M. Labuzeta, S. F. B. Morse, M. Gatu-Johnson, R. D. Petrasso, J. A. Frenje, J. Murray, B. Serrato, D. Guzman, C. Shuldberg, M. Farrell,  and C. Deeney, Nature Physics , 1 (2024), publisher: Nature Publishing Group.
  • Churnetski et al. (2024) K. Churnetski, K. M. Woo, W. Theobald, C. Stoeckl, L. Ceurvorst, V. Gopalaswamy, H. Rinderknecht, P. V. Heuer, J. Knauer, C. Forrest, I. Igumenshchev, S. Ivancic, M. Michalko, R. Shah, A. Lees, R. Bahukutumbi, R. Betti, C. Thomas, S. Regan, J. Kunimune, C. Wink, P. Adrian, M. Gatu Johnson,  and J. Frenje, Three-Dimensional Reconstruction of Implosion Stagnation in Laser Direct Drive on Omega, preprint (SSRN, 2024).
  • (19) S. Atzeni and J. Meyer-ter Vehn, The Physics of Inertial Fusion: Beam Plasma Interaction, Hydrodynamics, Hot Dense Matter, Oxford Science Publications No. 125 (Clarendon Press ; Oxford University Press).
  • Hurricane et al. (2023) O. Hurricane, P. Patel, R. Betti, D. Froula, S. Regan, S. Slutz, M. Gomez,  and M. Sweeney, Reviews of Modern Physics 95, 025005 (2023).
  • Daughton et al. (2023) W. Daughton, B. J. Albright, S. M. Finnegan, B. M. Haines, J. L. Kline, J. P. Sauppe,  and J. M. Smidt, Physics of Plasmas 30, 012704 (2023), arXiv:2207.00093 [physics].
  • Fan et al. (2016) Z. Fan, J. Liu, B. Liu, C. Yu,  and X. T. He, Physics of Plasmas 23, 010703 (2016).
  • Fraley et al. (1974) G. S. Fraley, E. J. Linnebur, R. J. Mason,  and R. L. Morse, The Physics of Fluids 17, 474 (1974).
  • Wu et al. (2018) D. Wu, X. T. He, W. Yu,  and S. Fritzsche, High Power Laser Science and Engineering 6, e50 (2018).
  • Spears et al. (2008) B. Spears, D. Hicks, C. Velsko, M. Stoyer, H. Robey, D. Munro, S. Haan, O. Landen, A. Nikroo,  and H. Huang, Journal of Physics: Conference Series 112, 022003 (2008).
  • (26) S. Atzeni and A. Caruso,  .
  • Daligault and Simoni (2019) J. Daligault and J. Simoni, Physical Review E 100, 043201 (2019), publisher: American Physical Society.
  • Temporal et al. (2012) M. Temporal, V. Brandon, B. Canaud, J. Didelez, R. Fedosejevs,  and R. Ramis, Nuclear Fusion 52, 103011 (2012).