Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: CC Zero
arXiv:2311.00701v2 [physics.flu-dyn] 08 Mar 2024

The Moving Discontinuous Galerkin Method with Interface Condition Enforcement for the Simulation of Hypersonic, Viscous Flows

Eric J. Ching, Andrew D. Kercher, Andrew Corrigan Laboratories for Computational Physics and Fluid Dynamics, U.S. Naval Research Laboratory, 4555 Overlook Ave SW, Washington, DC 20375
Abstract

The moving discontinuous Galerkin method with interface condition enforcement (MDG-ICE) is a high-order, r𝑟ritalic_r-adaptive method that treats the grid as a variable and weakly enforces the conservation law, constitutive law, and corresponding interface conditions in order to implicitly fit high-gradient flow features. In this paper, we develop an optimization solver based on the Levenberg-Marquardt algorithm that features an anisotropic, locally adaptive penalty method to enhance robustness and prevent cell degeneration in the computation of hypersonic, viscous flows. Specifically, we incorporate an anisotropic grid regularization based on the mesh-implied metric that inhibits grid motion in directions with small element length scales, an element shape regularization that inhibits nonlinear deformations of the high-order elements, and a penalty regularization that penalizes degenerate elements. Additionally, we introduce a procedure for locally scaling the regularization operators in an adaptive, elementwise manner in order to maintain grid validity. We apply the proposed MDG-ICE formulation to two- and three-dimensional test cases involving viscous shocks and/or boundary layers, including Mach 17.6 hypersonic viscous flow over a circular cylinder and Mach 5 hypersonic viscous flow over a sphere, which are very challenging test cases for conventional numerical schemes on simplicial grids. Even without artificial dissipation, the computed solutions are free from spurious oscillations and yield highly symmetric surface heat-flux profiles.

keywords:
Discontinuous Galerkin method; Interface condition enforcement; MDG-ICE; Implicit shock fitting; Anisotropic curvilinear r𝑟ritalic_r-adaptivity; Mesh adaptation
footnotetext:
                    DISTRIBUTION STATEMENT A. Approved for public release. Distribution is unlimited.

1 Introduction

The moving discontinuous Galerkin finite element method with interface condition enforcement (MDG-ICE) is an implicit shock fitting method capable of handling complex shock dynamics (Cor18, ; Ker20, ; Ker20_LS, ). The method is a unique variation of the well-known discontinuous Galerkin (DG) method (Ree73, ; Coc00, ). Specifically, neighboring elements are not coupled through interfacial, single-valued numerical fluxes; instead, the conservation law and interface conditions (known as the generalized Rankine-Hugoniot jump conditions (Maj12, )) are directly discretized, and the grid is treated as a variable. By simultaneously solving for the flow field and discrete geometry, MDG-ICE is able to compute highly accurate high-order solutions without artificial dissipation as the grid points are automatically adjusted to fit shocks and resolve smooth regions of the flow with sharp gradients. This has significant advantages over traditional shock capturing approaches, such as artificial viscosity and limiting; the former introduces low-order errors into high-order approximations that can excessively smear discontinuous and high-gradient features (and even cause movement of such features that may then contaminate prediction of other flow structures (Chi18, )), while the latter can obstruct iterative convergence, fail to provide sufficient stabilization, and be used only for certain approximation orders and/or element types. Furthermore, since MDG-ICE adapts the grid to satisfy the weak form, grid interfaces are naturally repositioned to fit a priori unknown shocks with arbitrary topology, overcoming key limitations of explicit shock fitting methods ((Mor02, ; Sal09, ; Sal11, )). Note that in the inviscid case, shocks are fit exactly along grid interfaces, while in the viscous setting, high-aspect-ratio cells form to resolve viscous shocks, which are sharp (yet smooth) features, via anisotropic curvilinear r𝑟ritalic_r-adaptivity. In previous work, MDG-ICE was shown to achieve not only extremely sharp, oscillation-free viscous-shock profiles, but also significantly higher accuracy than standard DG schemes in boundary-layer problems (Ker20, ). Another form of implicit shock fitting is the high-order implicit shock tracking (HOIST) framework developed by Zahr and Persson (Zah18, ) and improved in (Zah20, ; Shi22, ; Hua22, ; Hua23, ), which also treats the discrete geometry as a variable while retaining a standard discontinuous Galerkin method. High-quality solutions to inviscid flows with discontinuities on coarse grids have been achieved using HOIST. Additonally, a variant of MDG-ICE has been developed by Luo et al. (Luo21, ).

Figure 1.1 presents the temperature fields and corresponding meshes for hypersonic viscous flow over a half-cylinder at Mach 5 and Reynolds number 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT on two-dimensional triangular grids. A DG(𝒫1subscript𝒫1\mathcal{P}_{1}caligraphic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) solution (left), where 𝒫psubscript𝒫𝑝\mathcal{P}_{p}caligraphic_P start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT denotes the space of polynomials of total degree p𝑝pitalic_p on simplicial elements, with artificial viscosity is compared with an isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution (right) without any additional stabilization (Ker20, ). The corresponding meshes are also displayed. Unlike in the MDG-ICE solution, the shock in the DG solution is noticeably smeared, and spurious oscillations are visible both at the shock and in the shock layer. The resulting surface heating profile in the MDG-ICE solution is highly symmetric, which, despite the relatively simple flow conditions, is nevertheless encouraging since surface heat-flux predictions in state-of-the-practice finite-volume simulations of external hypersonic flows are known to deterioriate considerably when using simplicial grids (Nom04, ; Gno04, ). In particular, finite-volume schemes are extremely sensitive to misalignment between the grid and the shock and boundary layer, which generates errors in vorticity and entropy that propagate downstream and negatively impact the accuracy of computed surface quantities. Shock-tailored quadrilateral/hexahedral grids are typically needed to obtain heat-flux profiles that do not exhibit noticeable nonphysical asymmetries (Can09, ). Constructing such grids for large-scale geometries often requires significant time and user effort (Nom04, ; Can09, ), which is exacerbated when performing parametric studies.

Refer to caption
Figure 1.1: Comparison of DG(𝒫1subscript𝒫1\mathcal{P}_{1}caligraphic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) (with artificial viscosity) (left) and MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) (right) solutions to two-dimensional Mach 5 flow over a cylinder at Re=104Resuperscript104\mathrm{Re}=10^{4}roman_Re = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT (Ker20, ). The corresponding meshes are also displayed.

In contrast, implicit shock-fitting methods, such as MDG-ICE, automatically produce shock-aligned grids as part of the solution process, thereby mitigating the need for shock-tailored grids composed of cubiod elements, which significantly reduces the burden associated with generating a suitable grid for high-speed flow calculations. However, a major difficulty encountered in MDG-ICE calculations is frequent cell degeneration (i.e., the determinant of the geometric Jacobian becomes negative), particularly as curved, high-aspect-ratio cells form to resolve sharp viscous features. Figure 1.2 displays the final mesh for an MDG-ICE solution to the same cylinder problem, but with a higher Reynolds number (Ker20, ). Degenerate cells were treated via longest-edge refinement, leading to the generation of “sliver” elements that introduce unnecessary degrees of freedom. Furthermore, although adequate for this relatively simple flow, in problems of moderately greater complexity, this strategy either requires an inordinate number of refinements or, worse, simply fails to recover a valid grid. Given this significant bottleneck precluding the application of MDG-ICE to more complicated, larger-scale configurations, the primary objective of this work is to develop an optimization solver that improves the robustness of the Levenberg-Marquardt algorithm employed in (Ker20, ) and is capable of producing high-order approximations of hypersonic viscous flows in two and three dimensions without introducing low-order errors associated with artificial stabilization. The key feature of the solver is an anisotropic, locally adaptive penalty technique. In particular, we leverage the mesh-implied metric commonly employed in output-based anisotropic mesh adaptation (Fid07, ; Oli08, ), which encodes information about the local element size and orientation (even on curved, anisotropic grids), to create an anisotropic grid regularization that inhibits grid motion in directions with small element length scales. This is similar to the inverse-volume scaling introduced by Zahr et al. (Zah20, ) and employed in (Ker20_LS, ), but with element anisotropy explicitly taken into account. Furthermore, we incorporate additional regularization operators and develop an adaptive elementwise regularization strategy that locally adjusts regularization parameters as needed to maintain grid validity. We also employ increment limiting as described in (Cez15, ), where the increment at each iteration is dynamically reduced to mitigate excessive changes in pressure and/or density. The improved solver, paired with a full parallelization of our MDG-ICE implementation, enables consideration of significantly more challenging hypersonic test cases. We specifically employ simplicial grids that are initially coarse with respect to the final high-gradient features in order to demonstrate the potential of MDG-ICE to greatly alleviate the burden of mesh generation on the user. Finally, we note that although incorporating artificial dissipation into the formulation in some capacity (e.g., applied only during intermediate iterations or minimally in the final solution) may be beneficial, the present study aims to aggressively test the underlying MDG-ICE formulation without the aid of additional stabilization and, as a secondary goal, identify in what ways stabilization would be most useful; incorporation of such dissipation mechanisms is left for future work.

Refer to caption
Figure 1.2: Final mesh for the isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution to two-dimensional Mach 5 flow over a cylinder at Re=105Resuperscript105\mathrm{Re}=10^{5}roman_Re = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT (Ker20, ).

The remainder of this paper is organized as follows. Sections 2 and 3 briefly summarize the governing equations considered in this study and the basic MDG-ICE formulation, respectively. The improved optimization solver, which is the primary contribution of this work, is then detailed in Section 4. Section 5presents solutions for a variety of challenging test cases, including two- and three-dimensional hypersonic viscous flows over blunt bodies in which surface heat flux is a target quantity. We conclude with final remarks and recommendations for future work.

2 Governing equations

Let ΩdΩsuperscript𝑑\Omega\subset\mathbb{R}^{d}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT denote the domain, which can be either a space-time domain, Ωd=dx+1Ωsuperscript𝑑subscript𝑑𝑥1\Omega\subset\mathbb{R}^{d=d_{x}+1}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d = italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + 1 end_POSTSUPERSCRIPT, or a spatial domain, Ωd=dxΩsuperscript𝑑subscript𝑑𝑥\Omega\subset\mathbb{R}^{d=d_{x}}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d = italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. Consider the following nonlinear conservation law governing the vector of m𝑚mitalic_m state variables, y:Ωm:𝑦Ωsuperscript𝑚y:\Omega\rightarrow\mathbb{R}^{m}italic_y : roman_Ω → blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT:

(y,xy)=0 in Ω,𝑦subscript𝑥𝑦0 in Ω\nabla\cdot\mathcal{F}(y,\nabla_{x}y)=0\text{ in }\Omega,∇ ⋅ caligraphic_F ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = 0 in roman_Ω , (2.1)

where :m×m×dxm×d:superscript𝑚superscript𝑚subscript𝑑𝑥superscript𝑚𝑑\mathcal{F}:\mathbb{R}^{m}\times\mathbb{R}^{m\times d_{x}}\rightarrow\mathbb{R% }^{m\times d}caligraphic_F : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT × blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d end_POSTSUPERSCRIPT is the flux and x()subscript𝑥\nabla_{x}(\cdot)∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( ⋅ ) denotes the spatial gradient,

xy=(yx1,,yxdx).subscript𝑥𝑦𝑦subscript𝑥1𝑦subscript𝑥subscript𝑑𝑥\nabla_{x}y=\left(\frac{\partial y}{\partial x_{1}},\ldots,\frac{\partial y}{% \partial x_{d_{x}}}\right).∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y = ( divide start_ARG ∂ italic_y end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG , … , divide start_ARG ∂ italic_y end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG ) . (2.2)

In the case of a space-time domain (i.e., d=dx+1𝑑subscript𝑑𝑥1d=d_{x}+1italic_d = italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + 1), the space-time flux is given by

(y,xy)=(1x(y,xy),,dxx(y,xy),y),𝑦subscript𝑥𝑦superscriptsubscript1𝑥𝑦subscript𝑥𝑦superscriptsubscriptsubscript𝑑𝑥𝑥𝑦subscript𝑥𝑦𝑦\mathcal{F}(y,\nabla_{x}y)=\left(\mathcal{F}_{1}^{x}(y,\nabla_{x}y),\ldots,% \mathcal{F}_{d_{x}}^{x}(y,\nabla_{x}y),y\right),caligraphic_F ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = ( caligraphic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) , … , caligraphic_F start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) , italic_y ) , (2.3)

where x:m×m×dxm×dx:superscript𝑥superscript𝑚superscript𝑚subscript𝑑𝑥superscript𝑚subscript𝑑𝑥\mathcal{F}^{x}:\mathbb{R}^{m}\times\mathbb{R}^{m\times d_{x}}\rightarrow% \mathbb{R}^{m\times d_{x}}caligraphic_F start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT × blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the spatial flux, defined as the difference between the convective flux and the viscous flux,

x(y,xy)=c(y)v(y,xy).superscript𝑥𝑦subscript𝑥𝑦superscript𝑐𝑦superscript𝑣𝑦subscript𝑥𝑦\mathcal{F}^{x}(y,\nabla_{x}y)=\mathcal{F}^{c}(y)-\mathcal{F}^{v}(y,\nabla_{x}% y).caligraphic_F start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = caligraphic_F start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT ( italic_y ) - caligraphic_F start_POSTSUPERSCRIPT italic_v end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) .

The divergence operator in Equation (2.1) is then the space-time divergence operator, given by

(y,xy)=xx(y,xy)+ty.𝑦subscript𝑥𝑦subscript𝑥superscript𝑥𝑦subscript𝑥𝑦𝑡𝑦\nabla\cdot\mathcal{F}(y,\nabla_{x}y)=\nabla_{x}\cdot\mathcal{F}^{x}(y,\nabla_% {x}y)+\frac{\partial}{\partial t}y.∇ ⋅ caligraphic_F ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⋅ caligraphic_F start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) + divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG italic_y . (2.4)

In the case of a spatial domain (i.e., d=dx𝑑subscript𝑑𝑥d=d_{x}italic_d = italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT), the flux is simply the spatial flux, and the divergence operator is the spatial divergence operator. In this study, we consider the viscous Burgers equation and the compressible Navier-Stokes equations.

2.1 One-dimensional viscous Burgers equation

The one-dimensional viscous Burgers equation involves a single state variable, y:Ω1:𝑦Ωsuperscript1y:\Omega\rightarrow\mathbb{R}^{1}italic_y : roman_Ω → blackboard_R start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT. The convective and viscous fluxes are given by

c(y)=12y2,v(y,xy)=μxy,formulae-sequencesuperscript𝑐𝑦12superscript𝑦2superscript𝑣𝑦subscript𝑥𝑦𝜇subscript𝑥𝑦\mathcal{F}^{c}\left(y\right)=\frac{1}{2}y^{2},\quad\mathcal{F}^{v}\left(y,% \nabla_{x}y\right)=\mu\nabla_{x}y,caligraphic_F start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT ( italic_y ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , caligraphic_F start_POSTSUPERSCRIPT italic_v end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = italic_μ ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y , (2.5)

where μ+𝜇subscript\mu\in\mathbb{R}_{+}italic_μ ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the viscosity. The spatial flux is then written as

x(y,xy)=(12y2μxy),superscript𝑥𝑦subscript𝑥𝑦12superscript𝑦2𝜇subscript𝑥𝑦\mathcal{F}^{x}\left(y,\nabla_{x}y\right)=\left(\frac{1}{2}y^{2}-\mu\nabla_{x}% y\right),caligraphic_F start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_μ ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) , (2.6)

2.2 Compressible Navier-Stokes equations

The vector of state variables, y:Ωm:𝑦Ωsuperscript𝑚y:\Omega\rightarrow\mathbb{R}^{m}italic_y : roman_Ω → blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT, where m=dx+2𝑚subscript𝑑𝑥2m=d_{x}+2italic_m = italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + 2, is given by

y=(ρ,ρv1,,ρvdx,ρE),𝑦𝜌𝜌subscript𝑣1𝜌subscript𝑣subscript𝑑𝑥𝜌𝐸y=\left(\rho,\rho v_{1},\ldots,\rho v_{d_{x}},\rho E\right),italic_y = ( italic_ρ , italic_ρ italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_ρ italic_v start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_ρ italic_E ) , (2.7)

where ρ:Ω+:𝜌Ωsubscript\rho:\Omega\rightarrow\mathbb{R}_{+}italic_ρ : roman_Ω → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the density, v=(v1,,vdx)𝑣subscript𝑣1subscript𝑣subscript𝑑𝑥v=\left(v_{1},\ldots,v_{d_{x}}\right)italic_v = ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_v start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) is the velocity vector, with v:mdx:𝑣superscript𝑚superscriptsubscript𝑑𝑥v:\mathbb{R}^{m}\rightarrow\mathbb{R}^{d_{x}}italic_v : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, and E:m+:𝐸superscript𝑚subscriptE:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}italic_E : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the specific total energy. The i𝑖iitalic_ith spatial component of the convective flux is written as

ic(y)=(ρvi,ρviv1+Pδi1,,ρvivdx+Pδidx,ρHvi),superscriptsubscript𝑖𝑐𝑦𝜌subscript𝑣𝑖𝜌subscript𝑣𝑖subscript𝑣1𝑃subscript𝛿𝑖1𝜌subscript𝑣𝑖subscript𝑣subscript𝑑𝑥𝑃subscript𝛿𝑖subscript𝑑𝑥𝜌𝐻subscript𝑣𝑖\mathcal{F}_{i}^{c}\left(y\right)=\left(\rho v_{i},\rho v_{i}v_{1}+P\delta_{i1% },\ldots,\rho v_{i}v_{d_{x}}+P\delta_{id_{x}},\rho Hv_{i}\right),caligraphic_F start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT ( italic_y ) = ( italic_ρ italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_ρ italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_P italic_δ start_POSTSUBSCRIPT italic_i 1 end_POSTSUBSCRIPT , … , italic_ρ italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT + italic_P italic_δ start_POSTSUBSCRIPT italic_i italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_ρ italic_H italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (2.8)

where P:m+:𝑃superscript𝑚subscriptP:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}italic_P : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the pressure, δijsubscript𝛿𝑖𝑗\delta_{ij}italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the Kronecker delta, and H=(ρE+P)/ρ𝐻𝜌𝐸𝑃𝜌H=\left(\rho E+P\right)/\rhoitalic_H = ( italic_ρ italic_E + italic_P ) / italic_ρ is the specific total enthalpy, with H:m+:𝐻superscript𝑚subscriptH:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}italic_H : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT. In this work, we assume a calorically perfect gas, such that

P=(γ1)(ρE12i=1dxρvivi),𝑃𝛾1𝜌𝐸12superscriptsubscript𝑖1subscript𝑑𝑥𝜌subscript𝑣𝑖subscript𝑣𝑖P=\left(\gamma-1\right)\left(\rho E-\frac{1}{2}\sum_{i=1}^{d_{x}}\rho v_{i}v_{% i}\right),italic_P = ( italic_γ - 1 ) ( italic_ρ italic_E - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ρ italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (2.9)

where γ𝛾\gammaitalic_γ is the specific heat ratio, set to 1.41.41.41.4. Note that the assumption of a single-species calorically perfect gas is appropriate for evaluating the capability of a numerical method to handle sharp, high-gradient features such as strong shocks and boundary layers, although more complex gas and transport models are typically employed in realistic hypersonic applications (Gno99, ).

The i𝑖iitalic_ith spatial component of the viscous flux is written as

iν(y,xy)=(0,τ1i,,τdxi,j=1dxτijvjqi),superscriptsubscript𝑖𝜈𝑦subscript𝑥𝑦0subscript𝜏1𝑖subscript𝜏subscript𝑑𝑥𝑖superscriptsubscript𝑗1subscript𝑑𝑥subscript𝜏𝑖𝑗subscript𝑣𝑗subscript𝑞𝑖\mathcal{F}_{i}^{\nu}\left(y,\nabla_{x}y\right)=\left(0,\tau_{1i},\ldots,\tau_% {d_{x}i},\sum_{j=1}^{d_{x}}\tau_{ij}v_{j}-q_{i}\right),caligraphic_F start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = ( 0 , italic_τ start_POSTSUBSCRIPT 1 italic_i end_POSTSUBSCRIPT , … , italic_τ start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_τ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (2.10)

where q:m×m×dxdx:𝑞superscript𝑚superscript𝑚subscript𝑑𝑥superscriptsubscript𝑑𝑥q:\mathbb{R}^{m}\times\mathbb{R}^{m\times d_{x}}\rightarrow\mathbb{R}^{d_{x}}italic_q : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT × blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the heat flux and τ:m×m×dxdx×dx:𝜏superscript𝑚superscript𝑚subscript𝑑𝑥superscriptsubscript𝑑𝑥subscript𝑑𝑥\tau:\mathbb{R}^{m}\times\mathbb{R}^{m\times d_{x}}\rightarrow\mathbb{R}^{d_{x% }\times d_{x}}italic_τ : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT × blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the viscous stress tensor. The heat flux is expanded as q=κxT𝑞𝜅subscript𝑥𝑇q=-\kappa\nabla_{x}Titalic_q = - italic_κ ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_T, where κ:m+:𝜅superscript𝑚subscript\kappa:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}italic_κ : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the thermal conductivity and T:m+:𝑇superscript𝑚subscriptT:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}italic_T : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the temperature, calculated as T=P/(ρR)𝑇𝑃𝜌𝑅T=P/(\rho R)italic_T = italic_P / ( italic_ρ italic_R ), with R=287𝑅287R=287italic_R = 287 denoting the specific gas constant. The i𝑖iitalic_ith spatial component of the viscous stress tensor is given by

τi=μ(v1xi+vix1δi123j=1dxvjxj,,vdxxi+vixdxδidx23j=1dxvjxj),subscript𝜏𝑖𝜇subscript𝑣1subscript𝑥𝑖subscript𝑣𝑖subscript𝑥1subscript𝛿𝑖123superscriptsubscript𝑗1subscript𝑑𝑥subscript𝑣𝑗subscript𝑥𝑗subscript𝑣subscript𝑑𝑥subscript𝑥𝑖subscript𝑣𝑖subscript𝑥subscript𝑑𝑥subscript𝛿𝑖subscript𝑑𝑥23superscriptsubscript𝑗1subscript𝑑𝑥subscript𝑣𝑗subscript𝑥𝑗\tau_{i}=\mu\left(\frac{\partial v_{1}}{\partial x_{i}}+\frac{\partial v_{i}}{% \partial x_{1}}-\delta_{i1}\frac{2}{3}\sum_{j=1}^{d_{x}}\frac{\partial v_{j}}{% \partial x_{j}},\ldots,\frac{\partial v_{d_{x}}}{\partial x_{i}}+\frac{% \partial v_{i}}{\partial x_{d_{x}}}-\delta_{id_{x}}\frac{2}{3}\sum_{j=1}^{d_{x% }}\frac{\partial v_{j}}{\partial x_{j}}\right),italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_μ ( divide start_ARG ∂ italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG + divide start_ARG ∂ italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - italic_δ start_POSTSUBSCRIPT italic_i 1 end_POSTSUBSCRIPT divide start_ARG 2 end_ARG start_ARG 3 end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT divide start_ARG ∂ italic_v start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG , … , divide start_ARG ∂ italic_v start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG + divide start_ARG ∂ italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG - italic_δ start_POSTSUBSCRIPT italic_i italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG 2 end_ARG start_ARG 3 end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT divide start_ARG ∂ italic_v start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ) , (2.11)

where μ:m+:𝜇superscript𝑚subscript\mu:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}italic_μ : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the dynamic viscosity obtained from Sutherland’s law,

μ=μ0T0+CT+C(TT0)32.𝜇subscript𝜇0subscript𝑇0𝐶𝑇𝐶superscript𝑇subscript𝑇032\mu=\mu_{0}\frac{T_{0}+C}{T+C}\left(\frac{T}{T_{0}}\right)^{\frac{3}{2}}.italic_μ = italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_C end_ARG start_ARG italic_T + italic_C end_ARG ( divide start_ARG italic_T end_ARG start_ARG italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT . (2.12)

u0+subscript𝑢0subscriptu_{0}\in\mathbb{R}_{+}italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is the reference dynamic viscosity, and T0+subscript𝑇0subscriptT_{0}\in\mathbb{R}_{+}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT and C+𝐶subscriptC\in\mathbb{R}_{+}italic_C ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT are reference temperatures. Nondimensionalizing Equation (2.12) using freestream quantities, denoted ()subscript\left(\cdot\right)_{\infty}( ⋅ ) start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, yields

μ*=1+C*T*+C*(T*)32,superscript𝜇1superscript𝐶superscript𝑇superscript𝐶superscriptsuperscript𝑇32\mu^{*}=\frac{1+C^{*}}{T^{*}+C^{*}}\left(T^{*}\right)^{\frac{3}{2}},italic_μ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = divide start_ARG 1 + italic_C start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG start_ARG italic_T start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT + italic_C start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG ( italic_T start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT divide start_ARG 3 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT , (2.13)

where μ*=μ/μsuperscript𝜇𝜇subscript𝜇\mu^{*}=\mu/\mu_{\infty}italic_μ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_μ / italic_μ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, T*=T/Tsuperscript𝑇𝑇subscript𝑇T^{*}=T/T_{\infty}italic_T start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_T / italic_T start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, and C*=C/Tsuperscript𝐶𝐶subscript𝑇C^{*}=C/T_{\infty}italic_C start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = italic_C / italic_T start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. In this study, we use C=110𝐶110C=110italic_C = 110 K and T=293.15subscript𝑇293.15T_{\infty}=293.15italic_T start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 293.15 K. The thermal conductivity is computed as

k=cpμPr,𝑘subscript𝑐𝑝𝜇Prk=\frac{c_{p}\mu}{\mathrm{Pr}},italic_k = divide start_ARG italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_μ end_ARG start_ARG roman_Pr end_ARG , (2.14)

where Pr=0.72Pr0.72\mathrm{Pr}=0.72roman_Pr = 0.72 is the Prandtl number and cp=γR/(γ1)subscript𝑐𝑝𝛾𝑅𝛾1c_{p}=\gamma R/\left(\gamma-1\right)italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_γ italic_R / ( italic_γ - 1 ) is the specific heat capacity at constant pressure.

Two important nondimensional quantities that characterize compressible, viscous flows are the Reynolds number, ReRe\mathrm{Re}roman_Re, and Mach number, MaMa\mathrm{Ma}roman_Ma, defined as

Re=ρ|v|Lμ,Re𝜌𝑣𝐿𝜇\mathrm{Re}=\frac{\rho\lvert v\rvert L}{\mu},roman_Re = divide start_ARG italic_ρ | italic_v | italic_L end_ARG start_ARG italic_μ end_ARG , (2.15)

where L+𝐿subscriptL\in\mathbb{R}_{+}italic_L ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT is a characteristic length scale, and

Ma=|v|c,Ma𝑣𝑐\mathrm{Ma}=\frac{\lvert v\rvert}{c},roman_Ma = divide start_ARG | italic_v | end_ARG start_ARG italic_c end_ARG , (2.16)

where c=γP/ρ𝑐𝛾𝑃𝜌c=\sqrt{\gamma P/\rho}italic_c = square-root start_ARG italic_γ italic_P / italic_ρ end_ARG is the speed of sound, with c:m+:𝑐superscript𝑚subscriptc:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}italic_c : blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT, respectively.

3 Moving discontinuous Galerkin method with interface condition enforcement

In this subsection, we briefly review the MDG-ICE formulation for a system governed by a set of conservation laws and a generalized constitutive law. Further details reagarding the formulation and its application can be found in (Ker20, ), and an alternative least-squares formulation is described in (Ker20_LS, ).

Let ΩΩ\Omegaroman_Ω be partitioned by 𝒯𝒯\mathcal{T}caligraphic_T, which consists of cells κ𝜅\kappaitalic_κ. Furthermore, we define the set \mathcal{E}caligraphic_E of interfaces such that ϵ=κ𝒯κsubscriptitalic-ϵsubscript𝜅𝒯𝜅\bigcup_{\epsilon\in\mathcal{E}}=\bigcup_{\kappa\in\mathcal{T}}\partial\kappa⋃ start_POSTSUBSCRIPT italic_ϵ ∈ caligraphic_E end_POSTSUBSCRIPT = ⋃ start_POSTSUBSCRIPT italic_κ ∈ caligraphic_T end_POSTSUBSCRIPT ∂ italic_κ. The normal over each interface ϵitalic-ϵ\epsilon\in\mathcal{E}italic_ϵ ∈ caligraphic_E is denoted n:ϵd:𝑛italic-ϵsuperscript𝑑n:\epsilon\rightarrow\mathbb{R}^{d}italic_n : italic_ϵ → blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT. For space-time domains, nx:ϵdx:subscript𝑛𝑥italic-ϵsuperscriptsubscript𝑑𝑥n_{x}:\epsilon\rightarrow\mathbb{R}^{d_{x}}italic_n start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT : italic_ϵ → blackboard_R start_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT denotes the spatial normal.

3.1 Strong and weak formulations

Consider the following conservation law, constitutive law, and associated interface conditions in strong form:

(y,σ)=0𝑦𝜎0\displaystyle\nabla\cdot\mathcal{F}\left(y,\sigma\right)=0∇ ⋅ caligraphic_F ( italic_y , italic_σ ) = 0 in κκ𝒯,in 𝜅for-all𝜅𝒯\displaystyle\textup{ in }\kappa\qquad\forall\kappa\in\mathcal{T},in italic_κ ∀ italic_κ ∈ caligraphic_T , (3.1)
σG(y)xy=0𝜎𝐺𝑦subscript𝑥𝑦0\displaystyle\sigma-G\left(y\right)\nabla_{x}y=0italic_σ - italic_G ( italic_y ) ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y = 0 in κκ𝒯,in 𝜅for-all𝜅𝒯\displaystyle\textup{ in }\kappa\qquad\forall\kappa\in\mathcal{T},in italic_κ ∀ italic_κ ∈ caligraphic_T , (3.2)
n(y,σ)=0\displaystyle\left\llbracket n\cdot\mathcal{F}\left(y,\sigma\right)\right% \rrbracket=0⟦ italic_n ⋅ caligraphic_F ( italic_y , italic_σ ) ⟧ = 0 on ϵϵ,on italic-ϵfor-allitalic-ϵ\displaystyle\textup{ on }\epsilon\qquad\forall\epsilon\in\mathcal{E},on italic_ϵ ∀ italic_ϵ ∈ caligraphic_E , (3.3)
{{G(y)}}ynx=0\displaystyle\left\{\!\!\left\{G\left(y\right)\right\}\!\!\right\}\left% \llbracket y\otimes n_{x}\right\rrbracket=0{ { italic_G ( italic_y ) } } ⟦ italic_y ⊗ italic_n start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⟧ = 0 on ϵϵ,on italic-ϵfor-allitalic-ϵ\displaystyle\textup{ on }\epsilon\qquad\forall\epsilon\in\mathcal{E},on italic_ϵ ∀ italic_ϵ ∈ caligraphic_E , (3.4)

where σ:Ωm×dx:𝜎Ωsuperscript𝑚subscript𝑑𝑥\sigma:\Omega\rightarrow\mathbb{R}^{m\times d_{x}}italic_σ : roman_Ω → blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is an auxiliary variable, G(y)m×dx×m×dx𝐺𝑦superscript𝑚subscript𝑑𝑥𝑚subscript𝑑𝑥G(y)\in\mathbb{R}^{m\times d_{x}\times m\times d_{x}}italic_G ( italic_y ) ∈ blackboard_R start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT × italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the homogeneity tensor that satisfies G(y)xy=v(y,xy)=xyv(y,xy)xy𝐺𝑦subscript𝑥𝑦superscript𝑣𝑦subscript𝑥𝑦superscriptsubscriptsubscript𝑥𝑦𝑣𝑦subscript𝑥𝑦subscript𝑥𝑦G(y)\nabla_{x}y=\mathcal{F}^{v}\left(y,\nabla_{x}y\right)=\mathcal{F}_{\nabla_% {x}y}^{v}\left(y,\nabla_{x}y\right)\nabla_{x}yitalic_G ( italic_y ) ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y = caligraphic_F start_POSTSUPERSCRIPT italic_v end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) = caligraphic_F start_POSTSUBSCRIPT ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_v end_POSTSUPERSCRIPT ( italic_y , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y ) ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y (assuming the viscous flux is linear with respect to the spatial gradient of the state), and {{}}\left\{\!\!\left\{\cdot\right\}\!\!\right\}{ { ⋅ } } and delimited-⟦⟧\llbracket\cdot\rrbracket⟦ ⋅ ⟧ denote the average and jump operators, respectively. The interface condition (3.3), which corresponds to the conservation law (3.1), is known as the jump or generalized Rankine-Hugoniot conditions (Maj12, ). The interface condition (3.4), derived in (Ker20, ), is associated with the constitutive law (3.2) and constrains the continuity of the state variable at the interface.

Let the solution spaces Y𝑌Yitalic_Y and ΣΣ\Sigmaroman_Σ be the broken Sobolev spaces

Y𝑌\displaystyle Yitalic_Y =\displaystyle== {y[L2(Ω)]m|κ𝒯,y|κ[H1(κ)]m},\displaystyle\left\{y\in\left[L^{2}\left(\Omega\right)\right]^{m\hphantom{% \times d_{x}}}\bigl{|}\forall\kappa\in\mathcal{T},\>\>\hphantom{\nabla_{x}% \cdot}\left.y\right|_{\kappa}\in\left[H^{1}\left(\kappa\right)\right]^{m}% \right\},{ italic_y ∈ [ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Ω ) ] start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT | ∀ italic_κ ∈ caligraphic_T , italic_y | start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT ∈ [ italic_H start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( italic_κ ) ] start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT } , (3.5)
ΣΣ\displaystyle\Sigmaroman_Σ =\displaystyle== {σ[L2(Ω)]m×dx|κ𝒯,xσ|κ[L2(Ω)]m}.\displaystyle\left\{\sigma\in\left[L^{2}\left(\Omega\right)\right]^{m\times d_% {x}}\bigl{|}\forall\kappa\in\mathcal{T},\left.\nabla_{x}\cdot\sigma\right|_{% \kappa}\in\left[L^{2}\left(\Omega\right)\right]^{m}\right\}.{ italic_σ ∈ [ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Ω ) ] start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | ∀ italic_κ ∈ caligraphic_T , ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⋅ italic_σ | start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT ∈ [ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Ω ) ] start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT } . (3.6)

The MDG-ICE weak formulation is then obtained by integrating Equations (3.1)-(3.4) against separate test functions: find (y,σ)Y×Σ𝑦𝜎𝑌Σ\left(y,\sigma\right)\in Y\times\Sigma( italic_y , italic_σ ) ∈ italic_Y × roman_Σ such that

0=0absent\displaystyle 0=0 = κ𝒯((y,σ),vy)κsubscript𝜅𝒯subscript𝑦𝜎subscript𝑣𝑦𝜅\displaystyle\;\;\;\;\sum_{\kappa\in\mathcal{T}}\left(\nabla\cdot\mathcal{F}% \left(y,\sigma\right),v_{y}\right)_{\kappa}∑ start_POSTSUBSCRIPT italic_κ ∈ caligraphic_T end_POSTSUBSCRIPT ( ∇ ⋅ caligraphic_F ( italic_y , italic_σ ) , italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT
+κ𝒯(σG(y)xy,vσ)κsubscript𝜅𝒯subscript𝜎𝐺𝑦subscript𝑥𝑦subscript𝑣𝜎𝜅\displaystyle+\sum_{\kappa\in\mathcal{T}}\left(\sigma-G\left(y\right)\nabla_{x% }y,v_{\sigma}\right)_{\kappa}+ ∑ start_POSTSUBSCRIPT italic_κ ∈ caligraphic_T end_POSTSUBSCRIPT ( italic_σ - italic_G ( italic_y ) ∇ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT
ϵ(n(y,σ),wy)ϵ\displaystyle-\sum_{\epsilon\in\mathcal{E}}\left(\left\llbracket n\cdot% \mathcal{F}\left(y,\sigma\right)\right\rrbracket,w_{y}\right)_{\epsilon}- ∑ start_POSTSUBSCRIPT italic_ϵ ∈ caligraphic_E end_POSTSUBSCRIPT ( ⟦ italic_n ⋅ caligraphic_F ( italic_y , italic_σ ) ⟧ , italic_w start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT
ϵ({{G(y)}}ynx,wσ)ϵ(vy,vσ,wy,wσ)Vy×Vσ×Wy×Wσ.\displaystyle-\sum_{\epsilon\in\mathcal{E}}\left(\left\{\!\!\left\{G\left(y% \right)\right\}\!\!\right\}\left\llbracket y\otimes n_{x}\right\rrbracket,w_{% \sigma}\right)_{\epsilon}\qquad\forall\left(v_{y},v_{\sigma},w_{y},w_{\sigma}% \right)\in V_{y}\times V_{\sigma}\times W_{y}\times W_{\sigma}.- ∑ start_POSTSUBSCRIPT italic_ϵ ∈ caligraphic_E end_POSTSUBSCRIPT ( { { italic_G ( italic_y ) } } ⟦ italic_y ⊗ italic_n start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⟧ , italic_w start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT ∀ ( italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) ∈ italic_V start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT . (3.7)

where the test spaces are Vy=[L2(Ω)]msubscript𝑉𝑦superscriptdelimited-[]superscript𝐿2Ω𝑚V_{y}=\left[L^{2}\left(\Omega\right)\right]^{m}italic_V start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = [ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Ω ) ] start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT and Vσ=[L2(Ω)]m×dsubscript𝑉𝜎superscriptdelimited-[]superscript𝐿2Ω𝑚𝑑V_{\sigma}=\left[L^{2}\left(\Omega\right)\right]^{m\times d}italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = [ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Ω ) ] start_POSTSUPERSCRIPT italic_m × italic_d end_POSTSUPERSCRIPT, with Wysubscript𝑊𝑦W_{y}italic_W start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT and Wσsubscript𝑊𝜎W_{\sigma}italic_W start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT defined as the corresponding single-valued trace spaces. Note that numerical flux functions are not employed in the current MDG-ICE formulation, although with slight modifications, it is straightforward to include them (Cor18, ).

To treat the grid as a variable, the weak formulation (3.7) is transformed from physical to reference space. Let u:Ω^Ω:𝑢^ΩΩu:\hat{\Omega}\rightarrow\Omegaitalic_u : over^ start_ARG roman_Ω end_ARG → roman_Ω be a continuous, invertible mapping from the reference domain, Ω^^Ω\hat{\Omega}over^ start_ARG roman_Ω end_ARG, to the physical domain, ΩΩ\Omegaroman_Ω. Ω^^Ω\hat{\Omega}over^ start_ARG roman_Ω end_ARG is assumed to be partitioned by 𝒯^^𝒯\hat{\mathcal{T}}over^ start_ARG caligraphic_T end_ARG, such that Ω^¯=κ^𝒯^κ^¯¯^Ωsubscript^𝜅^𝒯¯^𝜅\overline{\hat{\Omega}}=\cup_{\hat{\kappa}\in\hat{\mathcal{T}}}\overline{\hat{% \kappa}}over¯ start_ARG over^ start_ARG roman_Ω end_ARG end_ARG = ∪ start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG ∈ over^ start_ARG caligraphic_T end_ARG end_POSTSUBSCRIPT over¯ start_ARG over^ start_ARG italic_κ end_ARG end_ARG. Furthermore, let ^^\hat{\mathcal{E}}over^ start_ARG caligraphic_E end_ARG denote the set of interfaces, ϵ^^italic-ϵ\hat{\epsilon}over^ start_ARG italic_ϵ end_ARG, such that ϵ^^ϵ^=κ^𝒯^κ^subscript^italic-ϵ^^italic-ϵsubscript^𝜅^𝒯^𝜅\cup_{\hat{\epsilon}\in\hat{\mathcal{E}}}\hat{\epsilon}=\cup_{\hat{\kappa}\in% \hat{\mathcal{T}}}\partial\hat{\kappa}∪ start_POSTSUBSCRIPT over^ start_ARG italic_ϵ end_ARG ∈ over^ start_ARG caligraphic_E end_ARG end_POSTSUBSCRIPT over^ start_ARG italic_ϵ end_ARG = ∪ start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG ∈ over^ start_ARG caligraphic_T end_ARG end_POSTSUBSCRIPT ∂ over^ start_ARG italic_κ end_ARG, and let U=[H1(Ω^)]d𝑈superscriptdelimited-[]superscript𝐻1^Ω𝑑U=\left[H^{1}\left(\hat{\Omega}\right)\right]^{d}italic_U = [ italic_H start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ( over^ start_ARG roman_Ω end_ARG ) ] start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, the dsuperscript𝑑\mathbb{R}^{d}blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT-valued Sobolev space over Ω^^Ω\hat{\Omega}over^ start_ARG roman_Ω end_ARG. The solution and test spaces are also now assumed to be defined over reference space. We then define a provisional state operator, e~:Y×Σ×U(Vy×Vσ×Wy×Wσ)*:~𝑒𝑌Σ𝑈superscriptsubscript𝑉𝑦subscript𝑉𝜎subscript𝑊𝑦subscript𝑊𝜎\tilde{e}:Y\times\Sigma\times U\rightarrow\left(V_{y}\times V_{\sigma}\times W% _{y}\times W_{\sigma}\right)^{*}over~ start_ARG italic_e end_ARG : italic_Y × roman_Σ × italic_U → ( italic_V start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT for (y,σ,u)Y×Σ×U𝑦𝜎𝑢𝑌Σ𝑈\left(y,\sigma,u\right)\in Y\times\Sigma\times U( italic_y , italic_σ , italic_u ) ∈ italic_Y × roman_Σ × italic_U, as

e~(y,σ,u)=(vy,vσ,wy,wσ)~𝑒𝑦𝜎𝑢subscript𝑣𝑦subscript𝑣𝜎subscript𝑤𝑦subscript𝑤𝜎maps-toabsent\displaystyle\tilde{e}\left(y,\sigma,u\right)=\left(v_{y},v_{\sigma},w_{y},w_{% \sigma}\right)\mapstoover~ start_ARG italic_e end_ARG ( italic_y , italic_σ , italic_u ) = ( italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) ↦ κ^𝒯^((cof(u))(y,σ),vy)κ^subscript^𝜅^𝒯subscriptcof𝑢𝑦𝜎subscript𝑣𝑦^𝜅\displaystyle\;\;\;\;\sum_{\hat{\kappa}\in\hat{\mathcal{T}}}\left(\left(% \operatorname{cof}\left(\nabla u\right)\nabla\right)\cdot\mathcal{F}\left(y,% \sigma\right),v_{y}\right)_{\hat{\kappa}}∑ start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG ∈ over^ start_ARG caligraphic_T end_ARG end_POSTSUBSCRIPT ( ( roman_cof ( ∇ italic_u ) ∇ ) ⋅ caligraphic_F ( italic_y , italic_σ ) , italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG end_POSTSUBSCRIPT
+κ^𝒯^(det(u)σG(y)(cof(u))xy,vσ)κ^subscript^𝜅^𝒯subscriptdet𝑢𝜎𝐺𝑦subscriptcof𝑢𝑥𝑦subscript𝑣𝜎^𝜅\displaystyle+\sum_{\hat{\kappa}\in\hat{\mathcal{T}}}\left(\operatorname{det}% \left(\nabla u\right)\sigma-G\left(y\right)\left(\operatorname{cof}\left(% \nabla u\right)\nabla\right)_{x}y,v_{\sigma}\right)_{\hat{\kappa}}+ ∑ start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG ∈ over^ start_ARG caligraphic_T end_ARG end_POSTSUBSCRIPT ( roman_det ( ∇ italic_u ) italic_σ - italic_G ( italic_y ) ( roman_cof ( ∇ italic_u ) ∇ ) start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_y , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG end_POSTSUBSCRIPT
ϵ^^(s(u)(y,σ),wy)ϵ^\displaystyle-\sum_{\hat{\epsilon}\in\hat{\mathcal{E}}}\left(\left\llbracket s% \left(\nabla u\right)\cdot\mathcal{F}\left(y,\sigma\right)\right\rrbracket,w_{% y}\right)_{\hat{\epsilon}}- ∑ start_POSTSUBSCRIPT over^ start_ARG italic_ϵ end_ARG ∈ over^ start_ARG caligraphic_E end_ARG end_POSTSUBSCRIPT ( ⟦ italic_s ( ∇ italic_u ) ⋅ caligraphic_F ( italic_y , italic_σ ) ⟧ , italic_w start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT over^ start_ARG italic_ϵ end_ARG end_POSTSUBSCRIPT
ϵ^^({{G(y)}}ys(u)x,wσ)ϵ^.\displaystyle-\sum_{\hat{\epsilon}\in\hat{\mathcal{E}}}\left(\left\{\!\!\left% \{G\left(y\right)\right\}\!\!\right\}\left\llbracket y\otimes s\left(\nabla u% \right)_{x}\right\rrbracket,w_{\sigma}\right)_{\hat{\epsilon}}.- ∑ start_POSTSUBSCRIPT over^ start_ARG italic_ϵ end_ARG ∈ over^ start_ARG caligraphic_E end_ARG end_POSTSUBSCRIPT ( { { italic_G ( italic_y ) } } ⟦ italic_y ⊗ italic_s ( ∇ italic_u ) start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⟧ , italic_w start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT over^ start_ARG italic_ϵ end_ARG end_POSTSUBSCRIPT . (3.8)

The state operator, e:Y×Σ×U(Vy×Vσ×Wy×Wσ)*:𝑒𝑌Σ𝑈superscriptsubscript𝑉𝑦subscript𝑉𝜎subscript𝑊𝑦subscript𝑊𝜎e:Y\times\Sigma\times U\rightarrow\left(V_{y}\times V_{\sigma}\times W_{y}% \times W_{\sigma}\right)^{*}italic_e : italic_Y × roman_Σ × italic_U → ( italic_V start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, is defined as

e(y,σ,u)=e~(y,σ,b(u)),𝑒𝑦𝜎𝑢~𝑒𝑦𝜎𝑏𝑢e\left(y,\sigma,u\right)=\tilde{e}\left(y,\sigma,b\left(u\right)\right),italic_e ( italic_y , italic_σ , italic_u ) = over~ start_ARG italic_e end_ARG ( italic_y , italic_σ , italic_b ( italic_u ) ) , (3.9)

which imposes geometric boundary conditions via the geometric projection operator, b(u)𝑏𝑢b(u)italic_b ( italic_u ).

The state equation in reference space is e(y,σ,u)=0,𝑒𝑦𝜎𝑢0e\left(y,\sigma,u\right)=0,italic_e ( italic_y , italic_σ , italic_u ) = 0 , such that the corresponding weak formulation in reference space is as follows: find (y,σ,u)Y×Σ×U𝑦𝜎𝑢𝑌Σ𝑈\left(y,\sigma,u\right)\in Y\times\Sigma\times U( italic_y , italic_σ , italic_u ) ∈ italic_Y × roman_Σ × italic_U such that

e(y,σ,u),(vy,vσ,wy,wσ)=0(vy,vσ,wy,wσ)Vy×Vσ×Wy×Wσ.formulae-sequence𝑒𝑦𝜎𝑢subscript𝑣𝑦subscript𝑣𝜎subscript𝑤𝑦subscript𝑤𝜎0for-allsubscript𝑣𝑦subscript𝑣𝜎subscript𝑤𝑦subscript𝑤𝜎subscript𝑉𝑦subscript𝑉𝜎subscript𝑊𝑦subscript𝑊𝜎\left\langle e\left(y,\sigma,u\right),\left(v_{y},v_{\sigma},w_{y},w_{\sigma}% \right)\right\rangle=0\qquad\forall\left(v_{y},v_{\sigma},w_{y},w_{\sigma}% \right)\in V_{y}\times V_{\sigma}\times W_{y}\times W_{\sigma}.⟨ italic_e ( italic_y , italic_σ , italic_u ) , ( italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) ⟩ = 0 ∀ ( italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) ∈ italic_V start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT . (3.10)

The solution is therefore given by (y,σ,b(u))Y×Σ×U𝑦𝜎𝑏𝑢𝑌Σ𝑈\left(y,\sigma,b\left(u\right)\right)\in Y\times\Sigma\times U( italic_y , italic_σ , italic_b ( italic_u ) ) ∈ italic_Y × roman_Σ × italic_U.

3.2 Discretization

To discretize the weak formulation (3.10), we choose discrete subspaces YhYsubscript𝑌𝑌Y_{h}\subset Yitalic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ⊂ italic_Y, ΣhΣsubscriptΣΣ\Sigma_{h}\subset\Sigmaroman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ⊂ roman_Σ, UhUsubscript𝑈𝑈U_{h}\subset Uitalic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ⊂ italic_U, Vy,hVysubscript𝑉𝑦subscript𝑉𝑦V_{y,h}\subset V_{y}italic_V start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT ⊂ italic_V start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT, Vσ,hVσsubscript𝑉𝜎subscript𝑉𝜎V_{\sigma,h}\subset V_{\sigma}italic_V start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT ⊂ italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, Wy,hWysubscript𝑊𝑦subscript𝑊𝑦W_{y,h}\subset W_{y}italic_W start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT ⊂ italic_W start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT, and Wσ,hWσsubscript𝑊𝜎subscript𝑊𝜎W_{\sigma,h}\subset W_{\sigma}italic_W start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT ⊂ italic_W start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT and define a discrete state operator

eh:Yh×Σh×Uhdim(Vy,h×Vσ,h×Wy,h×Wσ,h).:subscript𝑒subscript𝑌subscriptΣsubscript𝑈superscriptdimensionsubscript𝑉𝑦subscript𝑉𝜎subscript𝑊𝑦subscript𝑊𝜎e_{h}:Y_{h}\times\Sigma_{h}\times U_{h}\rightarrow\mathbb{R}^{\dim\left(V_{y,h% }\times V_{\sigma,h}\times W_{y,h}\times W_{\sigma,h}\right)}.italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT : italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT → blackboard_R start_POSTSUPERSCRIPT roman_dim ( italic_V start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT × italic_V start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT . (3.11)

For a simplicial grid,

Yhsubscript𝑌\displaystyle Y_{h}italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT =\displaystyle== {yY|κ^𝒯^,y|κ^[𝒫p]m},conditional-set𝑦𝑌formulae-sequencefor-all^𝜅^𝒯evaluated-at𝑦^𝜅superscriptdelimited-[]subscript𝒫𝑝𝑚\displaystyle\left\{y\in Y\,\middle|\,\forall\hat{\kappa}\in\hat{\mathcal{T}},% \left.y\right|_{\hat{\kappa}}\in\left[\mathcal{P}_{p}\right]^{m\hphantom{% \times d_{x}}}\right\},{ italic_y ∈ italic_Y | ∀ over^ start_ARG italic_κ end_ARG ∈ over^ start_ARG caligraphic_T end_ARG , italic_y | start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG end_POSTSUBSCRIPT ∈ [ caligraphic_P start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT } , (3.12)
ΣhsubscriptΣ\displaystyle\Sigma_{h}roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT =\displaystyle== {σΣ|κ^𝒯^,σ|κ^[𝒫p]m×dx},conditional-set𝜎Σformulae-sequencefor-all^𝜅^𝒯evaluated-at𝜎^𝜅superscriptdelimited-[]subscript𝒫𝑝𝑚subscript𝑑𝑥\displaystyle\left\{\sigma\in\Sigma\,\middle|\,\forall\hat{\kappa}\in\hat{% \mathcal{T}},\left.\sigma\right|_{\hat{\kappa}}\in\left[\mathcal{P}_{p}\right]% ^{m\times d_{x}}\right\},{ italic_σ ∈ roman_Σ | ∀ over^ start_ARG italic_κ end_ARG ∈ over^ start_ARG caligraphic_T end_ARG , italic_σ | start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG end_POSTSUBSCRIPT ∈ [ caligraphic_P start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT italic_m × italic_d start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUPERSCRIPT } , (3.13)

where 𝒫psubscript𝒫𝑝\mathcal{P}_{p}caligraphic_P start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the space of polynomials spanned by the monomials 𝒙βsuperscript𝒙𝛽\boldsymbol{x}^{\beta}bold_italic_x start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT with multi-index β0d𝛽superscriptsubscript0𝑑\beta\in\mathbb{N}_{0}^{d}italic_β ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT satisfying i=1dβipsuperscriptsubscript𝑖1𝑑subscript𝛽𝑖𝑝\sum_{i=1}^{d}\beta_{i}\leq p∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ italic_p. We set Vy,h=Yhsubscript𝑉𝑦subscript𝑌V_{y,h}=Y_{h}italic_V start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT = italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and Vσ,h=Σhsubscript𝑉𝜎subscriptΣV_{\sigma,h}=\Sigma_{h}italic_V start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT = roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT. Wy,hsubscript𝑊𝑦W_{y,h}italic_W start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT and Wσ,hsubscript𝑊𝜎W_{\sigma,h}italic_W start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT are selected to be the corresponding single-valued polynomial trace spaces. The discrete subspace Uhsubscript𝑈U_{h}italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT of geometric mappings is defined as

Uh={uU|κ^𝒯^,u|κ^[𝒫p]d}.subscript𝑈conditional-set𝑢𝑈formulae-sequencefor-all^𝜅^𝒯evaluated-at𝑢^𝜅superscriptdelimited-[]subscript𝒫𝑝𝑑U_{h}=\left\{u\in U\,\middle|\,\forall\hat{\kappa}\in\hat{\mathcal{T}},\left.u% \right|_{\hat{\kappa}}\in\left[\mathcal{P}_{p}\right]^{d}\right\}.italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = { italic_u ∈ italic_U | ∀ over^ start_ARG italic_κ end_ARG ∈ over^ start_ARG caligraphic_T end_ARG , italic_u | start_POSTSUBSCRIPT over^ start_ARG italic_κ end_ARG end_POSTSUBSCRIPT ∈ [ caligraphic_P start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT } . (3.14)

In this work, the polynomial degrees of Yhsubscript𝑌Y_{h}italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and ΣhsubscriptΣ\Sigma_{h}roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT are chosen to be equal; the polynomial degree of Uhsubscript𝑈U_{h}italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, however, may be different. The cases in which the polynomial degree of Uhsubscript𝑈U_{h}italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT is greater than, the same as, and less than the polynomial degrees of Yhsubscript𝑌Y_{h}italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and ΣhsubscriptΣ\Sigma_{h}roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT are referred to as superparametric, isoparametric, and subparametric, respectively.

4 Nonlinear solver

The nonlinear solver developed for this work is based on the Levenberg-Marquardt method that was previously employed (Ker20, ), in which the weak formulation is solved iteratively (in nondimensional form) using unconstrained optimization to minimize the objective function

J(y,σ,u)=12eh(y,σ,u)2𝐽𝑦𝜎𝑢12superscriptnormsubscript𝑒𝑦𝜎𝑢2J\left(y,\sigma,u\right)=\frac{1}{2}\left\|e_{h}\left(y,\sigma,u\right)\right% \|^{2}italic_J ( italic_y , italic_σ , italic_u ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (4.1)

by seeking a stationary point

J(y,σ,u)=eh(y,σ,u)*eh(y,σ,u)=0,𝐽𝑦𝜎𝑢superscriptsubscript𝑒superscript𝑦𝜎𝑢subscript𝑒𝑦𝜎𝑢0\nabla J\left(y,\sigma,u\right)=e_{h}^{\prime}\left(y,\sigma,u\right)^{*}e_{h}% \left(y,\sigma,u\right)=0,∇ italic_J ( italic_y , italic_σ , italic_u ) = italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) = 0 , (4.2)

where eh(y,σ,u)*:dim(Vy,h×Vσ,h×Wy,h×Wσ,h)Yh×Σh×Uh:superscriptsubscript𝑒superscript𝑦𝜎𝑢superscriptdimensionsubscript𝑉𝑦subscript𝑉𝜎subscript𝑊𝑦subscript𝑊𝜎subscript𝑌subscriptΣsubscript𝑈e_{h}^{\prime}\left(y,\sigma,u\right)^{*}:\mathbb{R}^{\dim\left(V_{y,h}\times V% _{\sigma,h}\times W_{y,h}\times W_{\sigma,h}\right)}\rightarrow Y_{h}\times% \Sigma_{h}\times U_{h}italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT : blackboard_R start_POSTSUPERSCRIPT roman_dim ( italic_V start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT × italic_V start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_y , italic_h end_POSTSUBSCRIPT × italic_W start_POSTSUBSCRIPT italic_σ , italic_h end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT → italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT is the adjoint operator. Given an initialization (y,σ,u)0subscript𝑦𝜎𝑢0\left(y,\sigma,u\right)_{0}( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the solution is repeatedly updated

(y,σ,u)i+1=(y,σ,u)i+Δ(y,σ,u)ii=0,1,2,formulae-sequencesubscript𝑦𝜎𝑢𝑖1subscript𝑦𝜎𝑢𝑖Δsubscript𝑦𝜎𝑢𝑖𝑖012\left(y,\sigma,u\right)_{i+1}=\left(y,\sigma,u\right)_{i}+\Delta\left(y,\sigma% ,u\right)_{i}\qquad i=0,1,2,\ldots( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT = ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + roman_Δ ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_i = 0 , 1 , 2 , … (4.3)

until (4.2) is satisfied to a given tolerance. The Levenberg-Marquardt method (Lev44, ; Mar63, ) is then employed to solve (4.2), which yields the increment

Δ(y,σ,u)=(eh(y,σ,u)*eh(y,σ,u)+Ih,λ(y,σ,u))1(eh(y,σ,u)*eh(y,σ,u)),Δ𝑦𝜎𝑢superscriptsuperscriptsubscript𝑒superscript𝑦𝜎𝑢superscriptsubscript𝑒𝑦𝜎𝑢subscript𝐼𝜆𝑦𝜎𝑢1superscriptsubscript𝑒superscript𝑦𝜎𝑢subscript𝑒𝑦𝜎𝑢\Delta\left(y,\sigma,u\right)=-\left(e_{h}^{\prime}\left(y,\sigma,u\right)^{*}% e_{h}^{\prime}\left(y,\sigma,u\right)+I_{h,\lambda}\left(y,\sigma,u\right)% \right)^{-1}\left(e_{h}^{\prime}\left(y,\sigma,u\right)^{*}e_{h}\left(y,\sigma% ,u\right)\right),roman_Δ ( italic_y , italic_σ , italic_u ) = - ( italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_y , italic_σ , italic_u ) + italic_I start_POSTSUBSCRIPT italic_h , italic_λ end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ) , (4.4)

where where Ih,λ(y,σ,u):(Yh×Σh×Uh)×(Yh×Σh×Uh):subscript𝐼𝜆𝑦𝜎𝑢subscript𝑌subscriptΣsubscript𝑈subscript𝑌subscriptΣsubscript𝑈I_{h,\lambda}\left(y,\sigma,u\right):\left(Y_{h}\times\Sigma_{h}\times U_{h}% \right)\times\left(Y_{h}\times\Sigma_{h}\times U_{h}\right)\rightarrow\mathbb{R}italic_I start_POSTSUBSCRIPT italic_h , italic_λ end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) : ( italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) × ( italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT × italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) → blackboard_R is a symmetric, positive-definite bilinear form that defines the choice of regularization

Ih,λ(y,σ,u)((δy,δσ,δu),(vy,vσ,vu))=(δy,λyvy)+(δσ,λσvσ)+(δu,λuvu).subscript𝐼𝜆𝑦𝜎𝑢𝛿𝑦𝛿𝜎𝛿𝑢subscript𝑣𝑦subscript𝑣𝜎subscript𝑣𝑢𝛿𝑦subscript𝜆𝑦subscript𝑣𝑦𝛿𝜎subscript𝜆𝜎subscript𝑣𝜎𝛿𝑢subscript𝜆𝑢subscript𝑣𝑢I_{h,\lambda}\left(y,\sigma,u\right)\left(\left(\delta y,\delta\sigma,\delta u% \right),\left(v_{y},v_{\sigma},v_{u}\right)\right)=\left(\delta y,\lambda_{y}v% _{y}\right)+\left(\delta\sigma,\lambda_{\sigma}v_{\sigma}\right)+\left(\delta u% ,\lambda_{u}v_{u}\right).italic_I start_POSTSUBSCRIPT italic_h , italic_λ end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ( ( italic_δ italic_y , italic_δ italic_σ , italic_δ italic_u ) , ( italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ) ) = ( italic_δ italic_y , italic_λ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) + ( italic_δ italic_σ , italic_λ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) + ( italic_δ italic_u , italic_λ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ) . (4.5)

with λysubscript𝜆𝑦\lambda_{y}italic_λ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT, λσsubscript𝜆𝜎\lambda_{\sigma}italic_λ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, and λusubscript𝜆𝑢\lambda_{u}italic_λ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT denoting nonnegative regularization coefficients for each solution variable. In practice, λysubscript𝜆𝑦\lambda_{y}italic_λ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT and λσsubscript𝜆𝜎\lambda_{\sigma}italic_λ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT are set to zero. In addition to the identity regularization (4.5), which ensures positive definiteness and prevents large-scale grid changes, we incorporate a Laplacian-type grid regularization,

Ih,λΔ(y,σ,u)((δy,δσ,δu),(vy,vσ,vu))=((bh(u)δu),λΔu(bh(u)vu)).superscriptsubscript𝐼𝜆Δ𝑦𝜎𝑢𝛿𝑦𝛿𝜎𝛿𝑢subscript𝑣𝑦subscript𝑣𝜎subscript𝑣𝑢superscriptsubscript𝑏𝑢𝛿𝑢subscript𝜆Δ𝑢superscriptsubscript𝑏𝑢subscript𝑣𝑢I_{h,\lambda}^{\Delta}\left(y,\sigma,u\right)\left(\left(\delta y,\delta\sigma% ,\delta u\right),\left(v_{y},v_{\sigma},v_{u}\right)\right)=-\left(\nabla\left% (b_{h}^{\prime}\left(u\right)\delta u\right),\lambda_{\Delta u}\nabla\left(b_{% h}^{\prime}\left(u\right)v_{u}\right)\right).italic_I start_POSTSUBSCRIPT italic_h , italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ end_POSTSUPERSCRIPT ( italic_y , italic_σ , italic_u ) ( ( italic_δ italic_y , italic_δ italic_σ , italic_δ italic_u ) , ( italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ) ) = - ( ∇ ( italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u ) italic_δ italic_u ) , italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT ∇ ( italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u ) italic_v start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ) ) . (4.6)

where λΔu0subscript𝜆Δ𝑢0\lambda_{\Delta u}\geq 0italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT ≥ 0 is the corresponding regularization coefficient and vuUhsubscript𝑣𝑢subscript𝑈v_{u}\in U_{h}italic_v start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ∈ italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, which introduces a compressibility effect into the grid motion. In previous work (Ker20_LS, ), following the approach of Zahr et al. (Zah20, ), the regularization factor λΔusubscript𝜆Δ𝑢\lambda_{\Delta u}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT was modified to incorporate a factor proportional to the inverse of the element volume, which locally stiffens small elements in an isotropic manner. Note that the regularizations (4.5) and (4.6) are not incorporated into the objective function (4.1).

The solver as hitherto described was employed in (Ker20, ). In the remainder of this section, we introduce enhancements to the solver that significantly improve its robustness.

4.1 Regularization terms

4.1.1 Anisotropic Laplacian-type grid regularization

In this work, we modify the regularization (4.6) to directly account for element anisotropy as

Ih,λΔ(y,σ,u)((δy,δσ,δu),(vy,vσ,vu))=((α/2bh(u)δu),λΔu(α/2bh(u)vu)),superscriptsubscript𝐼𝜆Δ𝑦𝜎𝑢𝛿𝑦𝛿𝜎𝛿𝑢subscript𝑣𝑦subscript𝑣𝜎subscript𝑣𝑢superscript𝛼2superscriptsubscript𝑏𝑢𝛿𝑢subscript𝜆Δ𝑢superscript𝛼2superscriptsubscript𝑏𝑢subscript𝑣𝑢I_{h,\lambda}^{\Delta}\left(y,\sigma,u\right)\left(\left(\delta y,\delta\sigma% ,\delta u\right),\left(v_{y},v_{\sigma},v_{u}\right)\right)=-\left(\nabla\left% (\mathcal{H}^{-\alpha/2}b_{h}^{\prime}(u)\delta u\right),\lambda_{\Delta u}% \nabla\left(\mathcal{H}^{-\alpha/2}b_{h}^{\prime}(u)v_{u}\right)\right),italic_I start_POSTSUBSCRIPT italic_h , italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ end_POSTSUPERSCRIPT ( italic_y , italic_σ , italic_u ) ( ( italic_δ italic_y , italic_δ italic_σ , italic_δ italic_u ) , ( italic_v start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ) ) = - ( ∇ ( caligraphic_H start_POSTSUPERSCRIPT - italic_α / 2 end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u ) italic_δ italic_u ) , italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT ∇ ( caligraphic_H start_POSTSUPERSCRIPT - italic_α / 2 end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_u ) italic_v start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ) ) , (4.7)

where α0𝛼0\alpha\geq 0italic_α ≥ 0 is a parameter and \mathcal{H}caligraphic_H, in the two-dimensional case, is an element-local, 2×2222\times 22 × 2, symmetric-positive-definite transformation matrix given by (Fid07, ; Chi21, )

=1/2=VΣVT=[e^1e^2][h100h2][e^1Te^2T].superscript12𝑉Σsuperscript𝑉𝑇delimited-[]missing-subexpressionmissing-subexpressionsubscript^𝑒1subscript^𝑒2matrixsubscript100subscript2delimited-[]missing-subexpressionsuperscriptsubscript^𝑒1𝑇missing-subexpressionmissing-subexpressionsuperscriptsubscript^𝑒2𝑇missing-subexpression\mathcal{H}=\mathcal{M}^{-1/2}=V\Sigma V^{T}=\left[\begin{array}[]{cc}\rule[-4% .30554pt]{0.5pt}{10.76385pt}&\rule[-4.30554pt]{0.5pt}{10.76385pt}\\ \widehat{e}_{1}&\widehat{e}_{2}\\ \rule[-4.30554pt]{0.5pt}{10.76385pt}&\rule[-4.30554pt]{0.5pt}{10.76385pt}\end{% array}\right]\begin{bmatrix}h_{1}&0\\[4.30554pt] 0&h_{2}\end{bmatrix}\left[\begin{array}[]{ccc}\rule[2.15277pt]{10.76385pt}{0.5% pt}&\widehat{e}_{1}^{T}&\rule[2.15277pt]{10.76385pt}{0.5pt}\\ \rule[2.15277pt]{10.76385pt}{0.5pt}&\widehat{e}_{2}^{T}&\rule[2.15277pt]{10.76% 385pt}{0.5pt}\end{array}\right].caligraphic_H = caligraphic_M start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT = italic_V roman_Σ italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = [ start_ARRAY start_ROW start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW end_ARRAY ] [ start_ARG start_ROW start_CELL italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ] [ start_ARRAY start_ROW start_CELL end_CELL start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_CELL start_CELL end_CELL end_ROW end_ARRAY ] . (4.8)

\mathcal{M}caligraphic_M is the metric implied by the mesh (Fid07, ), the columns of V𝑉Vitalic_V are the (orthonormal) left singular vectors, e^1subscript^𝑒1\widehat{e}_{1}over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and e^2subscript^𝑒2\widehat{e}_{2}over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, of the geometric Jacobian, J𝐽Jitalic_J, and ΣΣ\Sigmaroman_Σ is a diagonal matrix with the singular values, h1subscript1h_{1}italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and h2subscript2h_{2}italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, of J𝐽Jitalic_J along the main diagonal. \mathcal{H}caligraphic_H projects the unit circle to an ellipse with principal directions e^1subscript^𝑒1\widehat{e}_{1}over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and e^2subscript^𝑒2\widehat{e}_{2}over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and principal stretching magnitudes h1subscript1h_{1}italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and h2subscript2h_{2}italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (Chi21, ). Without loss of generality, we assume that the singular values are ordered such that h1h2subscript1subscript2h_{1}\leq h_{2}italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. αsuperscript𝛼\mathcal{H}^{-\alpha}caligraphic_H start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT can be expanded as

α=VΣαVT=[e^1e^2][h1α00h2α][ e^1Te^2T].superscript𝛼𝑉superscriptΣ𝛼superscript𝑉𝑇delimited-[]missing-subexpressionmissing-subexpressionsubscript^𝑒1subscript^𝑒2matrixsuperscriptsubscript1𝛼00superscriptsubscript2𝛼delimited-[] superscriptsubscript^𝑒1𝑇missing-subexpressionmissing-subexpressionsuperscriptsubscript^𝑒2𝑇missing-subexpression\mathcal{H}^{-\alpha}=V\Sigma^{-\alpha}V^{T}=\left[\begin{array}[]{cc}\rule[-4% .30554pt]{0.5pt}{10.76385pt}&\rule[-4.30554pt]{0.5pt}{10.76385pt}\\ \widehat{e}_{1}&\widehat{e}_{2}\\ \rule[-4.30554pt]{0.5pt}{10.76385pt}&\rule[-4.30554pt]{0.5pt}{10.76385pt}\end{% array}\right]\begin{bmatrix}h_{1}^{-\alpha}&0\\[4.30554pt] 0&h_{2}^{-\alpha}\end{bmatrix}\left[\begin{array}[]{ccc}\rule[2.15277pt]{10.76% 385pt}{0.5pt}&\widehat{e}_{1}^{T}&\rule[2.15277pt]{10.76385pt}{0.5pt}\\ \rule[2.15277pt]{10.76385pt}{0.5pt}&\widehat{e}_{2}^{T}&\rule[2.15277pt]{10.76% 385pt}{0.5pt}\end{array}\right].caligraphic_H start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT = italic_V roman_Σ start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = [ start_ARRAY start_ROW start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW end_ARRAY ] [ start_ARG start_ROW start_CELL italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] [ start_ARRAY start_ROW start_CELL end_CELL start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL over^ start_ARG italic_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_CELL start_CELL end_CELL end_ROW end_ARRAY ] . (4.9)

As such, the modified regularization (4.7) limits grid motion in directions with small element length scales while allowing for greater changes in directions with larger length scales, aiding in the formation of non-degenerate high-aspect-ratio elements to resolve thin viscous structures. For simplicity, αsuperscript𝛼\mathcal{H}^{-\alpha}caligraphic_H start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT is evaluated at the centroid of the given element. The metric described here is a powerful tool that can be used not only for anisotropic grid regularization but also for remeshing via a metric-based grid generator (Fid07, ) in order to maintain high-quality grids, which will be the subject of future work. Moreover, we employ additional regularization operators that further inhibit element degeneration, specifically 𝒫1subscript𝒫1\mathcal{P}_{1}caligraphic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT grid regularization and penalty grid regularization, which are described below. Similar forms of these two regularization operators were employed in the mesh untangling algorithm by Toulorge et al. (Tou13, ).

4.1.2 𝒫1subscript𝒫1\mathcal{P}_{1}caligraphic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT grid regularization

To penalize excessive element curvature, the following regularization term is included in the objective function (4.1):

J1(u)=12h1βλ1uΠ1u2,subscript𝐽1𝑢12superscriptsubscript1𝛽subscript𝜆1superscriptnorm𝑢subscriptΠ1𝑢2J_{1}(u)=\frac{1}{2}h_{1}^{\beta}\lambda_{1}\left|\left|u-\Pi_{1}u\right|% \right|^{2},italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_u ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | | italic_u - roman_Π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_u | | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (4.10)

where λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the corresponding regularization coefficient, h1subscript1h_{1}italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is obtained from the mesh-implied metric, Π1subscriptΠ1\Pi_{1}roman_Π start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT denotes the projection to a (multi-)linear subspace of Uhsubscript𝑈U_{h}italic_U start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, and β>0𝛽0\beta>0italic_β > 0 is a parameter. The h1βsuperscriptsubscript1𝛽h_{1}^{\beta}italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT scaling allows for increased curvature of elements with small length scales, which is important for resolving curved shocks.

4.1.3 Penalty grid regularization

Finally, a penalty regularization term is included in the objective function (4.1):

Jb(u)=12λbf(u)2,subscript𝐽𝑏𝑢12subscript𝜆𝑏superscriptnorm𝑓𝑢2J_{b}\left(u\right)=\frac{1}{2}\lambda_{b}\left|\left|f\left(u\right)\right|% \right|^{2},italic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( italic_u ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | | italic_f ( italic_u ) | | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (4.11)

where λbsubscript𝜆𝑏\lambda_{b}italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the corresponding regularization coefficient and f𝑓fitalic_f is defined as

f(u)=max{0,𝒥bdet(u)},𝑓𝑢0subscript𝒥𝑏det𝑢f\left(u\right)=\max\left\{0,\mathcal{J}_{b}-\operatorname{det}\left(\nabla u% \right)\right\},italic_f ( italic_u ) = roman_max { 0 , caligraphic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - roman_det ( ∇ italic_u ) } , (4.12)

with 𝒥bsubscript𝒥𝑏\mathcal{J}_{b}caligraphic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT represents a desired lower bound on the Jacobian determinant (e.g., 1010superscript101010^{-10}10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT). The regularization (4.11) penalizes invalid elements (i.e., elements for which the determinant of the geometric Jacobian is nonpositive). Note that the non-differentiability of f(u)𝑓𝑢f\left(u\right)italic_f ( italic_u ) at det(u)=𝒥bdet𝑢subscript𝒥𝑏\operatorname{det}\left(\nabla u\right)=\mathcal{J}_{b}roman_det ( ∇ italic_u ) = caligraphic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is not found to cause any noticeable issues; furthermore, it can easily be made differentiable by squaring the RHS of (4.12). In previous work (Cor18, ), barrier grid regularization was found to frequently lead to solver stagnation, although such regularization may nevertheless be worthy of future investigation.

4.1.4 Global scaling

In this work, the regularization coefficients λusubscript𝜆𝑢\lambda_{u}italic_λ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT, λΔusubscript𝜆Δ𝑢\lambda_{\Delta u}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT, λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and λbsubscript𝜆𝑏\lambda_{b}italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT are scaled by the residual magnitude, eh(y,σ,u)normsubscript𝑒𝑦𝜎𝑢\left\|e_{h}\left(y,\sigma,u\right)\right\|∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥. Early in the simulation, when the residual is large and spurious transients may appear in the solution, the higher regularization prevents excessive grid changes that would otherwise occur. As the solution converges and these transients disappear, the lower regularization encourages greater adjustments to the grid to facilitate resolution of high-gradient features. To prevent rank deficiency of the linear system at low residuals, we select a small, positive number as a lower bound on λusubscript𝜆𝑢\lambda_{u}italic_λ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT (e.g., 107superscript10710^{-7}10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT).

4.2 Adaptive, elementwise regularization

Even with manual tuning, the aforementioned global, static scaling of the regularization terms is not sufficiently robust for preventing cell degeneration, especially as highly anisotropic, curved elements form to resolve multidimensional, high-gradient features. Local grid operations, whether topology-preserving (e.g., vertex smoothing (Ala14, )) or topology-changing (e.g., edge refinement, collapse (Loh08, ), and swapping (Ala14, )), may be beneficial if used sparingly; however, such operations often fail to recover valid cells and will inhibit iterative convergence of the solver since they are not intrinsic to the formulation (i.e., they are applied as a “postprocessing” step at the end of the iteration). This obstruction of convergence is exacerbated in the viscous setting, wherein high-aspect-ratio elements are gradually compressed to better resolve thin viscous structures. Mesh deformation techniques have also been developed to improve mesh quality and untangle invalid elements, typically in the context of a posteriori high-order mesh generation. Examples include elasticity-based methods (Xie13, ; Mox16, ; Mar21, ), in which a set of elasticity equations is solved to obtain a mesh-displacement field, and mesh optimization algorithms (Tou13, ; Gar15, ), which seek to minimize element distortion and constrain invalid Jacobians via regularization operators similar to (4.10) and (4.11). Although more reliable for recovering valid elements and potentially useful for occasionally “resetting” the grid in case convergence begins to stall, these mesh-deformation techniques can be expensive and again disrupt solver convergence, especially because they need to be engaged frequently in the presence of curved, high-gradient features. Incorporating regularization operators directly into the objective function (4.1) eliminates, or at least mitigates, obstruction of iterative convergence, but, as previously mentioned, global, static scaling of the regularization terms does not provide sufficient robustness.

In light of the above, we propose a strategy for maintaining grid validity that can reliably recover valid cells while avoiding disruption of solver convergence. In particular, when a cell becomes invalid, the iteration is restarted (i.e., the solution is “rolled back”) and the regularization coefficients are automatically adjusted in an elementwise fashion until the solver can proceed with a valid grid. In this work, λΔusubscript𝜆Δ𝑢\lambda_{\Delta u}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT, λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and λbsubscript𝜆𝑏\lambda_{b}italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT are dynamically scaled; however, if the aspect ratio is high (e.g., h2/h1>10subscript2subscript110h_{2}/h_{1}>10italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > 10), then λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is left unchanged in order to facilitate resolution of thin, curved features using curved cells. The regularization coefficients are then successively decreased to their base values, unless the element once again becomes invalid. The rollback of the solution is crucial for ensuring a valid grid at every iteration. Although our formulation does not necessarily diverge in the presence of (slightly) negative Jacobian determinants, maintaining grid validity at every iteration is important for the following reasons: (a) allowing the solver to proceed with invalid cells can result in pollution of solution accuracy and lower-quality grids that may drive the solution towards a less-optimal local minimum; (b) in our experience, once the solver proceeds with negative Jacobian determinants, it can be extremely difficult to eliminate them; and (c) if negative Jacobian determinants become too large in magnitude, the solver can easily diverge or at least produce clearly erroneous results. Furthermore, although the intermittent scaling of the penalty grid regularization can cause abrupt increases in the objective function (specifically, Jbsubscript𝐽𝑏J_{b}italic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT), this is not a major concern since we are ultimately interested in minimizing eh(y,σ,u)normsubscript𝑒𝑦𝜎𝑢\left\|e_{h}\left(y,\sigma,u\right)\right\|∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥. This adaptive, elementwise regularization strategy is found to be critical for preventing cell degeneration in the vicinity of viscous shocks and boundary layers at hypersonic flow conditions. With such high-gradient features, zero to three rollbacks per iteration are typically required to preserve grid validity. Note that if the solution is rolled back, only the regularization terms for the corresponding cells need to be recomputed. The implementation details of this strategy will be provided in Section 4.4.

4.3 Increment limiting

The consideration of very strong shocks without artificial dissipation often leads to undershoots and overshoots in pressure and other quantities during intermediate iterations. To mitigate these instabilities, which can otherwise cause solver divergence, we employ an increment-limiting strategy used in standard implicit-DG solvers (Cez15, ), in which the increment is scaled by a factor no greater than unity such that the maximum change in pressure and density at the integration points is less than a user-specified fraction, denoted 𝖿𝖿\mathsf{f}sansserif_f (e.g., 𝖿=10𝖿10\mathsf{f}=10sansserif_f = 10%). The implementation details will be provided in Section 4.4.

We have also experimented with projecting the state in troubled elements to a first-order approximation (𝒫0subscript𝒫0\mathcal{P}_{0}caligraphic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT), but this can hinder iterative convergence by creating a cycle in which first-order projections are followed by eventual reappearance of undershoots/overshoots, which then causes additional first-order projections, and so on. Applying sophisticated limiters that, for example, nominally preserve order of accuracy to troubled elements during intermediate iterations may also be useful, but is not considered in this work.

4.4 Line search

The locally adaptive regularization and increment limiting discussed in Sections 4.2 and 4.3, respectively are incorporated into a simple line search method, as presented in Algorithm 1. It should be noted that decreasing eh(y,σ,u)normsubscript𝑒𝑦𝜎𝑢\left\|e_{h}\left(y,\sigma,u\right)\right\|∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥ and maintaining grid validity (i.e., minimizing Jbsubscript𝐽𝑏J_{b}italic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT) are often at odds with each other since penalizing small Jacobian determinants can hinder resolution of high-gradient features, especially those which are curved. Penalty regularization is thus activated only when necessary and in a local, gradual fashion in order to avoid interfering with reduction of eh(y,σ,u)normsubscript𝑒𝑦𝜎𝑢\left\|e_{h}\left(y,\sigma,u\right)\right\|∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥ and potential convergence towards a less optimal local minimum. The local regularization coefficients λΔuκsuperscriptsubscript𝜆Δ𝑢𝜅\lambda_{\Delta u}^{\kappa}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT, λ1κsuperscriptsubscript𝜆1𝜅\lambda_{1}^{\kappa}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT, and λbκsuperscriptsubscript𝜆𝑏𝜅\lambda_{b}^{\kappa}italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT are then successively decreased every few steps towards their base values and 𝒥bκsuperscriptsubscript𝒥𝑏𝜅\mathcal{J}_{b}^{\kappa}caligraphic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT is reset, provided that the given element remains valid. Recommended base values for the regularization coefficients and other parameters are given in Table 1. We also find that even though scaling the regularization coefficients by eh(y,σ,u)normsubscript𝑒𝑦𝜎𝑢\left\|e_{h}\left(y,\sigma,u\right)\right\|∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥ is often beneficial, sometimes, at low values of eh(y,σ,u)normsubscript𝑒𝑦𝜎𝑢\left\|e_{h}\left(y,\sigma,u\right)\right\|∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥, the grid has already repositioned itself enough to resolve the anisotropic structures in the flow, at which point anything more than marginal adjustments to the grid can slow down residual convergence. To minimize grid motion in this situation, λusubscript𝜆𝑢\lambda_{u}italic_λ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT and λΔusubscript𝜆Δ𝑢\lambda_{\Delta u}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT are automatically scaled in a global fashion by factors of 100 for a prescribed number of iterations (e.g., 30) when repeatedly low values of the increment factor (e.g., ω<0.02)\omega<0.02)italic_ω < 0.02 ) are accompanied by at least one rollback, which serves as a reliable indicator of said scenario.

Note that modifying the hard-coded values in Algorithm 1 (e.g., setting Nsearchsubscript𝑁searchN_{\mathrm{search}}italic_N start_POSTSUBSCRIPT roman_search end_POSTSUBSCRIPT to five and scaling λΔuκsuperscriptsubscript𝜆Δ𝑢𝜅\lambda_{\Delta u}^{\kappa}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT by a factor of ten) can potentially yield better performance; nevertheless, the chosen values work sufficiently well for the problems considered in this study, and we remark that the primary concern of this work is robustness in terms of maintaining grid validity. Future work will explore strategies to accelerate iterative convergence, as well as well-established methods to automatically adjust the global values of λusubscript𝜆𝑢\lambda_{u}italic_λ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT and/or λΔusubscript𝜆Δ𝑢\lambda_{\Delta u}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT at each step (Tra12, ; Zah20, ). Furthermore, similar results can likely be obtained with locally adaptive barrier functions (as opposed to penalty functions), although as previously mentioned, a naive strategy may lead to frequent solver stagnation. In addition, though not employed here, it may be useful to activate penalty regularization over a localized region of elements, instead of just one element, since doing so for only one element can lead to movement of grid points that then causes degeneration in neighboring cells.

Finally, we remark that the full system can also be explicitly formulated as a constrained-optimization problem with inequality constraints det(u)>0det𝑢0\operatorname{det}\left(\nabla u\right)>0roman_det ( ∇ italic_u ) > 0, where Lagrange multipliers corresponding to those constraints are introduced as solution variables. Although future work may explore a Lagrange-multiplier approach, the proposed unconstrained-optimization formulation is appealing due to its simplicity and smaller system size.

Algorithm 1 Line search algorithm for iteration i+1𝑖1i+1italic_i + 1. ω1𝜔1\omega\leq 1italic_ω ≤ 1 is a positive factor that scales the increment. The κ𝜅\kappaitalic_κ superscripts denote element-local quantities. 𝒥minκsuperscriptsubscript𝒥𝜅\mathcal{J}_{\min}^{\kappa}caligraphic_J start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT is the minimum Jacobian determinant for cell κ𝜅\kappaitalic_κ. The compressible Navier-Stokes system is assumed (Section 2.2).
(y,σ,u)isubscript𝑦𝜎𝑢𝑖\left(y,\sigma,u\right)_{i}( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
converged \leftarrow False
while not converged do
     Compute Δ(y,σ,u)iΔsubscript𝑦𝜎𝑢𝑖\Delta\left(y,\sigma,u\right)_{i}roman_Δ ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
     ω1,ωa0,ωb1formulae-sequence𝜔1formulae-sequencesubscript𝜔𝑎0subscript𝜔𝑏1\omega\leftarrow 1,\>\omega_{a}\leftarrow 0,\>\omega_{b}\leftarrow 1italic_ω ← 1 , italic_ω start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ← 0 , italic_ω start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ← 1
     (y,σ,u)*(y,σ,u)isuperscript𝑦𝜎𝑢subscript𝑦𝜎𝑢𝑖\left(y,\sigma,u\right)^{*}\leftarrow\left(y,\sigma,u\right)_{i}( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ← ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
     Nsearch5subscript𝑁search5N_{\mathrm{search}}\leftarrow 5italic_N start_POSTSUBSCRIPT roman_search end_POSTSUBSCRIPT ← 5
     for k0𝑘0k\leftarrow 0italic_k ← 0 to Nsearchsubscript𝑁searchN_{\mathrm{search}}italic_N start_POSTSUBSCRIPT roman_search end_POSTSUBSCRIPT do
         if k>0𝑘0k>0italic_k > 0 then
              ω(ωa+ωb)/2𝜔subscript𝜔𝑎subscript𝜔𝑏2\omega\leftarrow(\omega_{a}+\omega_{b})/2italic_ω ← ( italic_ω start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) / 2
         end if
         (y,σ,u)i+1(y,σ,u)i+ωΔ(y,σ,u)isubscript𝑦𝜎𝑢𝑖1subscript𝑦𝜎𝑢𝑖𝜔Δsubscript𝑦𝜎𝑢𝑖\left(y,\sigma,u\right)_{i+1}\leftarrow\left(y,\sigma,u\right)_{i}+\omega% \Delta\left(y,\sigma,u\right)_{i}( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT ← ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ω roman_Δ ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
         𝖥ρmaxΩ{|ρi+1ρi|/ρi}subscript𝖥𝜌subscriptΩsubscript𝜌𝑖1subscript𝜌𝑖subscript𝜌𝑖\mathsf{F}_{\rho}\leftarrow\max_{\Omega}\left\{\left|\rho_{i+1}-\rho_{i}\right% |/\rho_{i}\right\}sansserif_F start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT ← roman_max start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT { | italic_ρ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT - italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | / italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT }
         𝖥PmaxΩ{|P(yi+1)P(yi)|/P(yi)}subscript𝖥𝑃subscriptΩ𝑃subscript𝑦𝑖1𝑃subscript𝑦𝑖𝑃subscript𝑦𝑖\mathsf{F}_{P}\leftarrow\max_{\Omega}\left\{\left|P\left(y_{i+1}\right)-P\left% (y_{i}\right)\right|/P\left(y_{i}\right)\right\}sansserif_F start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ← roman_max start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT { | italic_P ( italic_y start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT ) - italic_P ( italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) | / italic_P ( italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) }
         if eh(y,σ,u)i+1/eh(y,σ,u)*<1normsubscript𝑒subscript𝑦𝜎𝑢𝑖1normsubscript𝑒superscript𝑦𝜎𝑢1\left\|e_{h}\left(y,\sigma,u\right)_{i+1}\right\|/\left\|e_{h}\left(y,\sigma,u% \right)^{*}\right\|<1∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT ∥ / ∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ∥ < 1 and 𝖥ρ𝖿subscript𝖥𝜌𝖿\mathsf{F}_{\rho}\leq\mathsf{f}sansserif_F start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT ≤ sansserif_f and 𝖥P𝖿subscript𝖥𝑃𝖿\mathsf{F}_{P}\leq\mathsf{f}sansserif_F start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ≤ sansserif_f then
              if ω=1𝜔1\omega=1italic_ω = 1 then
                  break
              end if
              (y,σ,u)*(y,σ,u)i+1superscript𝑦𝜎𝑢subscript𝑦𝜎𝑢𝑖1\left(y,\sigma,u\right)^{*}\leftarrow\left(y,\sigma,u\right)_{i+1}( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ← ( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT
              ωaωsubscript𝜔𝑎𝜔\omega_{a}\leftarrow\omegaitalic_ω start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ← italic_ω
         else
              ωbωsubscript𝜔𝑏𝜔\omega_{b}\leftarrow\omegaitalic_ω start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ← italic_ω
         end if
         if ωa=0subscript𝜔𝑎0\omega_{a}=0italic_ω start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = 0 and Nsearch=5subscript𝑁search5N_{\mathrm{search}}=5italic_N start_POSTSUBSCRIPT roman_search end_POSTSUBSCRIPT = 5 then
              NsearchNsearch+10subscript𝑁searchsubscript𝑁search10N_{\mathrm{search}}\leftarrow N_{\mathrm{search}}+10italic_N start_POSTSUBSCRIPT roman_search end_POSTSUBSCRIPT ← italic_N start_POSTSUBSCRIPT roman_search end_POSTSUBSCRIPT + 10 \triangleright Could not reduce eh(y,σ,u)normsubscript𝑒𝑦𝜎𝑢\left\|e_{h}\left(y,\sigma,u\right)\right\|∥ italic_e start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_y , italic_σ , italic_u ) ∥; increase Nsearchsubscript𝑁searchN_{\mathrm{search}}italic_N start_POSTSUBSCRIPT roman_search end_POSTSUBSCRIPT
         end if
     end for
     (y,σ,u)i+1(y,σ,u)*subscript𝑦𝜎𝑢𝑖1superscript𝑦𝜎𝑢\left(y,\sigma,u\right)_{i+1}\leftarrow\left(y,\sigma,u\right)^{*}( italic_y , italic_σ , italic_u ) start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT ← ( italic_y , italic_σ , italic_u ) start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT
     converged \leftarrow True
     for each element κ𝜅\kappaitalic_κ do
         if 𝒥minκ,i+1<𝒥bκsuperscriptsubscript𝒥𝜅𝑖1superscriptsubscript𝒥𝑏𝜅\mathcal{J}_{\min}^{\kappa,i+1}<\mathcal{J}_{b}^{\kappa}caligraphic_J start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ , italic_i + 1 end_POSTSUPERSCRIPT < caligraphic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT then
              converged \leftarrow False
              λΔuκ10λΔuκsuperscriptsubscript𝜆Δ𝑢𝜅10superscriptsubscript𝜆Δ𝑢𝜅\lambda_{\Delta u}^{\kappa}\leftarrow 10\lambda_{\Delta u}^{\kappa}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT ← 10 italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT, λbκ10λbκsuperscriptsubscript𝜆𝑏𝜅10superscriptsubscript𝜆𝑏𝜅\lambda_{b}^{\kappa}\leftarrow 10\lambda_{b}^{\kappa}italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT ← 10 italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT, λ1κ100λ1κsuperscriptsubscript𝜆1𝜅100superscriptsubscript𝜆1𝜅\lambda_{1}^{\kappa}\leftarrow 100\lambda_{1}^{\kappa}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT ← 100 italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT
              𝒥bκ10𝒥minκ,isuperscriptsubscript𝒥𝑏𝜅10superscriptsubscript𝒥𝜅𝑖\mathcal{J}_{b}^{\kappa}\leftarrow 10\mathcal{J}_{\min}^{\kappa,i}caligraphic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ end_POSTSUPERSCRIPT ← 10 caligraphic_J start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_κ , italic_i end_POSTSUPERSCRIPT
         end if
     end for
end while
Table 1: Recommended base values for regularization coefficients and other parameters
λusubscript𝜆𝑢\lambda_{u}italic_λ start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT λΔusubscript𝜆Δ𝑢\lambda_{\Delta u}italic_λ start_POSTSUBSCRIPT roman_Δ italic_u end_POSTSUBSCRIPT λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT λbsubscript𝜆𝑏\lambda_{b}italic_λ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT 𝒥bsubscript𝒥𝑏\mathcal{J}_{b}caligraphic_J start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT α𝛼\alphaitalic_α β𝛽\betaitalic_β
107superscript10710^{-7}10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 102superscript10210^{2}10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 1010superscript101010^{-10}10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT 11-1- 1 3333

5 Results

We apply MDG-ICE, equipped with the enhanced nonlinear solver described in Section 4, to compute viscous flows with high-gradient features. Unsteady solutions are obtained using a space-time discretization. All grids are generated using Gmsh (Geu09, ). The simulations in this study are performed using the JENRE® Multiphysics Framework employed in previous work (Cor18, ; Ker20, ; Ker20_LS, ) with the modifications and extensions described here. The three-dimensional solutions are computed using hybrid shared- and distributed-memory parallelism, the latter of which was not implemented in previous work. In the present study, we employ a sparse, direct LDLT solver provided by the MUMPS (MUltifrontal Massively Parallel Solver) package (Ame01, ; Ame19, ) via the PETSc (Portable, Extensible Toolkit for Scientific Computation) software library (Bal97, ; Bal23, ). Future work will improve the efficiency of solving the linear system by leveraging techniques such as domain decomposition and multigrid algorithms.

5.1 Space-time Burgers viscous shock formation

This section presents results for space-time Burgers viscous shock formation, previously computed using MDG-ICE in (Ker20, ). The initial condition is given by

y(x,t=0)=12πtssin(2πx)+y,𝑦𝑥𝑡012𝜋subscript𝑡𝑠2𝜋𝑥subscript𝑦y(x,t=0)=\frac{1}{2\pi t_{s}}\sin\left(2\pi x\right)+y_{\infty},italic_y ( italic_x , italic_t = 0 ) = divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG roman_sin ( 2 italic_π italic_x ) + italic_y start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT , (5.1)

where ts=0.5subscript𝑡𝑠0.5t_{s}=0.5italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0.5 is the time of shock formation and y=0.2subscript𝑦0.2y_{\infty}=0.2italic_y start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.2 is the freestream velocity. The computational domain is Ω=[0,1]×[0,1]Ω0101\Omega=\left[0,1\right]\times\left[0,1\right]roman_Ω = [ 0 , 1 ] × [ 0 , 1 ]. Inflow and outflow boundary conditions are applied at the left and right boundaries, respectively. The solution is initialized by extruding the initial condition (5.1) in the temporal direction. We employ continuation in μ𝜇\muitalic_μ, i.e., a series of problems is solved in which μ𝜇\muitalic_μ is successively decreased, with the final solution of a given problem used as the initial condition of the subsequent one. In previous work (Ker20, ), subparametric MDG-ICE(𝒫5/𝒫1subscript𝒫5subscript𝒫1\mathcal{P}_{5}/\mathcal{P}_{1}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) solutions were computed, also with μ𝜇\muitalic_μ continuation. Specifically, viscosities of μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=5×104𝜇5superscript104\mu=5\times 10^{-4}italic_μ = 5 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT were considered. The initial grid consisted of 200 linear triangular elements obtained by splitting each element of a uniform 10×10101010\times 1010 × 10 quadrilateral mesh into two triangles, resulting in a regular topology. Here, we compute subparametric MDG-ICE(𝒫5/𝒫2subscript𝒫5subscript𝒫2\mathcal{P}_{5}/\mathcal{P}_{2}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) solutions for a more challenging series of viscosities: μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. We employ two different starting grids: the first is the same as in (Ker20, ) (i.e., a regular grid with 200 triangular elements), while the second is an irregular grid with 242 triangular elements. A key difference with (Ker20, ) is that the elements are of quadratic geometric order, significantly increasing the difficulty of achieving solution convergence while maintaining a valid grid.

5.1.1 Initial mesh topology: Regular

Figure 5.1a presents the space-time initial condition and final solutions for μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, with the corresponding grids superimposed. Figure 5.2 shows the corresponding one-dimensional profiles at t=0.025𝑡0.025t=0.025italic_t = 0.025 and t=0.975𝑡0.975t=0.975italic_t = 0.975. MDG-ICE automatically adjusts the grid geometry to resolve the viscous shock as a sharp yet smooth profile without relying on artificial stabilization or mesh-topology modification. As the viscosity is decreased, the viscous shock becomes thinner and the aspect ratios of the nearby elements are increased via anisotropic space-time r𝑟ritalic_r-adaptivity. The solutions are free from spurious oscillations. The anisotropic, locally adaptive penalty technique described in Section (4) enable excellent resolution of the thin viscous shock while maintaining grid validity.

Refer to caption
(a) Initial condition.
Refer to caption
(b) μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT solution.
Refer to caption
(c) μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT solution.
Refer to caption
(d) μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT solution.
Figure 5.1: The space-time Burgers initial condition and final solutions for μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, with the corresponding grids superimposed. The initial grid consists of 200 triangular elements of quadratic geometric order with a regular topology. The initial condition is given in Equation (5.1).
Refer to caption
(a) μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT solution.
Refer to caption
(b) μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT solution.
Refer to caption
(c) μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT solution.
Figure 5.2: One-dimensional profiles at t=0.025𝑡0.025t=0.025italic_t = 0.025 and t=0.975𝑡0.975t=0.975italic_t = 0.975 for μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT for space-time Burgers viscous shock formation. The initial grid consists of 200 triangular elements of quadratic geometric order with a regular topology. The initial condition is given in Equation (5.1).

The nonlinear convergence history is given in Figure 5.3. The initial residual magnitudes for the μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT simulations are relatively small since the solver is restarted from the corresponding higher-viscosity solution.

Refer to caption
Figure 5.3: Nonlinear convergence history for space-time Burgers viscous shock formation. The initial grid consists of 200 triangular elements of quadratic geometric order with a regular topology. The initial condition is given in Equation (5.1).

5.1.2 Initial mesh topology: Irregular

The initial 242-element irregular grid, along with the initial condition, is displayed in Figure 5.4a. The final solutions and grids for μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT are given in Figures 5.4b5.4c, and 5.4d, respectively. Figure 5.5 presents the corresponding one-dimensional profiles at t=0.025𝑡0.025t=0.025italic_t = 0.025 and t=0.975𝑡0.975t=0.975italic_t = 0.975. Similar to the regular-topology case, MDG-ICE automatically modifies the location, size, and orientation of elements to resolve the viscous shock via anisotropic space-time r𝑟ritalic_r-adaptivity. Additional grid adjustments are introduced accordingly as the viscosity is decreased and the shock becomes more difficult to resolve. The final solutions are well-resolved and free from spurious oscillations. The nonlinear convergence history is displayed in Figure 5.6.

Refer to caption
(a) Initial condition.
Refer to caption
(b) μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT solution.
Refer to caption
(c) μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT solution.
Refer to caption
(d) μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT solution.
Figure 5.4: The space-time Burgers initial condition and final solutions for μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, with the corresponding grids superimposed. The initial grid consists of 242 triangular elements of quadratic geometric order with an irregular topology. The initial condition is given in Equation (5.1).
Refer to caption
(a) μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT solution.
Refer to caption
(b) μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT solution.
Refer to caption
(c) μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT solution.
Figure 5.5: One-dimensional profiles at t=0.025𝑡0.025t=0.025italic_t = 0.025 and t=0.975𝑡0.975t=0.975italic_t = 0.975 for μ=103𝜇superscript103\mu=10^{-3}italic_μ = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, μ=104𝜇superscript104\mu=10^{-4}italic_μ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and μ=105𝜇superscript105\mu=10^{-5}italic_μ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT for space-time Burgers viscous shock formation. The initial grid consists of 242 triangular elements of quadratic geometric order with an irregular topology. The initial condition is given in Equation (5.1).
Refer to caption
Figure 5.6: Nonlinear convergence history for space-time Burgers viscous shock formation. The initial grid consists of 242 triangular elements of quadratic geometric order with an irregular topology. The initial condition is given in Equation (5.1).

5.2 Mach 17.6 flow over two-dimensional cylinder

Next, we compute steady hypersonic viscous flow over a circular half-cylinder in two spatial dimensions. The freestream Mach number and Reynolds number (based on the cylinder radius) are 17.6 and 376,930376930376,930376 , 930, respectively. This problem is a common benchmark case for evaluating the ability of numerical techniques to predict hypersonic flows. With conventional finite volume techniques (Nom04, ; Gno04, ), significant asymmetries in the surface heat-flux profile were observed on both regular and irregular simplicial grids. Note that fairly symmetric heating results have been obtained on simplicial grids with DG schemes equipped with smooth artificial viscosity (Bar10, ; Chi18, ). In this work, we aim to use the proposed MDG-ICE formulation to achieve symmetric solutions without artificial dissipation. We employ two different starting grids: the first is a regular grid with 392 elements, while the second is an irregular grid with 526 elements. Freestream conditions are imposed at the inflow boundary, defined as the ellipse (x/6)2+(y/3)2=1superscript𝑥62superscript𝑦321\left(\nicefrac{{x}}{{6}}\right)^{2}+\left(\nicefrac{{y}}{{3}}\right)^{2}=1( / start_ARG italic_x end_ARG start_ARG 6 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( / start_ARG italic_y end_ARG start_ARG 3 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1. Extrapolation is employed at the outflow boundary, and the cylinder boundary, a half-circle of unit radius, is an isothermal no-slip wall with temperature Twall=2.5Tsubscript𝑇wall2.5subscript𝑇T_{\mathrm{wall}}=2.5T_{\infty}italic_T start_POSTSUBSCRIPT roman_wall end_POSTSUBSCRIPT = 2.5 italic_T start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT.

5.2.1 Initial mesh topology: Regular

We employ continuation in both the Mach number and the Reynolds number, starting with an isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution corresponding to Ma=5,Re=500formulae-sequenceMa5Re500\mathrm{Ma}=5,\mathrm{Re}=500roman_Ma = 5 , roman_Re = 500. This MDG-ICE solution is initialized with a conventional DG solution stabilized with elementwise-constant artificial viscosity of the form described in (Joh20, ). Keeping Re fixed at 500, the Mach number is consecutively increased by increments of one until the target value of 17.6 is reached. These intermediate solutions are not fully converged to a stationary point. Note that the Ma=17.6,Re=500formulae-sequenceMa17.6Re500\mathrm{Ma}=17.6,\mathrm{Re}=500roman_Ma = 17.6 , roman_Re = 500 solution was computed in previous work (Chi22_ICCFD, ) prior to development of the adaptive, elementwise regularization strategy (Section 4.2) and the increment-limiting procedure (Section 4.3); cell degeneration was treated then using a standard cell-collapse algorithm (Loh08, ), resulting in a total of 388 elements. The 388-element mesh and temperature field for this Ma=17.6,Re=500formulae-sequenceMa17.6Re500\mathrm{Ma}=17.6,\mathrm{Re}=500roman_Ma = 17.6 , roman_Re = 500 solution are presented in Figure 5.7. The elements in the vicinity of the shock are anisotropic yet still valid. The MDG-ICE solution is free from oscillations even in the absence of artificial dissipation.

Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Figure 5.7: The MDG-ICE solution computed using 388 𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT isoparametric triangle elements for two-dimensional Mach 17.6 flow over a cylinder at Re=500Re500\mathrm{Re}=500roman_Re = 500. The initial grid has a regular topology.

The Reynolds number is then consecutively increased by factors of approximately two until the target value of 376,930376930376,930376 , 930 is reached. At this Reynolds number, we increase the polynomial degrees of Yhsubscript𝑌Y_{h}italic_Y start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and ΣhsubscriptΣ\Sigma_{h}roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT from four to five in order to eliminate slight asymmetries in the surface heat-flux profile, as will be shown later in this section. The final mesh and temperature, Mach, and pressure fields are given in Figure 5.8. Considerable jumps in temperature and pressure are observed. MDG-ICE automatically adjusts the grid geometry in order to resolve the viscous shock and boundary layer, which become sharper as the Reynolds number is increased. Highly anisotropic elements are observed at the shock and in the boundary layer, yet grid validity is maintained as a result of the anisotropic Laplacian-type regularization and the adaptive, elementwise regularization strategy. The solution is well-resolved and free from spurious artifacts. Despite the initially regular nature of the mesh, there is undoubtedly very strong misalignment between the grid and the shock and boundary layer. Figure 5.9 displays the nonlinear convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution. The residual starts at a relatively small value since the simulation is restarted from an isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution. Future work will focus on incorporating stabilization mechanisms (at least during early/intermediate iterations) and/or space-time marching in order to accelerate convergence and circumvent the need for continuation in Mach number and Reynolds number, which itself can be considered akin to (global) artificial dissipation. Nevertheless, this demonstrates how MDG-ICE can be naturally employed for parametric studies in which the flow conditions are varied; other grid adaptation strategies may need to incorporate coarsening techniques to maintain efficiently refined grids.

Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Refer to caption
(c) Pressure field
Figure 5.8: The MDG-ICE solution computed using 388 subparametric 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT triangle elements for two-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has a regular topology.
Refer to caption
Figure 5.9: Nonlinear convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution to Mach 17.6 flow over a two-dimensional cylinder. The initial grid has a regular topology.

Figure 5.10 presents the isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) and subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) predictions of the pressure coefficient and Stanton number, defined as

Cp=PP12ρv2subscript𝐶𝑝𝑃subscript𝑃12subscript𝜌superscriptsubscript𝑣2C_{p}=\frac{P-P_{\infty}}{\frac{1}{2}\rho_{\infty}v_{\infty}^{2}}italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = divide start_ARG italic_P - italic_P start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ρ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG

and

Ch=qncpρv(Tt,Twall),subscript𝐶subscript𝑞𝑛subscript𝑐𝑝subscript𝜌subscript𝑣subscript𝑇𝑡subscript𝑇wallC_{h}=\frac{q_{n}}{c_{p}\rho_{\infty}v_{\infty}\left(T_{t,\infty}-T_{\mathrm{% wall}}\right)},italic_C start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = divide start_ARG italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ( italic_T start_POSTSUBSCRIPT italic_t , ∞ end_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT roman_wall end_POSTSUBSCRIPT ) end_ARG ,

respectively, where qnsubscript𝑞𝑛q_{n}italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the normal heat flux and Ttsubscript𝑇𝑡T_{t}italic_T start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the stagnation temperature. The pressure profiles for both MDG-ICE solutions are highly symmetric, but the heat-flux profile for the isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution exhibits a slight nonphysical cusp at the stagnation point. Nevertheless, this cusp is eliminated with p𝑝pitalic_p-refinement. The stagnation-point Stanton number in the MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution is approximately 0.0077. Note that there exists some variation in the stagnation-point Stanton number reported in the literature; for example, 0.0085 in (Gno04, ), 0.0076 in (Kit13_2, ), and 0.0082 in (Bar10, ).

Refer to caption
(a) Pressure coefficient profile.
Refer to caption
(b) Stanton number profile.
Figure 5.10: Surface profiles of pressure coefficient and Stanton number for the MDG-ICE solution computed using 388 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT triangle elements for for two-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has a regular topology.

5.2.2 Initial mesh topology: Irregular

Continuation in the Mach and Reynolds numbers is once again employed, starting with Ma=5,Re=100formulae-sequenceMa5Re100\mathrm{Ma}=5,\mathrm{Re}=100roman_Ma = 5 , roman_Re = 100. The 526-element mesh and temperature field for an intermediate MDG-ICE solution at Ma=5,Re=100formulae-sequenceMa5Re100\mathrm{Ma}=5,\mathrm{Re}=100roman_Ma = 5 , roman_Re = 100 are presented in Figure 5.11. The temperature field is smooth. At this low Reynolds number, only slight grid repositioning is required to resolve the flow.

Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Figure 5.11: The MDG-ICE solution computed using 526 𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT isoparametric triangle elements for for two-dimensional Mach 17.6 flow over a cylinder at Re=100Re100\mathrm{Re}=100roman_Re = 100. The initial grid has an irregular topology.

Global p𝑝pitalic_p-refinement of the state and auxiliary-variable approximations is again employed. The final subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution at Ma=17.6,Re=376,930formulae-sequenceMa17.6Re376930\mathrm{Ma}=17.6,\mathrm{Re}=376,930roman_Ma = 17.6 , roman_Re = 376 , 930 is given in Figure 5.12. The long, thin elements oriented orthogonal to the shock are a direct consequence of two main factors associated with the initially irregular nature of the grid: (a) moreso than in the regular case, incrementing the Mach and Reynolds numbers (in the absence of additional stabilization) causes transient artifacts to appear upstream of the shock, which then induce appreciable grid motion; (b) the compression of the grid at the shock and boundary layer is less likely to pull in outlying elements accordingly than in the regular case. Note that starting with intermediate Reynolds numbers, α𝛼\alphaitalic_α is set to zero in order to alleviate excessive stiffening of the anisotropic, shock-orthogonal elements due to the Laplacian regularization (4.7). Although it is not clear how detrimental this type of element may be in more complex configurations (if at all), the formation of such elements can be mitigated by, as previously discussed, the use of artificial dissipation during intermediate iterations and/or metric-based remeshing. Nevertheless, the thin viscous structures remain sharp and free from spurious oscillations.

Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Refer to caption
(c) Pressure field
Figure 5.12: The MDG-ICE solution computed using 526 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT subparametric triangle elements for two-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has an irregular topology.

Figure 5.9 displays the nonlinear convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution. The initial residual is already low since the simulation is restarted from an isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution.

Refer to caption
Figure 5.13: Nonlinear convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution to Mach 17.6 flow over a two-dimensional cylinder. The initial grid has an irregular topology.

Figure 5.14 presents the surface profiles of pressure coefficient and Stanton number for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution. The stagnation-point Stanton number is approximately 0.0077. Despite the evidently asymmetric grid and extremely strong grid-shock misalignment, the surface profiles are highly symmetric and agree well with those in Figure (5.10).

Refer to caption
(a) Pressure coefficient profile.
Refer to caption
(b) Stanton number profile.
Figure 5.14: Surface profiles of pressure coefficient and Stanton number obtained with the MDG-ICE solution computed using 526 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT subparametric triangle elements for two-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has an irregular topology.

5.3 Mach 17.6 flow over three-dimensional cylinder

This test case is the three-dimensional version of the previous configuration. The consideration of three spatial dimensions allows for the emergence of instabilities and asymmetries that are naturally suppressed in the two-dimensional case. In particular, it is with a three-dimensional domain that nonphysical artifacts in the heat-flux profiles obtained with finite volume schemes are most prominent (Gno04, ; Nom04, ). The two-dimensional domain is extruded one layer of cells in the homogeneous direction by two (nondimensional) units. Slip conditions are applied at the symmetry boundaries. Unlike in the previous subsection, the inflow boundary is defined as the circle x2+y2=6.52superscript𝑥2superscript𝑦2superscript6.52x^{2}+y^{2}=6.5^{2}italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 6.5 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The reason for this modification is that as the Mach number is increased during the continuation strategy, spurious transients appear upstream of the shock and interact with the boundary, causing noticeable instabilities that can be dampened by positioning the inflow boundary far away from the shock. Although also present in the two-dimensional simulations, these instabilities are exacerbated in the three-dimensional setting. Again, the use of artificial dissipation would likely resolve this issue. As in the two-dimensional case, we employ two different starting grids: the first is a regular grid with 3024 elements, while the second is an irregular grid with 3183 elements.

5.3.1 Initial mesh topology: Regular

Figure 5.15 provides a clipped, three-dimensional perspective of the initial grid. An isoparametric DG(𝒫3subscript𝒫3\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution at Ma=14,Re=100formulae-sequenceMa14Re100\mathrm{Ma}=14,\mathrm{Re}=100roman_Ma = 14 , roman_Re = 100 is used as the initial condition for the MDG-ICE continuation with superparametric 𝒫3/𝒫4subscript𝒫3subscript𝒫4\mathcal{P}_{3}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT cells. A positivity-preserving and entropy-based linear-scaling limiter (Lv15, ; Jia18, ; Chi22, ) is employed to help maintain stability in the DG solution. The temperature field and initial grid for the DG solution along the z=2𝑧2z=2italic_z = 2 symmetry plane are presented in Figure 5.16. Note the coarseness of the grid with respect to the expected high-gradient features at the target conditions. After reaching Ma=17.6,Re=376,930formulae-sequenceMa17.6Re376930\mathrm{Ma}=17.6,\mathrm{Re}=376,930roman_Ma = 17.6 , roman_Re = 376 , 930, two global p𝑝pitalic_p-refinements of the state and auxiliary-variable approximations are performed. The final subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution is displayed in Figure 5.16. Just as in the two-dimensional case, the grid is automatically adapted to resolve the viscous shock and boundary layer while maintaining grid validity. Figure 5.18 zooms in on the shock layer along the stagnation line. Extremely high-aspect-ratio elements at the shock and boundary layer are observed. In particular, the cells at the shock are almost visually indistinguishable. Furthermore, in the temperature field, although the smoothness of the boundary layer can be discerned, the viscous shock resembles a truly discontinuous feature.

Refer to caption
Figure 5.15: Clipped, three-dimensional perspective of the initial grid with a regular topology.
Refer to caption
Figure 5.16: Isoparametric DG(𝒫3subscript𝒫3\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution at Ma=14,Re=100formulae-sequenceMa14Re100\mathrm{Ma}=14,\mathrm{Re}=100roman_Ma = 14 , roman_Re = 100 on 3024 tetrahedral cells along the z=2𝑧2z=2italic_z = 2 symmetry plane. This solution is used as the initial condition for the MDG-ICE continuation strategy. The initial grid has a regular topology.
Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Refer to caption
(c) Pressure field
Figure 5.17: The MDG-ICE solution (shown along the z=2𝑧2z=2italic_z = 2 symmetry plane) computed using 3024 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT subparametric tetrahedral elements for three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has a regular topology.
Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Figure 5.18: Zoomed-in view of the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution (shown along the z=2𝑧2z=2italic_z = 2 symmetry plane) to three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has a regular topology.

The convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution is given in Figure 5.19. The residual magnitude starts at a relatively small value since the solution is restarted from an isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) calculation. Figure 5.20 shows the stagnation-line profiles of temperature and pressure. The stagnation point is located at y=1𝑦1y=1italic_y = 1. The stagnation-line quantities are plotted on a per-cell basis, i.e., only points within the same cell are connected. Note that it is very difficult to capture the interior of the extremely thin viscous shock using the employed line sampler, which probes the solution at discrete points. This explains why the shock is presented as a discontinuity in Figure 5.20. These results further confirm that the solution is free from spurious oscillations, despite the substantial gradient across the shock.

Refer to caption
Figure 5.19: Nonlinear convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution to Mach 17.6 flow over a three-dimensional cylinder. The initial grid has a regular topology.
Refer to caption
(a) Temperature.
Refer to caption
(b) Pressure.
Figure 5.20: Stagnation-line profiles of temperature and pressure obtained with the MDG-ICE solution computed using 3024 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT subparametric tetrahedral elements for three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The stagnation point is located at y=1𝑦1y=1italic_y = 1. The stagnation-line quantities are plotted on a per-cell basis, i.e., only points within the same cell are connected. The exact stagnation-point temperature, T=2.5𝑇2.5T=2.5italic_T = 2.5, is marked with the symbol ×\times×. The initial grid has a regular topology.

Figure 5.21 gives the surface pressure and heat flux evaluated at all degrees of freedom corresponding to ΣhsubscriptΣ\Sigma_{h}roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution. The pressure profile is essentially perfectly symmetric. Though very slight asymmetries are still present in the heat-flux profile, it is nevertheless highly symmetric. The stagnation-point Stanton number in the MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution is approximately 0.0077.

Refer to caption
(a) Pressure coefficient profile.
Refer to caption
(b) Stanton number profile.
Figure 5.21: Surface profiles of pressure coefficient and Stanton number for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution to three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. Pressure and heat-flux values at all auxiliary-variable degrees of freedom along the cylinder wall are shown. The initial grid has a regular topology.

5.3.2 Initial mesh topology: Irregular

A clipped, three-dimensional view of the initial grid is presented in Figure 5.22. In order to reduce the computational cost and memory footprint of the linear solver, the initial grid is made finer in the expected shock-layer region than near the inflow boundary. An isoparametric DG(𝒫3subscript𝒫3\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution at Ma=14,Re=200formulae-sequenceMa14Re200\mathrm{Ma}=14,\mathrm{Re}=200roman_Ma = 14 , roman_Re = 200 is used as the initial condition for the MDG-ICE continuation with superparametric 𝒫3/𝒫4subscript𝒫3subscript𝒫4\mathcal{P}_{3}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT cells. Figure 5.23 presents the temperature field and initial grid for the DG solution along the z=2𝑧2z=2italic_z = 2 symmetry plane. Given the irregular nature of the grid, continuation in Reynolds number, especially at higher Reynolds numbers, leads to the appearance of temperature undershoots that are especially difficult to dampen in this problem in the absence of artificial dissipation. To mitigate these instabilities, we employ the alternative least-squares MDG-ICE formulation with optimal test functions (Ker20_LS, ) based on the discontinuous Petrov-Galerkin methodology by Demkowicz and Gopalakrishnan (Dem10, ; Dem11, ; Dem15_20, ). This formulation is found to be significantly more robust (in terms of preventing spurious undershoots/overshoots in flow quantities) than the standard MDG-ICE formulation. The anisotropic, locally adaptive penalty method described in Section 4 is incorporated in an analogous manner. Furthermore, the irregular nature of the grid induces larger grid motion in the shock layer, especially in response to spurious transients as the Reynolds number is increased, which can lead to excessively anisotropic grid interfaces along the cylinder surface. To prevent the appearance of such thin grid interfaces along the cylinder surface, we freeze the cylinder surface grid in the geometric projection operator, b(u)𝑏𝑢b(u)italic_b ( italic_u ). Although it is not yet clear how detrimental the presence of unnecessarily high-aspect-ratio grid interfaces along a domain boundary may be in more complex configurations, the formation of such grid interfaces can be mitigated by the use of artificial dissipation (which can also alleviate the aforementioned instabilities) during intermediate iterations and/or metric-based remeshing.

Refer to caption
Figure 5.22: Clipped, three-dimensional perspective of the initial grid with an irregular topology.
Refer to caption
Figure 5.23: Isoparametric DG(𝒫3subscript𝒫3\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution at Ma=14,Re=200formulae-sequenceMa14Re200\mathrm{Ma}=14,\mathrm{Re}=200roman_Ma = 14 , roman_Re = 200 on 3183 tetrahedral cells along the z=2𝑧2z=2italic_z = 2 symmetry plane. This solution is used as the initial condition for the MDG-ICE continuation strategy. The initial grid has an irregular topology.

After reaching Ma=17.6,Re=376,930formulae-sequenceMa17.6Re376930\mathrm{Ma}=17.6,\mathrm{Re}=376,930roman_Ma = 17.6 , roman_Re = 376 , 930, two global p𝑝pitalic_p-refinements of the state and auxiliary-variable approximations are performed, and we revert to the standard MDG-ICE formulation since the alternative least-squares MDG-ICE formulation with optimal test functions (Ker20_LS, ) is no longer needed at this point. Figure 5.23 displays the final subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution. The grid is ostensibly more irregular than the previous grid and, as in the two-dimensional case, features long, thin elements oriented orthogonal to the shock. Nevertheless, the shock and boundary layer remain well-resolved. Figure 5.25 zooms in on the shock layer along the stagnation line. Again, the cells at the shock are almost visually indistinguishable, and the viscous shock resembles a discontinuous feature. The solution is free from spurious oscillations.

Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Refer to caption
(c) Pressure field
Figure 5.24: The MDG-ICE solution (shown along the z=2𝑧2z=2italic_z = 2 symmetry plane) computed using 3183 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT subparametric tetrahedral elements for three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has an irregular topology.
Refer to caption
(a) Mesh
Refer to caption
(b) Temperature field
Figure 5.25: Zoomed-in view of the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution (shown along the z=2𝑧2z=2italic_z = 2 symmetry plane) to three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The initial grid has an irregular topology.

The convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution is given in Figure 5.26. The residual magnitude starts at a relatively small value since the solution is restarted from an isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) calculation.

The stagnation-line profiles of temperature and pressure are given in Figure 5.27. The stagnation point is located at y=1𝑦1y=1italic_y = 1. The stagnation-line quantities are plotted on a per-cell basis, which means only points within the same cell are connected. Again, it is very difficult to capture the interior of the highly anisotropic viscous shock using the employed line sampler, which probes the solution at discrete points. This explains why the shock resembles a true discontinuity in Figure 5.27. The shock and boundary layer are well-resolved.

Refer to caption
Figure 5.26: Nonlinear convergence history for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution to Mach 17.6 flow over a three-dimensional cylinder. The initial grid has an irregular topology.
Refer to caption
(a) Temperature.
Refer to caption
(b) Pressure.
Figure 5.27: Stagnation-line profiles of temperature and pressure obtained with the MDG-ICE solution computed using 3183 𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT subparametric tetrahedral elements for three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. The stagnation point is located at y=1𝑦1y=1italic_y = 1. The stagnation-line quantities are plotted on a per-cell basis, i.e., only points within the same cell are connected. The exact stagnation-point temperature, T=2.5𝑇2.5T=2.5italic_T = 2.5, is marked with the symbol ×\times×. The initial grid has an irregular topology.

Figure 5.28 gives the surface pressure and heat flux evaluated at all degrees of freedom corresponding to ΣhsubscriptΣ\Sigma_{h}roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT for the subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution. Excellent symmetry is observed in the pressure and heat-flux profiles. Though very slight asymmetries can be discerned in the latter, these asymmetries are considerably smaller than those observed in finite-volume predictions (Gno04, ; Nom04, )). The stagnation-point Stanton number in the MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solution is approximately 0.0077.

Refer to caption
(a) Pressure coefficient profile.
Refer to caption
(b) Stanton number profile.
Figure 5.28: Surface profiles of pressure coefficient and Stanton number for the isoparametric MDG-ICE(𝒫4subscript𝒫4\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) and subparametric MDG-ICE(𝒫5/𝒫4subscript𝒫5subscript𝒫4\mathcal{P}_{5}/\mathcal{P}_{4}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) solutions to three-dimensional Mach 17.6 flow over a cylinder at Re=376,930Re376930\mathrm{Re}=376,930roman_Re = 376 , 930. Pressure and heat-flux values at all auxiliary-variable degrees of freedom along the cylinder wall are shown. The initial grid has an irregular topology.

5.4 Mach 5 flow over three-dimensional sphere

In our final test case, we compute steady Mach 5 laminar flow over a hemisphere in three spatial dimensions. Only one quarter of the hemisphere, which is of unit radius, is considered due to the memory footprint and computational cost of the linear solver. The Reynolds number (based on the sphere radius) is 3.775×1063.775superscript1063.775\times 10^{6}3.775 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. The sphere boundary is an isothermal no-slip wall with temperature Twall=1.308Tsubscript𝑇wall1.308subscript𝑇T_{\mathrm{wall}}=1.308T_{\infty}italic_T start_POSTSUBSCRIPT roman_wall end_POSTSUBSCRIPT = 1.308 italic_T start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. These conditions are similar to those considered by Blottner (Blo90, ), who employed an axisymmetric solver, and in the 2022 High-Fidelity CFD Workshop (Fis21, ; Mur23, ), except the Reynolds number here is four times higher. At the freestream conditions by Blottner (Blo90, ), large-scale asymmetries and mesh imprinting have been observed in three-dimensional finite volume and finite element solutions (Mur23, ; Kir10, ; Cou18, ), even with structured hexahedral grids. Freestream conditions are imposed at the inflow boundary, defined as a quarter hemisphere with radius 2.3 times larger than that of the spherical body. Extrapolation is applied at the outflow boundary.

Continuation in Reynolds number is employed. Figure 5.29 presents the initial 7650-cell tetrahedral grid and an isoparametric DG(𝒫2subscript𝒫2\mathcal{P}_{2}caligraphic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) solution at Ma=5,Re=100formulae-sequenceMa5Re100\mathrm{Ma}=5,\mathrm{Re}=100roman_Ma = 5 , roman_Re = 100, which is used to initialize the MDG-ICE continuation with isoparametric 𝒫3subscript𝒫3\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT elements. The sphere surface and the y=0𝑦0y=0italic_y = 0 and z=0𝑧0z=0italic_z = 0 planes are displayed. The freestream flow is in the x𝑥-x- italic_x-direction. Mesh imprinting along the sphere boundary and a highly diffused shock are observed in the DG solution. Given the cost of the LDLT linear solver, to maximize efficiency, the initial mesh is generated such that the azimuthal resolution is higher near the stagnation line than elsewhere. Furthermore, the wall-normal resolution is decreased near the inflow boundary. We also find that as the Reynolds number is increased, due to higher gradients at the shock than at the boundary layer, the solver tends to move cells from the near-wall region to the vicinity of the shock. To alleviate this loss of boundary-layer resolution, a very thin layer of cells is constructed adjacent to the wall. This issue can likely be resolved with localized artificial dissipation (to reduce gradients at the shock), a robust remeshing strategy, and/or local p𝑝pitalic_p-refinement, all of which will be pursued in the future. Nevertheless, we note that building an initial grid with additional resolution near the boundary surface is significantly simpler than constructing a grid with the required near-wall resolution while simultaneously aligning the grid interfaces with shocks, whose locations are generally unknown a priori. With our planned advancements to the MDG-ICE formulation (including a more efficient linear-solver strategy), we will consider the full hemispherical domain, as well as an initially irregular topology, in future work.

Refer to caption
(a) Mesh
Refer to caption
(b) Pressure field
Figure 5.29: Isoparametric DG(𝒫2subscript𝒫2\mathcal{P}_{2}caligraphic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) solution at Ma=5,Re=100formulae-sequenceMa5Re100\mathrm{Ma}=5,\mathrm{Re}=100roman_Ma = 5 , roman_Re = 100 on 7650 tetrahedral cells. This solution is used as the initial condition for the MDG-ICE continuation strategy. The sphere surface and the y=0𝑦0y=0italic_y = 0 and z=0𝑧0z=0italic_z = 0 planes are displayed. The freestream flow is in the x𝑥-x- italic_x-direction.

Successive increases in the Reynolds number leads to the appearance of temperature undershoots that are especially difficult to dampen in this problem in the absence of artificial dissipation. As in Section 5.3.2, to reduce these instabilities, we employ the alternative least-squares MDG-ICE formulation with optimal test functions (Ker20_LS, ). After reaching the final Reynolds number, we perform two global p𝑝pitalic_p-refinements of the state and auxiliary-variable approximations while retaining the alternative least-squares MDG-ICE formulation. Figure 5.30 displays the final subparametric MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution at Ma=5,Re=3.775×106formulae-sequenceMa5Re3.775superscript106\mathrm{Ma}=5,\mathrm{Re}=3.775\times 10^{6}roman_Ma = 5 , roman_Re = 3.775 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. To reduce the number of unnecessary elements in the freestream region, we project the inflow boundary to a quarter hemisphere of radius 2.15. Cell collapses (Loh08, ) are then performed to remove invalid cells, resulting in 7161 elements. The final grid is shown in Figure 5.30a. The surface mesh is noticeably coarse. The pressure field and surface heat-flux profile, presented in Figures 5.30b and 5.30c, respectively, are free from mesh imprinting and spurious oscillations. Figure 5.30d provides a zoomed-in view of the temperature field along the z=0𝑧0z=0italic_z = 0 plane. The temperature gradient in the boundary layer can be observed. The viscous shock resembles a true discontinuity, and the cells resolving the shock cannot be individually discerned. Nevertheless, grid validity is maintained. Note the closer proximity of the shock to the boundary layer than in the cylinder problem, despite a lower Mach number, which helps explain the aforementioned issue in which the solver moves grid points from the boundary layer to the shock.

Refer to caption
(a) Mesh.
Refer to caption
(b) Pressure field.
Refer to caption
(c) Surface heat flux.
Refer to caption
(d) Zoomed-in view of temperature profile along z=0𝑧0z=0italic_z = 0 plane.
Figure 5.30: Final grid and pressure field for the subparametric MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution to hypersonic flow over a sphere at Ma=5,Re=3.775×106formulae-sequenceMa5Re3.775superscript106\mathrm{Ma}=5,\mathrm{Re}=3.775\times 10^{6}roman_Ma = 5 , roman_Re = 3.775 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. In Figures 5.30a and 5.30b, the y=0𝑦0y=0italic_y = 0 and z=0𝑧0z=0italic_z = 0 planes and the sphere wall are displayed. The freestream flow is in the x𝑥-x- italic_x-direction.

The convergence history for the subparametric MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution is presented in Figure 5.31. The initial residual is already low since we restart from a subparametric MDG-ICE(𝒫4/𝒫3subscript𝒫4subscript𝒫3\mathcal{P}_{4}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) calculation. Figure 5.32 display profiles of temperature and pressure along the stagnation line. The stagnation point is located at x=1𝑥1x=-1italic_x = - 1. The stagnation-line quantities are plotted on a per-cell basis, such that only points within the same cell are connected. Again, it is very difficult to capture the interior of the extremely thin viscous shock using the employed line sampler, which explains why the shock is presented as a discontinuity. These results further exemplify the sharpness of the shock profile and the absence of spurious artifacts in the solution.

Refer to caption
Figure 5.31: Nonlinear convergence history for the subparametric MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution to Mach 5 flow over a sphere.
Refer to caption
(a) Temperature.
Refer to caption
(b) Pressure.
Figure 5.32: Stagnation-line profiles of temperature and pressure obtained with the MDG-ICE solution computed using 7161 𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT subparametric triangle elements for three-dimensional Mach 5 flow over a sphere at Re=3.775×106Re3.775superscript106\mathrm{Re}=3.775\times 10^{6}roman_Re = 3.775 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. The stagnation point is located at x=1𝑥1x=-1italic_x = - 1. The stagnation-line quantities are plotted on a per-cell basis, such that only points within the same cell are connected. The exact stagnation-point temperature, T=1.308𝑇1.308T=1.308italic_T = 1.308, is marked with the symbol ×\times×.

Figure 5.33 presents the streamwise variation of surface pressure and heat flux evaluated at all degrees of freedom corresponding to ΣhsubscriptΣ\Sigma_{h}roman_Σ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT for subparametric MDG-ICE(𝒫4/𝒫3subscript𝒫4subscript𝒫3\mathcal{P}_{4}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) and MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solutions. The surface pressure is essentially perfectly symmetric. Noticeable asymmetries in the heat flux near the stagnation point for the MDG-ICE(𝒫4/𝒫3subscript𝒫4subscript𝒫3\mathcal{P}_{4}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution are observed. These asymmetries are significantly reduced in the MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution, resulting in a highly symmetric heat-flux profile. Note that interaction between the symmetry boundaries near the stagnation point may be a partial source of the small asymmetries. The stagnation-point Stanton number in the MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solution is approximately 0.00167, which agrees well with the value of 0.00159 obtained using the correlation by Fay and Riddell (Fay58, ). These results demonstrate the ability of MDG-ICE, combined with the enhanced optimization solver introduced in this work, to produce very accurate surface heating predictions even in the presence of considerable misalignment between the grid and the high-gradient features.

Refer to caption
(a) Pressure coefficient profile.
Refer to caption
(b) Stanton number profile.
Figure 5.33: Streamwise variation of surface pressure and surface heat flux for the subparametric MDG-ICE(𝒫4/𝒫3subscript𝒫4subscript𝒫3\mathcal{P}_{4}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) and MDG-ICE(𝒫5/𝒫3subscript𝒫5subscript𝒫3\mathcal{P}_{5}/\mathcal{P}_{3}caligraphic_P start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT / caligraphic_P start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) solutions to three-dimensional Mach 5 flow over a sphere at Re=3.775×106Re3.775superscript106\mathrm{Re}=3.775\times 10^{6}roman_Re = 3.775 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. Pressure and heat-flux values at all auxiliary-variable degrees of freedom along the sphere wall are shown. The stagnation point is located at x=1𝑥1x=-1italic_x = - 1.

6 Conclusions and future work

We introduced an enhanced optimization solver for the moving discontinuous Galerkin method with interface condition enforcement (MDG-ICE), which automatically fits discontinuous and high-gradient features without a priori information via curvilinear r𝑟ritalic_r-adaptivity in both space and time. Specifically, to overcome the major bottleneck of frequent cell degeneration, we developed an anisotropic, locally adaptive penalty technique for the Levenberg-Marquardt method previously employed in (Ker20, ). The proposed MDG-ICE formulation, without any explicit stabilization mechanisms, was applied to three test cases involving sharp yet smooth gradients: Burgers viscous shock formation in space-time, Mach 17.6 viscous flow over a circular half-cylinder in two and three dimensions, and Mach 5 flow over a three-dimensional sphere. We used simplicial grids in order to evaluate the ability of the developed MDG-ICE formulation to obtain symmetric surface heat-flux profiles in the presence of strong misalignment between the grid and both the shock and boundary layer, which is considerably challenging for conventional numerical methods. Oscillation-free solutions and highly symmetric heating predictions were achieved.

Although the results presented in this work are very promising and demonstrate how MDG-ICE can significantly alleviate the burden of mesh generation on the user, additional advancements are needed in order to more fully realize its potential. Specifically, we plan to incorporate localized artificial dissipation (either only during intermediate iterations or in minimal amounts in the final solution), time marching via space-time MDG-ICE, local p𝑝pitalic_p-adaptivity, automatic selection and adjustment of regularization and other solver parameters (Tra12, ; Zah20, ), and further improvements to the optimization solver that were originally proposed for the baseline Levenberg-Marquardt method (Tra12, ). In addition, we will integrate metric-based mesh regeneration/adaptation strategies into the solver, potentially utilizing the mesh-implied metric naturally produced by an MDG-ICE solution, which should already yield small element length scales at high-gradient features. These developments will likely accelerate convergence, circumvent the need for continuation in Mach number and Reynolds number, and allow for even coarser initial grids. We will also explore approximating strong viscous shocks as inviscid discontinuities as an intrinsic feature of the solver (i.e., without the use of ad hoc strategies); layers of high-aspect-ratio cells would then no longer be required since discontinuous shocks can simply be fit along grid interfaces, as in the inviscid setting (Cor18, ). Finally, the efficiency of solving the linear system will be improved.

Acknowledgments

This work is sponsored by the Office of Naval Research through the Naval Research Laboratory 6.1 Computational Physics Task Area and by Dr. Eric Marineau of the Hypersonic Aerothermodynamics, High-Speed Propulsion and Materials Program of ONR Code 35.

References

  • (1) A. Corrigan, A. Kercher, D. Kessler, A moving discontinuous Galerkin finite element method for flows with interfaces, International Journal for Numerical Methods in Fluids 89 (9) (2019) 362–406. doi:10.1002/fld.4697.
  • (2) A. Kercher, A. Corrigan, D. Kessler, The moving discontinuous Galerkin finite element method with interface condition enforcement for compressible viscous flows, International Journal for Numerical Methods in Fluids 93 (5) (2021) 1490–1519. doi:10.1002/fld.4939.
  • (3) A. Kercher, A. Corrigan, A least-squares formulation of the moving discontinuous Galerkin finite element method with interface condition enforcement, Computers & Mathematics with Applications 95 (2021) 143–171. doi:10.1016/j.camwa.2020.09.012.
  • (4) W. H. Reed, T. Hill, Triangular mesh methods for the neutron transport equation, Tech. rep., Los Alamos Scientific Lab., N. Mex.(USA) (1973).
  • (5) B. Cockburn, G. Karniadakis, C.-W. Shu, The development of discontinuous Galerkin methods, in: Discontinuous Galerkin Methods, Springer, 2000, pp. 3–50.
  • (6) A. Majda, Compressible fluid flow and systems of conservation laws in several space variables, Springer Science & Business Media, 2012. doi:10.1007/978-1-4612-1116-7.
  • (7) E. Ching, Y. Lv, P. Gnoffo, M. Barnhardt, M. Ihme, Shock capturing for discontinuous Galerkin methods with application to predicting heat transfer in hypersonic flows, Journal of Computational Physics 376 (2018) 54–75. doi:10.1016/j.jcp.2018.09.016.
  • (8) G. Moretti, Thirty-six years of shock fitting, Computers & Fluids 31 (4) (2002) 719–723. doi:10.1016/S0045-7930(01)00072-X.
  • (9) M. Salas, A shock-fitting primer, CRC Press, 2009.
  • (10) M. Salas, A brief history of shock-fitting, in: Computational Fluid Dynamics 2010, Springer, 2011, pp. 37–53.
  • (11) M. Zahr, P.-O. Persson, An optimization-based approach for high-order accurate discretization of conservation laws with discontinuous solutions, Journal of Computational Physics 365 (2018) 105–134. doi:10.1016/j.jcp.2018.03.029.
  • (12) M. Zahr, A. Shi, P.-O. Persson, Implicit shock tracking using an optimization-based high-order discontinuous Galerkin method, Journal of Computational Physics 410 (2020) 109385. doi:10.1016/j.jcp.2020.109385.
  • (13) A. Shi, P.-O. Persson, M. J. Zahr, Implicit shock tracking for unsteady flows by the method of lines, Journal of Computational Physics 454 (2022) 110906.
  • (14) T. Huang, M. Zahr, A robust, high-order implicit shock tracking method for simulation of complex, high-speed flows, Journal of Computational Physics 454 (2022) 110981.
  • (15) T. Huang, C. J. Naudet, M. J. Zahr, High-order implicit shock tracking boundary conditions for flows with parametrized shocks, Journal of Computational Physics (2023) 112517.
  • (16) H. Luo, G. Absillis, R. Nourgaliev, A moving discontinuous Galerkin finite element method with interface condition enforcement for compressible flows, Journal of Computational Physics 445 (2021) 110618.
  • (17) I. Nompelis, T. Drayna, G. Candler, Development of a hybrid unstructured implicit solver for the simulation of reacting flows over complex geometries, in: 34th AIAA Fluid Dynamics Conference and Exhibit, 2004, p. 2227, AIAA-2004-2227. doi:10.2514/6.2004-2227.
  • (18) P. Gnoffo, J. White, Computational aerothermodynamic simulation issues on unstructured grids, in: 37th AIAA Thermophysics Conference, 2004, p. 2371, AIAA-2004-2371. doi:10.2514/6.2004-2371.
  • (19) G. Candler, D. Mavriplis, L. Trevino, Current status and future prospects for the numerical simulation of hypersonic flows, in: AIAA (Ed.), 47th AIAA Aerospace Sciences Meeting including The New Horizons Forum and Aerospace Exposition, 2009, AIAA-2009-0153. doi:10.2514/6.2009-0153.
  • (20) K. J. Fidkowski, A simplex cut-cell adaptive method for high-order discretizations of the compressible Navier-Stokes equations, Ph.D. thesis, Massachusetts Institute of Technology (2007).
  • (21) T. A. Oliver, A high-order, adaptive, discontinuous Galerkin finite element method for the Reynolds-Averaged Navier-Stokes equations, Ph.D. thesis, Massachusetts Institute of Technology (2008).
  • (22) M. Ceze, K. J. Fidkowski, Constrained pseudo-transient continuation, International Journal for Numerical Methods in Engineering 102 (11) (2015) 1683–1703.
  • (23) P. A. Gnoffo, Planetary-entry gas dynamics, Annual Review of Fluid Mechanics 31 (1) (1999) 459–494.
  • (24) K. Levenberg, A method for the solution of certain non-linear problems in least squares, Quarterly of applied mathematics 2 (2) (1944) 164–168. doi:10.1090/qam/10666.
  • (25) D. Marquardt, An algorithm for least-squares estimation of nonlinear parameters, Journal of the society for Industrial and Applied Mathematics 11 (2) (1963) 431–441. doi:10.1137/0111030.
  • (26) E. Ching, M. Ihme, Efficient projection kernels for discontinuous Galerkin simulations of disperse multiphase flows on arbitrary curved elements, Journal of Computational Physics 435 (2021) 110266. doi:10.1016/j.jcp.2021.110266.
  • (27) T. Toulorge, C. Geuzaine, J. Remacle, J. Lambrechts, Robust untangling of curvilinear meshes, Journal of Computational Physics 254 (2013) 8–26.
  • (28) F. Alauzet, A changing-topology moving mesh technique for large displacements, Engineering with Computers 30 (2) (2014) 175–200.
  • (29) R. Löhner, Applied CFD Techniques, J. Wiley & Sons, 2008.
  • (30) Z. Xie, R. Sevilla, O. Hassan, K. Morgan, The generation of arbitrary order curved meshes for 3D finite element analysis, Computational Mechanics 51 (2013) 361–374.
  • (31) D. Moxey, D. Ekelschot, Ü. Keskin, S. Sherwin, J. Peiró, High-order curvilinear meshing using a thermo-elastic analogy, Computer-Aided Design 72 (2016) 130–139.
  • (32) J. Marcon, A. Garai, M. Denison, S. Murman, An adjoint elasticity solver for high-order mesh deformation, in: AIAA Scitech 2021 Forum, 2021, p. 1238.
  • (33) A. Gargallo-Peiró, X. Roca, J. Peraire, J. Sarrate, Distortion and quality measures for validating and generating high-order tetrahedral meshes, Engineering with Computers 31 (3) (2015) 423–437.
  • (34) M. K. Transtrum, J. P. Sethna, Improvements to the Levenberg-Marquardt algorithm for nonlinear least-squares minimization, arXiv preprint arXiv:1201.5885 (2012).
  • (35) C. Geuzaine, J.-F. Remacle, Gmsh: A three-dimensional finite element mesh generator with built-in pre- and post-processing facilities, International Journal for Numerical Methods in Engineering (79(11)) (2009) 1310–1331.
  • (36) P. Amestoy, I. S. Duff, J. Koster, J.-Y. L’Excellent, A fully asynchronous multifrontal solver using distributed dynamic scheduling, SIAM Journal on Matrix Analysis and Applications 23 (1) (2001) 15–41.
  • (37) P. Amestoy, A. Buttari, J.-Y. L’Excellent, T. Mary, Performance and scalability of the block low-rank multifrontal factorization on multicore architectures, ACM Transactions on Mathematical Software 45 (2019) 2:1–2:26.
  • (38) S. Balay, W. D. Gropp, L. C. McInnes, B. F. Smith, Efficient management of parallelism in object oriented numerical software libraries, in: E. Arge, A. M. Bruaset, H. P. Langtangen (Eds.), Modern Software Tools in Scientific Computing, Birkhäuser Press, 1997, pp. 163–202.
  • (39) S. Balay, S. Abhyankar, M. F. Adams, S. Benson, J. Brown, P. Brune, K. Buschelman, E. Constantinescu, L. Dalcin, A. Dener, V. Eijkhout, J. Faibussowitsch, W. D. Gropp, V. Hapla, T. Isaac, P. Jolivet, D. Karpeev, D. Kaushik, M. G. Knepley, F. Kong, S. Kruger, D. A. May, L. C. McInnes, R. T. Mills, L. Mitchell, T. Munson, J. E. Roman, K. Rupp, P. Sanan, J. Sarich, B. F. Smith, S. Zampini, H. Zhang, H. Zhang, J. Zhang, PETSc/TAO users manual, Tech. Rep. ANL-21/39 - Revision 3.20, Argonne National Laboratory (2023). doi:10.2172/1968587.
  • (40) G. Barter, D. Darmofal, Shock capturing with PDE-based artificial viscosity for DGFEM: Part I. formulation, Journal of Computational Physics 229 (5) (2010) 1810–1827. doi:10.1016/j.jcp.2009.11.010.
  • (41) R. F. Johnson, A. D. Kercher, A conservative discontinuous Galerkin discretization for the chemically reacting Navier-Stokes equations, Journal of Computational Physics 423 (2020) 109826. doi:10.1016/j.jcp.2020.109826.
  • (42) E. Ching, A. Kercher, A. Corrigan, Anisotropic mesh modifications for the moving discontinuous Galerkin method with interface condition enforcement for robust simulations of high-speed viscous flows, in: Eleventh International Conference on Computational Fluid Dynamics (ICCFD11), Maui, Hawaii, 2022, ICCFD11-0305.
  • (43) K. Kitamura, E. Shima, Towards shock-stable and accurate hypersonic heating computations: A new pressure flux for AUSM-family schemes, Journal of Computational Physics 245 (2013) 62–83.
  • (44) Y. Lv, M. Ihme, Entropy-bounded discontinuous Galerkin scheme for Euler equations, Journal of Computational Physics 295 (2015) 715–739.
  • (45) Y. Jiang, H. Liu, Invariant-region-preserving DG methods for multi-dimensional hyperbolic conservation law systems, with an application to compressible Euler equations, Journal of Computational Physics 373 (2018) 385–409.
  • (46) E. Ching, R. Johnson, A. Kercher, Positivity-preserving and entropy-bounded discontinuous Galerkin method for the chemically reacting, compressible Euler equations. Part I: The one-dimensional case, arXiv preprint arXiv:2211.16254 https://arxiv.org/abs/2211.16254 (2022).
  • (47) L. Demkowicz, J. Gopalakrishnan, A class of discontinuous Petrov–Galerkin methods. Part I: The transport equation, Computer Methods in Applied Mechanics and Engineering 199 (23-24) (2010) 1558–1572. doi:10.1016/j.cma.2010.01.003.
  • (48) L. Demkowicz, J. Gopalakrishnan, A class of discontinuous Petrov–Galerkin methods. II. Optimal test functions, Numerical Methods for Partial Differential Equations 27 (1) (2011) 70–105. doi:10.1002/num.20640.
  • (49) L. Demkowicz, J. Gopalakrishnan, Discontinuous Petrov-Galerkin (DPG) method, Tech. Rep. 15-20, ICES, retrieved from https://www.oden.utexas.edu/media/reports/2015/1520.pdf (October 2015).
  • (50) F. Blottner, Accurate Navier-Stokes results for the hypersonic flow over a spherical nosetip, Journal of spacecraft and Rockets 27 (2) (1990) 113–122. doi:10.2514/3.26115.
  • (51) T. Fisher, High fidelity CFD workshop 2021: High speed steady advanced case: Blottner sphere, NASA Langley Research Center Turbulence Modeling Resource https://turbmodels.larc.nasa.gov/highfidelitycfd{_}workshop2022.html (2021).
  • (52) A. Murphy, R. Agarwal, Computational analysis of laminar steady hypersonic flow past Blottner sphere using ANSYS Fluent, in: AIAA AVIATION 2023 Forum, 2023, AIAA-2023-3847.
  • (53) B. Kirk, S. Bova, R. Bond, The influence of stabilization parameters in the SUPG finite element method for hypersonic flows, in: 48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition, 2010, AIAA-2010-1183.
  • (54) B. L. S. Couchman, Comparison of heat flux predictions over the Blottner cylinder and sphere., Tech. rep., Sandia National Lab. (SNL-NM), Albuquerque, NM (United States) (2018).
  • (55) J. Fay, F. Riddell, Theory of stagnation point heat transfer in dissociated air, Journal of the Aerospace Sciences 25 (2) (1958) 73–85.