Skip to main content

Boundary Closures for Sixth-Order Energy-Stable Weighted Essentially Non-Oscillatory Finite-Difference Schemes

  • Chapter
Advances in Applied Mathematics, Modeling, and Computational Science

Part of the book series: Fields Institute Communications ((FIC,volume 66))

Abstract

A general strategy was presented in 2009 by Yamaleev and Carpenter (J. Comput. Phys. 228(11):4248–4272, 2009; J. Comput. Phys. 228(8):3025–3047, 2009), for constructing Energy Stable Weighted Essentially Non-Oscillatory (ESWENO) finite-difference schemes on periodic domains. Fisher et al. (J. Comput. Phys. 230(10):3727–3752, 2011) provided boundary closures for the fourth-order ESWENO scheme that maintain, the WENO stencil biasing properties and satisfy the summation-by-parts (SBP) operator convention, thereby ensuring stability in an L 2 norm. Herein, the general capability of finite-domain schemes is extended by providing closures for the sixth-order case. Third-order and fifth-order boundary closures are developed that are stable in diagonal and block norms, respectively, and achieve fourth- and sixth-order global accuracy for hyperbolic systems. A novel set of nonuniform flux interpolation points is necessary near the boundaries to simultaneously achieve (1) accuracy, (2) the SBP convention, and (3) WENO stencil biasing mechanics. Complete implementation details for the diagonal-norm sixth-order operator are provided as well as examples that demonstrate shock capturing and multi-domain capabilities.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Stencil biasing across block interfaces is still an area of active development. Maturity is approximately equivalent to that of state-of-the-are finite-element WENO formulations.

  2. 2.

    The r={L2,L1,R1,R2} nomenclature identifies the origin of the data relative to the interface position.

  3. 3.

    Note that the dimensions of the even indexed Λ i are (N×N), while the odd-indexed Λ i have dimensions (N+1×N+1). This is the result of the difference matrix Δ having dimensions [N×(N+1)].

  4. 4.

    Equation (15) describes the role of the parameter ε in the WENO stencil biasing mechanics. The parameter determines the amplitude of oscillation that the biasing mechanics can detect. Using permits larger oscillations in the solution.

  5. 5.

    For applications where the solution changes drastically during simulation, ε can be rescaled as necessary.

References

  1. R. Borges, M. Carmona, B. Costa, and W. S. Don. An improved weighted essentially non-oscillatory scheme for hyperbolic conservation laws. Journal of Computational Physics, 227(6):3191–3211, 2008.

    Article  MathSciNet  MATH  Google Scholar 

  2. M. H. Carpenter, D. Gottlieb, and S. Abarbanel. Time-stable boundary conditions for finite-difference schemes solving hyperbolic systems: Methodology and application to high-order compact schemes. Journal of Computational Physics, 111(2):220–236, 1994.

    Article  MathSciNet  MATH  Google Scholar 

  3. M. H. Carpenter and C. A. Kennedy. Fourth-order 2n-storage Runge-Kutta schemes. Technical Report TM 109112, NASA, 1994.

    Google Scholar 

  4. M. H. Carpenter, J. Nordström, and D. Gottlieb. A stable and conservative interface treatment of arbitrary spatial accuracy. Journal of Computational Physics, 148(2):341–365, 1999.

    Article  MathSciNet  MATH  Google Scholar 

  5. M. H. Carpenter, J. Nordström, and D. Gottlieb. Revisiting and extending interface penalties for multi-domain summation-by-parts operators. Technical Report TM 214892, NASA, 2007.

    Google Scholar 

  6. M. H. Carpenter, J. Nordström, and D. Gottlieb. Revisiting and extending interface penalties for multi-domain summation-by-parts operators. Journal of Scientific Computing, 45(1–3):118–150, 2010.

    Article  MathSciNet  MATH  Google Scholar 

  7. J. Casper and M. H. Carpenter. Computational considerations for the simulation of shock-induced sound. SIAM Journal on Scientific Computing, 19(3):813–828, 1998.

    Article  MathSciNet  MATH  Google Scholar 

  8. B. Cockburn, C. Johnson, C.-W. Shu, and E. Tadmor. Advanced Numerical Approximation of Nonlinear Hyperbolic Equations. Springer, Berlin, 1998.

    MATH  Google Scholar 

  9. G. Erlebacher, M. Hussaini, and C.-W. Shu. Interaction of a shock with a longitudinal vortex. Journal of Fluid Mechanics, 337:129–153, 1997.

    Article  MathSciNet  MATH  Google Scholar 

  10. T. C. Fisher, M. H. Carpenter, N. K. Yamaleev, and S. H. Frankel. Boundary closures for fourth-order energy stable weighted essentially non-oscillatory finite-difference schemes. Journal of Computational Physics, 230(10):3727–3752, 2011.

    Article  MathSciNet  MATH  Google Scholar 

  11. B. Gustaffson. High Order Finite Difference Methods for Time Dependent PDE. Springer, Berlin, 2008.

    Google Scholar 

  12. B. Gustafsson. The convergence rate for difference approximations to mixed initial boundary value problems. Mathematics of Computation, 29:396–406, 1975.

    Article  MathSciNet  MATH  Google Scholar 

  13. G. Jiang and C.-W. Shu. Efficient implementation of weighted ENO schemes. Journal of Computational Physics, 126(1):202–228, 1996.

    Article  MathSciNet  MATH  Google Scholar 

  14. A. Kitson, R. I. McLachlan, and N. Robidoux. Skew-adjoint finite difference methods on nonuniform grids. New Zealand Journal of Mathematics, 32(2):139–159, 2003.

    MathSciNet  MATH  Google Scholar 

  15. H.-O. Kreiss and G. Scherer. Finite element and finite difference methods for hyperbolic partial differential equations. In Mathematical Aspects of Finite Elements in Partial Differential Equations, pages 195–212. Academic Press, San Diego, 1974.

    Google Scholar 

  16. P. Lax and M. Mock. The computation of discontinuous solutions of linear hyperbolic equations. Communications in Pure and Applied Mathematics, 31:423–430, 1978.

    Article  MathSciNet  MATH  Google Scholar 

  17. M. Martin, E. Taylor, M. Wu, and V. Weris. A bandwidth-optimized WENO scheme for the effective direct numerical simulation of compressible turbulence. Journal of Computational Physics, 220(1):270–289, 2006.

    Article  MATH  Google Scholar 

  18. K. Mattsson. Summation by parts operators for finite difference approximations of second-derivatives with variable coefficients. Journal of Scientific Computing, 29(1):1–33, 2011.

    Google Scholar 

  19. R. I. McLachlan and N. Robidoux. EQUADIFF 99. In Antisymmetry, Pseudospectral Methods, and Conservative PDEs, pages 994–999. World Scientific, Singapore, 2000.

    Google Scholar 

  20. J. Nordström. Conservative finite difference formulations, variable coefficients, energy estimates and artificial dissipation. Journal of Scientific Computing, 29(3):375–404, 2006.

    Article  MathSciNet  MATH  Google Scholar 

  21. J. Nordström and M. H. Carpenter. Boundary and interface conditions for high-order finite-difference methods applied to the Euler and Navier-Stokes equations. Journal of Computational Physics, 148(2):621–645, 1999.

    Article  MathSciNet  MATH  Google Scholar 

  22. J. Nordström, J. Gong, E. van der Weide, and M. Svärd. A stable and conservative high order multi-block method for the compressible Navier-stokes equation. Journal of Computational Physics, 228(24):9020–9035, 2009.

    Article  MathSciNet  MATH  Google Scholar 

  23. C.-W. Shu. Essentially non-oscillatory and weighted essentially non-oscillatory schemes for hyperbolic conservation laws. In Advanced Numerical Approximation of Nonlinear Hyperbolic Equations: Lecture Notes in Mathematics, volume 1697, pages 325–432. Springer, Berlin, 1998.

    Chapter  Google Scholar 

  24. B. Strand. Summation by parts for finite difference approximations for d/dx. Journal of Computational Physics, 110(1):47–67, 1994.

    Article  MathSciNet  MATH  Google Scholar 

  25. M. Svärd and J. Nordström. On the order of accuracy for difference approximations of initial-boundary value problems. Journal of Computational Physics, 218(1):333–352, 2006.

    Article  MathSciNet  MATH  Google Scholar 

  26. N. K. Yamaleev and M. H. Carpenter. A systematic methodology for constructing high-order energy stable WENO schemes. Journal of Computational Physics, 228(11):4248–4272, 2009.

    Article  MathSciNet  MATH  Google Scholar 

  27. N. K. Yamaleev and M. H. Carpenter. Third-order energy stable WENO scheme. Journal of Computational Physics, 228(8):3025–3047, 2009.

    Article  MathSciNet  MATH  Google Scholar 

  28. S. Zhang, S. Jiang, Y. Zhang, and C. Shu. The mechanism of sound generation in the interaction between a shock wave and two counter-rotating vortices. Physics of Fluids, 21:076101, 2009.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mark H. Carpenter .

Editor information

Editors and Affiliations

Appendix

Appendix

1.1 8.1 Implementation of WENO

The nonlinear WENO scheme is defined as

(67)

The fluxes are constructed by using the formula

$$ \bar{f}_{j} = + {\bar{w}}^{L2}_{j} {\bar{f}}^{L2}_{j} + {\bar{w}}^{L1}_{j} {\bar{f}}^{L1}_{j} + {\bar{w}}^{R1}_{j} {\bar{f}}^{R1}_{j} + {\bar{w}}^{R2}_{j} {\bar{f}}^{R2}_{j}, \quad 7 \le j \le N-6. $$
(68)

The expressions for \(\bar{f}_{j}\) at the six boundary closure points are slightly more complicated but conform to the same matrix conventions,

$$ \bar{f}_{j} = \sum_{i=1}^{n_b} {\bar{w}}^{i}_{j} {\bar{f}}^{i}_{j}, \quad 1 \le j \le6 $$
(69)

where n b is the number of flux terms in each boundary flux.

The nonlinear weights \(\bar{w}^{(r)}\) are defined by

$$ \everymath{\displaystyle} \begin{array} {l} \bar{w}_{j}^{(r)}=\frac{\bar{\alpha}_{j}^{(r)}}{\sum_r \bar{\alpha}_{j}^{(r)}} ; \qquad \bar{ \alpha}_{j}^{(r)}=\bar{d}_{j}^{(r)} \biggl(1+ \frac{\bar{\tau }_{j}}{\varepsilon+ \bar{\beta}_{j}^{(r)}} \biggr) , \quad \forall r, \ 1 \le j \le N-1, \\\noalign{\vspace{8pt}} \bar{w}^{(r)} = \mathrm{Diag}\bigl[\bar{w}^{(r)}_0, \ldots,\bar{w}^{(r)}_N\bigr] , \quad \forall r. \end{array} $$
(70)

The coefficients \(\bar{d}_{j}^{(r)}\) are the target weights of the candidate stencils, and \(\bar{\beta}_{j}^{(r)}\) and \(\bar{\tau}_{j}\) are the smoothness indicators.

The smoothness indicators \(\bar{\beta}_{j}^{(r)}\) have the same form at all flux points, although they may not all be defined near the boundaries. Experimentation has determined that the choice of f affects the solution smoothness and dissipation properties, and that the best practice is to base the smoothness indicator on the characteristic variables rather than the characteristic fluxes. The interior smoothness indicators

$$ \begin{array} {rcl} \bar{\beta}_{j}^{L2}&=& (f_{j-3}-3 f_{j-2} + 2 f_{j-1} )^2 + (f_{j-3}-2 f_{j-2} + f_{j-1} )^2 ; \\\noalign{\vspace{8pt}} \bar{\beta}_{j}^{L1}&=& (-f_{j-1} + f_{j-0} )^2 + (f_{j-2}-2 f_{j-1} + f_{j-0} )^2 ; \\\noalign{\vspace{8pt}} \bar{\beta}_{j}^{R1}&=& (-f_{j-1} + f_{j-0} )^2 + (f_{j-1}-2 f_{j-0} + f_{j+1} )^2 ; \\\noalign{\vspace{8pt}} \bar{\beta}_{j}^{R2}&=& (-2 f_{j-0} + 3 f_{j+1} - f_{j+2} )^2 + (f_{j-0}-2 f_{j+1} + f_{j+2} )^2; \end{array} \quad 7 \le j \le N-6. $$
(71)

The indicators \(\bar{\beta}^{L3}\) and \(\bar{\beta}^{R3}\) (needed only near boundaries) as well as the \(\bar{\beta}^{r}\) stencil coefficients used near the boundaries are included in Appendix 8.3.

To guarantee that the downwind stencil weight does not exceed that of the central or upwind weights, the downwind smoothness indicator is modified by using the expression

$$ \bar{\beta}_{j}^d= \biggl( \frac{1}{n_s}\sum_r {\bigl[\bar{\beta }_{j}^{(r)}\bigr]}^k \biggr)^{1/k} $$
(72)

where k is an even integer, and n s is the number of distinct interpolants contributing to the flux f j .

An additional stencil biasing parameter, \(\bar{\tau}_{j}\), is needed for the ESWENO scheme. Here, \(\bar{\tau}_{j}\) is a quadratic function of the sum of fourth-order undivided differences that are available on the stencil for \(\bar{f}_{j}\) (see Fig. 1),

(73)

At the points \(\bar{x}_{1}\) and \(\bar{x}_{2}\), \(\bar{\tau}_{j}\) is constructed from a single fourth-order undivided difference, biased toward the interior of the domain as

$$ \bar{\tau}_{j}= ( - f_{1} + 4 f_{2} -6 f_{3 } +4 f_{4} - f_{5})^2. $$
(74)

The \(\bar{\tau}_{j}\) is biased in a mirrored fashion at \(\bar {x}_{N-2}\) and \(\bar{x}_{N-1}\).

The parameter ε is a function of the number of points in the discretization:

$$ \varepsilon=\max\bigl(\|f_0\|,\bigl\|f_0'\bigr\| \bigr)_{x \neq x_d} ({\delta x} )^4, \quad{\delta x}= \frac{1}{N} $$
(75)

where ∥f 0∥ and ∥f 0∥′ represent a norm of the flux and the gradient of the flux, respectively, as determined using initial conditions but excluding points near discontinuities.Footnote 5

1.2 8.2 Recipe

Consider the Euler equations U t +F(U) x =0. A recipe for calculating the gradient term F(U) x using a sixth-order ESWENO scheme is summarized below.

  1. 1.

    Construct the convective flux F at solution points (x) according to the governing differential equations.

  2. 2.

    Construct the flux Jacobian matrices: \(A = \frac{\partial F}{\partial U}\) at the flux points (\(\bar{\mathbf{x}}\)) by using the Roe-averaged variables that are formed from nearest neighbor solution point data. Form the eigenvector decomposition A=SϒS −1, where S is the matrix of right eigenvectors and ϒ is the diagonal matrix of eigenvalues.

  3. 3.

    For the flux point \(\bar{x}_{j}\), use the eigenvector matrix \(\bar{S}^{-1}_{j}\) to transform the solution and the flux at all points x into characteristic form \(U_{c} = \bar{S}^{-1} U ; F_{c}(U) = \bar{S}^{-1} F(U) = \varUpsilon U_{c}\).

    Remark Forming the characteristic variables and fluxes is only necessary for nodal points that are located in the stencil of a given flux point. The presentation herein assumes that the transformed fluxes and variables are available at all points but can be reduced to improve performance.

  4. 4.

    Form the Lax-Friedrichs characteristic fluxes \(\mathbf{f}_{\mathbf{c}}^{\pm}=\frac{1}{2} (F_{c} \pm\varUpsilon_{max} U_{c} )\), where ϒ max is a diagonal matrix of the maximum local eigenvalues contained within each stencil.

  5. 5.

    Perform interpolations on each candidate stencil, .

  6. 6.

    Calculate the stencil biasing parameters:

    1. a.

      Calculate \(\bar{\beta}^{r}\) for each flux point according to (71), except at the end points.

    2. b.

      Calculate \(\bar{\tau}_{j}\) according to (73) and (74).

      Remark The smoothness indicators are calculated by using the characteristic variables U c in the stencil of the flux point.

  7. 7.

    Calculate and normalize the weights using the stencil biasing parameters and the target weights in (70) (see (15) in text).

  8. 8.

    Calculate Λ i from the weights by using (88) through (91) for the diagonal-norm scheme.

  9. 9.

    Modify the diagonal of Λ i to be smoothly positive according to (32).

  10. 10.

    Calculate the WENO flux from the weights and candidate interpolations by using (68) (see (20) in the text).

  11. 11.

    Calculate the energy stable flux \(\bar{\boldsymbol{\psi}}\) from (33).

  12. 12.

    Reconstruct the fluxes in characteristic space, \(\bar{f}_{j}=\bar{f}^{+}_{j}+\bar{f}_{j}^{-}\) and \(\psi_{j}=\psi_{j}^{+}+\psi_{j}^{-}\), and then transform the fluxes back to physical space by using \(\hat{f}_{j} = S (\bar{f}_{j} + \psi_{j} )\).

  13. 13.

    After calculating the ESWENO interpolation at all flux points, calculate the gradient using the inverse of the norm, as in (34).

Remark

Note that all of the presented equations are for the forward-propagating waves, \(\bar{f}_{j}^{+}\). The equations for interpolation of backward-propagating waves (\(\bar{f}_{j}^{-}\)) are found in exactly the same manner, except that the downwind stencil in (72) is the far left instead of the far right stencil. The Λ i terms are to be negative instead of positive; therefore, we use

$$ {[\hat{\lambda}_j]}_i= - \frac{1}{2} \Bigl(\sqrt{ \bigl({[\lambda_j]}_i \bigr)^2 + \delta_i^2}+{[ \lambda_j]}_i \Bigr) , \quad i={0,5}, \ j={0,N}. $$
(76)

1.3 8.3 Sixth-Order Diagonal-Norm Operator:

1.3.1 8.3.1 Differentiation Matrix

The upper left matrix quadrant of the target SBP differentiation operator can be written in the form

(77)

The matrix elements in the lower right quadrant are structurally equivalent relative to the outflow boundary, but with the opposite sign. The diagonal mass matrix is

The skew-symmetric matrix follows immediately from .

The distribution the flux points \(\bar{\mathbf{x}}\) as described in Lemma 1 follows immediately from the norm and can be written as

$$ \bar{\mathbf{x}} = \biggl[0,\bar{x}_{1}, \ldots, \bar{x}_{5}, \frac{11 {\delta x}}{2}, \ldots, \biggl(1- \frac{11 {\delta x}}{2} \biggr), (1-\bar{x}_{5} ), \ldots, (1-\bar{x}_{1} ), 1 \biggr]^T $$
(78)

where

$$ \everymath{\displaystyle} \begin{array}{l} \bar{x}_{1} = \frac{13649{\delta x}}{43200} ;\qquad \bar{x}_{2} = \frac{36857{\delta x}}{21600} ;\qquad \bar{x}_{3} = \frac{4201{\delta x}}{1800} ; \\\noalign{\vspace{7pt}} \bar{x}_{4} = \frac{77207{\delta x}}{21600} ;\qquad \bar{x}_{5} = \frac{193799{\delta x}}{43200} . \end{array} $$
(79)

Then, the vector \({\varDelta }\bar{\mathbf{x}}\) is the reciprocal of the diagonal elements of . The product is trivially the identity vector 1 because is a diagonal matrix. Note that all integer coefficients are exact. Case should be taken to ensure that the rational coefficients retain the full working precision of the coding language (e.g. 64 bit arithmetic).

1.3.2 8.3.2 Interpolation Operators

The tri-diagonal interpolation matrices , r={L4,L3,L2,L1,R1,R2,R3,R4} for the ESWENO3-6-3 scheme, presented in compressed format using the arrays I r , r={L4,L3,L2,L1,R1,R2,R3,R4}, are given by

(80)
(81)

The interpolants I r , r={R1,R2,R3,R4} are mirror images of I r , r={L1,L2,L3,L4}. For example, the interpolants I r , r={R1,R2} are given by

$$ I^{R1}=\left ( \begin{array}{c@{\quad}c@{\quad}c} 0 & 0 & 0 \\\noalign{\vspace{2pt}} \frac{15151}{28800} & \frac{13649}{21600} & -\frac{13649}{86400} \\\noalign{\vspace{5pt}} \frac{269}{2160} & \frac{2507}{2400} & -\frac{3653}{21600} \\\noalign{\vspace{5pt}} \frac{763}{1600} & \frac{5129}{7200} & -\frac{109}{576} \\\noalign{\vspace{5pt}} \frac{127}{450} & \frac{18601}{21600} & -\frac{3097}{21600} \\\noalign{\vspace{5pt}} \frac{29401}{86400} & \frac{5}{6} & -\frac{15001}{86400} \\\noalign{\vspace{5pt}} \frac{1}{3} & \frac{5}{6} & -\frac{1}{6} \\ \vdots& \vdots& \vdots\\ \frac{1}{3} & \frac{5}{6} & -\frac{1}{6} \\\noalign{\vspace{5pt}} \frac{8999}{28800} & \frac{18601}{21600} & -\frac{15001}{86400} \\\noalign{\vspace{5pt}} \frac{931}{2160} & \frac{5129}{7200} & -\frac{3097}{21600} \\\noalign{\vspace{5pt}} \frac{2083}{14400} & \frac{2507}{2400} & -\frac{109}{576} \\\noalign{\vspace{5pt}} \frac{967}{1800} & \frac{13649}{21600} & -\frac{3653}{21600} \\\noalign{\vspace{5pt}} 0 & 0 & 0 \\ 0 & 0 & 0 \end{array} \right ), \qquad I^{R2}=\left ( \begin{array}{c@{\quad}c@{\quad}c} 0 & 0 & 0 \\\noalign{\vspace{2pt}} \frac{38191}{17280} & -\frac{18751}{10800} & \frac{15151}{28800} \\\noalign{\vspace{5pt}} \frac{10211}{7200} & -\frac{11723}{21600} & \frac{269}{2160} \\\noalign{\vspace{5pt}} \frac{30859}{14400} & -\frac{11663}{7200} & \frac{763}{1600} \\\noalign{\vspace{5pt}} \frac{36889}{21600} & -\frac{4277}{4320} & \frac{127}{450} \\\noalign{\vspace{5pt}} \frac{53401}{28800} & -\frac{25801}{21600} & \frac{29401}{86400} \\\noalign{\vspace{5pt}} \frac{11}{6} & -\frac{7}{6} & \frac{1}{3} \\ \vdots& \vdots& \vdots\\ \frac{11}{6} & -\frac{7}{6} & \frac{1}{3} \\\noalign{\vspace{5pt}} \frac{31079}{17280} & -\frac{11999}{10800} & \frac{8999}{28800} \\\noalign{\vspace{5pt}} \frac{4813}{2400} & -\frac{31027}{21600} & \frac{931}{2160} \\\noalign{\vspace{5pt}} \frac{7097}{4800} & -\frac{4487}{7200} & \frac{2083}{14400} \\\noalign{\vspace{5pt}} 0 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & 0 \end{array} \right ) . $$
(82)

Reconstructing from the data I r is accomplished by populating the appropriate diagonals in the interpolation matrices. The middle column of I r coincides with the diagonal of , r={L4,L3,L2,L1,R1,R2,R3,R4} that is shifted {−4,−3,−2,−1,0,1,2,3} relative to the main diagonal.

1.3.3 8.3.3 Target Weights

The target weights for the five interpolation stencils are

$$ \hbox{\small$\displaystyle \mathbf{d}_{\text{3-6-3}}=\left ( \begin{array}{c@{\ }c@{\ }c@{\ }c@{\ }c@{\ }c} 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & \frac{14400}{15151} & \frac{3475559}{5786318410} \\\noalign{\vspace{7pt}} 0 & 0 & 0 & \frac{17609}{29224} & \frac{34899909}{19653140} & -\frac {720681767}{494416620} \\\noalign{\vspace{7pt}} 0 & 0 & -\frac{113713}{74988} & \frac{364030091}{204342300} & \frac {6384599}{6180300} & -\frac{73697}{247212} \\\noalign{\vspace{7pt}} 0 & \frac{86153}{1039608} & -\frac{140213393}{604921905} & \frac {903169}{758765} & -\frac{948239}{9439656} & \frac{15}{254} \\\noalign{\vspace{7pt}} -\frac{15025}{2221158} & \frac{6395825773}{172578423705} & -\frac {367011523}{8390397630} & \frac{72016320}{134993999} & \frac {190085760}{441044401} & \frac{1440}{29401} \\\noalign{\vspace{7pt}} 0 & 0 & \frac{1}{20} & \frac{9}{20} & \frac{9}{20} & \frac{1}{20} \\ \vdots& \vdots& \vdots& \vdots& \vdots& \vdots\\ 0 & 0 & \frac{1}{20} & \frac{9}{20} & \frac{9}{20} & \frac{1}{20} \\\noalign{\vspace{7pt}} 0 & 0 & \frac{1440}{29401} & \frac{190085760}{441044401} & \frac {72016320}{134993999} & -\frac{367011523}{8390397630} \\\noalign{\vspace{7pt}} 0 & 0 & \frac{15}{254} & -\frac{948239}{9439656} & \frac {903169}{758765} & -\frac{140213393}{604921905} \\\noalign{\vspace{7pt}} 0 & 0 & -\frac{73697}{247212} & \frac{6384599}{6180300} & \frac {364030091}{204342300} & -\frac{113713}{74988} \\\noalign{\vspace{7pt}} 0 & \frac{58297}{735192} & -\frac{720681767}{494416620} & \frac {34899909}{19653140} & \frac{17609}{29224} & 0 \\\noalign{\vspace{7pt}} -\frac{15025}{2537142} & \frac{13296322573}{242239975305} & \frac {3475559}{5786318410} & \frac{14400}{15151} & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 & 0 \\ \end{array} \right )$.} $$
(83)

For brevity, the last two columns are excluded from the table. Recognizing the relationship between columns 4 and 7 (and 5 and 6), columns 6 and 7 can be recovered from the mirror images of columns 1 and 2, respectively. Note that the WENO/ESWENO3-6-3 scheme requires the inclusion of two extra stencils near each boundary.

1.3.4 8.3.4 Smoothness Indicators Coefficients

The smoothness indicators \(\bar{\beta}^{r}\) measure the smoothness of the data used by each interpolant. This is accomplished by comparing the L 2 of the undivided first and second derivatives, evaluated at the flux point \({\bar{x}}_{j}\).

$$ \bar{\beta}^{r} = {\bigl[{\delta x}f_{x}^{r}({\bar{x}}_{j})\bigr]}^2 + {\bigl[{\delta x}^2 f_{2x}^{r}({ \bar{x}}_{j} )\bigr]}^2 , \quad1 \le j \le N-1. $$
(84)

The leading order truncation term is identical in all cases. For example, the interior terms

$$ \everymath{\displaystyle} \begin{array}{rcl} \bar{\beta}^{L2} &=& {f_{x}}^2 {\delta x}^2 + \biggl({f_{2x}}^2 - \frac{23}{12} f_{x}f_{3x} \biggr){\delta x}^4 + (+ 3 f_{2x}f_{3x} - 2 f_{x}f_{4x} ) {\delta x}^5 + O\bigl({\delta x}^6\bigr), \\\noalign{\vspace{9pt}} \bar{\beta}^{L1} &=& {f_{x}}^2 {\delta x}^2 + \biggl({f_{2x}}^2 + \frac{ 1}{12} f_{x}f_{3x} \biggr){\delta x}^4 + (+ 1 f_{2x}f_{3x} ) {\delta x}^5 + O\bigl({\delta x}^6\bigr), \\\noalign{\vspace{9pt}} \bar{\beta}^{R1} &=& {f_{x}}^2 {\delta x}^2 + \biggl({f_{2x}}^2 + \frac{ 1}{12} f_{x}f_{3x} \biggr){\delta x}^4 + (- 1 f_{2x}f_{3x} ) {\delta x}^5 + O\bigl({\delta x}^6\bigr), \\\noalign{\vspace{9pt}} \bar{\beta}^{R2} &=& {f_{x}}^2 {\delta x}^2 + \biggl({f_{2x}}^2 - \frac{23}{12} f_{x}f_{3x} \biggr){\delta x}^4 + (- 3 f_{2x}f_{3x} + 2 f_{x}f_{4x} ) {\delta x}^5 + O\bigl({\delta x}^6\bigr) . \end{array} $$
(85)

The coefficients needed for the first derivative term \({\delta x} \frac{\partial}{\partial x}\) in the smoothness indicators throughout the spatial domain can be expressed as follows.

(86)
(87)

The coefficients needed for the first derivative term \({\delta x} \frac{\partial}{\partial x}\) in the rightward biased stencils (r=R1,R2,R3,R4) are the mirror images of those used for leftward biased stencils (r=L1,L2,L3,L4).

The second derivative stencil coefficients are always

$${\delta x}^2 f_{2x}^{r}({\bar{x}}_{j}) = f_{r+j-1}^{r} - 2 f_{r+j}^{r} + f_{r+j+1}^{r} ; \quad1 \le j \le N-1 , \forall r . $$

1.3.5 8.3.5 Energy-Stable Terms

The stabilization matrix Diag(Λ 0) is identically zero for the ESWENO3-6-3 scheme (as it is with every consistent scheme). The nonzero terms of the diagonal matrices Λ i , i=1,5 are

(88)
(89)
(90)
(91)
$$ \hbox{\small$\displaystyle \mathrm{Diag} (\varLambda_5 )= \left ( \begin{array}{c} 0 \\ 0 \\ 0 \\ (370193 \bar{w}^{L4}_{5}-422857 \bar{w}^{R4}_{1}) / 172800 \\\noalign{\vspace{6pt}} \begin{array}{l} (5850 \bar{w}^{L4}_{6}-5011 \bar{w}^{R4}_{2}) / 2700 \\\noalign{\vspace{3pt}} \quad{} -(69251 \bar{w}^{R4}_{3}) / 28800 \\ \end{array}\\\noalign{\vspace{4pt}} 0 \\[-4pt] \vdots\\ 0 \\ (69251 \bar{w}^{L4}_{N-3}) / 28800 \\\noalign{\vspace{6pt}} (5011 \bar{w}^{L4}_{N-2}-5850 \bar{w}^{R4}_{N-6}) / 2700 \\\noalign{\vspace{6pt}} (422857 \bar{w}^{L4}_{N-1}-370193 \bar{w}^{R4}_{N-5}) / 172800 \\\noalign{\vspace{6pt}} 0 \\ 0 \\ 0 \end{array} \right )$.} $$
(92)

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer Science+Business Media New York

About this chapter

Cite this chapter

Carpenter, M.H., Fisher, T.C., Yamaleev, N.K. (2013). Boundary Closures for Sixth-Order Energy-Stable Weighted Essentially Non-Oscillatory Finite-Difference Schemes. In: Melnik, R., Kotsireas, I. (eds) Advances in Applied Mathematics, Modeling, and Computational Science. Fields Institute Communications, vol 66. Springer, Boston, MA. https://doi.org/10.1007/978-1-4614-5389-5_6

Download citation

Publish with us

Policies and ethics