Log in

Construction of Approximate Entropy Measure-Valued Solutions for Hyperbolic Systems of Conservation Laws

  • Published:
Foundations of Computational Mathematics Aims and scope Submit manuscript

“There is no theory for the initial value problem for compressible flows in two space dimensions once shocks show up, much less in three space dimensions. This is a scientific scandal and a challenge.”

P. D. Lax, 2007 Gibbs Lecture [48].

Abstract

Entropy solutions have been widely accepted as the suitable solution framework for systems of conservation laws in several space dimensions. However, recent results in De Lellis and Székelyhidi Jr (Ann Math 170(3):1417–1436, 2009) and Chiodaroli et al. (2013) have demonstrated that entropy solutions may not be unique. In this paper, we present numerical evidence that state-of-the-art numerical schemes need not converge to an entropy solution of systems of conservation laws as the mesh is refined. Combining these two facts, we argue that entropy solutions may not be suitable as a solution framework for systems of conservation laws, particularly in several space dimensions. We advocate entropy measure-valued solutions, first proposed by DiPerna, as the appropriate solution paradigm for systems of conservation laws. To this end, we present a detailed numerical procedure which constructs stable approximations to entropy measure-valued solutions, and provide sufficient conditions that guarantee that these approximations converge to an entropy measure-valued solution as the mesh is refined, thus providing a viable numerical framework for systems of conservation laws in several space dimensions. A large number of numerical experiments that illustrate the proposed paradigm are presented and are utilized to examine several interesting properties of the computed entropy measure-valued solutions.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27

Similar content being viewed by others

Notes

  1. We have tested at least three types of schemes, TeCNO scheme of [27], the high-resolution HLLC scheme of [44] and the finite volume scheme of [31], and obtained similar non-convergence and instability results as presented above. We strongly suspect that any numerical method will not converge or be stable with respect to perturbations in the initial data for this particular example.

References

  1. L. Ambrosio, N. Gigli and G. Savaré. Gradient Flows. Birkhäuser Basel, 2005.

    MATH  Google Scholar 

  2. E. J. Balder. Lectures on Young Measures. Université de Paris-Dauphine, 1995.

  3. J. Ball. A version of the fundamental theorem for Young measures. In PDEs and Continuum Models of Phase Transitions (M. Rascle, D. Serre and M. Slemrod, eds.), Lecture Notes in Physics, vol. 344, Springer, 1989. 207–215.

  4. T. J. Barth. Numerical methods for gas-dynamics systems on unstructured meshes. In An Introduction to Recent Developments in Theory and Numerics of Conservation Laws, Lecture Notes in Computational Science and Engineering volume 5, Springer, Berlin. Eds: D. Kroner, M. Ohlberger, and Rohde, C., 1999, 195–285.

    Chapter  Google Scholar 

  5. S. Bianchini and A. Bressan. Vanishing viscosity solutions of nonlinear hyperbolic systems. Ann. of Math. (2) 161 (2005), no. 1, 223–342.

    Article  MathSciNet  MATH  Google Scholar 

  6. P. Billingsley. Probability and Measure 3rd ed. John Wiley & Sons Inc., 1995.

  7. P. Billingsley. Convergence of Probability Measures. John Wiley & Sons, Inc, 2008.

  8. Y. Brenier and C. De Lellis and L. Székelyhidi Jr. Weak-Strong Uniqueness for Measure-Valued Solutions. Comm. Math. Phys., 305 (2), 2011, 351–361.

    Article  MathSciNet  MATH  Google Scholar 

  9. A. Bressan, G. Crasta and B. Piccoli. Well-posedness of the Cauchy problem for \(n\times n\) systems of conservation laws. Memoirs of the AMS, 146 (694), 2000.

  10. Central Station: high-resolution non-oscillatory central schemes for non-linear conservation laws and related problems, www.cscamm.umd.edu/centpack/publications/.

  11. G. Q. Chen and J. Glimm. Kolmogorov’s theory of turbulence and inviscid limit of the Navier-Stokes equations in \(\mathbb{R}^3\). Comm. Math. Phys. 310 (1), 2012, 267–283.

    Article  MathSciNet  MATH  Google Scholar 

  12. B. Cockburn, F. Coquel and P. G. LeFloch. Convergence of the finite volume method for multidimensional conservation laws. SIAM J. Numer. Anal., 32 (3), 1995, 687–705.

    Article  MathSciNet  MATH  Google Scholar 

  13. B. Cockburn, C. Johnson, C. -W. Shu and E. Tadmor. Advanced Numerical Approximation of Nonlinear Hyperbolic Equations. Lecture notes in Mathematics 1697, 1997 C.I.M.E. course in Cetraro, Italy, June 1997 (A. Quarteroni ed.), Springer Verlag 1998.

  14. B. Cockburn and C-W. Shu. TVB Runge-Kutta local projection discontinuous Galerkin finite element method for conservation laws. II. General framework. Math. Comput., 52, 1989, 411–435.

    MATH  Google Scholar 

  15. M. G. Crandall and A. Majda. Monotone difference approximations for scalar conservation laws. Math. Comput. 34, 1980, 1–21.

    Article  MathSciNet  MATH  Google Scholar 

  16. C. Dafermos. Hyperbolic conservation laws in continuum physics. Springer, Berlin, 2000.

    Book  MATH  Google Scholar 

  17. C. De Lellis, L. Székelyhidi Jr. The Euler equations as a differential inclusion. Ann. of Math. (2) 170 (2009), no. 3, 1417–1436.

    Article  MathSciNet  MATH  Google Scholar 

  18. E. Chiodaroli, C. De Lellis, O. Kreml. Global ill-posedness of the isentropic system of gas dynamics. Preprint, 2013.

  19. C. DeLellis and L. Székelyhidi Jr. On the admissibility criteria for the weak solutions of Euler equations. Arch. Rational Mech. Anal. 195 (2010), 225-260.

    Article  MathSciNet  MATH  Google Scholar 

  20. S. Demoulini and D. M. A. Stuart and A. E. Tzavaras. Weak-strong uniqueness of dissipative measure-valued solutions for polyconvex elastodynamics. Archives of Rational Mechanics and Analysis, 205 (3), 2012, 927–961.

    Article  MATH  Google Scholar 

  21. B. Depres, G. Poette and D. Lucor. Uncertainty quantification for systems of conservation laws. J. Comput. Phys. 228 (2009), no. 7, 2443–2467.

    Article  MathSciNet  MATH  Google Scholar 

  22. R. J. DiPerna. Measure valued solutions to conservation laws. Arch. Rational Mech. Anal., 88(3), 1985, 223–270.

    Article  MathSciNet  MATH  Google Scholar 

  23. R. J. DiPerna and A. Majda. Oscillations and concentrations in weak solutions of the incompressible fluid equations. Comm. Math. Phys. 108 (4), 1987, 667–689.

    Article  MathSciNet  MATH  Google Scholar 

  24. R. E. Edwards. Functional Analysis. Theory and Applications. Holt, Rinehart and Winston, Inc. (1965).

  25. U. S. Fjordholm, S. Lanthaler and S. Mishra. Statistical solutions of hyperbolic conservation laws I. Theory, In preparation, 2015.

  26. U. S. Fjordholm. S. Mishra and E. Tadmor. ENO reconstruction and ENO interpolation are stable. FoCM 13 (2), 2013, 139–159.

    MathSciNet  MATH  Google Scholar 

  27. U. S. Fjordholm, S. Mishra and E. Tadmor. Arbitrary order accurate essentially non-oscillatory entropy stable schemes for systems of conservation laws. SIAM J. Num. Anal 50 (2), 2012, 544–573.

    Article  MATH  Google Scholar 

  28. U. S. Fjordholm. High-order accurate entropy stable numerical schemes for hyperbolic conservation laws. ETH Zürich dissertation Nr. 21025, 2013.

  29. U. S. Fjordholm, S. Mishra and E. Tadmor. Computation of measure valued solutions. Acta Numerica. In preparation, 2016.

  30. G. B. Folland. Real Analysis. John Wiley & Sons Inc, 1999.

  31. F. Fuchs, A. McMurry, S. Mishra, N. H. Risebro and K. Waagan. Approximate Riemann solver based high-order finite volume schemes for the MHD equations in multi-dimensions. Comm. Comput. Phys 9, 2011, 324–362.

    Article  Google Scholar 

  32. H. Frid and I-S. Liu. Oscillation waves in Riemann problems for phase transitions. Quart. Appl. Math. 56 (1), 1998, 115–135.

    Article  MathSciNet  MATH  Google Scholar 

  33. H. Frid and I-S. Liu. Oscillation waves in Riemann problems inside elliptic regions for conservation laws of mixed type. Z. Angew. Math. Phys. 46 (1995), no. 6, 913–931.

    Article  MathSciNet  MATH  Google Scholar 

  34. S. Gottlieb, C.-W. Shu and E. Tadmor. High order time discretizations with strong stability properties. SIAM. Review 43, 2001, 89–112.

    Article  MathSciNet  MATH  Google Scholar 

  35. J. Glimm. Solutions in the large for nonlinear hyperbolic systems of equations. Comm. Pure Appl. Math. 18 (4), 1965, 697-715.

    Article  MathSciNet  MATH  Google Scholar 

  36. J. Glimm, J. Grove and Y. Zhang, Numerical Calculation of Rayleigh-Taylor and Richtmyer-Meshkov Instabilities for Three Dimensional Axisymmetric flows in Cylindrical and Spherical Geometries. Los Alamos Laboratory, Report# LA-UR99-6796, 1999.

  37. E. Godlewski and P.A. Raviart. Hyperbolic Systems of Conservation Laws. Mathematiques et Applications, Ellipses Publ., Paris (1991).

  38. A. Harten. High resolution schemes for hyperbolic conservation laws J. Comput. Phys. 49, 1983, 357–393.

    Article  MathSciNet  MATH  Google Scholar 

  39. A. Harten, B. Engquist, S. Osher and S. R. Chakravarty. Uniformly high order accurate essentially non-oscillatory schemes, III. J. Comput. Phys. 71 (2), 1987, 231–303.

    Article  MathSciNet  MATH  Google Scholar 

  40. A. Hiltebrand and S. Mishra. Entropy stable shock capturing streamline diffusion space-time discontinuous Galerkin (DG) methods for systems of conservation laws. Num. Math., . 126 (1), 2014, 103-151.

    Article  MATH  Google Scholar 

  41. J. Jaffre, C. Johnson and A. Szepessy. Convergence of the discontinuous Galerkin finite element method for hyperbolic conservation laws. Math. Model. Meth. Appl. Sci., 5(3), 1995, 367–386.

    Article  MathSciNet  MATH  Google Scholar 

  42. G.-S. Jiang and C.-W. Shu. Efficient implementation of weighted ENO schemes. J. Comput. Phys., 126(1), 1996, 202–228.

    Article  MathSciNet  MATH  Google Scholar 

  43. C. Johnson and A. Szepessy. On the convergence of a finite element method for a nonlinear hyperbolic conservation law. Math. Comput., 49 (180), 1987, 427–444.

    Article  MathSciNet  MATH  Google Scholar 

  44. R. Käppeli, S. C. Whitehouse, S. Scheidegger, U.-L. Pen and M. Liebendörfer. FISH: A Three-dimensional Parallel Magnetohydrodynamics Code for Astrophysical Applications. The Astrophysical Journal Supplement, 2011, 195, 20.

    Article  Google Scholar 

  45. S. N. Kruzkhov. irst order quasilinear equations in several independent variables. USSR Math. Sbornik., 10 (2), 1970, 217–243.

    Article  Google Scholar 

  46. N.N. Kuznetsov. Accuracy of some approximate methods for computing the weak solutions of a first-order quasi-linear equation. USSR Comput. Math. and Math. Phys. 16 (1976), 105–119.

    Article  MATH  Google Scholar 

  47. P. D. Lax. Hyperbolic systems of conservation laws II. Comm. Pure Appl. Math. 10 (4), 1957, 537-566.

    Article  MathSciNet  MATH  Google Scholar 

  48. P. D. Lax. Mathematics and Physics. Bull. AMS 45(1), 2007, 135-152.

    Article  MathSciNet  MATH  Google Scholar 

  49. P. G. LeFloch, J. M. Mercier and C. Rohde. Fully discrete entropy conservative schemes of arbitrary order. SIAM J. Numer. Anal., 40 (5), 2002, 1968–1992.

    Article  MathSciNet  MATH  Google Scholar 

  50. R. J. LeVeque. Finite volume methods for hyperbolic problems. Cambridge university press, Cambridge, 2002.

    Book  MATH  Google Scholar 

  51. H. Lim, Y. Yu, J. Glimm, X. L. Li and D. H. Sharp. Chaos, transport and mesh convergence for fluid mixing. Act. Math. Appl. Sin., 24 (3), 2008, 355–368.

    Article  MathSciNet  MATH  Google Scholar 

  52. S. Mishra and C. Schwab. Sparse tensor multi-level Monte Carlo finite volume methods for hyperbolic conservation laws with random initial data. Math. Comput., 81(180), 2012, 1979–2018.

    Article  MATH  Google Scholar 

  53. S. Mishra, Ch. Schwab and J. Sukys. Multi-level Monte Carlo finite volume methods for nonlinear systems of conservation laws in multi-dimensions. J. Comput. Phys 231 (8), 2012, 3365–3388.

    Article  MathSciNet  MATH  Google Scholar 

  54. S. Mishra, Ch. Schwab and J. Sukys. Monte Carlo and multi-level Monte Carlo finite volume methods for uncertainty quantification in nonlinear systems of balance laws. Uncertainty Quantification in Computational Fluid Dynamics. Lecture Notes in Computational Science and Engineering Volume 92, 2013, 225–294

  55. S. Mishra, N.H. Risebro, Ch. Schwab and S. Tokareva Numerical solution of scalar conservation laws with random flux functions. Research report 2012-35, SAM ETH Zürich.

  56. J. Munkres. Algorithms for the Assignment and Transportation Problems Journal of the Society for Industrial and Applied Mathematics, 5 (1), 1957, 32–38.

    Article  MathSciNet  MATH  Google Scholar 

  57. B. Perthame and E. Tadmor. A kinetic equation with kinetic entropy functions for scalar conservation laws. Communications in Mathematical Physics 136, 1991, 501-517.

    Article  MathSciNet  MATH  Google Scholar 

  58. S. Schochet. Examples of measure-valued solutions Comm. Par. Diff. Eqns. 14 (5), 1989, 545–575.

    Article  MathSciNet  MATH  Google Scholar 

  59. M. Schonbeck. Convergence of solutions to nonlinear dispersion equations. Comm. Par. Diff. Eqns. 7 (8), 1982, 959–1000.

    Article  Google Scholar 

  60. E. Tadmor. The numerical viscosity of entropy stable schemes for systems of conservation laws, I. Math. Comp. 49, 1987, 91–103.

    Article  MathSciNet  MATH  Google Scholar 

  61. E. Tadmor. Entropy stability theory for difference approximations of nonlinear conservation laws and related time-dependent problems. Act. Numerica,, 2003, 451-512.

  62. E. Tadmor. Convergence of spectral methods for nonlinear conservation laws. SIAM Journal on Numerical Analysis 26 (1989), 30–44.

    Article  MathSciNet  MATH  Google Scholar 

  63. L. Tartar. Compensated compactness and applications to partial differential equations. Nonlinear analysis and mechanics: Heriot-Watt Symposium, Vol. IV, Pitman, 1979, 136–212.

  64. C. Villani. Topics in Optimal Transportation. American Mathematical Society, Graduate Studies in Mathematics, Vol. 58 (2013)

Download references

Acknowledgments

S.M. and R.K. were supported in part by ERC STG. No. 306279, SPARCCLE. E.T. was supported in part by NSF Grants DMS10-08397, RNMS11-07444 (KI-Net) and ONR Grant N00014-1512094. Many of the computations were performed at CSCS Lugano through Project s345. SM thanks Prof. Christoph Schwab (ETH Zurich) for several helpful comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Eitan Tadmor.

Additional information

Communicated by Wolfgang Dahmen.

Appendices

Appendix 1: Young Measures

We provide here a very short introduction to Young measures. The reader may wish to consult [7, 30] on the theory of Radon measures and probability measures and [2, 3] on the theory of Young measures.

1.1 Probability Measures

A.1.1 We denote by \(\mathcal {M}(\mathbb {R}^N)\) the set of finite Radon measures on \(\mathbb {R}^N\), which are inner regular Borel measures \(\mu \) with finite total variation \(|\mu |(\mathbb {R}^N)\). Let \(C_0(\mathbb {R}^N)\) be the space of continuous real-valued functions on \(\mathbb {R}^N\) which vanish at infinity, equipped with the supremum norm. Then it can be shown (see, e.g., [30, Section 7.3]) that \(\mathcal {M}(\mathbb {R}^N)\) can be identified with the dual space of \(C_0(\mathbb {R}^N)\) through the pairing \(\left\langle \mu , g\right\rangle = \int _{\mathbb {R}^N} g(\xi )\ \hbox {d}\mu (\xi )\). We do not distinguish between these two equivalent definitions of \(\mathcal {M}\). By a slight abuse of notation, we shall sometimes write \(\left\langle \mu , g(\xi )\right\rangle = \int _{\mathbb {R}^N}g(\xi )\ \hbox {d}\mu (\xi ).\) We will be particularly interested in the pairing \(\left\langle \mu , {{\mathrm{id}}}\right\rangle = \int _{\mathbb {R}^N}\xi \ \hbox {d}\mu (\xi )\) between \(\mu \) and the identity function \({{\mathrm{id}}}(\xi ) = \xi \).

A.1.2 The duality between \(C_0(\mathbb {R}^N)\) and \(\mathcal {M}(\mathbb {R}^N)\) induces a weak* topology on \(\mathcal {M}(\mathbb {R}^N)\), that of weak* convergence. A sequence \(\mu ^n\in \mathcal {M}(\mathbb {R}^N)\) converges weak* to \(\mu \in \mathcal {M}(\mathbb {R}^N)\) provided \(\left\langle \mu ^n, g\right\rangle \rightarrow \left\langle \mu , g\right\rangle \) for all \(g\in C_0(\mathbb {R}^N)\). (This is also called weak or vague convergence, see [7, 30].)

A.1.3 The set of probability measures on \(\mathbb {R}^N\) is the subset

$$\begin{aligned} \mathcal {P}(\mathbb {R}^N) := \left\{ \mu \in \mathcal {M}(\mathbb {R}^N)\ :\ \mu \geqslant 0,\ \mu (\mathbb {R}^N) = 1\right\} . \end{aligned}$$

Let \(\mathcal {P}^p(\mathbb {R}^N) \subset \mathcal {P}(\mathbb {R}^N)\) for \(p\in [1,\infty )\) denote the set of probability measures \(\mu \) such that \(\left\langle \mu , |\xi |^p\right\rangle < \infty \). For \(\mu ,\rho \in \mathcal {P}^p(\mathbb {R}^N)\) the Wasserstein metric \(W_p\) is defined as

$$\begin{aligned} W_p(\mu ,\rho ) := \inf \left\{ \int _{\mathbb {R}^N\times \mathbb {R}^N}|\xi -\zeta |^p\ \hbox {d}\pi (\xi ,\zeta )\ :\ \pi \in \varPi (\mu ,\rho )\right\} ^{1/p}, \end{aligned}$$

where \(\varPi (\mu ,\rho )\) is the set of probability measures on \(\mathbb {R}^N\times \mathbb {R}^N\) with marginals \(\mu \) and \(\rho \):

It can be shown that \(W_p\) for any p metrizes the topology of weak convergence on \(\mathcal {P}^p(\mathbb {R}^N)\) (see [1, Proposition 7.1.5] or [64, Chapter 7]).

A.1.4 Let \(\mu , \rho \in \mathcal {P}(\mathbb {R})\), and let \(F, G : \mathbb {R}\rightarrow [0,1]\) be their distribution functions,

$$\begin{aligned} F(x) := \mu ((-\infty ,x]), \quad G(y) := \rho ((-\infty ,y]). \end{aligned}$$

Then it can be shown that

$$\begin{aligned} W_p(\mu , \rho ) = \left( \int _0^1 \left| F^{-1}(s) - G^{-1}(s)\right| ^p\ \hbox {d}s\right) ^{1/p}, \end{aligned}$$

see [64, p. 75]. This gives rise to an efficient algorithm for computing the Wasserstein distance between discrete probability distributions. Let \(x_1, \ldots , x_n\) and \(y_1, \ldots , y_n\) be random numbers drawn from the probability distributions \(\mu \) and \(\rho \), respectively, and define the discrete distributions \(\mu _n := (\delta _{x_1} + \cdots + \delta _{x_n})/n\) and \(\rho _n := (\delta _{y_1} + \cdots + \delta _{y_n})/n\). By the law of large numbers, we have \(\mu _n \rightarrow \mu \) and \(\rho _n \rightarrow \rho \) weak* as \(n\rightarrow \infty \), almost surely. Moreover, their distribution functions are

$$\begin{aligned} F_n(x) = \frac{\#\{x_j\ :\ x_j \leqslant x\}}{n}, \quad G_n(y) = \frac{\#\{y_j\ :\ y_j \leqslant y\}}{n}. \end{aligned}$$

Hence, if the sequences \(x_j\) and \(y_j\) are sorted in increasing order, then

$$\begin{aligned} W_p(\mu _n, \rho _n)^p = \int _0^1 \left| F_n^{-1}(s) - G_n^{-1}(s)\right| ^p\ \hbox {d}s = \frac{1}{n}\sum _{j=1}^n |x_j-y_j|^p. \end{aligned}$$

The latter expression is very easy to implement on a computer.

The analogous problem when \(\mu ,\rho \in \mathcal {P}(\mathbb {R}^N)\) is more complex, but can be solved in \(O(n^3)\) time using the so-called Hungarian algorithm, see [56].

1.2 Young Measures

A.2.1 A Young measure from \(D\subset \mathbb {R}^k\) to \(\mathbb {R}^N\) is a function which maps \(z\in D\) to a probability measure on \(\mathbb {R}^N\). More precisely, a Young measure is a weak* measurable map \(\nu : D \rightarrow \mathcal {P}(\mathbb {R}^N)\), that is, the map** \(z\mapsto \left\langle \nu (z), g\right\rangle \) is Borel measurable for every \(g\in C_0(\mathbb {R}^N)\). We denote the image of \(z\in D\) under \(\nu \) by \(\nu _z:= \nu (z) \in \mathcal {P}(\mathbb {R}^N)\). The set of all Young measures from D into \(\mathbb {R}^N\) is denoted by \(\mathbf {Y}(D,\mathbb {R}^N)\). When \(N=1\) we write \(\mathbf {Y}(D) := \mathbf {Y}(D,\mathbb {R})\).

A.2.2 A Young measure \(\nu \in \mathbf {Y}(D,\mathbb {R}^N)\) is uniformly bounded if there is a compact set \(K\subset \mathbb {R}^N\) such that \({{\mathrm{supp}}}\nu _z\subset K\) for all \(z\in D\). Note that if \(\nu \) is atomic, \(\nu = \delta _u\), then \(\nu \) is uniformly bounded if and only if \(\Vert u\Vert _{L^\infty (D)} < \infty \).

A.2.3 If \(u:\mathbb {R}^k\rightarrow \mathbb {R}^N\) is any measurable function, then \(\nu _{z} := \delta _{u(z)}\) defines a Young measure, and we have \(u(z) = \left\langle \nu _z, {{\mathrm{id}}}\right\rangle \) for every \(z\). Conversely, we will say that a given Young measure \(\nu \) is atomic if it can be written as \(\nu = \delta _{u}\) for a measurable function u.

A.2.4 Two topologies on \(\mathbf {Y}(D,\mathbb {R}^N)\) arise naturally in the study of Young measures: those of weak* and strong convergence. A sequence \(\nu ^n \in \mathbf {Y}(D,\mathbb {R}^N)\) converges weak* to \(\nu \in \mathbf {Y}(D,\mathbb {R}^N)\) if \(\left\langle \nu ^n, g\right\rangle \overset{*}{\rightharpoonup }\left\langle \nu , g\right\rangle \) in \(L^\infty (D)\) for all \(g\in C_0(\mathbb {R}^N)\), that is,

$$\begin{aligned} \int _D \varphi (z) \left\langle \nu ^n_z, g\right\rangle \ \hbox {d}z \rightarrow \int _D \varphi (z) \left\langle \nu _z, g\right\rangle \ \hbox {d}z \quad \forall \ \varphi \in L^1(D). \end{aligned}$$

We say that \(\nu ^n\in \mathbf {Y}(D,\mathbb {R}^N)\) converges strongly to \(\nu \in \mathbf {Y}(D,\mathbb {R}^N)\) if

$$\begin{aligned} \bigl \Vert W_p(\nu ^n,\nu )\bigr \Vert _{L^p(D)} \rightarrow 0 \end{aligned}$$

for some \(p\in [1,\infty )\). If \(\nu \) is atomic, \(\nu = \delta _u\) for some \(u:D\rightarrow \mathbb {R}^N\), then \(\nu ^n \rightarrow \nu \) strongly if and only if

$$\begin{aligned} \int _D \int _{\mathbb {R}^N} |\xi - u(z)|^p\ \hbox {d}\nu ^n_z(\xi ) \hbox {d}z \rightarrow 0. \end{aligned}$$

A.2.5 The fundamental theorem of Young measures was first introduced by Tartar for \(L^\infty \)-bounded sequences [63] and then generalized by Schonbek [59] and Ball [3] for sequences of measurable functions. We provide a further generalization: Every sequence \(\nu ^n\in \mathbf {Y}(D,\mathbb {R}^N)\) which does not “leak mass at infinity” (condition (40)) has a weak* convergent subsequence:

Theorem 13

Let \(\nu ^n \in \mathbf {Y}(D,\mathbb {R}^N)\) for \(n\in \mathbb {N}\) be a sequence of Young measures. Then there exists a subsequence \(\nu ^m\) which converges weak* to a nonnegative measure-valued function \(\nu :D\rightarrow \mathcal {M}_+(\mathbb {R}^N)\) in the sense that

  1. (i)

    \(\left\langle \nu ^m_z, g\right\rangle \overset{*}{\rightharpoonup }\left\langle \nu , g\right\rangle \) in \(L^\infty (D)\) for all \(g\in C_0(\mathbb {R}^N)\),

and moreover satisfies

  1. (ii)

    \(\Vert \nu _z\Vert _{\mathcal {M}(\mathbb {R}^N)} \leqslant 1\) for a.e. \(z\in D\);

  2. (iii)

    If \(K\subset \mathbb {R}^N\) is closed and \({{\mathrm{supp}}}\nu ^n_z\subset K\) for a.e. \(z\in D\) and n large, then \({{\mathrm{supp}}}\nu _z\subset K\) for a.e. \(z\in D\).

Suppose further that for every bounded, measurable \(E \subset D\), there is a nonnegative \(\kappa \in C(\mathbb {R}^N)\) with \(\lim _{|\xi |\rightarrow \infty }\kappa (\xi )=\infty \) such that

$$\begin{aligned} \sup _n \int _E \left\langle \nu ^n_z, \kappa \right\rangle \ \hbox {d}z < \infty . \end{aligned}$$
(40)

Then

  1. (iv)

    \(\Vert \nu _z\Vert _{\mathcal {M}(\mathbb {R}^N)}=1\) for a.e. \(z\in D\),

whence \(\nu \in \mathbf {Y}(D,\mathbb {R}^N)\).

Proof

The proof is a generalization of Ball [3].

Denote by \(L_{w}^\infty (D;\mathcal {M}(\mathbb {R}^N))\) the set of weak* measurable functions \(\mu :D\rightarrow \mathcal {M}(\mathbb {R}^N)\), equipped with the norm

$$\begin{aligned} \Vert \mu \Vert _{\infty ,\mathcal {M}}:= \mathop {\hbox {ess sup}}\limits _{z\in D}\Vert \mu _z\Vert _\mathcal {M}. \end{aligned}$$

From the fact that \(C_0(\mathbb {R}^N)\) is separable, it can be shown (see [24, Theorem 8.18.2]) that \(L_{w}^\infty (D;\mathcal {M}(\mathbb {R}^N))\) is isometrically isomorphic to the dual of \(L^1(D;C_0(\mathbb {R}^N))\). The sequence \(\mu ^n\) is bounded in \(L_{w}^\infty (D;\mathcal {M}(\mathbb {R}^N))\) since \(\Vert \mu ^n\Vert _{\infty ,\mathcal {M}}\equiv 1\), and hence there is a \(\mu \in L_{w}^\infty (D;\mathcal {M}(\mathbb {R}^N))\) and a weak* convergent subsequence \(\mu ^m\) of \(\mu ^n\) such that \(\left\langle \mu ^m, \varPsi \right\rangle _{\infty ,\mathcal {M}}\rightarrow \left\langle \mu , \varPsi \right\rangle _{\infty ,\mathcal {M}}\), or equivalently,

$$\begin{aligned} \int _{D} \left\langle \mu ^m_z, \varPsi (z,\cdot )\right\rangle \hbox {d}z\rightarrow \int _{D} \left\langle \mu _z, \varPsi (z,\cdot )\right\rangle \hbox {d}z\quad \text {as }m\rightarrow \infty \end{aligned}$$

for all \(\varPsi \in L^1(D;C_0(\mathbb {R}^N))\). In particular, letting \(\varPsi (z,\xi )=\varphi (z)g(\xi )\) for \(\varphi \in L^1(D)\) and \(g\in C_0(\mathbb {R}^N)\), we obtain (i). We claim that \(\mu _z\geqslant 0\) for a.e. \(z\in D\). If not, then there would be a nonnegative \(\varPsi \in L^1(D;C_0(\mathbb {R}^N))\) such that \(\int _D\left\langle \mu _z, \varPsi (z,\cdot )\right\rangle \ \hbox {d}z< 0\). But then

$$\begin{aligned} 0 > \int _{D} \left\langle \mu _z, \varPsi (z,\cdot )\right\rangle \hbox {d}z= \lim _{m\rightarrow \infty } \int _D \left\langle \mu ^m_z, \varPsi (z,\cdot )\right\rangle \hbox {d}z\geqslant 0 \end{aligned}$$

(since \(\mu ^m_z\geqslant 0\) for all \(z\)), a contradiction.

(ii) follows from the weak* lower semi-continuity of the norm \(\Vert \cdot \Vert _{\infty ,\mathcal {M}}\). To see that (iii) holds, let \(g\in C_0(\mathbb {R}^N)\) be such that \(g\bigr |_{K} = 0\). Since \(\mu ^m \rightarrow K\) in measure, it follows that \(\left\langle \mu ^m, g\right\rangle \rightarrow 0\) in measure (that is, \(|\{z\in D:|\left\langle \mu ^m_z, g\right\rangle |>\delta \}| \rightarrow 0\) for all \(\delta >0\)). Hence,

$$\begin{aligned} \int _D \varphi (z)\left\langle \mu _z, g\right\rangle \hbox {d}z= \lim _m \int _D \varphi (z)\left\langle \mu ^m_z, g\right\rangle \hbox {d}z= 0 \end{aligned}$$

and therefore \(\left\langle \mu _z, g\right\rangle = 0\) for a.e.  \(z\in D\). This is precisely (ii).

Assume now that (40) holds. Fix a set \(E\subset D\) of finite, nonzero Lebesgue measure |E|, and denote the average integral over E as \(-\!\!\!\!\!\!\int _E = \frac{1}{|E|}\int _E\). For every \(R>0\), we define

$$\begin{aligned} \theta _R(\xi ) = {\left\{ \begin{array}{ll} 1 &{} \kappa (\xi ) \leqslant R \\ 1+R-\kappa (\xi ) &{} R<\kappa (\xi )\leqslant R+1 \\ 0 &{} R+1 < \kappa (\xi ). \end{array}\right. } \end{aligned}$$

Then \(\theta _R \in C_0(\mathbb {R}^N)\), so

$$\begin{aligned} \lim _m-\!\!\!\!\!\!\int _E \left\langle \mu ^m_z, \theta _R\right\rangle \hbox {d}z= -\!\!\!\!\!\!\int _E \left\langle \mu _z, \theta _R\right\rangle \hbox {d}z\leqslant -\!\!\!\!\!\!\int _E \Vert \mu _z\Vert _\mathbb {R}\hbox {d}z\leqslant 1, \end{aligned}$$

the last inequality following from the fact that \(\Vert \mu _z\Vert _\mathbb {R}\leqslant 1\) for all \(z\). Conversely,

$$\begin{aligned} 0 \leqslant -\!\!\!\!\!\!\int _E \left( 1-\left\langle \mu ^m_z, \theta _R\right\rangle \right) \ \hbox {d}z= -\!\!\!\!\!\!\int _E \left\langle \mu ^m_z, 1-\theta _R\right\rangle \ \hbox {d}z\leqslant \frac{1}{R}-\!\!\!\!\!\!\int _E \left\langle \mu ^m_z, \kappa \right\rangle \ \hbox {d}z, \end{aligned}$$

so (40) gives

$$\begin{aligned} 1&\leqslant \lim _{R\rightarrow \infty }\lim _m-\!\!\!\!\!\!\int _E \left\langle \mu ^m_z, \theta _R\right\rangle \hbox {d}z+ \lim _{R\rightarrow \infty } \sup _m \frac{1}{R}-\!\!\!\!\!\!\int _E \left\langle \mu ^m_z, \kappa \right\rangle \ \hbox {d}z\\&= \lim _{R\rightarrow \infty }-\!\!\!\!\!\!\int _E \left\langle \mu _z, \theta _R\right\rangle \hbox {d}z\\&\leqslant -\!\!\!\!\!\!\int _E \Vert \mu _z\Vert _{\mathcal {M}(\mathbb {R}^N)} \hbox {d}z\leqslant 1, \end{aligned}$$

whence \(-\!\!\!\!\!\!\int _E \Vert \mu _z\Vert _{\mathcal {M}(\mathbb {R}^N)}\ \hbox {d}z= 1\). Since \(E\subset D\) is arbitrary, (iv) follows. \(\square \)

A.2.6 An important special case of (40) is when \(\kappa (\xi ) = |\xi |^p\) for \(1\leqslant p < \infty \), which translates to the \(L^p\) bound

$$\begin{aligned} \sup _n\int _D \left\langle \mu ^n, |\xi |^p\right\rangle \ \hbox {d}z < \infty . \end{aligned}$$

The case \(p=\infty \) translates to the support of \(\nu ^n_z\) lying in a compact set \(K\subset \mathbb {R}^N\) for a.e. z and all n. Part (iii) of Theorem 13 then holds for all \(g\in C(\mathbb {R}^N)\), and condition (40) is automatically satisfied for any such \(\kappa \). The latter is the original form of the theorem given by Tartar [63].

1.3 Random Fields and Young Measures

A.3.1 If \((\varOmega ,\mathcal {F},P)\) is a probability space, \(D\subset \mathbb {R}^k\) is a Borel set and \(u : \varOmega \times D \rightarrow \mathbb {R}^N\) is a random field (i.e., a jointly measurable function), then we can define its law by

$$\begin{aligned} \nu _{z}(F) := P\left( u(z) \in F\right) = P\left( \left\{ \omega \ :\ u(\omega ,z) \in F\right\} \right) \end{aligned}$$
(41a)

for Borel subsets \(F\subset \mathbb {R}^N\) of phase space, or equivalently,

$$\begin{aligned} \left\langle \nu _z, g\right\rangle := \int _\varOmega g(u(\omega ,z))\ \hbox {d}P(\omega ) \end{aligned}$$
(41b)

for \(g\in C_0(\mathbb {R}^N)\). This defines a Young measure:

Proposition 1

If \(u:\varOmega \times D \rightarrow \mathbb {R}^N\) is jointly measurable, then (41) defines a Young measure from D to \(\mathbb {R}^N\).

Proof

First of all, for fixed \(z\in D\) the set \(\bigl \{\omega : u(\omega ,z) \in U\bigr \}\) is P-measurable for Borel sets U. Indeed, if \(w(\omega ) := u(\omega ,z)\) denotes the \(z\)-section of the measurable function \((\omega ,y) \mapsto u(\omega ,y)\), then \(\bigl \{\omega : u(\omega ,z) \in U\bigr \} = w^{-1}(U)\) is measurable.

We need to show that the definition of \(\nu \) is independent of the choice of map** in the equivalence classes of map**s from \(\varOmega \times D \rightarrow \mathbb {R}^N\). Let \(\hat{u}, \tilde{u}: \varOmega \times D \rightarrow \mathbb {R}^N\) be two map**s such that \(\hat{u}(\omega ,z) = \tilde{u}(\omega ,z)\) for \(P\times \lambda \)-a.e. \((\omega ,z)\). We apply Tonelli’s theorem to find that

$$\begin{aligned} 0 = \int _{\varOmega \times D} \mathbbm {1}_{\{\hat{u}\ne \tilde{u}\}}(\omega ,z)\ d(P\times \lambda )(\omega ,z) = \int _D P(\{\hat{u}(z) \ne \tilde{u}(z) \})\ \hbox {d}z. \end{aligned}$$

Hence, \(P(\hat{u}(z) \ne \tilde{u}(z)) = 0\) for a.e. \(z\in D\), so for every Borel set \(U\subset \mathbb {R}^N\),

$$\begin{aligned} P\left( \hat{u}(z) \in U\right) = P\left( \tilde{u}(z) \in U\right) \end{aligned}$$

for a.e. \(z\in D\).

Finally, \(\nu \) is weak* measurable since

$$\begin{aligned} \left\langle \nu _z, g\right\rangle = \int _{\mathbb {R}^N}g(\xi )\ \hbox {d}\nu _z(\xi ) = \int _{\varOmega }g(u(\omega ,z))\ \hbox {d}P(\omega ), \end{aligned}$$

which is measurable in \(z\) for any \(g\in C_0(\mathbb {R}^N)\). \(\square \)

A.3.2 It is well known that every measure on \(\mathbb {R}^N\) can be realized as the law of a random variable. Here we show that for every Young measure \(\nu \), there is always a random field with law \(\nu \).

Proposition 2

For every Young measure \(\nu \in \mathbf {Y}(D,\mathbb {R}^N)\), there exist a probability space \((\varOmega , \mathcal {F}, P)\) and a Borel measurable function \(u : \varOmega \times D \rightarrow \mathbb {R}^N\) such that u has law \(\nu \), i.e., for all Borel sets E,

$$\begin{aligned} \nu _z(E) = P(u(\omega ,z) \in E). \end{aligned}$$

In particular, we can choose \((\varOmega , \mathcal {F}, P)\) to be the Borel \(\sigma \)-algebra on \(\varOmega =[0,1)\) with Lebesgue measure.

Proof

The method of proof is standard, see, e.g., [6, Theorem 5.3].

We assume that \(N=1\). The generalization to \(N>1\) is straightforward but tedious. For \(n\in \mathbb {N}\) and \(j\in \mathbb {Z}\), we set

$$\begin{aligned} F_n^j := {\left\{ \begin{array}{ll} (-\infty , -2^{n}) &{} \text {if }j = -2^{2n} \\ {\big [2^{-n}(j-1), 2^{-n}j\big )} &{} \text {if } j = -2^{2n}+1, \ldots , 2^{2n} \\ {[2^n, \infty )} &{} \text {if } j = 2^{2n}+1\\ \emptyset &{} \text {otherwise.} \end{array}\right. } \end{aligned}$$

Let \(p_n^j(z) := \sum _{l\leqslant j} \nu _z(F_n^l)\). Note that \(p_n^j : \mathbb {R}\rightarrow [0,1]\) is measurable for all nj, and that \(0\leqslant p_n^{-j}\leqslant \dots \leqslant p_n^j=1\) for j large enough. Choose any \(\xi _n^j \in F_n^j\), and for \(\omega \in \varOmega :=[0,1)\), define

$$\begin{aligned} u_n(\omega ,z) := \xi _n^j \quad \text {for { j} such that } p_n^{j-1}(z) \leqslant \omega < p_n^{j}. \end{aligned}$$

We claim that \(u_n\) is measurable on the product \(\sigma \)-algebra between \(\mathcal {F}\) and the Borel \(\sigma \)-algebra on D. Each function \(u_n\) takes only finitely many values \(\xi _n^j\), so it suffices to show that \(u_n^{-1}(\{\xi _n^j\})\) is measurable for every \(\xi _n^j\). Indeed,

$$\begin{aligned} u_n^{-1}(\{\xi _n^j\})&= \Bigl \{(\omega ,z)\in \varOmega \times D\ :\ p_n^j(z) \leqslant \omega < p_n^{j+1}(z) \Bigr \} \\&= \Bigl (\varOmega \times D\Bigr ) \cap \Bigl \{(\omega ,z)\in \mathbb {R}\times D\ :\ p_n^j(z) \leqslant \omega \Bigr \} \\&\quad \cap \Bigl \{(\omega ,z)\in \mathbb {R}\times D\ :\ \omega < p_n^{j+1}(z) \Bigr \}, \end{aligned}$$

the intersection between the epigraph of \(p_n^j\) and the hypograph of \(p_n^{j+1}\), which are measurable by the measurability of the functions \(p_n^j\) and \(p_n^{j+1}\).

Because the partition \(\{F_m^j\}_{j\in \mathbb {Z}}\) is a refinement of \(\{F_n^j\}_{j\in \mathbb {Z}}\) whenever \(m>n\), it follows that \(|u_n(\omega ,z) - u_m(\omega ,z)| < \mathrm{diam}(F_n^j) = 2^{-n}\) for any \((\omega ,z)\) whenever mn are large enough. Hence, \(u_n\) converges pointwise to some function \(u : \varOmega \times D \rightarrow \mathbb {R}\), which is measurable by the measurability of each \(u_n\).

Finally, for every \(g\in C_0(\mathbb {R})\) and almost every \(z\in D\), we have by Lebesgue’s dominated convergence theorem

$$\begin{aligned} \int _\varOmega g(u(\omega ,z))\ \hbox {d}P(\omega )&= \lim _n \int _\varOmega g(u_n(\omega ,z))\ \hbox {d}P(\omega ) \\&= \lim _n \sum _j \nu _z(F_n^j)g(\xi _n^j) = \int _\mathbb {R}g(\xi )\ \hbox {d}\nu _z(\xi ). \end{aligned}$$

Hence, \(u(\cdot ,z)\) has law \(\nu _z\). \(\square \)

Appendix 2: Proof of Theorem 11

Proof

For any random field \(\zeta : \varOmega \rightarrow L^1(\mathbb {R}^d \times \mathbb {R}_+) \cap L^{\infty }(\mathbb {R}^d \times \mathbb {R}_+)\) on \((\varOmega ,\mathcal {F},P)\), we denote the expectation with respect to the probability measure P as

$$\begin{aligned} \mathbb {E}(\zeta ) := \int \limits _{\varOmega } \zeta (\omega ) \hbox {d}P(\omega ). \end{aligned}$$

For \(1 \leqslant k \leqslant M\), denote

$$\begin{aligned} \begin{aligned} G(\omega )&= \int _{\mathbb {R}_+}\int _{\mathbb {R}^d} \psi (x,t) g\left( u^{{\Delta x}}(\omega ;x,t)\right) \hbox {d}x \hbox {d}t, \\ G_k(\omega )&= \int _{\mathbb {R}_+}\int _{\mathbb {R}^d} \psi (x,t) g\left( u^{{\Delta x},k}(\omega ;x,t)\right) \hbox {d}x \hbox {d}t. \end{aligned} \end{aligned}$$
(42)

Henceforth, we suppress the \(\omega \)-dependence of G and \(G_k\) for notational convenience. The \(L^2(P)\) error in the approximation can be written as

$$\begin{aligned}&\mathbb {E}\left( \left( \mathbb {E}(G) - \frac{1}{M}\sum \limits _{k=1}^M G_k \right) ^2\right) = \mathbb {E}\left( \frac{1}{M^2}\left( \sum \limits _{k=1}^M (\mathbb {E}(G) - G_k) \right) ^2 \right) , \\&\quad = \mathbb {E}\left( \frac{1}{M^2} \left( \sum \limits _{k=1}^M \bigl (\mathbb {E}(G) - G_k\bigr )^2 + 2\sum _{k=1}^{M}\sum _{l \ne k} \bigl (\mathbb {E}(G) - G_k\bigr )\bigl (\mathbb {E}(G) - G_l\bigr )\right) \right) \\&\quad = \underbrace{\frac{1}{M^2}\sum \limits _{k=1}^M \mathbb {E}\left( \bigl (\mathbb {E}(G) - G_k\bigr )^2\right) }_{=:\ T_1} + \frac{2}{M^2} \sum _{k=1}^{M}\sum _{l \ne k}\underbrace{\mathbb {E}\Bigl ( \bigl (\mathbb {E}(G) - G_k\bigr ) \bigl (\mathbb {E}(G) - G_l\bigr )\Bigr )}_{=:\ T^{kl}_2}. \end{aligned}$$

As \(u^{{\Delta x},1}, \ldots , u^{{\Delta x},M}\) are independent and identically distributed, it follows from the definition of \(G_k\) that \(G_1,\ldots ,G_M\) are independent and identically distributed random variables. Hence, \(\mathbb {E}(G_k) = \mathbb {E}(G)\) and \(\mathbb {E}(G_kG_l) = \mathbb {E}(G_k)\mathbb {E}(G_l)\) for all kl. Consequently, a simple calculation shows that \(T^{kl}_2=0\) for all \(1 \leqslant k,l \leqslant M\) and \(k \ne l\).

The fact that \(G_1,\ldots ,G_M\) are independent and identically distributed yields

$$\begin{aligned} T_1 = \frac{1}{M} \left( \mathbb {E}(G^2) - \mathbb {E}(G)^2\right) . \end{aligned}$$

Hence,

In conclusion, the sample mean

$$\begin{aligned} \frac{1}{M} \sum _{k=1}^{M} \int _{\mathbb {R}_+}\int _{\mathbb {R}^d} \psi (x,t) g\left( u^{{\Delta x},k}(x,t)\right) \ \hbox {d}x\hbox {d}t \end{aligned}$$

converges to the corresponding ensemble average, \(\int _{\mathbb {R}_+}\int _{\mathbb {R}^d} \psi (x,t) \left\langle \nu ^{{\Delta x}}_{x,t}, g\right\rangle \ \hbox {d}x\hbox {d}t\) in \(L^2(\varOmega ;P)\), with a convergence rate of \(\frac{1}{\sqrt{M}}\). Taking a subsequence \(M'\rightarrow \infty \), the convergence also holds P-almost surely. \(\square \)

Appendix 3: Time Continuity of Approximations

From the time integration procedure (17b), we can show that the approximate MV solutions are time continuous. Consequently, the initial data is attained in a certain sense, and moreover, it is meaningful to evaluate the MV solution at a specific time t.

We state the theorem without proof, since the results are straightforward generalizations of “deterministic” counterparts.

Theorem 14

Let \(\psi \in C_c^1(\mathbb {R})\) and assume that (19a) and (19b) are satisfied. Let \(\nu ^{\Delta x}\) be generated by Algorithm 5. Then the functions

$$\begin{aligned} \varPsi ^{\Delta x}(t) := \int _\mathbb {R}\psi (x)\left\langle \nu ^{\Delta x}_{(x,t)}, {{\mathrm{id}}}\right\rangle \ \hbox {d}x \end{aligned}$$

and

$$\begin{aligned} \varPsi (t) := \int _\mathbb {R}\psi (x) \left\langle \nu _{(x,t)}, {{\mathrm{id}}}\right\rangle \ \hbox {d}x \end{aligned}$$

are Hölder continuous with exponent \(\gamma := \frac{r-1}{r}\) and with constant independent of \({\Delta x}\), and \(\varPsi ^{\Delta x}(t) \rightarrow \varPsi (t)\) as \({\Delta x}\rightarrow 0\) for a.e. \(t\in [0,T]\). Moreover,

$$\begin{aligned} \varPsi (0) = \lim _{t\rightarrow 0} \varPsi (t) = \int _\mathbb {R}\psi (x) \left\langle \sigma _{x}, {{\mathrm{id}}}\right\rangle \ \hbox {d}x. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fjordholm, U.S., Käppeli, R., Mishra, S. et al. Construction of Approximate Entropy Measure-Valued Solutions for Hyperbolic Systems of Conservation Laws. Found Comput Math 17, 763–827 (2017). https://doi.org/10.1007/s10208-015-9299-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10208-015-9299-z

Keywords

Mathematics Subject Classification

Navigation