Skip to main content
Log in

Globally Hyperbolic Moment Model of Arbitrary Order for One-Dimensional Special Relativistic Boltzmann Equation

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

This paper extends the model reduction method by the operator projection to the one-dimensional special relativistic Boltzmann equation. The derivation of arbitrary order globally hyperbolic moment system is built on our careful study of two families of the complicate Grad type orthogonal polynomials depending on a parameter. We derive their recurrence relations, calculate their derivatives with respect to the independent variable and parameter respectively, and study their zeros and coefficient matrices in the recurrence formulas. Some properties of the moment system are also proved. They include the eigenvalues and their bound as well as eigenvectors, hyperbolicity, characteristic fields, linear stability, and Lorentz covariance. A semi-implicit numerical scheme is presented to solve a Cauchy problem of our hyperbolic moment system in order to verify the convergence behavior of the moment method. The results show that the solutions of our hyperbolic moment system converge to the solution of the special relativistic Boltzmann equation as the order of the hyperbolic moment system increases.

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

Access this article

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

Similar content being viewed by others

References

  1. Anderson, J.L.: Relativistic Boltzmann theory and Grad method of moments. In: Carmeli, M., Fickler, S.I., Witten, L. (eds.) Relativity, pp. 109–124. Springer, Berlin (1970)

    Chapter  Google Scholar 

  2. Anderson, J.L.: Relativistic Grad polynomials. J. Math. Phys. 15, 1116–1119 (1974)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Anderson, J.L., Witting, H.R.: A relativistic relaxation-time model for the Boltzmann equation. Physica 74, 466–488 (1974)

    Article  ADS  Google Scholar 

  4. Au, J.D., Torrilhon, M., Weiss, W.: The shock tube study in extended thermodynamics. Phys. Fluids 13, 2423–2432 (2001)

    Article  ADS  MATH  Google Scholar 

  5. Cai, Z., Fan, Y., Li, R.: Globally hyperbolic regularization of Grad’s moment system in one dimensional space. Commun. Math. Sci. 11, 547–571 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cai, Z., Fan, Y., Li, R.: Globally hyperbolic regularization of Grad’s moment system. Commun. Pure Appl. Math. 67, 464–518 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  7. Cai, Z., Fan, Y., Li, R.: A framework on moment model reduction for kinetic equation. SIAM J. Appl. Math. 75, 2001–2023 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  8. Cai, Z., Li, R.: Numerical regularized moment method of arbitrary order for Boltzmann-BGK equation. SIAM J. Sci. Comput. 32, 2875–2907 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  9. Cai, Z., Li, R., Wang, Y.: Numerical regularized moment method for high Mach number flow. Commun. Comput. Phys. 11, 1415–1438 (2012)

    Article  MathSciNet  Google Scholar 

  10. Cercignani, C.: The Boltzmann Equation and Its Applications. Springer, Berlin (1988)

    Book  MATH  Google Scholar 

  11. Cercignani, C., Kremer, G.M.: The Relativistic Boltzmann Equation: Theory and Applications. Birkhauser, Berlin (2002)

    Book  MATH  Google Scholar 

  12. Chapman, S., Cowling, T.G.: The Mathematical Theory of Non-uniform Gases, 3rd edn. Cambridge University Press, Cambridge (1991)

    MATH  Google Scholar 

  13. Denicol, G.S., Niemi, H., Molnár, E., Rischke, D.H.: Derivation of transient relativistic fluid dynamics from the Boltzmann equation. Phys. Rev. D 85, 114047 (2012)

    Article  ADS  Google Scholar 

  14. Denicol, G.S., Kodama, T., Koide, T., Mota, P.: Stability and causality in relativistic dissipative hydrodynamics. J. Phys. G Nucl. Part. Phys. 35, 115102 (2008)

    Article  ADS  Google Scholar 

  15. Di, Y., Fan, Y., Li, R., Zheng, L.: Linear stability of hyperbolic moment models for Boltzmann equation. arXiv:1609.03669 (2016)

  16. Eckart, C.: The thermodynamics of irreversible processes. III. Relativistic theory of the simple fluid. Phys. Rev. 58, 919–924 (1940)

    Article  ADS  MATH  Google Scholar 

  17. Einstein, A.: Relativity: The Special and the General Theory. Three Rivers Press, New York (1995)

    MATH  Google Scholar 

  18. Fan, Y., Koellermeier, J., Li, J., Li, R., Torrilhon, M.: Model reduction of kinetic equations by operator projection. J. Stat. Phys. 162, 457–486 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Florkowski, W., Jaiswal, A., Maksymiuk, E., Ryblewski, R., Strickland, M.: Relativistic quantum transport coefficients for second-order viscous hydrodynamics. Phys. Rev. C 91, 054907 (2015)

    Article  ADS  Google Scholar 

  20. Garcia-Perciante, A.L., Sandoval-Villalbazob, A., Garcia-Colin, L.S.: Generalized relativistic Chapman–Enskog solution of the Boltzmann equation. Phys. A 21, 5073–5079 (2008)

    Article  MathSciNet  Google Scholar 

  21. Grad, H.: On the kinetic theory of rarefied gases. Commun. Pure Appl. Math. 2, 331–407 (1949)

    Article  MathSciNet  MATH  Google Scholar 

  22. Grad, H.: Note on \(N\)-dimensional Hermite polynomials. Commun. Pure Appl. Math. 2, 325–330 (1949)

    Article  MathSciNet  MATH  Google Scholar 

  23. Groot, S.R.D., Leeuwen, W.A.V., Weert, C.G.V.: Relativistic Kinetic Theory: Principles and Applications. North-Holland Press, New York (1980)

    Google Scholar 

  24. Hiscock, W.A., Lindblom, L.: Stability and causality in dissipative relativistic fluids. Ann. Phys. 151, 466–496 (1983)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Hiscock, W.A., Lindblom, L.: Generic instabilities in first-order dissipative relativistic fluid theories. Phys. Rev. D 31, 725–733 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  26. Hiscock, W.A., Lindblom, L.: Linear plane waves in dissipative relativistic fluids. Phys. Rev. D 35, 3723–3732 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  27. Hiscock, W.A., Lindblom, L.: Nonlinear pathologies in relativistic heat-conducting fluid theories. Phys. Lett. A 131, 509–513 (1988)

    Article  ADS  Google Scholar 

  28. Hiscock, W.A., Olson, T.S.: Effects of frame choice on nonlinear dynamics in relativistic heat-conducting fluid theories. Phys. Lett. A 141, 125–130 (1989)

    Article  ADS  Google Scholar 

  29. Israel, W.: Relativistic kinetic theory of a simple gas. J. Math. Phys. 4, 1163–1181 (1963)

    Article  ADS  MathSciNet  Google Scholar 

  30. Israel, W., Stewart, J.M.: Thermodynamics of nonstationary and transient effects in a relativistic gas. Phys. Lett. A 58, 213–215 (1976)

    Article  ADS  Google Scholar 

  31. Israel, W., Stewart, J.M.: Transient relativistic thermodynamics and kinetic theory. Ann. Phys. 118, 341–372 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  32. Israel, W., Stewart, J.M.: On transient relativistic thermodynamics and kinetic theory II. Proc. R. Soc. Lond. A 365, 43–52 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  33. Jaiswal, A.: Relativistic third-order dissipative fluid dynamics from kinetic theory. Phys. Rev. C 88, 021903(R) (2013)

    Article  ADS  Google Scholar 

  34. Jüttner, F.: Das Maxwellsche gesetz der geschwindigkeitsverteilung in der relativtheorie. Ann. Physik und Chemie 339, 856–882 (1911)

    Article  ADS  MATH  Google Scholar 

  35. Koellermeier, J., Torrilhon, M.: Hyperbolic moment equations using quadrature based projection methods. AIP Conf. Proc. 1628, 626–633 (2014)

    Article  ADS  MATH  Google Scholar 

  36. Koellermeier, J., Schaerer, R., Torrilhon, M.: A framework for hyperbolic approximation of kinetic equations using quadrature-based projection methods. Kinet. Relat. Models 7, 531–549 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  37. Kranyš, M.: Kinetic derivation of nonstationary general relativistic thermodynamics. Nuovo Cim. 8B, 417–441 (1972)

    Article  ADS  Google Scholar 

  38. Landau, L.D., Lifshitz, E.M.: Fluid Mechanics, 2nd edn. Pergamon Press, New York (1987)

    MATH  Google Scholar 

  39. Lichnerowicz, A., Marrot, R.: Propriétés statistiques des ensembles de particules en relativité restreite. C. R. Acad. Sci. Paris 210, 759–761 (1940)

    MathSciNet  MATH  Google Scholar 

  40. Marle, C.: Modèle cinétique pour l’établissement des lois de la conduction de la chaleur et de la viscosité en theorié de la relativité. C.R. Acad. Sci. Paris 260, 6539–6541 (1965)

    MathSciNet  MATH  Google Scholar 

  41. Mieussens, L.: Discrete velocity model and implicit scheme for the BGK equation of rarefied gas dynamics. Math. Models Methods Appl. Sci. 10, 1121–1149 (2000)

    MathSciNet  MATH  Google Scholar 

  42. Rhebergen, S., Bokhove, O., van der Vegt, J.J.W.: Discontinuous Galerkin finite element methods for hyperbolic nonconservative partial differential equations. J. Comput. Phys. 227, 1887–1922 (2008)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Shen, J., Tang, T., Wang, L.: Spectral Methods: Algorithms, Analysis and Applications. Springer, Berlin (2011)

    Book  MATH  Google Scholar 

  44. Stewart, J.M.: On transient relativistic thermodynamics and kinetic theory. Proc. R. Soc. Lond. A 357, 59–75 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  45. Struchtrup, H.: Projected moments in relativistic kinetic theory. Phys. A 253, 555–593 (1998)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work was partially supported by the Special Project on High-performance Computing under the National Key R&D Program (No. 2016YFB0200603), Science Challenge Project (No. JCKY2016212A502), and the National Natural Science Foundation of China (Nos. 91330205, 91630310, 11421101).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Huazhong Tang.

Appendices

Appendix 1: Proofs in Section 2

1.1 Proof of Theorem 2.1

Proof

For the nonnegative distribution f(xpt), which is not identically zero, using (2.3) gives

$$\begin{aligned} T^{\alpha \alpha }>0, \alpha =0,1;\quad T^{00}+T^{11}{\pm }2T^{01}={c}\int _{\mathbb {R}}\big (p^{0}{\pm } p^{1}\big )^2f\frac{dp}{p^{0}}>0, \end{aligned}$$

which implies the first inequality in (2.22).

Using the definition of \(\Delta ^{\alpha \beta }\) in (2.6) and the tensor decomposition of \(T^{\alpha \beta }\) in (2.5) gives (2.23), which is a quadratic equation with respect to u. The first inequality in (2.22) tells us that (2.23) has two different solutions whose product is equal to \({c^{2}}\), while one of them with a smaller absolute value is (2.24).

Using further (2.3) gives

$$\begin{aligned} {N^{0}-c^{-1}u N^{1}=c\int _{\mathbb {R}}\big (p^{0}-c^{-1}up^{1}\big )f\frac{dp}{p^{0}}}>0, \end{aligned}$$

i.e. the second inequality in (2.22), and then using the tensor decomposition of \(N^\alpha \) in (2.4) gives

$$\begin{aligned} \rho ={c^{-1}m\frac{N^{0}-c^{-1}u N^{1}}{\sqrt{1-c^{-2}u^2}}}>0. \end{aligned}$$

Moreover, using the second identity in (2.17) and (2.3) can give (2.26).

The inequality \({E\ge mc^{2}}\) holds because

$$\begin{aligned} E=U_{\alpha }p^{\alpha }={(1-c^{-2}u^2)^{-\frac{1}{2}}\left( c\sqrt{m^2c^2+p^2}-up\right) }>0, \end{aligned}$$

and

$$\begin{aligned} {E^2-m^2c^4=(1-c^{-2}u^2)^{-1}\left( u\sqrt{m^2c^2+p^2}-cp\right) ^2=\left( \frac{c^2}{U^{0}}p_{{\langle }1{\rangle }}\right) ^2}\ge 0. \end{aligned}$$

Thus, it holds

$$\begin{aligned} T^{00}-c^{-1}uT^{01}-c^{2}\rho= & {} \frac{T^{00}- 2c^{-1}u T^{01}+ c^{-2}u^2 T^{11}}{1-c^{-2}u^2} -c^{2}\rho \\= & {} c^{-1}\int _{\mathbb {R}}E\big (E-mc^2\big )f\frac{dp^1}{p^{0}}>0, \end{aligned}$$

which gives the third inequality in (2.22), and implies that \(G(\theta ^{-1})-\theta >1\) for \(\theta \in (0,+\infty )\).

On the other hand, one has

$$\begin{aligned} \lim _{\theta \rightarrow 0}\left( G(\theta ^{-1})-\theta \right) =1,\ \lim _{\theta \rightarrow +\infty }\left( G(\theta ^{-1})-\theta \right) =\lim _{\theta \rightarrow +\infty }\theta = +\infty , \end{aligned}$$

and

$$\begin{aligned} \frac{\partial (G(\theta ^{-1})-\theta )}{\partial \theta } =&-\theta ^{-2} \left( G(\theta ^{-1})^2-3G(\theta ^{-1})\theta +\theta ^{2}-1\right) =:\tilde{\psi }(G(\theta ^{-1}),\theta ). \end{aligned}$$

Because

$$\begin{aligned} 0&<c^{-1}\int _{\mathbb {R}}(E-mc^2)f^{(0)}\frac{dp}{p^{0}}=-m^{-1}\rho (G(\theta ^{-1})-2\theta -1),\\ 0&<c^{-1}\int _{\mathbb {R}}(E-mc^2)^2f^{(0)}\frac{dp}{p^{0}}=\rho c^{2}(2G(\theta ^{-1}) - 3\theta -2),\\ 0&<c^{-1}\int _{\mathbb {R}}(E-mc^2)^3f^{(0)}\frac{dp}{p^{0}}=-\rho mc^{4}((4 - \theta )G(\theta ^{-1}) - 5\theta -4), \end{aligned}$$

one obtains

$$\begin{aligned} \frac{3}{2}\theta +1< G(\theta ^{-1})< {\left\{ \begin{array}{ll} \min \left\{ 2\theta +1,(4-\theta )^{-1}(5\theta +4)\right\} ,&{} 0<\theta <4,\\ 2\theta +1,&{} \theta \ge 4, \end{array}\right. } \end{aligned}$$

which is equivalent to the following inequality

$$\begin{aligned} \frac{3}{2}\theta +1&<G(\theta ^{-1})< {\left\{ \begin{array}{ll} (4-\theta )^{-1}(5\theta +4), &{} 0<\theta <1,\\ 2\theta +1, &{} \theta \ge 1. \end{array}\right. } \end{aligned}$$

Thus, one has

$$\begin{aligned} \tilde{\psi }(G(\theta ^{-1}),\theta )> {\left\{ \begin{array}{ll} \tilde{\psi }\left( 2\theta +1,\theta \right)>\theta ^{3}(\theta -1)>0, &{}\theta \ge 1,\\ \\ \tilde{\psi }\left( (4-\theta )^{-1}(5\theta +4),\theta \right)>(4-\theta )^{-2}\theta ^4(\theta +8)(1-\theta )>0, &{}\theta <1, \end{array}\right. } \end{aligned}$$

i.e.

$$\begin{aligned} \frac{\partial (G(\theta ^{-1})-\theta )}{\partial \theta }>0, \end{aligned}$$

which implies that \(G(\theta ^{-1})-\theta \) is a strictly monotonic function of \(\theta \) in the interval \((0,+\infty )\).

Thus (2.26) has a unique solution in the interval \((0,+\infty )\). The proof is completed. \(\square \)

1.2 Proof of Theorem 2.2

Proof

Under Theorem 2.1, for the nonnegative distribution f(xpt), which is not identically zero, one obtains \(\{\rho ,u,\theta \}\) satisfying

$$\begin{aligned} \rho>0, \quad |u|<{c}, \quad \theta >0. \end{aligned}$$
(8.1)

Due to the last equations in (2.7) and (2.20), one obtains

$$\begin{aligned} {\Pi =-\int _{\mathbb {R}}\Delta _{\alpha \beta }p^{\alpha }p^{\beta }f\frac{dp}{p^{0}}-c^{2}\rho \theta =c^{-1}\int _{\mathbb {R}}\big (E^2-m^2c^4\big )f\frac{dp}{p^{0}}-\rho c^{2}\theta >-\rho c^{2}\theta }, \end{aligned}$$

which completes the proof. \(\square \)

Appendix 2: Proofs in Section 3

1.1 Proof of Theorem 3.2

Proof

  1. (i)

    For \(k\le n+2\), taking the inner product with respect to \(\omega ^{(0)}\) between the polynomials \(P_{k}^{(0)}(x;\zeta )\) and \((x^2-1)P_{n}^{(1)}(x;\zeta )\) gives

    $$\begin{aligned}&\left( (x^2-1)P_{n}^{(1)},P_{n+2}^{(0)}\right) _{\omega ^{(0)}}=\left( c_{n}^{(1)}x^{n+2},P_{n+2}^{(0)}\right) _{\omega ^{(0)}}= \frac{c_{n}^{(1)}}{c_{n+2}^{(0)}}\left( P_{n+2}^{(0)},P_{n+2}^{(0)}\right) _{\omega ^{(0)}}=r_{n+1},\\&\left( (x^2-1)P_{n}^{(1)},P_{n+1}^{(0)}\right) _{\omega ^{(0)}}\\&\quad = \left( c_{n}^{(1)}\left( x^{n+2}-\sum _{i=1}^{n+2}x_{i,n+2}^{(0)}x^{n+1}+ \left( \sum _{i=1}^{n+2}x_{i,n+2}^{(0)}-\sum _{i=1}^{n}x_{i,n}^{(1)}\right) x^{n+1}\right) ,P_{n+1}^{(0)}\right) _{\omega ^{(0)}}\\&\quad =r_{n+1}\left( P_{n+2}^{(0)},P_{n+1}^{(0)}\right) _{\omega ^{(0)}}+ q_{n}\left( P_{n+1}^{(0)},P_{n+1}^{(0)}\right) _{\omega ^{(0)}}=q_{n},\\&\left( (x^2-1)P_{n}^{(1)},P_{n+1}^{(0)}\right) _{\omega ^{(0)}}\\&\quad =\left( P_{n}^{(1)},c_{n+1}^{(0)} \left( x^{n+1}-\sum _{i=1}^{n+1}x_{i,n+1}^{(1)}x^{n}+ \left( \sum _{i=1}^{n+1}x_{i,n+1}^{(1)}-\sum _{i=1}^{n+1}x_{i,n+1}^{(0)}\right) x^{n}\right) \right) _{\omega ^{(1)}}\\&\quad =p_{n+1}\left( P_{n}^{(1)},P_{n+1}^{(1)}\right) _{\omega ^{(1)}}+\frac{c_{n+1}^{(0)}}{c_{n}^{(1)}}\left( \sum _{i=1}^{n+1}x_{i,n+1}^{(1)}- \sum _{i=1}^{n+1}x_{i,n+1}^{(0)}\right) \left( P_{n}^{(1)},P_{n}^{(1)}\right) _{\omega ^{(1)}}\\&\quad =\frac{c_{n+1}^{(0)}}{c_{n}^{(1)}} \sum _{i=1}^{n+1} \left( x_{i,n+1}^{(1)}-x_{i,n+1}^{(0)}\right) =q_{n}, \\&\left( (x^2-1)P_{n}^{(1)},P_{n}^{(0)}\right) _{\omega ^{(0)}} =\left( P_{n}^{(1)},P_{n}^{(0)}\right) _{\omega ^{(1)}}= \left( P_{n}^{(1)},c_{n}^{(0)}x^{n}\right) _{\omega ^{(1)}}\\&\qquad \qquad \qquad \qquad \qquad \qquad \quad =p_{n}\left( P_{n}^{(1)},P_{n}^{(1)}\right) _{\omega ^{(1)}}= p_{n},\\&\left( (x^2-1)P_{n}^{(1)},P_{k}^{(0)}\right) _{\omega ^{(0)}} =\left( P_{n}^{(1)},P_{k}^{(0)}\right) _{\omega ^{(1)}}=0, \quad k\le n-1, \end{aligned}$$

    Substituting them into (3.3) gives (3.9).

  2. (ii)

    Taking the inner product with respect to \(\omega ^{(1)}\) between \(P_{n+1}^{(0)} (x;\zeta )\) and \(P_{k}^{(1)} (x;\zeta )\) with \(k\le n+1\)

    $$\begin{aligned} \left( P_{n+1}^{(0)},P_{n+1}^{(1)}\right) _{\omega ^{(1)}}&= \left( c_{n+1}^{(0)}x^{n+1},P_{n+1}^{(1)}\right) _{\omega ^{(1)}} =p_{n+1}\left( P_{n+1}^{(1)},P_{n+1}^{(1)}\right) _{\omega ^{(1)}}=p_{n+1},\\ \left( P_{n+1}^{(0)},P_{n}^{(1)}\right) _{\omega ^{(1)}}&= \left( P_{n+1}^{(0)},(x^2-1)P_{n}^{(1)}\right) _{\omega ^{(0)}}=q_{n}, \\ \left( P_{n+1}^{(0)},P_{n-1}^{(1)}\right) _{\omega ^{(1)}}&=\left( P_{n+1}^{(0)},(x^2-1)P_{n-1}^{(1)}\right) _{\omega ^{(0)}}= r_{n}\left( P_{n+1}^{(0)},P_{n+1}^{(0)}\right) _{\omega ^{(0)}}=r_{n},\\ \left( P_{n+1}^{(0)},P_{k}^{(1)}\right) _{\omega ^{(1)}}&=\left( P_{n+1}^{(0)},(x^2-1)P_{k}^{(1)}\right) _{\omega ^{(0)}}=0, \quad k\le n-2. \end{aligned}$$

    Similarly, substituting them into (3.3) gives (3.10).

  3. (iii)

    If using (3.6) to eliminate \(P_{n+2}^{(0)}\) and \(P_{n+1}^{(1)}\) in (3.9) and (3.10) respectively, then one obtains

    $$\begin{aligned} (x^2-1)P_{n}^{(1)}=\tilde{p}_{n}(x+\tilde{q}_{n})P_{n+1}^{(0)}+\tilde{r}_{n}P_{n}^{(0)},\quad P_{n+1}^{(0)}=\frac{1}{\tilde{\tilde{p}}_{n}}\big (x-\tilde{\tilde{q}}_{n}\big )P_{n}^{(1)}-\frac{a_{n-1}^{(1)}}{a_{n}^{(0)}}\tilde{\tilde{r}}_{n}P_{n-1}^{(1)}, \end{aligned}$$

    with

    $$\begin{aligned} \tilde{p}_{n}= & {} \frac{r_{n+1}}{a_{n+1}^{(0)}}=\frac{c_{n}^{(1)}}{c_{n+1}^{(0)}}=\frac{a_{n}^{(1)}}{p_{n+1}}=\tilde{\tilde{p}}_{n},\\ \tilde{q}_{n}= & {} \frac{1}{\tilde{p}_{n}}q_{n}-b_{n+1}^{(0)}=\sum _{i=1}^{n+1}x_{i,n+1}^{(0)}-\sum _{i=1}^{n}x_{i,n}^{(1)}=b_{n}^{(1)}-\tilde{p}_{n}q_{n}=\tilde{\tilde{q}}_{n},\\ \tilde{r}_{n}= & {} p_{n}-\tilde{p}_{n}a_{n}^{(0)}={p_{n}(1-\tilde{p}_{n}^{2})} =\frac{a_{n}^{(0)}}{a_{n-1}^{(1)}}\left( -r_{n}+\frac{1}{\tilde{p}_{n}}a_{n-1}^{(1)}\right) = \tilde{\tilde{r}}_{n}. \end{aligned}$$

    The proof is completed.

\(\square \)

1.2 Proof of Theorem 3.3

Proof

With the aid of definition and recurrence relation of the second kind modified Bessel function in (2.18) and (2.19), one has

$$\begin{aligned} \frac{\partial }{\partial \zeta }\omega ^{(\ell )}(x;\zeta )&=\frac{K_{2}(\zeta )+K_{0}(\zeta )-2xK_{1}(\zeta )}{2K_{1}(\zeta )}\left( \frac{1}{K_{1}(\zeta )}(x^2-1)^{\ell -\frac{1}{2}}\exp (-\zeta x)\right) \\&=\left( G(\zeta )-\zeta ^{-1}-x\right) \omega ^{(\ell )}(x;\zeta ). \end{aligned}$$

Taking the partial derivative of both sides of identities

$$\begin{aligned} \left( P_{n+1}^{(\ell )},P_{k}^{(\ell )}\right) _{\omega ^{(\ell )}}=\delta _{n+1,k}, k=0,\ldots ,n+1, \end{aligned}$$

with respect to \(\zeta \) and using (3.8) gives

$$\begin{aligned} \frac{\partial }{\partial \zeta } \left( P_{n+1}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}} =\,&2\left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}} +\left( G(\zeta )-\zeta ^{-1}\right) \left( P_{n+1}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}}\\&-\,\left( xP_{n+1}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}}\\ =\,&2\left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}} +\left( G(\zeta )-\zeta ^{-1}-b_{n+1}^{(\ell )}\right) =0,\\ \frac{\partial }{\partial \zeta } \left( P_{n+1}^{(\ell )},P_{n}^{(\ell )}\right) _{\omega ^{(\ell )}} =\,&\left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{n}^{(\ell )}\right) _{\omega ^{(\ell )}} +\left( P_{n+1}^{(\ell )},\frac{\partial }{\partial \zeta }P_{n}^{(\ell )}\right) _{\omega ^{(\ell )}}\\&+\,\left( G(\zeta )-\zeta ^{-1}\right) \left( P_{n+1}^{(\ell )},P_{n}^{(\ell )}\right) _{\omega ^{(\ell )}}-\left( xP_{n}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}}\\ =\,&\left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{n}^{(\ell )}\right) _{\omega ^{(\ell )}} -a_{n}^{(\ell )}=0,\\ \frac{\partial }{\partial \zeta } \left( P_{n+1}^{(\ell )},P_{k}^{(\ell )}\right) _{\omega ^{(\ell )}} =\,&\left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{k}^{(\ell )}\right) _{\omega ^{(\ell )}} +\left( P_{n+1}^{(\ell )},\frac{\partial }{\partial \zeta }P_{k}^{(\ell )}\right) _{\omega ^{(\ell )}}\\&\,+\left( G(\zeta )-\zeta ^{-1}\right) \left( P_{n+1}^{(\ell )},P_{k}^{(\ell )}\right) _{\omega ^{(\ell )}}-\left( xP_{k}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}}\\ =\,&\left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{k}^{(\ell )}\right) _{\omega ^{(\ell )}} =\,0,\quad k\le n-1. \end{aligned}$$

Thus one has

$$\begin{aligned} \left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{n+1}^{(\ell )}\right) _{\omega ^{(\ell )}}&=-\frac{1}{2}\left( G(\zeta )-\zeta ^{-1}-b_{n+1}^{(\ell )}\right) ,\\ \left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{n}^{(\ell )}\right) _{\omega ^{(\ell )}}&= a_{n}^{(\ell )},\quad \left( \frac{\partial }{\partial \zeta }P_{n+1}^{(\ell )},P_{k}^{(\ell )}\right) _{\omega ^{(\ell )}} =0,\quad k\le n-1. \end{aligned}$$

Because \(\frac{\partial P_{n+1}^{(\ell )}}{\partial \zeta }\) is a polynomial and its degree is not larger than \( n+1\), using (3.3) gives (3.17). The proof is completed. \(\square \)

1.3 Proof of Theorem 3.4

Proof

Similar to the proof of Theorem 3.3, one has

$$\begin{aligned} {\frac{\partial }{\partial x}\omega ^{(1)}(x;\zeta ) =x\omega ^{(0)}(x;\zeta )-\zeta \omega ^{(1)}(x;\zeta ).} \end{aligned}$$

Because the degrees of polynomials \(\frac{\partial P_{n+1}^{(0)}}{\partial x}\) and \((x^2-1)\frac{\partial P_{n}^{(1)}}{\partial x}+xP_{n}^{(1)}\) are not larger than n and \(n+1\), respectively, and

$$\begin{aligned}&\lim _{x\rightarrow +\infty }P_{i}^{(0)}(x;\zeta )P_{j}^{(1)}(x;\zeta )\omega ^{(1)}(x;\zeta )=0,\\&\quad \lim _{x\rightarrow 1}P_{i}^{(0)}(x;\zeta )P_{j}^{(1)}(x;\zeta )\omega ^{(1)}(x;\zeta )=0,\quad \forall i,j\in \mathbb {N}, \end{aligned}$$

one can calculate the expansion coefficients in (3.3) as follows

$$\begin{aligned} \left( \frac{\partial }{\partial x}P_{n+1}^{(0)},P_{n}^{(1)}\right) _{\omega ^{(1)}}=&\left( (n+1)c_{n+1}^{(0)}x^{n},P_{n}^{(1)}\right) _{\omega ^{(1)}}= \frac{n+1}{\tilde{p}_{n}}\left( P_{n}^{(1)},P_{n}^{(1)}\right) _{\omega ^{(1)}} =\frac{n+1}{\tilde{p}_{n}},\\ \left( \frac{\partial }{\partial x}P_{n+1}^{(0)},P_{n-1}^{(1)}\right) _{\omega ^{(1)}}=&\int _{1}^{+\infty }\frac{\partial }{\partial x}\left( P_{n+1}^{(0)}P_{n-1}^{(1)}\omega ^{(1)}\right) dx -\left( P_{n+1}^{(0)},(x^2-1)\frac{\partial }{\partial x}P_{n-1}^{(1)}\right) _{\omega ^{(0)}}\\&-\left( P_{n+1}^{(0)},xP_{n-1}^{(1)}\right) _{\omega ^{(0)}}+\zeta \left( P_{n+1}^{(0)},(x^2-1)P_{n-1}^{(1)}\right) _{\omega ^{(0)}} =\zeta r_{n},\\ \left( \frac{\partial }{\partial x}P_{n+1}^{(0)},P_{k}^{(1)}\right) _{\omega ^{(1)}}=&\int _{1}^{+\infty }\frac{\partial }{\partial x}\left( P_{n+1}^{(0)}P_{k}^{(1)}\omega ^{(1)}\right) dx -\left( P_{n+1}^{(0)},(x^2-1)\frac{\partial }{\partial x}P_{k}^{(1)}\right) _{\omega ^{(0)}}\\&-\left( P_{n+1}^{(0)},xP_{k}^{(1)}\right) _{\omega ^{(0)}} +\zeta \left( P_{n+1}^{(0)},(x^2-1)P_{k}^{(1)}\right) _{\omega ^{(0)}} \!=0,k{\le } n{-}2{,} \end{aligned}$$

and

$$\begin{aligned} \left( (x^2-1)\frac{\partial }{\partial x}P_{n}^{(1)}+xP_{n}^{(1)},P_{n+1}^{(0)}\right) _{\omega ^{(0)}}=&\left( (n+1)c_{n}^{(1)}x^{n+1},P_{n+1}^{(0)}\right) _{\omega ^{(0)}}\\ =&(n+1)\tilde{p}_{n}\left( P_{n+1}^{(0)},P_{n+1}^{(0)}\right) _{\omega ^{(0)}} =(n+1)\tilde{p}_{n},\\ \left( (x^2{-}1)\frac{\partial }{\partial x}P_{n}^{(1)}{+}xP_{n}^{(1)},P_{n}^{(0)}\right) _{\omega ^{(0)}}=&\int _{1}^{+\infty }\frac{\partial }{\partial x}\left( P_{n}^{(1)}P_{n}^{(0)}\omega ^{(1)}\right) dx{-}\left( \!P_{n}^{(1)},\frac{\partial }{\partial x}P_{n}^{(0)}\!\right) _{\omega ^{(1)}}\\&+\zeta \left( P_{n}^{(1)},P_{n}^{(0)}\right) _{\omega ^{(1)}}=\zeta \left( P_{n}^{(1)},P_{n}^{(0)}\right) _{\omega ^{(1)}}=\zeta p_{n},\\ \left( \!(x^2{-}1)\frac{\partial }{\partial x}P_{n}^{(1)}{+}xP_{n}^{(1)},P_{k}^{(0)}\right) _{\omega ^{(0)}}=&\int _{1}^{+\infty }\!\!\frac{\partial }{\partial x}\left( P_{n}^{(1)}P_{k}^{(0)}\omega ^{(1)}\right) dx{-}\left( P_{n}^{(1)},\frac{\partial }{\partial x}P_{k}^{(0)}\right) _{\omega ^{(1)}}\\&+\zeta \left( P_{n}^{(1)},P_{k}^{(0)}\right) _{\omega ^{(1)}}=0,k\le n-1.\\ \end{aligned}$$

The proof is completed. \(\square \)

1.4 Proof of Theorem 3.6

Proof

Substituting \(\{x_{i,n+1}^{(0)}\}_{i=1}^{n+1}\) into (3.14) gives

$$\begin{aligned} \left( (x_{i,n+1}^{(0)})^2-1\right) P_{n}^{(1)}(x_{i,n+1}^{(0)};\zeta ) =\tilde{r}_{n}P_{n}^{(0)}(x_{i,n+1}^{(0)};\zeta ){,} \end{aligned}$$

which implies that \(\tilde{r}_{n}\ne 0\). In fact, if assuming \(\tilde{r}_{n}=0\), then the above identity and the fact that \((x_{i,n+1}^{(0)})^2-1>0\) imply \(P_{n}^{(1)}(x_{i,n+1}^{(0)};\zeta )=0\), which contradicts with \(P_{n}^{(1)}\) being a polynomial of degree n.

Using Theorem 3.5 gives

$$\begin{aligned} \mathrm{sign}\left( P_{n}^{(1)}(x_{i,n+1}^{(0)};\zeta )P_{n}^{(1)}(x_{i+1,n+1}^{(0)};\zeta )\right) =\mathrm{sign}\left( P_{n}^{(0)}(x_{i,n+1}^{(0)};\zeta )P_{n}^{(0)}(x_{i+1,n+1}^{(0)};\zeta )\right) <0. \end{aligned}$$

Thus there exists at least one zero of the polynomial \(P_{n}^{(1)}\) in each subinterval \(\left( x_{i,n+1}^{(0)},x_{i+1,n+1}^{(0)}\right) \). The proof is completed. \(\square \)

1.5 Proof of Corollary 1

Proof

It is obvious that

$$\begin{aligned} p_{n}=\frac{c_{n}^{(0)}}{c_{n}^{(1)}}>0,\quad r_{n}=\frac{c_{n-1}^{(1)}}{c_{n+1}^{(0)}}>0,\quad \tilde{p}_{n}=\frac{c_{n}^{(1)}}{c_{n+1}^{(0)}}>0. \end{aligned}$$

Using Theorems 3.1 and 3.6 gives

$$\begin{aligned} \tilde{q}_{n}&=\sum _{i=1}^{n+1}x_{i,n+1}^{(0)}-\sum _{i=1}^{n}x_{i,n}^{(1)}=\sum _{i=1}^{n}\left( x_{i+1,n+1}^{(0)}-x_{i,n}^{(1)}\right) +x_{1,n+1}^{(0)}>0,\\ q_{n}&=\tilde{p}_{n}\left( \sum _{i=1}^{n+2}x_{i,n+2}^{(0)}-\sum _{i=1}^{n}x_{i,n}^{(1)}\right) =\tilde{p}_{n}\left( b_{n+1}^{(0)}+\tilde{q}_{n}\right) >0, \end{aligned}$$

which imply \(q_{n}>0\) and \(\tilde{q}_{n}>0\).

Comparing the coefficients of the nth order terms at two sides of (3.14) gives

$$\begin{aligned} \tilde{r}_{n}&=p_{n}^{-1} \left( \sum _{i=1}^{n}\sum _{j=i+1}^{n}x_{i,n}^{(1)}x_{j,n}^{(1)}-1- \sum _{i=1}^{n+1}\sum _{j=i+1}^{n+1}x_{i,n+1}^{(0)}x_{j,n+1}^{(0)}\right. \\&\quad + \left. \left( \sum _{i=1}^{n+1}x_{i,n+1}^{(0)}-\sum _{i=1}^{n}x_{i,n}^{(1)}\right) \sum _{i=1}^{n+1}x_{i,n+1}^{(0)}\right) \\&= p_{n}^{-1} \left( \sum _{i=1}^{n}\sum _{j=i+1}^{n}x_{i,n}^{(1)}x_{j,n}^{(1)}+\sum _{i=1}^{n+1}\sum _{j=i}^{n+1}x_{i,n+1}^{(0)}x_{j,n+1}^{(0)} -\sum _{i=1}^{n}x_{i,n}^{(1)}\sum _{i=1}^{n+1}x_{i,n+1}^{(0)}-1\right) \\&=p_{n}^{-1} \left( \sum _{i=1}^{n}x_{i+1,n+1}^{(0)}(x_{i+1,n+1}^{(0)}-x_{i,n}^{(1)})+(x_{1,n+1}^{(0)})^{2}-1\right. \\&\left. \quad +\sum _{i=0}^{n}\sum _{j=i+1}^{n}(x_{i+1,n+1}^{(0)}-x_{i,n}^{(1)})(x_{j+1,n+1}^{(0)}-x_{j,n}^{(1)})\right) , \end{aligned}$$

where \(x_{0,n}^{(1)}=0\).

Combining it with Theorem 3.6 gives \(\tilde{r}_{n}>0\). The proof is completed. \(\square \)

1.6 Proof of Corollary 3

Proof

Taking partial derivative of \(P_{n}^{(\ell )}(x_{i,n}^{(\ell )};\zeta )\) with respect to \(\zeta \) and using Theorem 3.3 gives

$$\begin{aligned}&\frac{\partial x_{i,n}^{(\ell )}}{\partial \zeta }=-\left( \frac{\partial P_{n}^{(\ell )}}{\partial x}(x_{i,n}^{(\ell )};\zeta )\right) ^{-1} \left( \frac{\partial P_{n}^{(\ell )}}{\partial \zeta }(x_{i,n}^{(\ell )};\zeta )\right) \\&=-a_{n-1}^{(\ell )}\left( \frac{\partial P_{n}^{(\ell )}}{\partial x}(x_{i,n}^{(\ell )};\zeta )\right) ^{-1}P_{n-1}^{(\ell )}(x_{i,n}^{(\ell )};\zeta ). \end{aligned}$$

Due to Theorem 3.5, one has

$$\begin{aligned} \mathrm{sign}(P_{n-1}^{(\ell )}(x_{i,n}^{(\ell )};\zeta ))=(-1)^{n+i}=\mathrm{sign}\left( \frac{\partial P_{n}^{(\ell )}}{\partial x}(x_{i,n}^{(\ell )};\zeta )\right) . \end{aligned}$$

Combining them completes the proof. \(\square \)

1.7 Proof of Lemma 1

Proof

According to the definition of \(Q_{2n}(x;\zeta )\) in (3.30), it is not difficult to know that \(Q_{2n}(x;\zeta )\) is an even function and a polynomial of degree 2n.

If taking x in (3.30) as the zero of \(P_{n+1}^{(0)}(x;\zeta )\), i.e. \(x=x_{i,n+1}^{(0)}\), \(i=1,\ldots ,{n+1}\), then one has

$$\begin{aligned} Q_{2n}(x_{i,n+1}^{(0)};\zeta )=P_{n+1}^{(0)}(-x_{i,n+1}^{(0)};\zeta )P_{n}^{(1)}(x_{i,n+1}^{(0)};\zeta ). \end{aligned}$$

Since

$$\begin{aligned} \mathrm{sign}\left( P_{n+1}^{(0)}(-x_{i,n+1}^{(0)};\zeta )\right) =(-1)^{n+1}, \quad i=1,\ldots , n+1, \end{aligned}$$

using Theorem 3.6 gives

$$\begin{aligned} \mathrm{sign}\left( Q_{2n}(x_{i,n+1}^{(0)};\zeta )Q_{2n}(x_{i+1,n+1}^{(0)};\zeta )\right) =\mathrm{sign}\left( P_{n}^{(1)}(x_{i,n+1}^{(0)};\zeta )P_{n}^{(1)}(x_{i+1,n+1}^{(0)};\zeta )\right) <0, \end{aligned}$$

for \(i=1,\ldots ,n\), which implies that there exists at least one zero of \(Q_{2n}(x;\zeta )\) in each subinterval \( (x_{i,n+1}^{(0)},x_{i+1,n+1}^{(0)})\), \(i=1,\ldots ,n\). Because \(Q_{2n}(x;\zeta )\) is an even polynomial of degree 2n, there exists exactly one zero of \(Q_{2n}(x;\zeta )\) in each subinterval \((x_{i,n+1}^{(0)},x_{i+1,n+1}^{(0)})\), \(i=1,\ldots ,n\). The proof is completed. \(\square \)

1.8 Proof of Lemma 2

Proof

According to the definition of \(Q_{2n}(x;\zeta )\) in (3.30), one has

$$\begin{aligned} \frac{\partial Q_{2n}}{\partial \zeta }(z_{i,n};\zeta )=\,&\frac{\partial P_{n+1}^{(0)}}{\partial \zeta }\Big |_{x=z_{i,n}} P_{n}^{(1)}(-z_{i,n};\zeta ) +\frac{\partial P_{n}^{(1)}}{\partial \zeta }\Big |_{x=z_{i,n}} P_{n+1}^{(0)}(-z_{i,n};\zeta )\\&+\frac{\partial P_{n+1}^{(0)}}{\partial \zeta } \Big |_{x=-z_{i,n}} P_{n}^{(1)}(z_{i,n};\zeta ) +\frac{\partial P_{n}^{(1)}}{\partial \zeta }\Big |_{x=-z_{i,n}} P_{n+1}^{(0)}(z_{i,n};\zeta ). \end{aligned}$$

Using Theorem 3.3 gives

$$\begin{aligned} \frac{\partial Q_{2n}}{\partial \zeta }(z_{i,n};\zeta )=\,&a_{n}^{(0)}\left( P_{n}^{(0)}(z_{i,n};\zeta )P_{n}^{(1)}(-z_{i,n};\zeta )+P_{n}^{(0)}(-z_{i,n};\zeta )P_{n}^{(1)}(z_{i,n};\zeta )\right) \\&+a_{n-1}^{(1)}\left( P_{n-1}^{(1)}(z_{i,n};\zeta )P_{n+1}^{(0)}(-z_{i,n};\zeta )+P_{n-1}^{(1)}(-z_{i,n};\zeta )P_{n+1}^{(0)}(z_{i,n};\zeta )\right) \\&+\left( \frac{1}{2}(b_{n+1}^{(0)}+b_{n}^{(1)})-(G(\zeta )-\zeta ^{-1})\right) Q_{2n}(z_{i,n};\zeta )\\ =\,&a_{n}^{(0)}\left( P_{n}^{(0)}(z_{i,n};\zeta )P_{n}^{(1)}(-z_{i,n};\zeta )+P_{n}^{(0)}(-z_{i,n};\zeta )P_{n}^{(1)}(z_{i,n};\zeta )\right) \\&+a_{n-1}^{(1)}\left( P_{n-1}^{(1)}(z_{i,n};\zeta )P_{n+1}^{(0)}(-z_{i,n};\zeta )+P_{n-1}^{(1)}(-z_{i,n};\zeta )P_{n+1}^{(0)}(z_{i,n};\zeta )\right) . \end{aligned}$$

Substituting (3.14) and (3.15) into it gives

$$\begin{aligned} \frac{\partial Q_{2n}}{\partial \zeta }(z_{i,n};\zeta ) =\,&\frac{2a_{n}^{(0)}}{\tilde{r}_{n}}\left( ((z_{i,n})^2-1)P_{n}^{(1)}(z_{i,n};\zeta )P_{n}^{(1)}(-z_{i,n};\zeta )\right. \\&\quad \left. + \tilde{p}_{n}z_{i,n}P_{n+1}^{(0)}(-z_{i,n};\zeta )P_{n}^{(1)}(z_{i,n};\zeta )\right) \\&\,-\frac{2a_{n}^{(0)}}{\tilde{r}_{n}}\left( P_{n+1}^{(0)}(z_{i,n};\zeta )P_{n+1}^{(0)}(-z_{i,n};\zeta )- \tilde{p}_{n}^{-1}z_{i,n}P_{n+1}^{(0)}(-z_{i,n};\zeta )P_{n}^{(1)}(z_{i,n};\zeta )\right) \\ =\,&2\frac{P_{n}^{(1)}(z_{i,n};\zeta )}{P_{n}^{(1)}(-z_{i,n};\zeta )} \frac{a_{n}^{(0)}}{\tilde{r}_{n}}\left( (\tilde{p}_{n}+ \tilde{p}_{n}^{-1})z_{i,n}P_{n}^{(1)}(-z_{i,n};\zeta )P_{n+1}^{(0)}(-z_{i,n};\zeta )\right. \\&\left. +(z_{i,n}^2-1)P_{n}^{(1)}(-z_{i,n};\zeta )^2 +P_{n+1}^{(0)} (-z_{i,n};\zeta )^2\right) \\ =\,&2\frac{P_{n}^{(1)}(z_{i,n};\zeta )}{P_{n}^{(1)}(-z_{i,n};\zeta )} \frac{a_{n}^{(0)}}{\tilde{r}_{n}}\left( (c_{n+1}^{(0)})^2\prod _{j=1}^{n+1}(z_{i,n}+x_{j,n+1}^{(0)}) \tilde{I}_{1}-(c_{n}^{(1)})^2\prod _{j=1}^{n}(z_{i,n}+x_{j,n}^{(1)}) \tilde{I}_{2}\right) \!, \end{aligned}$$

where

$$\begin{aligned} \tilde{I}_{1}:=\prod _{j=1}^{n+1}\left( z_{i,n}+x_{j,n+1}^{(0)}\right) -z_{i,n}\prod _{j=1}^{n}\left( z_{i,n}+x_{j,n}^{(1)}\right) , \end{aligned}$$
$$\begin{aligned} \tilde{I}_{2}:=z_{i,n}\prod _{j=1}^{n+1}\left( z_{i,n}+x_{j,n+1}^{(0)}\right) -(z_{i,n}^2-1)\prod _{j=1}^{n}\left( z_{i,n}+x_{j,n}^{(1)}\right) . \end{aligned}$$

Similarly, using Theorem 3.4 and (3.14)-(3.15) gives

$$\begin{aligned} \frac{\partial Q_{2n}}{\partial x}(z_{i,n};\zeta ) =\,&\zeta \frac{c_{n-1}^{(1)}}{c_{n+1}^{(0)}}\left( P_{n-1}^{(1)}(z_{i,n};\zeta )P_{n}^{(1)}(-z_{i,n};\zeta )-P_{n-1}^{(1)}(-z_{i,n};\zeta )P_{n}^{(1)}(z_{i,n};\zeta )\right) \\&+(z_{i,n}^2-1)^{-1}\zeta \frac{c_{n}^{(0)}}{c_{n}^{(1)}}\left( P_{n}^{(0)}(z_{i,n};\zeta )P_{n+1}^{(0)}(-z_{i,n};\zeta )\right. \\&\left. -P_{n}^{(0)}(-z_{i,n};\zeta )P_{n+1}^{(0)}(z_{i,n};\zeta )\right) \\ =\,&2\zeta \frac{P_{n}^{(1)}(z_{i,n};\zeta )}{P_{n}^{(1)}(-z_{i,n};\zeta )} \frac{a_{n}^{(0)}}{\tilde{r}_{n}}\left( z_{i,n}P_{n}^{(1)}(-z_{i,n};\zeta )^{2}\right. \\&\left. +(\tilde{p}_{n}+\tilde{p}_{n}^{-1})P_{n+1}^{(0)}(-z_{i,n};\zeta )P_{n}^{(1)}(-z_{i,n};\zeta )\right. \\&\left. +(z_{i,n}^2-1)^{-1}z_{i,n}P_{n+1}^{(0)}(-z_{i,n};\zeta )^2 \right) \\ =\,&2\zeta \frac{P_{n}^{(1)}(z_{i,n};\zeta )}{P_{n}^{(1)}(-z_{i,n};\zeta )}\\&\frac{a_{n}^{(0)}}{\tilde{r}_{n}}\left( (c_{n+1}^{(0)})^2\frac{\prod _{j=1}^{n+1}(z_{i,n}+x_{j,n+1}^{(0)})}{z_{i,n}^2-1} \tilde{I}_{2}-(c_{n}^{(1)})^2\prod _{j=1}^{n}(z_{i,n}+x_{j,n}^{(1)})\tilde{I}_{1}\right) . \end{aligned}$$

Using Theorem 3.6 gives

$$\begin{aligned} z_{i,n}+x_{j+1,n+1}^{(0)}>z_{i,n}+x_{j,n}^{(1)} ,\quad j=1,\ldots ,n, \\ z_{i,n}+x_{1,n+1}^{(0)}>z_{i,n}+1>z_{i,n}>z_{i,n}-1, \end{aligned}$$

for \(i=1,\ldots ,n\), which imply

$$\begin{aligned} \tilde{I}_{1}>0,\quad \tilde{I}_{2}>0. \end{aligned}$$

Thus one has

$$\begin{aligned} \prod _{j=1}^{n+1}&(z_{i,n}+x_{j,n+1}^{(0)})\tilde{I}_{1}-\prod _{j=1}^{n}(z_{i,n}+x_{j,n}^{(1)}) \tilde{I}_{2}\\ =&\left( \prod _{j=1}^{n+1}(z_{i,n}+x_{j,n+1}^{(0)})-(z_{i,n}-1)\prod _{j=1}^{n}(z_{i,n}+x_{j,n}^{(1)} )\right) \\&\times \left( \prod _{j=1}^{n+1}(z_{i,n}+x_{j,n+1}^{(0)})-(z_{i,n}+1)\prod _{j=1}^{n}(z_{i,n}+x_{j,n}^{(1)})\right) >0, \end{aligned}$$
$$\begin{aligned} \prod _{j=1}^{n+1}&(z_{i,n}+x_{j,n+1}^{(0)}) \tilde{I}_{2}-((z_{i,n})^2-1)\prod _{j=1}^{n}(z_{i,n} +x_{j,n}^{(1)})\tilde{I}_{1}\\ =&z_{i,n}{\left( \prod _{j=1}^{n+1}(z_{i,n}+x_{j,n+1}^{(0)})\tilde{I}_{1}-\prod _{j=1}^{n}(z_{i,n}+x_{j,n}^{(1)}) \tilde{I}_{2}\right) }\\&+2\prod _{j=1}^{n+1}(z_{i,n}+x_{j,n+1}^{(0)})\prod _{j=1}^{n}(z_{i,n}+x_{j,n}^{(1)})>0. \end{aligned}$$

Using Corollaries 1 and 2, and the above results gives (3.31).

The proof is completed. \(\square \)

1.9 Proof of Lemma 3

Proof

Taking partial derivative of \(Q_{2n}(z_{i,n};\zeta )\) with respect to \(\zeta \) gives

$$\begin{aligned} \frac{\partial z_{i,n}}{\partial \zeta }=-\left( \frac{\partial Q_{2n}}{\partial x}\Big |_{x=z_{i,n}}\right) ^{-1} \left( \frac{\partial Q_{2n}}{\partial \zeta }\Big |_{x=z_{i,n}}\right) , i=1,\ldots ,n. \end{aligned}$$

Using Lemma 2 completes the proof. \(\square \)

1.10 Proof of Theorem 3.7

Proof

Obviously, both vectors \(\mathbf {u}_{i,n}\) and \(\mathbf {v}_{i,n}\) defined in (3.35) are not zero at the same time, \(i=\pm 1,\ldots ,\pm n\). The nonzero eigenvalues and eigenvectors of the matrix pair \(\mathbf {A}^{0}_{n}\) and \(\mathbf {A}^{1}_{n}\) in (3.32) and (3.34) can be obtained with the aid of (3.28)-(3.29) and Lemma 1. Using Lemma 3 further gives (3.33).

In the following, let us discuss the eigenvector \(\mathbf {y}_{0,n}\). Multiplying (3.12) by \(P_{n+1}^{(0)}(-x;\zeta )\) gives

$$\begin{aligned}&(x^2-1)\mathbf {P}_{n-1}^{(1)}(x;\zeta )P_{n+1}^{(0)}(-x;\zeta )\nonumber \\&=\mathbf {J}_{n-1}\mathbf {P}_{n}^{(0)}(x;\zeta )P_{n+1}^{(0)}(-x;\zeta ) +r_{n}P_{n+1}^{(0)}(x;\zeta )P_{n+1}^{(0)}(-x;\zeta )\mathbf {e}_{n}. \end{aligned}$$
(9.1)

Transforming (9.1) by x to \(-x\) and then subtracting it from (9.1) and letting \(x=1\) gives as follows

$$\begin{aligned} 0=\mathbf {J}_{n-1}\left( \mathbf {P}_{n}^{(0)}(1;\zeta )P_{n+1}^{0}(-1;\zeta )-\mathbf {P}_{n}^{(0)}(-1;\zeta )P_{n+1}^{(0)}(1;\zeta )\right) =\mathbf {J}_{n-1}\mathbf {u}_{0,n}, \end{aligned}$$

which is a special case of (3.21) with \( \hat{\lambda } =0\).

The proof is completed. \(\square \)

Appendix 3: Proofs in Section 4

1.1 Proof of Lemma 4

Proof

  1. (i)

    Due to the definition of E and \(p_{{\langle }1{\rangle }}\), it is obvious that each component of \(\mathbf {\mathcal {P}}_{\infty }[u,\theta ]\) (resp. \(\mathbf {\mathcal {P}}_{M}[u,\theta ]\)) belongs to \(\mathbb {H}^{g^{(0)}_{[u,\theta ]}}\) (resp. \(\mathbb {H}^{g^{(0)}_{[u,\theta ]}}_{M}\)).

  2. (ii)

    The mathematical induction is used to prove that any element in the space \(\mathbb {H}^{g^{(0)}_{[u,\theta ]}}\) (resp. \(\mathbb {H}^{g^{(0)}_{[u,\theta ]}}_{M}\)) can be expressed as a linear combination of vectors in \(\mathbf {\mathcal {P}}_{\infty }[u,\theta ]\) (resp. \(\mathbf {\mathcal {P}}_{M}[u,\theta ]\)) . For \(M=1\), it is clear to have the linear combination

    $$\begin{aligned} p^{\alpha }g^{(0)}_{[u,\theta ]}\overset{(2.9)}{=}&\left( p^{{\langle }\alpha {\rangle }}+U^{\alpha }E\right) g^{(0)}_{[u,\theta ]} \overset{(2.8)}{=} \left( -(U^{0})^{-1}U^{1-\alpha }p_{{\langle }1{\rangle }}+U^{\alpha }E\right) g^{(0)}_{[u,\theta ]}\\ \overset{(3.5)}{=}&-(c_{0}^{(1)})^{-1}U^{1-\alpha }\tilde{P}_{0}^{(1)}[u,\theta ]+ (c_{1}^{(0)})^{-1}U^{\alpha }\tilde{P}_{1}^{(0)}[u,\theta ]\\&+(c_{0}^{(0)})^{-1}U^{\alpha }x_{1,1}^{(0)}\tilde{P}_{0}^{(0)}[u,\theta ], \end{aligned}$$

    where the decomposition of the particle velocity vector (2.9) has been used. Assume that the linear combination

    $$\begin{aligned} p^{\mu _{1}}p^{\mu _{2}}\ldots p^{\mu _{M}}g^{(0)}_{[u,\theta ]}=&\sum _{i=0}^{M}c_{i,0}^{\mu _{1},\ldots ,\mu _{M}}\tilde{P}_{i}^{(0)}[u,\theta ] +\sum _{i=0}^{M-1}c_{i,1}^{\mu _{1},\ldots ,\mu _{M}}\tilde{P}_{i}^{(1)}[u,\theta ],\\&\mu _{i}=0,1, i\in \mathbb {N}, i\le M, \quad c_{i,0}^{\mu _{1},\ldots ,\mu _{M}},c_{i,1}^{\mu _{1},\ldots ,\mu _{M}}\in \mathbb {R}, \end{aligned}$$

    holds. One has to show that \(p^{\mu _{1}}p^{\mu _{2}}\ldots p^{\mu _{M+1}}g^{(0)}_{[u,\theta ]}\) can be expressed as a linear combination of components of \(\mathbf {\mathcal {P}}_{M+1}[u,\theta ]\). Because

    $$\begin{aligned}&p^{\mu _{1}}p^{\mu _{2}}\ldots p^{\mu _{M+1}}g^{(0)}_{[u,\theta ]}\\&=\left( \sum _{i=0}^{M}c_{i,0}^{\mu _{1},\ldots ,\mu _{M}}\tilde{P}_{i}^{(0)}[u,\theta ] +\sum _{i=0}^{M-1}c_{i,1}^{\mu _{1},\ldots ,\mu _{M}}\tilde{P}_{i}^{(1)}[u,\theta ]\right) \\&\times \left( -(U^{0})^{-1}U^{1-\mu _{M+1}}p_{{\langle }1{\rangle }}+U^{\mu _{M+1}}E\right) \\&=\sum _{i=0}^{M}c_{i,0}^{\mu _{1},\ldots ,\mu _{M}}U^{\mu _{M+1}}E\tilde{P}_{i}^{(0)}[u,\theta ] -\sum _{i=0}^{M-1}c_{i,1}^{\mu _{1},\ldots ,\mu _{M}}U^{1-\mu _{M+1}}(E^2-1)P_{i}^{(1)}(E;\zeta )\\&\quad -\sum _{i=0}^{M}c_{i,0}^{\mu _{1},\ldots ,\mu _{M}}(U^{0})^{-1}U^{1-\mu _{M+1}}P_{i}^{(0)}(E;\zeta )p_{{\langle }1{\rangle }}+ \sum _{i=0}^{M-1}c_{i,1}^{\mu _{1},\ldots ,\mu _{M}}U^{\mu _{M+1}}\tilde{P}_{i}^{(1)}[u,\theta ], \end{aligned}$$

    one has

    $$\begin{aligned}&p^{\mu _{1}}p^{\mu _{2}}\ldots p^{\mu _{M+1}}g^{(0)}_{[u,\theta ]}\\&\quad =\sum _{i=0}^{M}c_{i,0}^{\mu _{1},\ldots ,\mu _{M}}U^{\mu _{M+1}} \left( a_{i-1}^{(0)}\tilde{P}_{i-1}^{(0)}[u,\theta ]+b_{i}^{(0)}\tilde{P}_{i}^{(0)}[u,\theta ]+a_{i}^{(0)}\tilde{P}_{i+1}^{(0)}[u,\theta ]\right) \\&\qquad -\sum _{i=0}^{M-1}c_{i,1}^{\mu _{1},\ldots ,\mu _{M}}U^{1-\mu _{M+1}} \left( p_{i}\tilde{P}_{i}^{(0)}[u,\theta ]+q_{i}\tilde{P}_{i+1}^{(0)}[u,\theta ]+r_{i+1}\tilde{P}_{i+2}^{(0)}[u,\theta ]\right) \\&\qquad -\sum _{i=0}^{M}c_{i,0}^{\mu _{1},\ldots ,\mu _{M}}U^{1-\mu _{M+1}} \left( r_{i-1}\tilde{P}_{i-2}^{(1)}[u,\theta ]+q_{i-1}\tilde{P}_{i-1}^{(1)}[u,\theta ]+p_{i}\tilde{P}_{i}^{(1)}[u,\theta ]\right) \\&\qquad +\sum _{i=0}^{M-1}c_{i,1}^{\mu _{1},\ldots ,\mu _{M}}U^{\mu _{M+1}} \left( a_{i-1}^{(1)}\tilde{P}_{i-1}^{(1)}[u,\theta ]+b_{i}^{(1)}\tilde{P}_{i+1}^{(1)}[u,\theta ]+a_{i}^{(1)}\tilde{P}_{i+1}^{(1)}[u,\theta ]\right) \\&\quad =:\sum _{i=0}^{M+1}c_{i,0}^{\mu _{1},\ldots ,\mu _{M+1}}\tilde{P}_{i}^{(0)}[u,\theta ] +\sum _{i=0}^{M}c_{i,1}^{\mu _{1},\ldots ,\mu _{M+1}}\tilde{P}_{i}^{(1)}[u,\theta ]{,} \end{aligned}$$

    by using the three-term recurrence relations (3.6), (3.9), and (3.10) for the orthogonal polynomials \(\{P_{n}^{(\ell )}(x;\zeta ), \ell =0,1\}\).

  3. (iii)

    Using (3.1) gives

    $$\begin{aligned} {\langle }\tilde{P}_{i}^{(\ell )}[u,\theta ],\tilde{P}_{j}^{(\ell )}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}}= \left( P_{i}^{(\ell )},P_{j}^{(\ell )}\right) _{ \omega ^{(\ell )}}=\delta _{i,j},\ \ell =0,1. \end{aligned}$$
    (10.1)

    Because of (2.9), one has

    $$\begin{aligned} \frac{dp}{p^{0}}=dp_{{\langle }1{\rangle }}\frac{-1+u(U^{0}E)^{-1}p_{{\langle }1{\rangle }}}{-up_{{\langle }1{\rangle }}+U^{0}E}=-\frac{dp_{{\langle }1{\rangle }}}{U^{0}E}, \quad E=\sqrt{\left( (U^{0})^{-1}p_{{\langle }1{\rangle }}\right) ^2+1}. \end{aligned}$$

    Thus it holds

    $$\begin{aligned} \nonumber {\langle }\tilde{P}_{i}^{(0)}[u,\theta ],\tilde{P}_{j}^{(1)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}}=&\int _{\mathbb {R}}g^{(0)}_{[u,\theta ]}P_{i}^{(0)}(E;\zeta )P_{j}^{(1)}(E;\zeta )(U^{0})^{-1}p_{{\langle }1{\rangle }}\frac{dp}{p^{0}}\\ =&-\int _{\mathbb {R}}g^{(0)}_{[u,\theta ]}P_{i}^{(0)}(E;\zeta )P_{j}^{(1)}(E;\zeta )(U^{0})^{-1}p_{{\langle }1{\rangle }}\frac{dp_{{\langle }1{\rangle }}}{U^{0}E}=0. \end{aligned}$$
    (10.2)

    Combining (i) and (ii) with (iii) completes the proof.

\(\square \)

1.2 Proof of Lemma 5

Proof

For \(s=t\) and x, it is clear to have

$$\begin{aligned} \frac{\partial E}{\partial s}=\frac{\partial u}{\partial s}\frac{1}{(1-u^2)}(U^{0})^{-1}p_{{\langle }1{\rangle }},\quad \frac{\partial \left( (U^{0})^{-1}p_{{\langle }1{\rangle }}\right) }{\partial s}=\frac{\partial u}{\partial s}\frac{1}{1-u^2}E. \end{aligned}$$

Using the above identities and (4.1) gives

$$\begin{aligned} \frac{\partial g^{(0)}_{[u,\theta ]}}{\partial s}=-\left( \frac{\partial \theta }{\partial s}\zeta ^2\left( G(\zeta )-\zeta ^{-1}-E\right) + \frac{\partial u}{\partial s}\frac{1}{\theta (1-u^2)}(U^{0})^{-1}p_{{\langle }1{\rangle }}\right) g^{(0)}_{[u,\theta ]}. \end{aligned}$$

The derivation rule of compound function gives

$$\begin{aligned} \frac{\partial \tilde{P}_{n}^{(0)}[u,\theta ]}{\partial s}=&\frac{\partial P_{n}^{(0)}}{\partial E}\frac{\partial E}{\partial s}g^{(0)}_{[u,\theta ]} -\zeta ^2\frac{\partial P_{n}^{(0)}}{\partial \zeta }\frac{\partial \theta }{\partial s}g^{(0)}_{[u,\theta ]} +P_{n}^{(0)}\frac{\partial g^{(0)}_{[u,\theta ]}}{\partial s},\\ \frac{\partial \tilde{P}_{n-1}^{(1)}[u,\theta ]}{\partial s}=&\frac{\partial P_{n-1}^{(1)}}{\partial E}\frac{\partial E}{\partial s}(U^{0})^{-1}p_{{\langle }1{\rangle }}g^{(0)}_{[u,\theta ]} -\zeta ^2\frac{\partial P_{n-1}^{(1)}}{\partial \zeta }\frac{\partial \theta }{\partial s}(U^{0})^{-1}p_{{\langle }1{\rangle }}g^{(0)}_{[u,\theta ]}\\&+P_{n-1}^{(1)}\frac{\partial \left( (U^{0})^{-1}p_{{\langle }1{\rangle }}\right) }{\partial s}g^{(0)}_{[u,\theta ]} +P_{n-1}^{(1)}(U^{0})^{-1}p_{{\langle }1{\rangle }}\frac{\partial g^{(0)}_{[u,\theta ]}}{\partial s}{.} \end{aligned}$$

Combining them and using Theorems 3.13.4 complete the proof. \(\square \)

1.3 Proof of Lemma 6

Proof

Using the three-term recurrence relations (3.7), (3.11), and (3.12) gives

$$\begin{aligned} E\mathbf {\mathcal {\tilde{P}}}_{M}[u,\theta ]&=\mathbf {A}^{0}_{M} \mathbf {\mathcal {\tilde{P}}}_{M}[u,\theta ]+a_{M}^{(0)}\tilde{P}_{M+1}^{(0)}[u,\theta ]\mathbf {e}_{2M+1}^{3} +a_{M-1}^{(1)}\tilde{P}_{M}^{(1)}[u,\theta ]\mathbf {e}_{2M+1}^{2}, \\ (U^{0})^{-1}p_{{\langle }1{\rangle }}\mathbf {\mathcal {\tilde{P}}}_{M}&=\mathbf {A}^{1}_{M}\mathbf {\mathcal {\tilde{P}}}_{M}[u,\theta ] +p_{M}\tilde{P}_{M}^{(1)}[u,\theta ]\mathbf {e}_{2M+1}^{3} +r_{M}\tilde{P}_{M+1}^{(0)}[u,\theta ]\mathbf {e}_{2M+1}^{2}, \end{aligned}$$

where \(\mathbf {e}_{2M+1}^{3}\) is the \((M+1)\)th column of the identity matrix of order \((2M+1)\). Thus one has

$$\begin{aligned} E\mathbf {\mathcal {P}}_{M}[u,\theta ]&= \mathbf {P}_{M}^{p}\mathbf {A}^{0}_{M}(\mathbf {P}_{M}^{p})^{T} \mathbf {\mathcal {P}}_{M}[u,\theta ]{+}a_{M}^{(0)}\tilde{P}_{M+1}^{(0)}[u,\theta ]\mathbf {e}_{2M+1}^{1} {+}a_{M-1}^{(1)}\tilde{P}_{M}^{(1)}[u,\theta ]\mathbf {e}_{2M+1}^{2},\\ (U^{0})^{-1}p_{{\langle }1{\rangle }}\mathbf {\mathcal {P}}_{M}&= \mathbf {P}_{M}^{p}\mathbf {A}^{1}_{M}(\mathbf {P}_{M}^{p})^{T}\mathbf {\mathcal {P}}_{M}[u,\theta ] +p_{M}\tilde{P}_{M}^{(1)}[u,\theta ]\mathbf {e}_{2M+1}^{1} +r_{M}\tilde{P}_{M+1}^{(0)}[u,\theta ]\mathbf {e}_{2M+1}^{2}. \end{aligned}$$

Combining them with (2.9) completes the proof. \(\square \)

1.4 Proof of Lemma 7

Proof

It is obvious that \(\Pi _{M}[u,\theta ]\) is a linear bounded operator and \(\Pi _{M}[u,\theta ]f\in \mathbb {H}^{g^{(0)}_{[u,\theta ]}}_{M}\) for all \(f\in \mathbb {H}^{g^{(0)}_{[u,\theta ]}}\).

For all \(f\in \mathbb {H}^{g^{(0)}_{[u,\theta ]}}_{M}\), besides (4.7), by using Lemma 4 one has

$$\begin{aligned} f=&\sum _{i=0}^{M}\tilde{f}_{i}^{0}\tilde{P}_{i}^{(0)}[u,\theta ]+ \sum _{j=0}^{M-1}\tilde{f}_{j}^{1}\tilde{P}_{j}^{(1)}[u,\theta ]. \end{aligned}$$

Taking respectively the inner product with \(\tilde{P}_{i}^{(0)}[u,\theta ]\) and \(\tilde{P}_{j}^{(1)}[u,\theta ]\) from both sides of the last equation gives

$$\begin{aligned} f_{i}^{0}={\langle }f,\tilde{P}_{i}^{(0)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}},i\le M,\quad f_{j}^{1}={\langle }f,\tilde{P}_{j}^{(1)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}},j\le M-1. \end{aligned}$$

Comparing them with the coefficients in (4.9) shows that \(\tilde{f}_{i}^{0}=f_{i}^{0}\), \(\tilde{f}_{j}^{1}=f_{j}^{1}\),

\(i=0,\ldots , M\), \(j=1,\ldots ,M-1\). The proof is completed. \(\square \)

Appendix 4: Proofs in Section 5

1.1 Proof of Lemma 8

Proof

It is obvious that for \(M=1\), the matrix \(\mathbf {D}_M\) is invertible because

\(\det (\mathbf {D}_M)=\rho \zeta ^2c_0^{(1)}(c_0^{(0)}c_1^{(0)}(1-u^2))^{-1}>0\). For \(M\ge 2\), according to the form of \(\mathbf {D}_M\) in Sect. 4.2, one has

$$\begin{aligned} \det (\mathbf {D}_{M})=\det (\mathbf {D}_{2})=\zeta ^3c_{2}^{(0)}c_{1}^{(1)}(x_{1,2}^{(0)}+x_{2,2}^{(0)})(\rho G(\zeta )+\Pi )\rho (c_{1}^{(0)}c_{0}^{(0)}(1-u^2))^{-1}. \end{aligned}$$

Using \(\Pi >-\rho \theta \) gives

$$\begin{aligned} \det (\mathbf {D}_{M})>\zeta ^3c_{2}^{(0)}c_{1}^{(1)}(x_{1,2}^{(0)}+x_{2,2}^{(0)})\rho ^2( G(\zeta )-\zeta ^{-1})(c_{1}^{(0)}c_{0}^{(0)}(1-u^2))^{-1}>0. \end{aligned}$$

The proof is completed. \(\square \)

1.2 Proof of Theorem 5.1

Proof

Consider the following generalized eigenvalue problem (2nd sense): Find a vector \(\mathbf {r}\) that obeys \(\lambda \mathbf {B}_{M}^{0}\mathbf {r}=\mathbf {B}_{M}^{1}\mathbf {r}\) or \( \lambda \mathbf {M}_{M}^{t}\mathbf {D}_{M}\mathbf {r}=\mathbf {M}_{M}^{x}{\mathbf {D}}_{M}\mathbf {r}\). Thanks to (4.5), this eigenvalue problem is equivalent to

$$\begin{aligned} (\lambda -u) \mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T}\mathbf {D}_{M}\mathbf {r}= (\lambda u-1) \mathbf {A}_{M}^{1}(\mathbf {P}_{M}^{p})^{T}\mathbf {D}_{M}\mathbf {r}. \end{aligned}$$

Because Theorem 3.7 tells us that \( \hat{\lambda } _{i,M}\) and \(\mathbf {y}_{i,M}\) satisfy

$$\begin{aligned} \hat{\lambda } _{i,M}\mathbf {A}_{M}^{0}\mathbf {y}_{i,M}=\mathbf {A}_{M}^{1}\mathbf {y}_{i,M},\ | \hat{\lambda } _{i,M}|<1, \end{aligned}$$

the scalar \(\lambda _{i,M}\) in (5.1) and vector \(\mathbf {r}_{i,M}\) in (5.2) solve the above generalized eigenvalue problem,

and satisfy

$$\begin{aligned} |\lambda _{i,M}|<\frac{1-u}{1-u}=1. \end{aligned}$$

The proof is completed. \(\square \)

1.3 Proof of Lemma 9

Proof

Because \(U^{0}\mathbf {M}_{M}^{t}-U^{1}\mathbf {M}_{M}^{x}=\mathbf {P}_{M}^{p}A_{M}^{0}(\mathbf {P}_{M}^{p})^{T}\) and the permutation matrix \(\mathbf {P}_{M}^{p}\) in (4.6) satisfies \(\mathbf {P}_{M}^{p}(\mathbf {P}_{M}^{p})^{T}=(\mathbf {P}_{M}^{p})^{T}\mathbf {P}_{M}^{p}={\mathbf {I}}\), two matrices \(U^{0}\mathbf {M}_{M}^{t}-U^{1}\mathbf {M}_{M}^{x}\) and \(\mathbf {A}_{M}^{0}\) are similar and thus have the same eigenvalues. The definition of \(\mathbf {A}_{M}^{0}\) in (3.20) tells us that the eigenvalues of \(\mathbf {A}_{M}^{0}\) are the zeros of \(P_{M+1}^{(0)}(x;\zeta )\) and \(P_{M}^{(1)}(x;\zeta )\) which are larger than one [43, Theorem 3.4], so the matrix \(U^{0}\mathbf {M}_{M}^{t}-U^{1}\mathbf {M}_{M}^{x}\) is positive definite.

Theorem 3.7 implies

$$\begin{aligned} \rho \left( (\mathbf {A}_{M}^{0})^{-\frac{1}{2}}\mathbf {A}_{M}^{1}(\mathbf {A}_{M}^{0})^{-\frac{1}{2}}\right) =\rho \left( (\mathbf {A}_{M}^{0})^{-1}\mathbf {A}_{M}^{1}\right) <1, \end{aligned}$$

where \(\rho (\cdot )\) is the spectral radius of the matrix. Then \([{\mathbf {I}}-\left( (U^{0}\mathbf {A}_{M}^{0})^{-\frac{1}{2}}U^{1}\mathbf {A}_{M}^{1}(U^{0}\mathbf {A}_{M}^{0})^{-\frac{1}{2}}\right) \) is positive definite, so the matrix \({\mathbf {M}}_{M}^{t}\) is positive definite. \(\square \)

1.4 Proof of Theorem 5.2

Proof

Lemmas 8 and 9 show that the matrix \(\mathbf {B}_{M}^{0}=\mathbf {M}^{t}_{M}\mathbf {D}_{M}\) is invertible, and Theorem 5.1 implies that \(\mathbf {B}_{M}\) is diagonalizable with real eigenvalues and the spectral radius of \(\mathbf {B}_{M}\) is less than one. The proof is completed. \(\square \)

1.5 Proof of Theorem 5.3

Proof

Because

$$\begin{aligned} \nabla _{\mathbf {W}_{M}}\lambda _{i,M}=\frac{1}{\left( 1-u \hat{\lambda } _{i,M}\right) ^{2}}\left( 0,1- \hat{\lambda } _{i,M}^2, -(1-u^2)\frac{\partial \hat{\lambda } _{i,M}}{\partial \theta },0\ldots ,0\right) ^{T},\ \end{aligned}$$

and \( \mathbf {r}_{i,M}=\mathbf {D}_{M}^{-1}\mathbf {P}_{M}^{p}\left( (\mathbf {u}_{i,M})^{T},(\mathbf {v}_{i,M})^{T}\right) ^{T}\), \(i=-M,\ldots , M\), one has

$$\begin{aligned} \nabla _{\mathbf {W}_{M}}\lambda _{i,M}\cdot \mathbf {r}_{i,M}=\frac{1}{\left( 1-u \hat{\lambda } _{i,M}\right) ^{2}}\left( \left( 1- \hat{\lambda } _{i,M}^2\right) \mathbf {d}_{2}\mathbf {P}_{2}^{p}\tilde{\mathbf {r}}_{i}^{M} -(1-u^2)\frac{\partial \hat{\lambda } _{i,M}}{\partial \theta }\mathbf {d}_{3}\mathbf {P}_{2}^{p}\tilde{\mathbf {r}}_{i}^{M}\right) , \end{aligned}$$
(11.1)

where \(\tilde{\mathbf {r}}_{i}^{M}=\left( (\mathbf {u}_{i,M}^{(3)})^{T},(\mathbf {v}_{i,M}^{(2)})^{T}\right) ^{T}\), \(\mathbf {u}_{i,M}^{(3)}\) and \(\mathbf {v}_{i,M}^{(2)}\) denote two vectors formed by first three and two components of \(\mathbf {u}_{i,M}\) and \(\mathbf {v}_{i,M}\) respectively, and \(\mathbf {d}_{2}\) and \(\mathbf {d}_{3}\) are the second and third row of \(\mathbf {D}_{2}^{-1}\), specifically

$$\begin{aligned} \mathbf {d}_{2}=&-\frac{G(\zeta )(1-u^2)}{(\rho G(\zeta )+\Pi )\sqrt{\zeta }}\left( 0,0,G(\zeta ),0,\sqrt{-G(\zeta )^{2}+3\zeta ^{-1}G(\zeta )+1}\right) , \\ \mathbf {d}_{3}=&\frac{1}{\rho \zeta (G(\zeta )^2\zeta ^2{-}3G(\zeta )\zeta {-}\zeta ^2{+}1)}\left( 0, \frac{\sqrt{G(\zeta ){-}2\zeta ^{-1}}(G(\zeta )^2\zeta ^2{-}3G(\zeta )\zeta {-}\zeta ^2{+}1)}{\sqrt{G(\zeta )^2\zeta ^2{-}3G(\zeta )\zeta {-}\zeta ^2{+}2}}, \right. \\&\left. \frac{(G(\zeta )\zeta -1)\tilde{n}^{1}G(\zeta )}{\sqrt{\zeta }(\rho G(\zeta )+\Pi )}, -\frac{\sqrt{2G(\zeta )^3\zeta ^2-7G(\zeta )^2\zeta -2G(\zeta )\zeta ^2+6G(\zeta )+\zeta }}{\sqrt{G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+2}},\right. \\&\left. \frac{(G(\zeta )-\zeta ^{-1})\sqrt{-G(\zeta )+3\zeta ^{-1}G(\zeta )+1}}{\rho G(\zeta )+\Pi }\right) . \end{aligned}$$

The identity (11.1) always holds, because \( \hat{\lambda } _{0,M}=0\) and \(\mathbf {u}_{0,M}\) and \(\mathbf {v}_{0,M}\) are given in (3.36).

The proof is completed. \(\square \)

1.6 Explanation of Remark 11

In fact, in order to judge by numerical experiments whether the sign of \( \nabla _{\mathbf {W}_{M}}\lambda _{i,M}\cdot \mathbf {r}_{i,M}\) is constant or not, (11.1) should be reformed. For \(i=\pm 1,\pm 2,\ldots ,\pm M\), Theorem 3.7 and (11.1) give

$$\begin{aligned} \frac{ \nabla _{\mathbf {W}_{M}}\lambda _{i,M}\cdot \mathbf {r}_{i,M}}{ -(1-u^2) } =&\left( \frac{z_{i,M}^2-1}{\rho G(\zeta )+\Pi } -\zeta \frac{\partial z_{i,M}}{\partial \zeta }\frac{1}{\rho (G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+1)}\right. \\&\quad \times \left. \left( G(\zeta )\zeta -1-\zeta z_{i,M}-\frac{(G(\zeta )\zeta -1)\tilde{n}^{1}\sqrt{z_{i,M}^2-1}}{\rho G(\zeta )+\Pi }\right) \right) P_{M}^{(1)}(-z_{i,M};\zeta ) \\&+\left( \frac{z_{i,M}^2-1}{\rho G(\zeta )+\Pi } + \zeta \frac{\partial z_{i,M}}{\partial \zeta }\frac{1}{\rho (G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+1)}\right. \\&\quad \times \left. \left( G(\zeta )\zeta -1+\zeta z_{i,M} + \frac{(G(\zeta )\zeta -1)\tilde{n}^{1}\sqrt{z_{i,M}^2-1}}{\rho G(\zeta )+\Pi }\right) \right) P_{M}^{(1)}(z_{i,M};\zeta ). \end{aligned}$$

Only a simple case is discussed in the following. As shown in Remark 2, at the local thermodynamic equilibrium, \(\Pi =0\) and \(n^\alpha =0\), thus one has

$$\begin{aligned} \frac{ \nabla _{\mathbf {W}_{M}}\lambda _{i,M}\cdot \mathbf {r}_{i,M} }{-(1-u^2)} =&\left( \frac{z_{i,M}^2-1}{\rho G(\zeta )} -\zeta \frac{\partial z_{i,M}}{\partial \zeta } \frac{G(\zeta )\zeta -1-\zeta z_{i,M} }{\rho (G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+1)}\right) P_{M}^{(1)}(-z_{i,M};\zeta ) \\&+\left( \frac{z_{i,M}^2-1}{\rho G(\zeta )} +\zeta \frac{\partial z_{i,M}}{\partial \zeta } \frac{G(\zeta )\zeta -1+\zeta z_{i,M} }{\rho (G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+1)}\right) P_{M}^{(1)}(z_{i,M};\zeta ). \end{aligned}$$

Using the term

$$\begin{aligned} \frac{\rho (G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+1)}{(z_{i,M}^2-1)(G(\zeta )\zeta -1) P_{M}^{(1)}(-z_{i,M};\zeta )}, \end{aligned}$$

to normalize the above identity and noting that

$$\begin{aligned} \mathrm{sign}(G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+1)=-\mathrm{sign}(x_{1,2}^{(0)}x_{2,2}^{(0)})<0{,} \end{aligned}$$

gives

$$\begin{aligned} \mathrm{sign}\left( \nabla _{\mathbf {W}_{M}}\lambda _{i,M}\cdot \mathbf {r}_{i,M}\right) =(-1)^{M}\mathrm{sign}\left( \hat{g}(z_{i,M};\zeta )\right) , \end{aligned}$$

where \(\hat{g}(z_{i,M};\zeta )\) is defined by

$$\begin{aligned} \hat{g}(z_{i,M};\zeta )=&\frac{G(\zeta )^2\zeta ^2-3G(\zeta )\zeta -\zeta ^2+1}{ G(\zeta )\left( G(\zeta )\zeta -1\right) }\left( 1+\frac{P_{M}^{(1)}(z_{i,M};\zeta )}{P_{M}^{(1)}(-z_{i,M};\zeta )}\right) \\&-\frac{\zeta }{(z_{i,M})^2-1} \left( 1-\frac{P_{M}^{(1)}(z_{i,M};\zeta )}{P_{M}^{(1)}(-z_{i,M};\zeta )}\right) \frac{\partial z_{i,M}}{\partial \zeta }\\&+\frac{\zeta ^2 z_{i,M}}{ \big ( (z_{i,M})^2-1\big ) \left( G(\zeta )\zeta -1\right) }\left( 1+\frac{P_{M}^{(1)}(z_{i,M};\zeta )}{P_{M}^{(1)}(-z_{i,M};\zeta )}\right) \frac{\partial z_{i,M}}{\partial \zeta }, \ i\ge 1, \end{aligned}$$

and \(\hat{g}(z_{i,M};\zeta ):=\hat{g}(z_{-i,M};\zeta )\) for \(i\le -1\). It is relatively easy to judge by numerical experiments whether the sign of \(\hat{g}(z_{i,M};\zeta )\) is constant or not. Figure 6 shows plots of \(\hat{g}(z_{1,4};\zeta )\) and \(\hat{g}(z_{1,7};\zeta )\) in terms of \(\zeta \). Similar to the special case of \(M=4\) and 7, our observation in numerical experiments is that the sign of \(\hat{g}(z_{1,M};\zeta )\) is not constant when \(M\ge 4\) so that both \(\lambda _{1,M}\) and \(\lambda _{-1,M}\) characteristic fields are neither linearly degenerate nor genuinely nonlinear when \(M\ge 4\). Such phenomenon is still not found in the case of \(M\le 3\).

Fig. 6
figure 6

Plots of \(\hat{g}(z_{1,M};\zeta )\) in terms of \(\zeta \) for \(M=4\) (left) and 7 (right)

1.7 Proof of Theorem 5.4

Proof

Because the matrix \(\mathbf {D}_M\) in (4.14) at \(\mathbf {W}_{M}=\mathbf {W}_{M}^{(0)}\) can be reformed as follows

$$\begin{aligned} \mathbf {D}_{M}=\begin{pmatrix} \mathbf {D}_{3\times 3}^{11} &{}\mathbf {D}_{3\times 2}^{12} &{} \mathbf {O}\\ \mathbf {O} &{} \mathbf {D}_{2\times 2}^{22} &{} \mathbf {O}\\ \mathbf {O} &{} \mathbf {O} &{} \mathbf {I}_{2M-4} \end{pmatrix}, \end{aligned}$$

and its inverse is given by

$$\begin{aligned} \mathbf {D}_{M}^{-1}=\begin{pmatrix} \left( \mathbf {D}_{3\times 3}^{11}\right) ^{-1} &{} -\left( \mathbf {D}_{3\times 3}^{11}\right) ^{-1}\mathbf {D}_{3\times 2}^{12}\left( \mathbf {D}_{2\times 2}^{22}\right) ^{-1} &{} \mathbf {O}\\ \mathbf {O} &{} \left( \mathbf {D}_{2\times 2}^{22}\right) ^{-1} &{} \mathbf {O}\\ \mathbf {O} &{} \mathbf {O} &{} \mathbf {I}_{2M-4} \end{pmatrix}, \end{aligned}$$

as well as

$$\begin{aligned} \tilde{\mathbf {D}}_{M}^{W}=\begin{pmatrix} \mathbf {O} &{}\mathbf {D}_{3\times 2}^{12} &{} \mathbf {O}\\ \mathbf {O} &{} \mathbf {D}_{2\times 2}^{22} &{} \mathbf {O}\\ \mathbf {O} &{} \mathbf {O} &{} \mathbf {I}_{2M-4} \end{pmatrix}, \end{aligned}$$

the product of \(\tilde{\mathbf {D}}_{M}^{W}\) and \(\mathbf {D}_{M}^{-1}\) is of the form

$$\begin{aligned} \tilde{\mathbf {D}}_{M}^{W}\mathbf {D}_{M}^{-1}=\begin{pmatrix} \mathbf {O}_{3\times 3}&{} \mathbf {D}_{3\times 2}^{12}\left( \mathbf {D}_{2\times 2}^{22}\right) ^{-1} &{} \mathbf {O}_{3\times (2M-4)}\\ \mathbf {O}_{2\times 3}&{} \mathbf {I}_{2}&{} \mathbf {O}_{2\times (2M-4)}\\ \mathbf {O}_{(2M-4)\times 3}&{}\mathbf {O}_{(2M-4)\times 2}&{}\mathbf {I}_{2M-4} \end{pmatrix}, \end{aligned}$$

where \(\mathbf {D}_{3\times 3}^{11}\) is the \(3\times 3\) subblock of \(\mathbf {D}_{2}\) in the upper left corner, \(\mathbf {D}_{3\times 2}^{12}\) denotes the \(3\times 2\) subblock of \(\mathbf {D}_{2}\) in the upper right corner, and \(\mathbf {D}_{2\times 2}^{22}\) is \(2\times 2\) subblock of \(\mathbf {D}_{2}\) in the bottom right corner. It is obvious that each eigenvalue of \(-\tilde{\mathbf {D}}_{M}^{W}\mathbf {D}_{M}^{-1}\) is non-positive, so does the matrix

$$\begin{aligned} \bar{\mathbf {Q}}_{M}:=-\frac{1}{\tau }\left( U^{0}\mathbf {M}_{M}^{t}-U^{1}\mathbf {M}_{M}^{x}\right) ^{\frac{1}{2}}\tilde{\mathbf {D}}_{M}^{W}\mathbf {D}_{M}^{-1}\left( U^{0}\mathbf {M}_{M}^{t}-U^{1}\mathbf {M}_{M}^{x}\right) ^{-\frac{1}{2}}. \end{aligned}$$

The matrix \( U^{0}\mathbf {M}_{M}^{t}-U^{1}\mathbf {M}_{M}^{x} \) can be written as follows

$$\begin{aligned} \begin{pmatrix} \mathbf {M}_{3\times 3}^{11}&{}\mathbf {M}_{3\times 2}^{12}&{}\mathbf {O}_{3,2M-4}\\ (\mathbf {M}_{3\times 2}^{12})^{T}&{}\mathbf {M}_{2\times 2}^{22}&{}\mathbf {M}_{2\times (2M-4)}^{23}\\ \mathbf {O}_{2M-4,3}&{}(\mathbf {M}_{2\times (2M-4)}^{23})^{T}&{}\mathbf {M}_{(2M-4)\times (2M-4)}^{33} \end{pmatrix}, \end{aligned}$$

where \(\mathbf {M}_{3\times 3}^{11}\) is the \(3\times 3\) subblock of \(\mathbf {P}_{2}^{p}\mathbf {A}_{2}^{0}(\mathbf {P}_{2}^{p})^{T}\) in the upper left corner, \(\mathbf {M}_{3\times 2}^{12}\) denotes the \(3\times 2\) subblock of \(\mathbf {P}_{2}^{p}\mathbf {A}_{2}^{0}(\mathbf {P}_{2}^{p})^{T}\) in the upper right corner, and \(\mathbf {M}_{2\times 2}^{22}\) is \(2\times 2\) subblock of \(\mathbf {P}_{2}^{p}\mathbf {A}_{2}^{0}(\mathbf {P}_{2}^{p})^{T}\) in the bottom right corner, the rest subblocks form the \((2M-2)\times (2M-2)\) bottom right corner of \(\mathbf {P}_{2}^{p}\mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T}\). Thus one has

$$\begin{aligned} \mathbf {M}_{D}:=(\mathbf {M}_{3\times 2}^{12})^{T}\mathbf {D}_{3\times 2}^{12}\left( \mathbf {D}_{2\times 2}^{22}\right) ^{-1} =-\left( \mathbf {D}_{3\times 2}^{12}\left( \mathbf {D}_{2\times 2}^{22}\right) ^{-1}\right) ^{T}\mathbf {M}_{3\times 3}^{11} \left( \mathbf {D}_{3\times 2}^{12}\left( \mathbf {D}_{2\times 2}^{22}\right) ^{-1}\right) , \end{aligned}$$

which is symmetric because \( \mathbf {M}_{3\times 3}^{11}\mathbf {D}_{3\times 2}^{12}\left( \mathbf {D}_{2\times 2}^{22}\right) ^{-1}+ \mathbf {M}_{3\times 2}^{12}=\mathbf {O}_{3\times 2}. \)

On the other hand, because the first three components of \(\mathbf {S}(\mathbf {W}_M)\) are zero, all elements in the first three rows and the first three columns of the matrix

$$\begin{aligned} \mathbf {Q}_{M}=-\frac{1}{\tau }\left( U^{0}\mathbf {M}_{M}^{t}-U^{1}\mathbf {M}_{M}^{x}\right) \tilde{\mathbf {D}}_{M}^{W}\mathbf {D}_{M}^{-1}, \end{aligned}$$

are zero, and the matrix \(\mathbf {Q}_{M}\) is of form

$$\begin{aligned} \mathbf {Q}_{M}=-\frac{1}{\tau } \begin{pmatrix} \mathbf {O}_{3,3}&{}\quad \mathbf {O}_{3,2}&{}\quad \mathbf {O}_{3,2M-4}\\ \mathbf {O}_{2,3}&{}\quad \mathbf {M}_{2\times 2}^{22}+\mathbf {M}_{D}&{}\quad \mathbf {M}_{2\times (2M-4)}^{23}\\ \mathbf {O}_{2M-4,3}&{}\quad (\mathbf {M}_{2\times (2M-4)}^{23})^{T}&{}\quad \mathbf {M}_{(2M-4)\times (2M-4)}^{33} \end{pmatrix}. \end{aligned}$$

Hence the matrix \(\mathbf {Q}_{M}\) is symmetric. It is obvious that \(\mathbf {Q}_{M}\) is congruent with \(\bar{\mathbf {Q}}_{M}\), so it is negative semi-definite.

Because both matrices \(\mathbf {D}_{M}\) and \(\mathbf {M}_{M}^{t}\) are invertible, and \(\mathbf {M}_{M}^{t}\) is positive definite, (5.4) is equivalent to

$$\begin{aligned} \det \left( i\omega \mathbf {I}-ik\mathbf {M}_{M}-\hat{\mathbf {Q}}_{M}\right) =0, \end{aligned}$$
(11.2)

where

$$\begin{aligned} \hat{\mathbf {Q}}_{M}:=&\left( \mathbf {M}_{M}^{t}\right) ^{-\frac{1}{2}}\mathbf {Q}_{M}\left( \mathbf {M}_{M}^{t}\right) ^{-\frac{1}{2}}, \end{aligned}$$

and

$$\begin{aligned} \mathbf {M}_{M}:=\left( \mathbf {M}_{M}^{t}\right) ^{-\frac{1}{2}}\mathbf {M}_{M}^{x}\left( \mathbf {M}_{M}^{t}\right) ^{-\frac{1}{2}}. \end{aligned}$$

It is obvious that the matrix \(\hat{\mathbf {Q}}_{M}\) is congruent with \(\mathbf {Q}_{M}\) and negative semi-definite, and \(\mathbf {M}_{M}\) is symmetric. Using Lemmas 1 and 2 in [15] completes the proof. \(\square \)

1.8 Proof of Lemma 10

Proof

  1. (i)

    Under the given Lorentz boost (x direction)

    $$\begin{aligned} t'=\gamma (v) (t-vx),\ x'=\gamma (v)(x-vt),\ \gamma (v)=(1-v^2)^{-\frac{1}{2}}, \end{aligned}$$

    where v is the relative velocity between frames in the x-direction, one has

    $$\begin{aligned}&(p^{0})'=\gamma (v)(p^{0}-p^{1}v),\quad (p^{1})'=\gamma (v)(p^{1}-p^{0}v), \\&(U^{0})'=\gamma (v)(U^{0}-U^{1}v),\quad (U^{1})'=\gamma (v)(U^{1}-U^{0}v). \end{aligned}$$

    Thus one further obtains

    $$\begin{aligned} E'= (U^{0})'(p^{0})'-(U^{1})'(p^{1})'=U^{0}p^{0}-U^{1}p^{1}=E, \end{aligned}$$

    and

    $$\begin{aligned} \left( \frac{p_{{\langle }1{\rangle }}}{U^{0}}\right) '=&\frac{-(p^{{\langle }1{\rangle }})'}{(U^{0})'}=-\frac{p^{{\langle }0{\rangle }}-p^{{\langle }1{\rangle }}v}{U^{1}-U^{0}v}=-\frac{(U^{0})^{-1}U^{1}p^{{\langle }1{\rangle }}-p^{{\langle }1{\rangle }}v}{U^{1}-U^{0}v}=\frac{p_{{\langle }1{\rangle }}}{U^{0}}, \\ \left( \frac{dp}{p^{0}}\right) '=&\frac{d(p^{1})'}{(p^{0})'}=\frac{dp^{0}-dp^{1}v}{p^{1}-p^{0}v}=\frac{(p^{0})^{-1}p^{1}dp^{1}-dp^{1}v}{p^{1}-p^{0}v}=\frac{dp}{p^{0}}. \end{aligned}$$

    Combining them with (4.9) gives that each component of \(\mathbf {f}_{M}\) is Lorentz invariant, so that the last \((2M-4)\) components of \(\mathbf {W}_{M}\) are also Lorentz invariant. From (2.17) and (2.20), it is not difficult to prove that \(\rho \) and \(\theta \) are Lorentz invariant. On the other hand, because

    $$\begin{aligned} \tilde{n}^{1}=\int _{\mathbb {R}}\frac{p^{{\langle }1{\rangle }}}{U^{0}}f\frac{dp}{p^{0}}, \end{aligned}$$

    the quantity \(\tilde{n}^{1}\) is Lorentz invariant. Moreover, one has

    $$\begin{aligned} \left( \frac{du}{1-u^2}\right) '=\frac{d(U^{1})'}{(U^{0})'}=\frac{dU^{0}-dU^{1}v}{U^{1}-U^{0}v}=\frac{(U^{0})^{-1}U^{1}dU^{1}-dU^{1}v}{U^{1}-U^{0}v}=\frac{dU^{1}}{U^{0}}=\frac{du}{1-u^2}. \end{aligned}$$

    Using the above results completes the proof of the first part.

  2. (ii)

    Because \(\mathbf {A}_{M}^{0}\) and \(\mathbf {A}_{M}^{1}\) only depend on \(\theta \), they are Lorentz invariant. The source term \(\mathbf {S}(\mathbf {W}_{M})\) in (4.21) can be rewritten into

    $$\begin{aligned} \mathbf {S}(\mathbf {W}_{M})=-\frac{1}{\tau }\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T}\left( \mathbf {f}_{M}-\mathbf {f}^{(0)}_{M}\right) , \end{aligned}$$

    which has been expressed in terms of the Lorentz covariant quantities. In fact, the general source term \(\mathbf {S}(\mathbf {W}_{M})\) in the moment system (4.20) is also Lorentz invariant. The proof is completed.

\(\square \)

1.9 Proof of Theorem 5.5

Proof

From the 3rd step in Sec. 4.2 and Lemma 10, one knows that \(\hat{\mathbf {D}}_{M}=\mathbf {D}_M(\mathbf {D}_{M}^{u})^{-1}\) can be expressed in terms of the Lorentz covariant quantities, so it is Lorentz invariant. Because

$$\begin{aligned} (\mathbf {M}_{M}^{t})'&=-\gamma (v)(U^{1}-U^{0}v)\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{1}(\mathbf {P}_{M}^{p})^{T}+\gamma (v)(U^{0}-U^{1}v)\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T},\\ (\mathbf {M}_{M}^{x})'&=-\gamma (v)(U^{0}-U^{1}v)\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{1}(\mathbf {P}_{M}^{p})^{T}+\gamma (v)(U^{1}-U^{0}v)\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T}, \end{aligned}$$

and

$$\begin{aligned} \left( \frac{\partial }{\partial t}\right) '=\gamma (v)\left( \frac{\partial }{\partial t}+v\frac{\partial }{\partial x}\right) ,\quad \left( \frac{\partial }{\partial x}\right) '=\gamma (v)\left( \frac{\partial }{\partial x}+v\frac{\partial }{\partial t}\right) , \end{aligned}$$

one has

$$\begin{aligned} \left( \mathbf {D}_{M}^{u}\frac{\partial \mathbf {W}_{M}}{\partial t}\right) '=\,&\mathrm{diag}\left\{ 1,\big ((U^{0})^2\big )',1,\ldots ,1\right\} \gamma (v)\left( \frac{\partial (\rho ,u',\theta ,\Pi ,\tilde{n}^{1},f_{3}^{0},\ldots ,f_{M-1}^{1})^{T}}{\partial t}\right. \\&+\left. v\frac{\partial (\rho ,u',\theta ,\Pi ,\tilde{n}^{1},f_{3}^{0},\ldots ,f_{M-1}^{1})^{T}}{\partial x}\right) \\ =\,&\mathrm{diag}\left\{ 1,\big ((U^{0}\big )^{-1})',1,\ldots ,1\right\} \gamma (v)\left( \frac{\partial (\rho ,(U^{1})',\theta ,\Pi ,\tilde{n}^{1},f_{3}^{0},\ldots ,f_{M-1}^{1})^{T}}{\partial t}\right. \\&+\left. v\frac{\partial (\rho ,(U^{1})',\theta ,\Pi ,\tilde{n}^{1},f_{3}^{0},\ldots ,f_{M-1}^{1})^{T}}{\partial x}\right) \\ =\,&\mathbf {D}_{M}^{u}\gamma (v)\left( \frac{\partial \mathbf {W}_{M}}{\partial t}+v\frac{\partial \mathbf {W}_{M}}{\partial x}\right) , \end{aligned}$$

where the last equal sign is derived by following the proof of Lemma 10. Similarly, one has

$$\begin{aligned} \left( \mathbf {D}_{M}^{u}\frac{\partial \mathbf {W}_{M}}{\partial x}\right) '=\mathbf {D}_{M}^{u}\gamma (v)\left( \frac{\partial \mathbf {W}_{M}}{\partial x}+v\frac{\partial \mathbf {W}_{M}}{\partial t}\right) . \end{aligned}$$

Thus it holds

$$\begin{aligned}&\left( \mathbf {B}_{M}^{0}\frac{\partial \mathbf {W}_{M}}{\partial t} + \mathbf {B}_{M}^{1}\frac{\partial \mathbf {W}_{M}}{\partial x}\right) '\\&\quad =\,(\mathbf {M}_{M}^{t})'\left( \mathbf {D}_{M}\frac{\partial \mathbf {W}_{M}}{\partial t}\right) '+(\mathbf {M}_{M}^{x})'\left( \mathbf {D}_{M}\frac{\partial \mathbf {W}_{M}}{\partial x}\right) '\\&\quad =\left( -(U^{1})'\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{1}(\mathbf {P}_{M}^{p})^{T}+ (U^{0})'P_{M}^{p}\mathbf {A}_{M}^{0}(P_{M}^{p})^{T}\right) \mathbf {D}_{M}\left( \gamma (v)\left( \frac{\partial \mathbf {W}_{M}}{\partial t}+v\frac{\partial \mathbf {W}_{M}}{\partial x}\right) \right) \\&\qquad +\,\left( -(U^{0})'\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{1}(\mathbf {P}_{M}^{p})^{T}+ (U^{1})'\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T}\right) \mathbf {D}_{M}\left( \gamma (v)\left( \frac{\partial \mathbf {W}_{M}}{\partial x}+v\frac{\partial \mathbf {W}_{M}}{\partial t}\right) \right) \\&\quad =\left( -U^{1}v\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{1}(\mathbf {P}_{M}^{p})^{T}+U^{0}\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T}\right) \mathbf {D}_{M}\frac{\partial \mathbf {W}_{M}}{\partial t}\\&\qquad +\,\left( -U^{0}\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{1}(\mathbf {P}_{M}^{p})^{T}+U^{1}\mathbf {P}_{M}^{p}\mathbf {A}_{M}^{0}(\mathbf {P}_{M}^{p})^{T}\right) \mathbf {D}_{M}\frac{\partial \mathbf {W}_{M}}{\partial x}\\&\quad =\mathbf {B}_{M}^{0}\frac{\partial \mathbf {W}_{M}}{\partial t} + \mathbf {B}_{M}^{1}\frac{\partial \mathbf {W}_{M}}{\partial x}. \end{aligned}$$

Combining it with Lemma 10 completes the proof. \(\square \)

Appendix 5: Proofs in Section 6

1.1 Proof of Lemma 11

Proof

Using Lemmas 4 and 7 gives

$$\begin{aligned} \tilde{f}g^{(0)}_{[u,\theta ]}=&\sum _{i=0}^{M}\tilde{f}_{i}^{0}\tilde{P}_{i}^{(0)}[u,\theta ]+\sum _{j=0}^{M-1}\tilde{f}_{j}^{1}\tilde{P}_{j}^{(1)}[u,\theta ],\\ \Pi _{M}[u,\theta ] f=&\sum _{i=0}^{M}f_{i}^{0}\tilde{P}_{i}^{(0)}[u,\theta ]+\sum _{j=0}^{M-1}f_{j}^{(1)}\tilde{P}_{j}^{(1)}[u,\theta ], \end{aligned}$$

where

$$\begin{aligned} f_{i}^{0}&={\langle }f,\tilde{P}_{i}^{(0)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}},\quad \tilde{f}_{i}^{0}={\langle }\tilde{f}g^{(0)}_{[u,\theta ]},\tilde{P}_{i}^{(0)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}},\quad i\le M,\\ f_{j}^{1}&={\langle }f,\tilde{P}_{j}^{(1)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}},\quad \tilde{f}_{j}^{1}={\langle }\tilde{f}g^{(0)}_{[u,\theta ]},\tilde{P}_{j}^{(1)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}},\quad j\le M-1. \end{aligned}$$

Therefore one has

$$\begin{aligned}&{\langle }\tilde{f}g^{(0)}_{[u,\theta ]},\Pi _{M}[u,\theta ]f{\rangle }_{g^{(0)}_{[u,\theta ]}}={\langle }\tilde{f}f,\Pi _{M}[u,\theta ]f{\rangle }_{f}\\&\quad =\sum _{i=0}^{M}f_{i}^{0}{\langle }\tilde{f}f,\tilde{P}_{i}^{(0)}[u,\theta ]{\rangle }_{f}+ \sum _{j=0}^{M-1}f_{j}^{1}{\langle }\tilde{f}f,\tilde{P}_{j}^{(1)}[u,\theta ]{\rangle }_{f}\\&\quad =\sum _{i=0}^{M}f_{i}^{0}{\langle }\tilde{f}g^{(0)}_{[u,\theta ]},\tilde{P}_{i}^{(0)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}}+ \sum _{j=0}^{M-1}f_{j}^{1}{\langle }\tilde{f}g^{(0)}_{[u,\theta ]},\tilde{P}_{j}^{(1)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}}\\&\quad =\sum _{i=0}^{M}f_{i}^{0}\tilde{f}_{i}^{0}+ \sum _{j=0}^{M-1}f_{j}^{1}\tilde{f}_{j}^{1}\\&\quad =\sum _{i=0}^{M}{\langle }f,\tilde{f}_{i}^{0}\tilde{P}_{i}^{(0)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}}+ \sum _{j=0}^{M-1}{\langle }f,\tilde{f}_{j}^{1}\tilde{P}_{j}^{(1)}[u,\theta ]{\rangle }_{g^{(0)}_{[u,\theta ]}}\\&\quad ={\langle }f,\tilde{f}g^{(0)}_{[u,\theta ]}{\rangle }_{g^{(0)}_{[u,\theta ]}}={\langle }f,\tilde{f}f{\rangle }_{f}. \end{aligned}$$

The proof is completed. \(\square \)

1.2 Proof of Lemma 12

Proof

Using Lemma 7 gives

$$\begin{aligned}&\Pi _{M}[u_{1},\theta _{1}]f=\sum _{i=0}^{M}f_{i}^{0}\tilde{P}_{i}^{(0)}[u_{1},\theta _{1}]+ \sum _{j=0}^{M-1}f_{j}^{1}\tilde{P}_{i}^{(1)}[u_{1},\theta _{1}],\\&\Pi _{M}[u_{1},\theta _{1}]\Pi _{M}[u_{2},\theta _{2}]f=\sum _{i=0}^{M}\tilde{f}_{i}^{0}\tilde{P}_{i}^{(0)}[u_{1},\theta _{1}]+ \sum _{j=0}^{M-1}\tilde{f}_{j}^{1}\tilde{P}_{j}^{(1)}[u_{1},\theta _{1}], \end{aligned}$$
$$\begin{aligned} f_{i}^{0}&={\langle }f,\tilde{P}_{i}^{(0)}[u_{1},\theta _{1}]{\rangle }_{g^{(0)}_{[u_{1},\theta _{1}]}}= {{\langle }f,{P}_{i}^{(0)}(u_{1},\zeta _{1})f{\rangle }_{f}},i\le M,\\ f_{j}^{1}&={\langle }f,\tilde{P}_{j}^{(1)}[u_{1},\theta _{1}]{\rangle }_{g^{(0)}_{[u_{1},\theta _{1}]}}= {{\langle }f,{P}_{j}^{(1)}(u_{1},\zeta _{1})(U_{0})_{1}^{-1}p_{{\langle }1{\rangle }}f{\rangle }_{f}},j\le M-1, \end{aligned}$$
$$\begin{aligned} \tilde{f}_{i}^{0}&={\langle }\Pi _{M}[u_{2},\theta _{2}]f,\tilde{P}_{i}^{(0)}[u_{1},\theta _{1}]{\rangle }_{g^{(0)}_{[u_{1},\theta _{1}]}}= {{\langle }\Pi _{M}[u_{2},\theta _{2}]f,{P}_{i}^{(0)}(u_{1},\zeta _{1})f{\rangle }_{f}},i\le M,\\ \tilde{f}_{j}^{1}&={\langle }\Pi _{M}[u_{2},\theta _{2}]f,\tilde{P}_{j}^{(1)}[u_{1},\theta _{1}]{\rangle }_{g^{(0)}_{[u_{1},\theta _{1}]}}\\&= {{\langle }\Pi _{M}[u_{2},\theta _{2}]f,{P}_{j}^{(1)}(u_{1},\zeta _{1})(U_{0})_{1}^{-1}p_{{\langle }1{\rangle }}f{\rangle }_{f}},j\le M-1{.} \end{aligned}$$

Because both \({P}_{i}^{(0)}(u_{1},\zeta _{1})f\) and \({P}_{j}^{(1)}(u_{1},\zeta _{1})(U_{0})_{1}^{-1}p_{{\langle }1{\rangle }}f\) belong to the space \(\mathbb {H}_{M}^{f}\), using Lemma 11 completes the proof. \(\square \)

1.3 Proof of Theorem 6.1

Proof

Because Eq. (6.3) is equivalent to

$$\begin{aligned} \left( \mathbf {I}+\frac{\Delta t}{\tau _{i}^{*}}\left( \mathbf {M}_{i,M}^{t*}\right) ^{-1}\left( U_{i}^{0*}\mathbf {M}_{i,M}^{t*}-U_{i}^{1*} \mathbf {M}_{i,M}^{x*}\right) \left( \mathbf {I}-\mathbf {D}_{i,M}^{f_{i}^{(0)*}}\right) \right) \mathbf {f}_{i,M}^{n+1}=\mathbf {f}_{i,M}^{*}, \end{aligned}$$

it is unconditionally stable if and only if the modulus of each eigenvalue of the matrix

$$\begin{aligned} \mathbf {I}+\frac{\Delta t}{\tau _{i}^{*}}\left( \mathbf {M}_{i,M}^{t*}\right) ^{-1}\left( U_{i}^{0*}\mathbf {M}_{i,M}^{t*}-U_{i}^{1*} \mathbf {M}_{i,M}^{x*}\right) \left( \mathbf {I}-\mathbf {D}_{i,M}^{f_{i}^{(0)*}}\right) , \end{aligned}$$

is not less than one. It is true if the real part of each eigenvalue of the matrix

$$\begin{aligned} \left( \mathbf {M}_{i,M}^{t*}\right) ^{-1}\left( U_{i}^{0*}\mathbf {M}_{i,M}^{t*}-U_{i}^{1*}\mathbf {M}_{i,M}^{x*}\right) \left( \mathbf {I}-\mathbf {D}_{i,M}^{f_{i}^{(0)*}}\right) =:\left( \mathbf {M}_{i,M}^{t*}\right) ^{-1}\bar{\mathbf {M}}_{D}^{*}, \end{aligned}$$
(12.1)

is non-negative.

In fact, thanks to (6.4), the characteristic polynomial of the upper triangular matrix \( \mathbf {I}-\mathbf {D}_{i,M}^{f_{i}^{(0)*}}\) is explicitly given by

$$\begin{aligned} 0=\det \left( \lambda \mathbf {I}-\left( \mathbf {I}-\mathbf {D}_{i,M}^{f_{i}^{(0)*}}\right) \right) = \det \left( (\lambda -1)\mathbf {I}+\mathbf {D}_{i,M}^{f^{(0)*}}\right) =\lambda (\lambda -1)^{2M}, \end{aligned}$$

and \(\bar{\mathbf {M}}_{D}^{*}\) is a symmetric matrix and congruent with

$$\begin{aligned} \left( U_{i}^{0*}\mathbf {M}_{i,M}^{t*}-U_{i}^{1*}\mathbf {M}_{i,M}^{x*}\right) ^{\frac{1}{2}} \left( \mathbf {I}-\mathbf {D}_{i,M}^{f_{i}^{(0)*}}\right) \left( U_{i}^{0*} \mathbf {M}_{i,M}^{t*}-U_{i}^{1*}\mathbf {M}_{i,M}^{x*}\right) ^{-\frac{1}{2}}, \end{aligned}$$

which is similar to the matrix \( \mathbf {I}-\mathbf {D}_{i,M}^{f_{i}^{(0)*}}\). Thus the matrix \(\bar{\mathbf {M}}_{D}^{*}\) is positive semi-definite and each eigenvalue of the matrix \( \left( \mathbf {M}_{i,M}^{t*}\right) ^{-1}\bar{\mathbf {M}}_{D}^{*}\) is non-negative because of the relation \( \left( \mathbf {M}_{i,M}^{t*}\right) ^{-1}\bar{\mathbf {M}}_{D}^{*} =\left( \mathbf {M}_{i,M}^{t*}\right) ^{-\frac{1}{2}} \big ( \left( \mathbf {M}_{i,M}^{t*}\right) ^{-\frac{1}{2}} \bar{\mathbf {M}}_{D}^{*} \left( \mathbf {M}_{i,M}^{t*}\right) ^{-\frac{1}{2}}\big ) \left( \mathbf {M}_{i,M}^{t*}\right) ^{\frac{1}{2}}\). The proof is completed. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kuang, Y., Tang, H. Globally Hyperbolic Moment Model of Arbitrary Order for One-Dimensional Special Relativistic Boltzmann Equation. J Stat Phys 167, 1303–1353 (2017). https://doi.org/10.1007/s10955-017-1773-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1773-3

Keywords

Mathematics Subject Classification

Navigation