Skip to main content
Log in

Symplectic Geometry and Spectral Properties of Classical and Quantum Coupled Angular Momenta

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We give a detailed study of the symplectic geometry of a family of integrable systems obtained by coupling two angular momenta in a non-trivial way. These systems depend on a parameter \(t \in [0,1]\) and exhibit different behaviors according to its value. For a certain range of values, the system is semitoric, and we compute some of its symplectic invariants. Even though these invariants have been known for almost a decade, this is to our knowledge the first example of their computation in the case of a non-toric semitoric system on a compact manifold. (The only invariant of toric systems is the image of the momentum map.) In the second part of the paper, we quantize this system, compute its joint spectrum and describe how to use this joint spectrum to recover information about the symplectic invariants.

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
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  • Alonso, J., Dullin, H.R., Hohloch, S.: Taylor Series and Twisting-Index Invariants of Coupled Spin-oscillators (2017). Preprint arXiv:1712.06402

  • Atiyah, M.F.: Convexity and commuting Hamiltonians. Bull. Lond. Math. Soc. 14, 1–15 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  • Bloch, A., Golse, F., Paul, T., Uribe, A.: Dispersionless Toda and Toeplitz operators. Duke Math. J. 117(1), 157–196 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  • Bolsinov, A.V., Fomenko, A.T.: Integrable Hamiltonian Systems. Chapman & Hall/CRC, Boca Raton (2004). Geometry, Topology, Classification, Translated from the 1999 Russian original

  • Boutet de Monvel, L., Guillemin, V.: The Spectral Theory of Toeplitz Operators, volume 99 of Annals of Mathematics Studies. Princeton University Press, Princeton (1981)

  • Charles, L.: Berezin–Toeplitz operators, a semi-classical approach. Comm. Math. Phys. 239(1–2), 1–28 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  • Charles, L.: Symbolic calculus for Toeplitz operators with half-form. J. Symplectic Geom. 4(2), 171–198 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  • Dullin, H.R.: Semi-global symplectic invariants of the spherical pendulum. J. Differ. Equ. 254(7), 2942–2963 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Eliasson, L.H.: Hamiltonian Systems with Poisson Commuting Integrals. PhD thesis, University of Stockholm (1984)

  • Garay, M.D., van Straten, D.: Classical and quantum integrability. Mosc. Math. J. 10(3), 519–545, 661 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Guillemin, V., Sternberg, S.: Convexity properties of the moment mapping. Invent. Math. 67, 491–513 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  • Hohloch, S., Palmer, J.: A family of compact semitoric systems with two focus-focus singularities. J. Geom. Mech. 10(3), 331–357 (2018). https://doi.org/10.3934/jgm.2018012

    Article  Google Scholar 

  • Kostant, B.: Quantization and unitary representations. Uspehi Mat. Nauk. 28(1(169)):163–225 (1973). Translated from the English (Lectures in Modern Analysis and Applications, III, pp. 87–208, Lecture Notes in Math., Vol. 170, Springer, Berlin, 1970) by A. A. Kirillov

  • Le Floch, Y.: Singular Bohr–Sommerfeld conditions for 1D Toeplitz operators: elliptic case. Comm. Partial Differ. Equ. 39(2), 213–243 (2014a)

    Article  MathSciNet  MATH  Google Scholar 

  • Le Floch, Y.: Singular Bohr–Sommerfeld conditions for 1D Toeplitz operators: hyperbolic case. Anal. PDE 7(7), 1595–1637 (2014b)

    Article  MathSciNet  MATH  Google Scholar 

  • Ma, X., Marinescu, G.: Toeplitz operators on symplectic manifolds. J. Geom. Anal. 18(2), 565–611 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Meyer, K.R., Hall, G.R., Offin, D.: Introduction to Hamiltonian Dynamical Systems and the \(N\)-Body Problem, volume 90 of Applied Mathematical Sciences, 2nd ed. Springer, New York (2009)

  • Pelayo, A.: Hamiltonian and symplectic symmetries: an introduction. Bull. Am. Math. Soc. (N.S.) 54(3), 383–436 (2017)

    Article  MathSciNet  Google Scholar 

  • Pelayo, Á., Vũ Ngọc, S.: Semitoric integrable systems on symplectic 4-manifolds. Invent. Math. 177(3), 571–597 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Pelayo, A., Vũ Ngọc, S.: Symplectic theory of completely integrable Hamiltonian systems. Bull. Am. Math. Soc. (N.S.) 48(3), 409–455 (2011a)

  • Pelayo, Á., Vũ Ngọc, S.: Constructing integrable systems of semitoric type. Acta Math. 206(1), 93–125 (2011b)

    Article  MathSciNet  MATH  Google Scholar 

  • Pelayo, Á., Vũ Ngọc, S.: Hamiltonian dynamics and spectral theory for spin-oscillators. Commun. Math. Phys. 309(1), 123–154 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Pelayo, Á., Ratiu, T., Vũ Ngọc, S.: The affine invariant of proper semitoric systems. Nonlinearity 30(11), 3993–4028 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Sadovskij, D., Zhilinskij, B.: Monodromy, diabolic points, and angular momentum coupling. Phys. Lett. A 256(4), 235–244 (1999)

    Article  MathSciNet  Google Scholar 

  • Sampson, J., Washnitzer, G.: A Künneth formula for coherent algebraic sheaves. Ill. J. Math. 3, 389–402 (1959)

    MATH  Google Scholar 

  • Schlichenmaier, M.: Berezin–Toeplitz quantization for compact Kähler manifolds. A review of results. Adv. Math. Phys. pages Art. ID 927280, 38 (2010)

  • Sepe, D., Vũ Ngọc, S.: Integrable systems, symmetries, and quantization. Lett. Math. Phys. 108(3), 499–571 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  • Souriau, J.-M.: Quantification géométrique. Comm. Math. Phys. 1, 374–398 (1966)

    MathSciNet  MATH  Google Scholar 

  • Toth, J.A., Zelditch, S.: \(L^p\) norms of eigenfunctions in the completely integrable case. Ann. Henri Poincaré 4(2), 343–368 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  • Vũ Ngọc, S.: Bohr–Sommerfeld conditions for integrable systems with critical manifolds of focus–focus type. Commun. Pure Appl. Math. 53(2), 143–217 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  • Vũ Ngọc, S.: On semi-global invariants for focus–focus singularities. Topology 42(2), 365–380 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  • Vũ Ngọc, S.: Moment polytopes for symplectic manifolds with monodromy. Adv. Math. 208(2), 909–934 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  • Williamson, J.: On the algebraic problem concerning the normal forms of linear dynamical systems. Am. J. Math. 58(1), 141–163 (1936)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

YLF was supported by the European Research Council Advanced Grant 338809. Part of this work was done during a visit of YLF to AP at University of California San Diego in April 2016, and he would like to thank the members of this university for their hospitality. AP is supported by NSF CAREER Grant DMS-1518420. He also received support from Severo Ochoa Program (Grant No. Severo Ochoa Group Grant Sev-2011-0087) at ICMAT in Spain. Part of this work was carried out at ICMAT and Universidad Complutense de Madrid. We thank Joseph Palmer for useful comments, and Jaume Alonso and Holger Dullin for pointing out a mistake in the computation of the coefficient \(a_1\) in the Taylor series invariant (Proposition 3.12) in an earlier version of this manuscript. We would also like to thank two anonymous referees for their very useful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yohann Le Floch.

Additional information

Communicated by Paul Newton.

Appendices

Appendix A: Proofs of Technical Results

Proof of Lemma 3.3

A standard computation shows that the symplectic form becomes

$$\begin{aligned} (\Psi ^{-1})^* \omega = \frac{2 i R_1 \mathrm{d}z \wedge d{\bar{z}}}{(1+|z|^2)^2} + \frac{2 i R_2 \mathrm{d}w \wedge d{\bar{w}}}{(1+|w|^2)^2}. \end{aligned}$$

Moreover, we deduce from Eq. (7) that

$$\begin{aligned} \frac{\partial J}{\partial z} = \frac{-2R_1 {\bar{z}}}{(1+|z|^2)^2}, \qquad \frac{\partial J}{\partial w} = \frac{2R_2 {\bar{w}}}{(1+|w|^2)^2}. \end{aligned}$$

Since J is real-valued, \(\frac{\partial J}{\partial {\bar{z}}} = \overline{\frac{\partial J}{\partial z}}\) and \(\frac{\partial J}{\partial {\bar{w}}} = \overline{\frac{\partial J}{\partial w}}\), and we obtain the desired result. Furthermore,

$$\begin{aligned} \frac{\partial H}{\partial z}= & {} \frac{1}{2} \left( \frac{-2{\bar{z}}}{(1+|z|^2)^2} \right. \\&\left. + \frac{(2w-{\bar{z}}|w|^2+{\bar{z}})(1+|z|^2) - {\bar{z}}(2(zw + {\bar{z}}{\bar{w}}) + (1-|z|^2)(|w|^2-1))}{(1+|z|^2)^2(1+|w|^2)} \right) , \end{aligned}$$

which yields, after simplification

$$\begin{aligned} \frac{\partial H}{\partial z} = \frac{w-2{\bar{z}}|w|^2-{\bar{z}}^2{\bar{w}}}{(1+|z|^2)^2(1+|w|^2)}. \end{aligned}$$

A similar computation shows that

$$\begin{aligned} \frac{\partial H}{\partial w} = \frac{z + {\bar{w}}-{\bar{w}}|z|^2-{\bar{z}}{\bar{w}}^2}{(1+|z|^2)(1+|w|^2)^2}, \end{aligned}$$

and we conclude by using the same argument as above. \(\square \)

Proof of Lemma 3.4

Let \(\lambda _1(z,w), \lambda _2(z,w)\) be as in the previous lemma, and let \((z,w) \in \Lambda _0 \setminus \{(0,0)\}\). Since \(z \ne 0\), we get, by multiplying both the numerator and the denominator in \(\lambda _1\) by \({\bar{z}}\), that

$$\begin{aligned} \lambda _1(z,w) = \frac{{\bar{z}} {\bar{w}} - 2 |z|^2 |w|^2 - zw |z|^2}{R_1 {\bar{z}}(1+|w|^2)}. \end{aligned}$$

But the second equation in (8) yields \({\bar{z}} {\bar{w}} = - zw - |w|^2 + |z|^2 |w|^2\), hence

$$\begin{aligned} \lambda _1(z,w) = \frac{-zw - |w|^2 - |z|^2 |w|^2 - zw |z|^2}{R_1 {\bar{z}}(1+|w|^2)} = -\frac{w(z+{\bar{w}})(1+|z|^2)}{R_1 {\bar{z}}(1+|w|^2)}. \end{aligned}$$

The first equation in (8) allows us to further simplify this expression and to obtain

$$\begin{aligned} \lambda _1(z,w) = -\frac{zw(z+{\bar{w}})(1+|z|^2)}{R_2 |w|^2 (1+|z|^2)} = -\frac{z(z+{\bar{w}})}{R_2 {\bar{w}}}. \end{aligned}$$

Now, since \(w \ne 0\), we have that

$$\begin{aligned} \lambda _2(z,w)= & {} \frac{{\bar{z}} |w|^2 + w |w|^2 - w |z|^2 |w|^2- z w^2 |w|^2}{R_2 |w|^2 (1+|z|^2)}\\= & {} \frac{{\bar{z}} |w|^2 - z w^2 |w|^2 + w |w|^2 (1 -|z|^2)}{R_2 |w|^2 (1+|z|^2)}. \end{aligned}$$

The second equation in (8) gives

$$\begin{aligned}&{\bar{z}} |w|^2 - z w^2 |w|^2 + w |w|^2 (1 -|z|^2) = {\bar{z}} |w|^2\\&\quad - z w^2 |w|^2 - z w^2 - {\bar{z}} |w|^2 = -zw^2(1+|w|^2), \end{aligned}$$

and thus, using the first one again, we finally obtain that \( \lambda _2(z,w) = -\frac{w^2}{R_1 {\bar{z}}}\). \(\square \)

Proof of Proposition 3.7

We start by diagonalizing A over \({\mathbb {C}} \). With our choice of parameters, it follows from Eq. (3) that

$$\begin{aligned} A = \begin{pmatrix} 0 &{} 0 &{} 0 &{} \frac{1}{2} \\ 0 &{} 0 &{} -\frac{1}{2} &{} 0 \\ 0 &{} -\frac{1}{5} &{} 0 &{} -\frac{1}{5} \\ \frac{1}{5} &{} 0 &{} \frac{1}{5} &{} 0 \end{pmatrix}. \end{aligned}$$

A has eigenvalues \(\lambda _1 = \frac{3+i}{10}, {\bar{\lambda }}_1, \lambda _2 = - \lambda _1, {\bar{\lambda }}_2\) with respective eigenvectors \(X_1, \overline{X_1}, X_2, \overline{X_2}\) where

$$\begin{aligned} X_1 = \frac{1}{2} \begin{pmatrix} 3 - i \\ -1 - 3i \\ 2i \\ 2 \end{pmatrix} , \qquad X_2 = \frac{1}{2} \begin{pmatrix} -3 + i \\ - 1 - 3i \\ -2i \\ 2 \end{pmatrix}. \end{aligned}$$

Let \(E_{\lambda }\) be the eigenspace associated with \(\lambda \), and let \(F = E_{\lambda _1} \oplus E_{{\bar{\lambda }}_1}\) and \(G = E_{\lambda _2} \oplus E_{{\bar{\lambda }}_2}\); then \(T_{m_0}M = F \oplus G\) and a real basis of F is given by

$$\begin{aligned} Y_1 = X_1 + \overline{X_1} = \begin{pmatrix} 3 \\ -1 \\ 0 \\ 2 \end{pmatrix}, \qquad Y_2 = -i(X_1 - \overline{X_1}) = \begin{pmatrix} -1 \\ -3 \\ 2 \\ 0 \end{pmatrix}. \end{aligned}$$

There exists a unique basis \((Z_1,Z_2)\) of G such that \((Y_1,Y_2,Z_1,Z_2)\) is a symplectic basis of \(T_{m_0}M\) [see for instance (Meyer et al. 2009, Lemma 3.2.3)]; in the latter, A will take the form displayed in Eq. (12). Since

$$\begin{aligned} G = \mathrm {Span}\left( \begin{pmatrix} -3 \\ -1 \\ 0 \\ 2 \end{pmatrix}, \begin{pmatrix} 1 \\ -3 \\ -2 \\ 0 \end{pmatrix} \right) , \end{aligned}$$

there exists \(a,b,c,d \in {\mathbb {R}} \) such that

$$\begin{aligned} Z_1 = \begin{pmatrix} -3a + b \\ -a - 3b \\ -2b \\ 2a \end{pmatrix} , \qquad Z_2 = \begin{pmatrix} -3c + d \\ -c - 3d \\ -2d \\ 2c \end{pmatrix}. \end{aligned}$$

Recall that \(\omega _{m_0} = - \mathrm{d}x_1 \wedge \mathrm{d}y_1 + \frac{5}{2} \mathrm{d}x_2 \wedge \mathrm{d}y_2\). Hence, we need to solve the system

$$\begin{aligned} {\left\{ \begin{array}{ll} 1 = \omega _{m_0}(Y_1,Z_1) = 6a + 18b \\ 0 = \omega _{m_0}(Y_1,Z_2) = 6c + 18d \\ 0 = \omega _{m_0}(Y_2,Z_1) = 18a - 6b \\ 1 = \omega _{m_0}(Y_2,Z_2) = 18c - 6d \end{array}\right. }. \end{aligned}$$

We find \(a = \frac{1}{60}, b = \frac{1}{20}, c = \frac{1}{20}\) and \(d = -\frac{1}{60}\), hence

$$\begin{aligned} Z_1 = \frac{1}{30} \begin{pmatrix} 0 \\ -5 \\ -3 \\ 1 \end{pmatrix} , \qquad Z_2 = \frac{1}{30} \begin{pmatrix} -5 \\ 0 \\ 1 \\ 3 \end{pmatrix}. \end{aligned}$$

Consequently, the matrix P of change of basis from the basis \({\mathcal {B}}\) associated with \((x_1,y_1,x_2,y_2)\) to the basis \((Y_1,Y_2,Z_1,Z_2)\) satisfies

$$\begin{aligned} P = \begin{pmatrix} 3 &{} -1 &{} 0 &{} -\frac{1}{6} \\ -1 &{} -3 &{} - \frac{1}{6} &{} 0 \\ 0 &{} 2 &{} -\frac{1}{10} &{} \frac{1}{30} \\ 2 &{} 0 &{} \frac{1}{30} &{} \frac{1}{10} \end{pmatrix}, \quad P^{-1} = \begin{pmatrix} \frac{1}{6} &{} 0 &{} \frac{1}{12} &{} \frac{1}{4} \\ 0 &{} -\frac{1}{6} &{} \frac{1}{4} &{} -\frac{1}{12} \\ -1 &{} -3 &{} -5 &{} 0 \\ -3 &{} 1 &{} 0 &{} 5 \end{pmatrix}, \end{aligned}$$

which yields the desired expression for the coordinates \(u_1,u_2,\xi _1,\xi _2\). Moreover, \(P^{-1}AP\) is as in Eq. (12), with \(\alpha = \frac{-3}{10}\) and \(\beta =\frac{1}{10}\); this gives

$$\begin{aligned} B = \begin{pmatrix} 1 &{} 0 \\ \frac{1}{10} &{} -\frac{3}{10} \end{pmatrix}^{-1} = \begin{pmatrix} 1 &{} 0 \\ \frac{1}{3} &{} -\frac{10}{3} \end{pmatrix}. \end{aligned}$$

This is not satisfactory since we want the lower right coefficient in this matrix to be positive. In order to obtain a B satisfying this requirement, it suffices to perform the symplectic change of coordinates \((u_1,u_2,\xi _1,\xi _2) \mapsto (-\xi _1,-\xi _2, u_1, u_2)\). \(\square \)

Appendix B: Critical Points of Corank One

We explain how to prove the claim about the critical points of corank one of \(F=(J,H_t)\) in the proof of Corollary 2.11, namely that they are non-degenerate for every \(t \in ]0,1]\). (The case \(t=0\) is clear since the system is toric up to vertical scaling.) It suffices to prove that for every E in the image of J, except the ones corresponding to critical points of corank two (which are the only ones for which \(dJ = 0\) for this system), the critical points of the restriction of \(H_t\) to the symplectic quotient \(M^{\text {red}}_E = J^{-1}(E) / S^1\) with respect to the action generated by J are non-degenerate, see Toth and Zelditch (2003, Definition 3) for a general explanation or Hohloch and Palmer (2018, Corollary 2.5) for the particular case of dimension 4. Coming back to our particular case, let \(E \in (-(R_1 + R_2), R_1 + R_2) \setminus \{ R_1 - R_2, R_2 - R_1 \}\); since the poles do not give rise to critical points on \(J^{-1}(E)\), we may work with cylindrical coordinates as in Sect. 3.5 [or as in Hohloch and Palmer (2018, Section 3.3)] where a similar computation is performed)

$$\begin{aligned} (x_j, y_j, z_j) = \left( \sqrt{1-z_j^2} \cos \theta _j, \sqrt{1-z_j^2} \sin \theta _j, z_j \right) , \quad j = 1,2. \end{aligned}$$

In these coordinates, \(H_t\) reads

$$\begin{aligned} H_t(\theta _1, z_1, \theta _2, z_2) = (1-t) z_1 + t \left( \sqrt{(1-z_1^2)(1-z_2^2)} \cos (\theta _1 - \theta _2) + z_1 z_2 \right) . \end{aligned}$$

Since \(z_2\) can be deduced from \(z_1\) on \(J^{-1}(E)\), namely

$$\begin{aligned} z_2 = \frac{E-R_1 z_1}{R_2}, \qquad \max \left( -1, \frac{E - R_2}{R_1} \right)< z_1 < \min \left( 1, \frac{E + R_2}{R_1} \right) \end{aligned}$$

and since the action of J preserves the angle \(\theta = \theta _1 - \theta _2\), we can use \((z_1, \theta )\) as coordinates on \(M^{\text {red}}_E\):

$$\begin{aligned} H_t(z_1,\theta )= & {} (1-t) z_1 + \frac{t z_1(E-R_1 z_1)}{R_2} + \frac{t \cos \theta }{R_2} \sqrt{P_E(z_1)}, \\ P_E(z_1)= & {} (1-z_1^2)\left( R_2^2 - (E - R_1 z_1)^2 \right) . \end{aligned}$$

The first derivatives of \(H_t\) read

$$\begin{aligned} \frac{\partial H_t}{\partial \theta }(z_1,\theta )= & {} \frac{-t \sqrt{P_E(z_1)}\sin \theta }{R_2}, \quad \frac{\partial H_t}{\partial z_1}(z_1,\theta )\\= & {} 1-t + \frac{tE}{R_2} - \frac{2 t R_1 z_1}{R_2} + \frac{t P_E'(z_1) \cos \theta }{2 R_2 \sqrt{P_E(z_1)}}. \end{aligned}$$

Hence if \((z_1^*,\theta ^*)\) is a critical point, then necessarily \(\theta ^* \in \{0, \pi \}\), and \(\frac{\partial ^2 H_t}{\partial \theta \partial z_1}(z_1^*,\theta ^*) = 0\). Let \(\varepsilon = \cos \theta ^* \in \{-1,1\}\); then one readily checks that

$$\begin{aligned} \frac{\partial ^2 H_t}{\partial \theta ^2}(z_1^*,\theta ^*)= & {} \frac{-\varepsilon t \sqrt{P_E(z_1^*)}}{R_2}, \quad \frac{\partial ^2 H_t}{\partial z_1^2}(z_1^*,\theta ^*) \\= & {} \frac{t}{4 R_2} \left( - 8 R_1 + \varepsilon \left( \frac{2 P_E''(z_1^*) P_E(z_1^*) - P_E'(z_1^*)^2}{P_E(z_1^*)^{3/2}} \right) \right) \end{aligned}$$

We claim that this last quantity has the sign of \(-\varepsilon \); this follows from the fact that

$$\begin{aligned} Q(E,z_1) = \frac{2 P_E''(z_1) P_E(z_1) - P_E'(z_1)^2}{P_E(z_1)^{3/2}} < - 8 R_1 \end{aligned}$$

for any \(E, z_1\) satisfying the above bounds. In order to prove this, one may check that Q is minimal at \((E,z_1) = (0,0)\); since \(Q(0,0) = -4 \left( R_2 + \frac{R_1^2}{R_2} \right) \) and since the function \(x > 0 \mapsto x + \frac{R_1^2}{x}\) is minimal at \(x = R_1\) with value \(2 R_1\), we obtain the desired result because \(R_2 > R_1\).

In fact, this analysis gives us the sign of the determinant of the Hessian of \(H_t\) at a critical point, so we can deduce from it that the corank one critical points are of elliptic–transverse type. Hence, if one is only interested in proving this, and not in finding a parametrization of the boundary of the image of the momentum map, this appendix constitutes a faster way to obtain Corollary 2.11.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Le Floch, Y., Pelayo, Á. Symplectic Geometry and Spectral Properties of Classical and Quantum Coupled Angular Momenta. J Nonlinear Sci 29, 655–708 (2019). https://doi.org/10.1007/s00332-018-9501-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-018-9501-y

Keywords

Mathematics Subject Classification

Navigation