Skip to main content
Log in

Lorentz invariant quantum concurrence for \(\textit{SU}(2) \otimes \textit{SU}(2)\) spin–parity states

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

    We’re sorry, something doesn't seem to be working properly.

    Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Abstract

The quantum concurrence of \(\textit{SU}(2) \otimes \textit{SU}(2)\) spin–parity states is shown to be invariant under \(\textit{SO}(1,3)\) Lorentz boosts and O(3) rotations when the density matrices are constructed in consonance with the covariant probabilistic distribution of Dirac massive particles. Similar invariance properties are obtained for the quantum purity and for the trace of unipotent density matrix operators. The reported invariance features—obtained in the scope of the \(\textit{SU}(2) \otimes \textit{SU}(2)\) corresponding to just one of the inequivalent representations enclosed by the \(\textit{SL}(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\) symmetry—set a more universal and kinematical-independent meaning for the quantum entanglement encoded in systems containing not only information about spin polarization but also the correlated information about intrinsic parity. Such a covariant framework is used for computing the Lorentz invariant spin–parity entanglement of spinorial particles coupled to a magnetic field, through which the extensions to more general Poincaré classes of spinor interactions are straightforwardly depicted.

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.

Similar content being viewed by others

Notes

  1. The free particle Dirac Hamiltonian in the coordinate space reads

    $$\begin{aligned} {H} \, \psi {(t,{\varvec{x}})} = i \frac{\partial \, \psi {(t,{\varvec{x}})}}{\partial t} = (-i{\varvec{\nabla }} \cdot {{\varvec{\alpha }}} + m {\beta }) \,\psi {(t,{\varvec{x}})} = (-i {\alpha }_i\partial ^i + m {\beta }) \,\psi {(t,{\varvec{x}})} = {\pm } E_p \,\psi {(t,{\varvec{x}})}, \end{aligned}$$

    where H expressed in natural units, i.e., with \(c = \hbar = 1\), the eigenvalues, \({\pm } E_p\), are expressed by \(E_p = \sqrt{p^2 + m^2}\), and the matrix operators \({{\varvec{\alpha }}} =({\alpha }_x,\,{\alpha }_y,\,{\alpha }_z)\) and \({\beta }\) satisfy the anticommuting relations, \( {\alpha }_i {\alpha }_j + {\alpha }_j {\alpha }_i = 2 \delta _{ij} {I}_4\), and \({\alpha }_i {\beta } + {\beta } {\alpha }_i =0\), for \(i,j = x,y,z\), with \( {\beta }^2 = {I}_4\), where \(I_N\) the N-dim identity matrix. These so-called Dirac matrices, \({\alpha }_i\) and \({\beta }\), in their standard (or Dirac) representation are decomposed into tensor products of Pauli matrices [17], as \({\alpha }_i = {\sigma }_x^{(P)} \otimes {\sigma }_i^{(S)}\), for \(i = x,y,z\) and \({\beta } = {\sigma }_z^{(P)} \otimes {I}_{2}^{(S)}\).

  2. Given that \(\gamma ^{{0}} S^{\dagger }\gamma ^{{0}} = S^{{-}{1}}\) , one notices that \({\overline{w}} \, w\) transforms as \({\overline{w}} \, w = w^{\dagger } \gamma ^{{0}} w \rightarrow w^{\dagger }\,S^{\dagger } \gamma ^{{0}} S \, w = w^{\dagger }\, \gamma ^{{0}}S^{{-}{1}} S \, w = w^{\dagger } \gamma ^{{0}} w = {\overline{w}} \, w\), which is therefore Lorentz invariant.

  3. Even if it is already globally preserved by the Lorentz invariant continuity equation which sets

    figure a
  4. The spinor\(\xi \) is an object dimensionally represented by given by \(({\varvec{dim}}(SU_{\xi }(2)),{\varvec{dim}}(SU_{\chi }(2))) =({\varvec{2}},{\varvec{1}})\).

References

  1. R.M. Gingrich, C. Adami, Phys. Rev. Lett. 89, 270402 (2002)

    Google Scholar 

  2. A. Peres, P.F. Scudo, D.R. Terno, Phys. Rev. Lett. 88, 230402 (2002)

    ADS  MathSciNet  Google Scholar 

  3. D. Ahn, H.J. Lee, Y.H. Moon, S.W. Hwang, Phys. Rev. A 67, 012103 (2003)

    ADS  MathSciNet  Google Scholar 

  4. H. Terashima, M. Ueda, Int. J. Quantum Inf. 1, 93 (2003)

    Google Scholar 

  5. S.D. Bartlett, D.R. Terno, Phys. Rev. A 71, 012302 (2005)

    ADS  MathSciNet  Google Scholar 

  6. T.F. Jordan, A. Shaji, E.C.G. Sudarshan, Phys. Rev. A 75, 022101 (2007)

    ADS  MathSciNet  Google Scholar 

  7. N. Friis, R.A. Bertlmann, M. Huber, B.C. Hiesmayr, Phys. Rev. A 81, 042114 (2010)

    ADS  Google Scholar 

  8. V. Palge, J. Dunningham, Phys. Rev. A 85, 042322 (2012)

    ADS  Google Scholar 

  9. J. Dunningham, V. Palge, V. Vedral, Phys. Rev. A 80, 044302 (2009)

    ADS  Google Scholar 

  10. V. Palge, V. Vedral, J.A. Dunningham, Phys. Rev. A 84, 044303 (2011)

    ADS  Google Scholar 

  11. R. Jozsa, D.S. Abrams, J.P. Dowling, C.P. Williams, Phys. Rev. Lett. 85, 2010 (2000)

    ADS  Google Scholar 

  12. V. Giovannetti, S. Lloyd, L. Maccone, Nature 412, 417 (2001)

    ADS  Google Scholar 

  13. U. Yurtsever, J.P. Dowling, Phys. Rev. A 65, 052317 (2002)

    ADS  Google Scholar 

  14. A. Kent, W.J. Munro, T.P. Spiller, Phys. Rev. A 84, 012326 (2011)

    ADS  Google Scholar 

  15. N. Friis, A.R. Lee, K. Truong, C. Sabín, E. Solano, G. Johansson, I. Fuentes, Phys. Rev. Lett. 110, 113602 (2013)

    ADS  Google Scholar 

  16. V.A.S.V. Bittencourt, A.E. Bernardini, M. Blasone, Phys. Rev. A 97, 032106 (2018)

    ADS  Google Scholar 

  17. A.E. Bernardini, S.S. Mizrahi, Physica Scr. 89, 075105 (2014)

    ADS  Google Scholar 

  18. V.A.S.V. Bittencourt, A.E. Bernardini, Ann. Phys. 364, 182 (2016)

    ADS  Google Scholar 

  19. V.A.S.V. Bittencourt, A.E. Bernardini, M. Blasone, Phys. Rev. A 93, 053823 (2016)

    ADS  Google Scholar 

  20. V.A.S.V. Bittencourt, A.E. Bernardini, Phys. Rev. B 95, 195145 (2017)

    ADS  Google Scholar 

  21. W. Greiner, Relativistic Quantum Mechanics: Wave Equations (Springer, Berlin, 2000)

    MATH  Google Scholar 

  22. P. Caban, J. Rembielinski, Phys. Rev. A 72, 012103 (2005)

    ADS  Google Scholar 

  23. H.P. Breuer, F. Petruccione, The Theory of Open Quantum Systems (Oxford University Press, New York, 2002)

    MATH  Google Scholar 

  24. R. Horodecki, P. Horodecki, M. Horodecki, K. Horodecki, Rev. Mod. Phys. 81, 865 (2009)

    ADS  Google Scholar 

  25. T.J. Osbourne, Quantum Inf. Comput. 7(3), 209–227 (2007)

    MathSciNet  Google Scholar 

  26. W.K. Wootters, Phys. Rev. Lett. 80, 2245 (1998)

    ADS  Google Scholar 

  27. W.K. Wootters, Quantum Inf. Comput. 1, 27 (2001)

    MathSciNet  Google Scholar 

  28. L. Henderson, V. Vedral, J. Phys. A Math. Gen. 34, 6899–6905 (2001)

    ADS  Google Scholar 

  29. T.J. Osborne, Phys. Rev. A 72, 022309 (2005)

    ADS  MathSciNet  Google Scholar 

  30. R.F. Streater, A.S. Wightman, PCT, Spin and Statistics, and All That (W. A. Benjamin Inc., New York, 1964)

    MATH  Google Scholar 

  31. S.N. Gupta, Proc. Phys. Soc. 63, 681 (1950)

    ADS  Google Scholar 

  32. K. Bleuler, Helv. Phys. Acta 23, 567 (1950)

    MathSciNet  Google Scholar 

  33. Yu. Chang-shui, C. Li, He-shan Song, Phys. Rev. A 77, 012305 (2008)

    ADS  MathSciNet  Google Scholar 

  34. S. Hill, W.K. Wootters, Phys. Rev. Lett. 78, 5022 (1997)

    ADS  Google Scholar 

  35. W.K. Tung, Group Theory (World Scientific Publishing, London, 2003)

    Google Scholar 

  36. P.M. Alsing, G.J. Milburn, Lorentz Invariance of Entanglement. arXiv:quant-ph/020305

  37. P. Caban, J. Rembielinski, M. Wlodarczyk, Phys. Rev. A 88, 022119 (2013)

    ADS  Google Scholar 

  38. H. Bauke, S. Ahrens, C.H. Keitel, R. Grobe, New J. Phys. 16, 043012 (2014)

    ADS  Google Scholar 

  39. K. Fujikawa, C.H. Oh, C. Zhang, Phys. Rev. D 90, 025028 (2014)

    ADS  Google Scholar 

  40. P.L. Saldanha, V. Vedral, New J. Phys. 14, 023041 (2012)

    ADS  Google Scholar 

  41. L.C. Céleri, V. Kiosses, D.R. Terno, Phys. Rev. A 94, 062115 (2016)

    ADS  Google Scholar 

  42. P. Caban, J. Rembielinski, Phys. Rev. A 74, 042103 (2006)

    ADS  MathSciNet  Google Scholar 

  43. S. Moradi, JETP Lett. 89, 50 (2009)

    ADS  Google Scholar 

  44. T. Choi, J. Hur, J. Kim, Phys. Rev. A 84, 012334 (2011)

    ADS  Google Scholar 

  45. T. Choi, J. Korean Phys. Soc. 62, 1085 (2013)

    ADS  Google Scholar 

  46. L. Fonda, G.C. Ghirardi, Symmetry Principles in Quantum Physics (Marcel Dekker Inc., New York, 1970)

    Google Scholar 

  47. S. Weinberg, Quantum Theory of Fields, vol. 1 (Cambridge University Press, New York, 1995)

    Google Scholar 

  48. R. Mosseri, R. Dandoloff, J. Phys. A Math. Gen. 34, 10243 (2001)

    ADS  Google Scholar 

  49. P. Lévay, J. Phys. A Math. Gen. 37, 1821 (2004)

    ADS  Google Scholar 

  50. V. Kiosses, J. Phys. A Math. Theor. 47, 405301 (2014)

    MathSciNet  Google Scholar 

  51. M. Czachor, Quantum Inf. Process. 9, 171 (2010)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

The work of AEB is supported by the Brazilian Agencies FAPESP (Grant 2018/03960-9) and CNPq (Grant 301000/2019-0). VASVB acknowledge the support from the Max Planck Gesellschaft through an independent Max Planck Research Group.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alex E. Bernardini.

Appendix: The group \(\textit{SU}(2)\otimes \textit{SU}(2)\) as a subgroup of \(\textit{SL}(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\) and the spin–parity intrinsic structure of Dirac bispinors

Appendix: The group \(\textit{SU}(2)\otimes \textit{SU}(2)\) as a subgroup of \(\textit{SL}(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\) and the spin–parity intrinsic structure of Dirac bispinors

The group representations of \(sl(2,{\mathbb {C}})\oplus sl(2,{\mathbb {C}})\)—the Lie algebra of the \(\textit{SL}(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\) Lie group—are irreducible, since they are tensor products of linear complex representations of \(sl(2,{\mathbb {C}})\). Given that \(\textit{SU}(2)\otimes \textit{SU}(2) \subset SL(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\), the unitary irreducible representations (irreps) of the \(\textit{SU}(2)\otimes \textit{SU}(2)\) are built through tensor products between unitary representations of \(\textit{SU}(2)\). Such an one-to-one correspondence with the \(\textit{SL}(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\) simply connected group imposes a single correspondence with the \(sl(2,{\mathbb {C}})\oplus sl(2,{\mathbb {C}})\) algebra. In fact, the existence of inequivalent representations of \(\textit{SU}(2) \otimes \textit{SU}(2)\) follows exactly from such one-to-one correspondences. The point is that inequivalent representations (for instance, the one for the chiral basis and another one for the spin–parity basis) do not correspond to the complete set of representations of \(\textit{SL}(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\), and therefore, each of them, independently, does not exhibit a one-to-one correspondence with the homogeneous Lorentz transformations that compose the algebra of the \(\textit{SO}(1,3)\) group. The chiral basis described by a double-doublet representation of \(\textit{SU}(2)\otimes \textit{SU}(2)\), for instance, maps a subset of transformations of the \(\textit{SO}(4) \equiv SO(3)\otimes SO(3)\) group, as for instance, those which include the double-covering rotations.

Turning back to our point, since the \(\textit{SU}(2)\otimes \textit{SU}(2)\) transformations can be mapped into a subset of \(\textit{SL}(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\), one may choose two inequivalent subsets of \(\textit{SU}(2)\) generators, such that \(\textit{SU}(2)\otimes \textit{SU}(2) \subset SL(2,{\mathbb {C}})\otimes \textit{SL}(2,{\mathbb {C}})\), with each group transformation generator having its own irrep. In an overall context, despite the effectiveness of the irreps of the Poincaré group, in the Lorentz covariant Hamiltonian formulation of quantum mechanics, one has to pay attention to the inclusion of masses in the relativistic formalism described by the Dirac equation. It requires the inclusion of the parity symmetry and the equalization of its role with the helicity (spin one-half projection, \(\hat{{\varvec{e}}}_p\cdot \hat{\varvec{\sigma }} \sim {\hat{\sigma }}_z\)) symmetry, as an accomplished \(\textit{SU}(2)\) symmetry. Under such conditions, one designates the fundamental representation of the \(SU_{\xi }(2)\) as a spinor-like object \(\xi \) described by \(({\pm },\,0)\), which transforms as a \(SU_{\xi }(2)\)doublet (2-dim) of parity quantum numbers, \({\pm }\), and as a singlet (1-dim), which is transparent under any \(SU_{\chi }(2)\) transformation. Reciprocally, the fundamental object of the \(SU_{\chi }(2)\), a typical spinor \(\chi \) described by (0, 1/2), transforms as a \(SU_{\chi }(2)\)doublet of spin quantum numbers, \({\pm } 1/2\), and as a singlet of the \(SU_{\xi }\).Footnote 4

The meaning of the above-mentioned quantities related to parity and helicity can be clarified by noticing that the total parity operator \({\hat{P}}\) acts on the direct product \(\left| {\pm } \right\rangle \otimes \left| \chi _s{({\varvec{p}})}\right\rangle \) as

$$\begin{aligned} {\hat{P}}\left( \left| {\pm } \right\rangle \otimes \left| \chi _s{({\varvec{p}})}\right\rangle \right) ={\pm } \left( \left| {\pm } \right\rangle \otimes \left| \chi _s(-{\varvec{p}})\right\rangle \right) , \end{aligned}$$

and, for instance, it corresponds to the Kronecker product of two operators, \({\hat{P}}^{(P)}\otimes {\hat{P}}^{(S)}\), where \({\hat{P}}^{(P)}\) is the intrinsic parity (with two eigenvalues, \({\hat{P}}^{(P)}\left| {\pm } \right\rangle ={\pm } \left| {\pm } \right\rangle \)) and \({\hat{P}}^{(S)}\) is the spatial parity (with \({\hat{P}}^{(S)}\chi _s \left( {\varvec{p}}\right) =\chi _s \left( -{\varvec{p}}\right) \)).

The above construction supports the identification of Dirac Hamiltonian eigenstates with electrons in the double-doublet irrep of the \(\textit{SU}(2) \otimes \textit{SU}(2)\). Spatial parity couples quantum states with \({\pm }\) parity and \({\pm } 1/2\) spin quantum numbers given that they are described by irreps of the Poincaré group [46]. To comprehend the covariant behavior of parity, one needs to consider the extended Poincaré group [35, 47] which also accounts for the helicity (spin projection) in a double-doublet \(\textit{SL}(2,{\mathbb {C}}) \otimes \textit{SL}(2,{\mathbb {C}})\) representation given by Dirac four-component spinors, the bispinors satisfying the Dirac equation. Reciprocally, the same assertion is applied for the understanding of the factorized covariant behavior of the helicity (spin projection) at such quantum states.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bernardini, A.E., Bittencourt, V.A.S.V. & Blasone, M. Lorentz invariant quantum concurrence for \(\textit{SU}(2) \otimes \textit{SU}(2)\) spin–parity states. Eur. Phys. J. Plus 135, 320 (2020). https://doi.org/10.1140/epjp/s13360-020-00323-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-020-00323-w

Navigation