Skip to main content

Particle Polarization, Spin Tensor, and the Wigner Distribution in Relativistic Systems

  • Chapter
  • First Online:
Strongly Interacting Matter under Rotation

Part of the book series: Lecture Notes in Physics ((LNP,volume 987))

Abstract

Particle spin polarization is known to be linked both to rotation (angular momentum) and magnetization of many particle systems. However, in the most common formulation of relativistic kinetic theory, the spin degrees of freedom appear only as degeneracy factors multiplying phase-space distributions. Thus, it is important to develop theoretical tools that allow to make predictions regarding the spin polarization of particles, which can be directly confronted with experimental data. Herein, we discuss a link between the relativistic spin tensor and particle spin polarization, and elucidate the connections between the Wigner function and average polarization. Our results may be useful for the theoretical interpretation of heavy-ion data on spin polarization of the produced hadrons.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 59.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 79.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Note that in the Weyl representation of the Clifford algebra a different explicit formula for the massive eigenspinors is typically used, but final results remain the same.

  2. 2.

    Massless particles have a positive energy for any nonvanishing momentum and the situation for them is quite similar.

  3. 3.

    The space-time region might be finite, with the fields going to zero at the boundary or infinite, as long as the fields decay fast enough to have a vanishing flux at infinity.

  4. 4.

    This is so in the case of an arbitrary original spin tensor which is antisymmetric only in the last two indices. For the canonical spin tensor that is totally antisymmetric, the number of independent components is 16+4=20.

  5. 5.

    Neither in a free theory nor in the standard model there are anomalies in the conservation laws for the four-momentum and angular momentum. However, one has to check on a case by case basis if this is so while dealing with a generic quantum field theory.

  6. 6.

    Note that in the Weyl representation of the Clifford algebra a different explicit formula for the massive eigenspinors is typically used, but the general conclusions remain the same.

  7. 7.

    At the classical level, the latter corresponds to rotation with respect to an internal axis of the extended object.

  8. 8.

    A very similar definition is used for the Pauli–Lubanski pseudovector. It follows the same construction procedure, but without mass in the denominator. Besides different physical dimensions, it is a very close concept which is well defined in the massless case. Since we focus on massive fields herein, we are not going to analyze it. It is useful to notice, however, that using the Pauli–Lubanski definition, one can follow the same steps in the massless case, obtaining the helicity distribution instead of the polarization one.

  9. 9.

    If one applies the normal ordering \(:\hat{J}_{\nu \rho }\hat{P}_\sigma :\) at the operator level there are two destruction operators on the left-hand side, which annihilate any single particle state. One would need at least two (anti)particle states to have a nonvanishing expectation value.

  10. 10.

    This is actually forbidden, since the wave function \(\psi _1\) is a regular distribution in momentum. However, one can have the spin density matrix factorized in a Hermitian \(2\times 2\) matrix times and arbitrarily sharp gaussian in the momentum. Such a strongly delocalized state is, for all practical purposes, equivalent to a momentum eigenstate.

  11. 11.

    Which can be expected, since the conventional two component spinors \(\chi \) in the negative frequency solutions of the Dirac equation are taken with the opposite eigenvalue of \(\sigma _z\), compared to the positive frequency solutions.

  12. 12.

    In order to exchange the order of the integrations and integrate by parts.

  13. 13.

    The hypothesis of an isolated system is important too. Being the integrand an observable, the flux over the light cone starting from the spatial boundary of an isolated system must be vanishing. Causality prevents the Wigner distribution to flow out of the light cone, as it would be a superluminal signal transfer, and the hypothesis of an isolated system prevents any signal to flow inside of the light cone. The flux over the light cone is therefore vanishing.

  14. 14.

    Compare with Eq. (5.50) for a quick check.

  15. 15.

    In general one needs the renormalized operators. For free fields this is just the normal ordering, that is, removing the vacuum expectation value. We always assume massive free Dirac fields and normal ordering in this section.

  16. 16.

    Being the sum of integrals of a real non-negative weight of the forms \(z^* z\).

References

  1. Busza, W., Rajagopal, K., van der Schee, W.: Ann. Rev. Nucl. Part. Sci. 68, 339 (2018)

    Article  ADS  Google Scholar 

  2. Romatschke, P., Romatschke, U.: https://doi.org/10.1017/9781108651998,arXiv:1712.05815 [nucl-th]

  3. Florkowski, W., Heller, M.P., Spalinski, M.: Rept. Prog. Phys. 81, (2018)

    Article  ADS  Google Scholar 

  4. Borsanyi, S., et al.: Wuppertal-Budapest collaboration. JHEP 1009, 073 (2010)

    Article  ADS  Google Scholar 

  5. Bazavov, A., et al.: HotQCD collaboration. Phys. Lett. B 795, 15 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  6. Miller, M.L., Reygers, K., Sanders, S.J., Steinberg, P.: Ann. Rev. Nucl. Part. Sci. 57, 205 (2007)

    Article  ADS  Google Scholar 

  7. Gelis, F., Iancu, E., Jalilian-Marian, J., Venugopalan, R.: Ann. Rev. Nucl. Part. Sci. 60, 463 (2010)

    Article  ADS  Google Scholar 

  8. Cooper, F., Frye, G.: Phys. Rev. D 10, 186 (1974)

    Article  ADS  Google Scholar 

  9. de Groot, S.R., van Leeuwen, W.A., van Weert, ChG: Relativistic Kinetic Theory. North-Holland, Amsterdam (1980)

    Google Scholar 

  10. Wang, Q.: Nucl. Phys. A 967, 225 (2017)

    Article  ADS  Google Scholar 

  11. Huang, X.G.: arXiv:2002.07549 [nucl-th]

  12. Becattini, F., Lisa, M.A.: arXiv:2003.03640 [nucl-ex]

  13. Becattini, F.: arXiv:2004.04050 [hep-th]

  14. Florkowski, W.: Phenomenology of Ultra-Relativistic Heavy-Ion Collisions. World Scientific, Singapore (2010)

    Book  Google Scholar 

  15. Itzykson, C., Zuber, J.B.: Quantum Field Theory. McGraw-Hill, New York (1980)

    MATH  Google Scholar 

  16. Florkowski, W., Kumar, A., Ryblewski, R.: Phys. Rev. C 98 (2018)

    Google Scholar 

  17. Weickgenannt, N., Sheng, X.L., Speranza, E., Wang, Q., Rischke, D.H.: Phys. Rev. D 100, 56018 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  18. Liu, Y.C., Mameda, K., Huang, X.G.: arXiv:2002.03753 [hep-ph]

  19. Hehl, F.W.: Rep. Math. Phys. 9, 55 (1976)

    Article  ADS  Google Scholar 

  20. Florkowski, W., Ryblewski, R., Kumar, A.: Prog. Part. Nucl. Phys. 108 (2019)

    Google Scholar 

  21. Weickgenannt, N., Speranza, E., Sheng, X.l., Wang, Q., Rischke, D.H.: arXiv:2005.01506 [hep-ph]

  22. Speranza, E., Weickgenannt, N.: arXiv:2007.00138 [nucl-th]

  23. Becattini, F., Chandra, V., Del Zanna, L., Grossi, E.: Ann. Phys. 338, 32 (2013)

    Article  ADS  Google Scholar 

  24. Becattini, F., Florkowski, W., Speranza, E.: Phys. Lett. B 789, 419 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  25. Mathisson, M.: Acta Phys. Pol. 6, 163 (1937)

    Google Scholar 

  26. Weyssenhoff, J., Raabe, A.: Acta Phys. Pol. 9, 7 (1947)

    Google Scholar 

Download references

Acknowledgements

We would like to thank F. Becattini, B. Friman, R. Ryblewski, and E. Speranza for insightful discussions. L.T. was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) through the CRC-TR 211 “Strong-interaction matter under extreme conditions” - project number 315477589–TRR 211. W.F. was supported in part by the Polish National Science Center Grants No. 2016/23/B/ST2/00717.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Leonardo Tinti .

Editor information

Editors and Affiliations

Appendix: Expectation Values of Creation and Destruction Operators

Appendix: Expectation Values of Creation and Destruction Operators

In this appendix, we show details of the calculations of the expectation values of creation and destruction operators. In particular, we find an interesting and intuitive link between the average (anti)particle number and the quantum fluctuations required for the mixed terms (\(\langle a^\dagger b^\dagger \rangle \) and \(\langle a b\rangle \)) to be nonvanishing.

The starting point is the density matrix (5.2), which reads

$$\begin{aligned} \rho = \sum _i \mathsf{P}_i \left| \psi _i \right\rangle \left\langle \psi _i\right| . \end{aligned}$$
(5.86)

All \(\mathsf{P}_i\)’s are classical probabilities

$$\begin{aligned} \sum _i \mathsf{P}_i =1. \end{aligned}$$
(5.87)

The states \(|\psi _i\rangle \) are proper quantum states, that is, they are normalized to one

$$\begin{aligned} \langle \psi _i|\psi _i\rangle =1, \qquad \forall i. \end{aligned}$$
(5.88)

The expectation value \(\mathcal{O}\) of any quantum operatorFootnote 15 \(\hat{\mathcal{O}}\) is a weighted average of the expectation values in the pure states, with the classical weights \(\mathsf{P}_i\)

$$\begin{aligned} \mathcal{O} = \mathrm{tr}\left( \phantom {\frac{}{}} \rho \, \hat{\mathcal{O}} \right) = \sum _i \mathsf{P}_i \, \mathrm{tr}\left( \phantom {\frac{}{}} \left| \psi _i \right\rangle \left\langle \psi _i\right| \hat{\mathcal{O}} \right) , \end{aligned}$$
(5.89)

therefore, our problem reduces to the expectation value in a generic pure state. The trace must be taken over a complete set of independent states (not necessarily quantum states that are normalized to one). Since we are interested in the expectation values of the creation and destruction operators of four-momentum and polarization eigenstates, the most convenient states are N-particle and \(\bar{N}\)-antiparticle ones. The trace is defined as an integration over the momentum degrees of freedom and a sum over discrete polarizations, namely

$$\begin{aligned} \begin{aligned} \mathrm{tr}\left( \cdots \phantom {\frac{}{}} \right)&= \sum _r\int \frac{d^3 p}{(2\pi )^3 2 E_\mathbf{p}} \langle p,r| \cdots |p,r\rangle +\\&\quad + \sum _{r,s}\int \frac{d^3 p}{(2\pi )^3 2 E_\mathbf{p}}\frac{d^3 q}{(2\pi )^3 2 E_\mathbf{q}} \langle p,r,q,s| \cdots |p,r,q,s\rangle + \cdots , \end{aligned} \end{aligned}$$
(5.90)

and so on, until exausting all the combinations of N particles and \(\bar{N}\) antiparticles. In the last formula the standard definition is used

$$\begin{aligned} |p,r\rangle = \sqrt{2 E_{p}}a^\dagger _r(\mathbf{p})|0\rangle , \end{aligned}$$
(5.91)

along with the analogous expressions for antiparticles and multiparticle states. The anticommutation relations have the form

$$\begin{aligned} \{a_r(\mathbf{p}),a^\dagger _s(\mathbf{p}^\prime )\}=\{b_r(\mathbf{p}),b^\dagger _s(\mathbf{p}^\prime )\} = (2\pi )^3 \delta _{rs}\,\delta ^3(\mathbf{p}-\mathbf{p}^\prime ), \end{aligned}$$
(5.92)

with the normalization

$$\begin{aligned} \langle p,r|q,s\rangle = 2E_\mathbf{p} (2 \pi )^3 \delta _{rs}\, \delta ^3(\mathbf{p}-\mathbf{p}^\prime ). \end{aligned}$$
(5.93)

It is convenient to introduce the compact notation for multiparticle states

$$\begin{aligned} \begin{aligned} \!\!\!\! |{\underline{p}},{\underline{r}};\bar{\underline{q}},\bar{\underline{s}}\rangle&=|p_1,r_1,p_2,r_2,\cdots p_N, r_N; {\bar{q}}_1,{\bar{s}}_1,{\bar{q}}_2,{\bar{s}}_2,\cdots {\bar{q}}_{\bar{N}},{\bar{s}}_{\bar{N}}\rangle ,\\ \int [d {\underline{p}}]^N [d {\underline{\bar{q}}}]^{\bar{N}}&= \int \!\!\!\frac{d^3 p_1}{(2\pi )^3 2 E_{\mathbf{p}_1}}\cdots \frac{d^3 p_N}{(2\pi )^3 2 E_{\mathbf{p}_N}}\frac{d^3 {\bar{q}}_1}{(2\pi )^3 2 E_{{\bar{\mathbf{q}}}_1}}\cdots \frac{d^3 {\bar{q}}_{\bar{N}}}{(2\pi )^3 2 E_{{\bar{\mathbf{q}}}_{\bar{N}}}}, \end{aligned} \end{aligned}$$
(5.94)

where the bar is used to distinguish antiparticle from particle variables. In this way the trace (5.90) can be written in a more compact form as

$$\begin{aligned} \mathrm{tr}\left( \phantom {\frac{(}{}{}}\cdots \right) = \sum _{N,{\bar{N}}}\sum _{{\underline{r}},{\underline{\bar{s}}}} \int \!\!\![d {\underline{p}}]^N [d {\underline{\bar{q}}}]^{\bar{N}} \langle {\underline{p}},{\underline{r}};\bar{\underline{q}},\bar{\underline{s}}| \cdots |{\underline{p}},{\underline{r}};\bar{\underline{q}},\bar{\underline{s}}\rangle . \end{aligned}$$
(5.95)

This compact notation is useful to write the generic quantum state \(|\psi \rangle \)

$$\begin{aligned} |\psi \rangle = \sum _{N,{\bar{N}}} \sum _{{\underline{r}},{\bar{\underline{s}}}}\int \!\!\![d {\underline{p}}]^N [d {\underline{\bar{q}}}]^{\bar{N}}\; \alpha _{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}})\, |{\underline{p}},{\underline{r}};\bar{\underline{q}},\bar{\underline{s}}\rangle , \end{aligned}$$
(5.96)

where the complex functions \(\alpha _{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}})\) are partial N-particle-\(\bar{N}\)-antiparticle wave functions in momentum space. The normalization reads

$$\begin{aligned} \begin{aligned} 1= \langle \psi |\psi \rangle&= \sum _{N,{\bar{N}}} \sum _{{\underline{r}},{\bar{\underline{s}}}}\int \!\!\![d {\underline{p}}]^N [d {\underline{\bar{q}}}]^{\bar{N}}\; \alpha ^*_{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}}) \alpha _{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}})=\\&= \sum _{N,{\bar{N}}} \Vert \alpha _{N,{\bar{N}}}\Vert ^2, \end{aligned} \end{aligned}$$
(5.97)

with\( \Vert \alpha _{N,{\bar{N}}}\Vert ^2\) being a shorthand notation for the (non-negativeFootnote 16) sum of integrals

$$\begin{aligned} \Vert \alpha _{N,{\bar{N}}}\Vert ^2=\sum _{{\underline{r}},{\bar{\underline{s}}}}\int \!\!\![d {\underline{p}}]^N [d {\underline{\bar{q}}}]^{\bar{N}}\; \alpha ^*_{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}}) \alpha _{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}}). \end{aligned}$$
(5.98)

The tensor product \(|\psi \rangle \langle \psi |\), that is, the projector on the quantum state \(|\psi \rangle \) reads

$$\begin{aligned} \begin{aligned} \!\!\!\! |\psi \rangle \langle \psi |&= \sum _{N,{\bar{N}}} \sum _{{\underline{r}},{\bar{\underline{s}}}} \sum _{N^\prime ,{\bar{N}}^\prime } \sum _{{\underline{r}}^\prime ,{\bar{\underline{s}}}^\prime }\int \!\!\![d {\underline{p}}]^N [d {\underline{\bar{q}}}]^{\bar{N}} [d {\underline{p}}^\prime ]^N [d {\underline{\bar{q}}}^\prime ]^{\bar{N}}\times \\&\qquad \times \alpha ^*_{N^\prime ,{\bar{N}}^\prime }({\underline{p}}^\prime ,{\underline{r}}^\prime ;{\underline{\bar{q}}}^\prime ,{\underline{\bar{s}}}^\prime )\alpha _{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}})\, |{\underline{p}},{\underline{r}};\bar{\underline{q}},\bar{\underline{s}}\rangle \langle {\underline{p}}^\prime ,{\underline{r}}^\prime ;\bar{\underline{q}}^\prime ,\bar{\underline{s}}^\prime |. \end{aligned} \end{aligned}$$
(5.99)

Making use of the normalization relations between the states, it is possible to write the trace in a pure state \(|\psi \rangle \) of an operator \(\hat{\mathcal{O}}\) in the compact form

$$\begin{aligned} \begin{aligned}&\!\!\!\!\mathrm{tr}\left( \phantom {\frac{}{}} \left| \psi \right\rangle \left\langle \psi \right| \hat{\mathcal{O}} \right) = \sum _{N,{\bar{N}}} \sum _{{\underline{r}},{\bar{\underline{s}}}} \sum _{N^\prime ,{\bar{N}}^\prime } \sum _{{\underline{r}}^\prime ,{\bar{\underline{s}}}^\prime }\int \!\!\![d {\underline{p}}]^N [d {\underline{\bar{q}}}]^{\bar{N}} [d {\underline{p}}^\prime ]^N [d {\underline{\bar{q}}}^\prime ]^{\bar{N}} \times \\&\times \alpha ^*_{N^\prime ,{\bar{N}}^\prime }({\underline{p}}^\prime ,{\underline{r}}^\prime ;{\underline{\bar{q}}}^\prime ,{\underline{\bar{s}}}^\prime )\alpha _{N,{\bar{N}}}({\underline{p}},{\underline{r}};{\underline{\bar{q}}},{\underline{\bar{s}}})\, \langle {\underline{p}}^\prime ,{\underline{r}}^\prime ;\bar{\underline{q}}^\prime ,\bar{\underline{s}}^\prime |\hat{\mathcal{O}} |{\underline{p}},{\underline{r}};\bar{\underline{q}},\bar{\underline{s}}\rangle . \end{aligned} \end{aligned}$$
(5.100)

There is a couple of results that can be immediately inferred from the last formula. The first one is that the expectation values of \(a^\dagger b^\dagger \) and ba, for any momentum and polarization combination, can be nonvanishing if and only if the quantum state of the system is in a superposition of states with different particle content. More precisely, only the quantum interference between states that differ exactly by a particle-antiparticle pair can give a nonvanishing contribution ( understanding that the integral over the partial wave functions can still simplify and give a vanishing result).

The second observation is that the expectation value of \(a^\dagger a\) and \(b^\dagger b\) can be simplified. The only combinations that can give a contribution are the ones between states with exactly the same number of particles and the same number of antiparticles. In the following computations, we consider only the term \(a^\dagger a\), understanding that the very same transformations hold for antiparticles.

As a particular case of (5.100) one can write the expectation value of \(a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime )\)

$$\begin{aligned} \begin{aligned}&\!\!\!\!\mathrm{tr}\left( \phantom {\frac{}{}} \left| \psi \right\rangle \left\langle \psi \right| a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) \right) = \sum _{N,{\bar{N}}}\sum _{{\underline{t}},{\underline{t}}^{\prime }}\sum _{{\bar{\underline{u}}},{\bar{\underline{u}}}^\prime }\int \!\!\![d {\underline{k}}]^N[d {\underline{k}}^{\prime }]^N [d {\underline{\bar{q}}}]^{\bar{N}}[d {\underline{\bar{q}}}^\prime ]^{\bar{N}} \times \\&\!\!\!\! \times \alpha ^*_{N,{\bar{N}}}({\underline{k}}^{\prime },{\underline{t}}^{\prime };{\underline{\bar{q}}}^\prime ,{\underline{\bar{u}}}^\prime )\alpha _{N,{\bar{N}}}({\underline{k}},{\underline{t}};{\underline{\bar{q}}},{\underline{\bar{u}}}) \langle {\underline{k}}^{\prime },{\underline{t}}^{\prime };{\underline{\bar{q}}}^{\prime },{\underline{\bar{u}}}^{\prime }|a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) |{\underline{k}},{\underline{t}};{\underline{\bar{q}}},{\underline{\bar{u}}}\rangle . \end{aligned} \end{aligned}$$
(5.101)

It is relatively simple to obtain the final formula by making use of the standard anticommutation relations

$$\begin{aligned} \begin{aligned} \!\!\!\!&\cdots a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) \sqrt{2E_{\mathbf{k}_j}}a^\dagger _{t_j}(\mathbf{k}_j)\cdots = \\ \!\!\!\!&\qquad \cdots a^\dagger _r(\mathbf{p}) \sqrt{2E_{\mathbf{k}_j}} \left( \{ a_s(\mathbf{p}^\prime ), a^\dagger _{t_j}(\mathbf{k}_j)\} - a^\dagger _{t_j}(\mathbf{k}_j) a_s(\mathbf{p}^\prime ) \right) \cdots = \\ \!\!\!\!&\cdots \sqrt{2E_{\mathbf{k}_j}}\left( a^\dagger _{t_j}(\mathbf{k}_j) a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) + a^\dagger _r(\mathbf{p}) (2\pi )^3\delta _{s t_j}\delta ^3(\mathbf{k}_j -\mathbf{p}^\prime ) \right) \cdots = \\ & \cdots \left[ \sqrt{2E_{\mathbf{k}_j}} a^\dagger _{t_j}(\mathbf{k}_j) \left( a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) \right) + \sqrt{2E_{\mathbf{p}^\prime }} (2\pi )^3\delta _{s t_j}\delta ^3(\mathbf{k}_j -\mathbf{p}^\prime ) a^\dagger _r(\mathbf{p})\right] \cdots =\\ \\ & =\cdots \left[ \sqrt{2E_{\mathbf{k}_j}} a^\dagger _{t_j}(\mathbf{k}_j) \left( a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) \right) + \right. \\&\qquad \left. + \sqrt{\frac{2E_{\mathbf{p}^\prime }}{2E_{\mathbf{p}}}} (2\pi )^3\delta _{s t_j}\delta ^3(\mathbf{k}_j -\mathbf{p}^\prime ) \sqrt{2E_{\mathbf{p}}}a^\dagger _r(\mathbf{p})\right] \cdots \end{aligned} \end{aligned}$$
(5.102)

In other words, even if \( a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) \) doesn’t commute with the creation operators, it is possible to “move it to the right”. However, each time we do that we have to add a new state, with a delta between the j’th degrees of freedom and the destruction operator \(a_s(\mathbf{p}^\prime )\), a numerical factor \((2\pi )^3\sqrt{E_{\mathbf{p}^\prime }/E_\mathbf{p}}\) and a substitution of the momentum and polarization at the j’th place with the ones related to the creation operator \( a^\dagger _r(\mathbf{p})\). After moving to the right all the particle creation operators, \( a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) \) commutes with the creation operators of the antiparticles (if present).

In the end, after making use of the normalization of the eigenstates we find

$$\begin{aligned} \begin{aligned}&\!\!\!\!\mathrm{tr}\left( \phantom {\frac{}{}} \left| \psi \right\rangle \left\langle \psi \right| a^\dagger _r(\mathbf{p})a_s(\mathbf{p}^\prime ) \right) = 0 +\\&\quad + \frac{1}{\sqrt{2E_\mathbf{p}2E_{\mathbf{p}^\prime }}} \sum _{{\bar{N}},N>0}\sum _{j+1}^N\sum _{{\underline{t}}-t_j}\sum _{{\bar{\underline{u}}}}\int \!\!\![d {\underline{k}}]^{(N-j)} [d {\underline{\bar{q}}}]^{\bar{N}}\times \\&\!\!\!\! \times \alpha ^*_{N,{\bar{N}}}({\underline{k}}-\mathbf{k}_j,\mathbf{p}, {\underline{t}}-t_j, r;{\underline{\bar{q}}},{\underline{\bar{u}}})\alpha _{N,{\bar{N}}}({\underline{k}}-\mathbf{k}_j,\mathbf{p}^\prime , {\underline{t}}-t_j, s;{\underline{\bar{q}}},{\underline{\bar{u}}}). \end{aligned} \end{aligned}$$
(5.103)

The notation \(\sum _{{\underline{t}}-t_j}\int d {[\underline{k}}]^{(N-j)}\) means that the integral and the sum is over all the particle degrees of freedom except for the j’th. In the similar way \(\alpha _{N,{\bar{N}}}({\underline{k}}-\mathbf{k}_j,\mathbf{p}, {\underline{t}}-t_j, r;{\underline{\bar{q}}},{\underline{\bar{u}}})\) is a shorthand notation for the (partial) wave function with the j’th degrees of freedom fixed to the momentum \(\mathbf{p}\) and polarization r.

The formula (5.103) has many interesting consequences. Besides the expected vanishing expectation value for purely antiparticle states, one can immediately check that the expectation value of \(a^\dagger _r(\mathbf{p})a_r(\mathbf{p}) \) is nonnegative, since it is a series of integrals and sums of squares. Moreover, as one could expect, it is linked to the average number of particles. Indeed, the expression

$$\begin{aligned} \sum _r \int \frac{d^3 p}{(2 \pi )^3} \mathrm{tr}\left( |\psi \rangle \langle \psi | a^\dagger _r(\mathbf{p})a_r(\mathbf{p}) \right) = \sum _N N \sum _{\bar{N}}\Vert \alpha _{N,{\bar{N}}}\Vert ^2, \end{aligned}$$
(5.104)

exactly gives the average number of particles in the state \(|\psi \rangle \) because of the normalization (5.97). More interestingly, the expectation value of \(a^\dagger _r(\mathbf{p})a_s(\mathbf{p})\) (same momentum, different polarization) performs the role of a momentum-dependent spin density matrix. The momentum integral of the trace is proportional to the average number of particles, but the matrix itself is sensitive to polarization in the rs indices and can be used to obtain the average number of particles, per momentum cell, for some polarization states.

All these arguments do not change if one reinserts the classical probabilities \(\mathsf{P}_i\) from (5.86) and deals with mixed states. The classical fluctuations do not change the properties of the spin density matrix, like the non-negative diagonal elements and normalization of the trace (after dividing by \((2\pi )^3\) and integrating over momentum, like for the pure states) does not change the average number of particles.

Rights and permissions

Reprints and permissions

Copyright information

© 2021 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Tinti, L., Florkowski, W. (2021). Particle Polarization, Spin Tensor, and the Wigner Distribution in Relativistic Systems. In: Becattini, F., Liao, J., Lisa, M. (eds) Strongly Interacting Matter under Rotation. Lecture Notes in Physics, vol 987. Springer, Cham. https://doi.org/10.1007/978-3-030-71427-7_5

Download citation

Publish with us

Policies and ethics