Skip to main content

Center of Mass, Spin Supplementary Conditions, and the Momentum of Spinning Particles

  • Chapter
  • First Online:
Equations of Motion in Relativistic Gravity

Part of the book series: Fundamental Theories of Physics ((FTPH,volume 179))

Abstract

We discuss the problem of defining the center of mass in general relativity and the so-called spin supplementary condition. The different spin conditions in the literature, their physical significance, and the momentum-velocity relation for each of them are analyzed in depth. The reason for the non-parallelism between the velocity and the momentum, and the concept of “hidden momentum”, are dissected. It is argued that the different solutions allowed by the different spin conditions are equally valid descriptions for the motion of a given test body, and their equivalence is shown to dipole order in curved spacetime. These different descriptions are compared in simple examples.

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 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover 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.

    One could say the same about the point \(x_{\mathrm{{CM}}}^{\alpha }(u)\), although one must bear in mind when transforming its coordinates to the new frame that it will still be the CM as measured by the specific observer \(u^{\alpha }\), and not the CM as measured in the new frame.

  2. 2.

    For the fastest spinning celestial body known to date, the pulsar PSR J1748-2446ad (rotation frequency 716 Hz, estimated radius \(a=16\) km), whose equatorial velocity is 0.23c, \(R_{\mathrm{Moller}}\simeq 0.1a\), see also the contribution by Giulini in this volume.

  3. 3.

    Such observers are said to be “kinematically comoving” (see [31, Sect. 6.1]).

  4. 4.

    To see the reason for this assumption, consider two observers \(u_{1}^{\alpha }=u_{\mathrm{lab}}^{\alpha }(x_{1})\) and \(u_{2}^{\alpha }=u_{\mathrm{lab}}^{\alpha }(x_{2})\), orthogonal to the same geodesic hypersurface \(\Sigma \). Let \(\{x^{\hat{\alpha }}\}\) and \(\{x^{\tilde{\alpha }}\}\), respectively, denote their normal coordinate systems, related by \(x^{\tilde{\alpha }}=\Lambda _{\ \hat{\beta }}^{\tilde{\alpha }}(x^{\hat{\beta }}-x_{2}^{\hat{\beta }})\) (where \(\Lambda _{\ \hat{\beta }}^{\tilde{\alpha }}\) is a function of \(x^{\hat{\beta }}\)). They will agree on the centroid position if \(x_{\mathrm{CM}}^{\tilde{\alpha }}(u_{2})=\Lambda _{\ \hat{\beta }}^{\tilde{\alpha }}(x_{\mathrm{CM}}^{\hat{\beta }}(u_{1})-x_{2}^{\hat{\beta }})\). From Eq. (4) we see that it is the case when \(\mathbf {d}x^{\hat{0}}=\mathbf {d}x^{\tilde{0}},\) \(x^{\tilde{i}}=\Lambda _{\ \hat{j}}^{\tilde{i}}(x^{\hat{j}}-x_{2}^{\hat{j}})\), with \(\Lambda _{\ j}^{\tilde{i}}\) a constant matrix. Due to the curvature, however, this cannot be exactly so; choosing \(\partial _{\tilde{\alpha }}|_{x_{2}}\simeq \partial _{\hat{\alpha }}|_{x_{2}}\), we have \(\Lambda _{\ \hat{\beta }}^{\tilde{\alpha }}=\delta _{\ \hat{\beta }}^{\hat{\alpha }}+\mathcal {O}(\Vert \mathbf {R}\Vert \hat{x}\hat{x}_{2})+\mathcal {O}(\Vert \mathbf {R}\Vert \hat{x}_{2}^{2})\), e.g. Eq. 11.12 of [32]. It follows that, for all observers within the body’s convex hull, \(\Vert x_{\mathrm{CM}}^{\tilde{\alpha }}(u_{2})-x_{\mathrm{CM}}^{\tilde{\alpha }}(u_{1})\Vert /a\lesssim \lambda \), \(\lambda =\Vert \mathbf {R}\Vert a^{2}\); hence \(x_{\mathrm{CM}}^{\tilde{\alpha }}(u_{2})\simeq x_{\mathrm{CM}}^{\tilde{\alpha }}(u_{1})\) if \(\lambda \ll 1\). This is, as expected, the condition that the metric \(g_{\hat{\alpha }\hat{\beta }}=\eta _{\hat{\alpha }\hat{\beta }}+\mathcal {O}(\Vert \mathbf {R}\Vert \hat{x}^{2})\) can be taken as nearly flat throughout the body.

  5. 5.

    The factor \(m/M^{2}\) (involving \(U^{\alpha }\) via m) can be determined by the normalization condition \(U^{\alpha }U_{\alpha }=~-1\).

  6. 6.

    We thank O. Semerák and A. Harte for discussions on these issues.

  7. 7.

    We thank O. Semerák for his input on this issue.

  8. 8.

    Only when \(\nabla _{\mathbf {U}}u^{\alpha }=\nabla _{\bar{\mathbf {U}}}\bar{u}^{\alpha }=0\) should one expect two different centroids of the same body to move parallel, even in flat spacetime, as explained in Sect. 3.1 (see also Fig. 2b). Otherwise (i.e. when \(P_{\mathrm{hidI}}^{\alpha }\ne 0\)) they can have an arbitrary relative motion, cf. Fig. 5.

  9. 9.

    In [13], where Eq. (37) was originally derived, \(F^{\alpha }\) was taken to be the spin-curvature force (40); it is however easy to check, following the derivation therein, that it holds for an arbitrary force, as long as Eq. (30) holds.

  10. 10.

    Since \(\tau ^{\alpha \beta }=0\), \(S^{\alpha }\) is Fermi-Walker transported for the MP condition, Eq. (35), and approximately so for the TD condition (cf. Eq. (7.11) of [26]); since Eq. (54) is of first order in v, \(\vec {S}\) can be taken constant therein.

  11. 11.

    E.g. the torque on an electric dipole in an uniform electromagnetic field, when \(z^{\alpha }\) is the common centroid given by the TD or the (non-helical) FMP condition.

  12. 12.

    To obtain (63) from (64) in the special case above, one uses \(d\tau =\gamma (\bar{U},P)d\bar{\tau }\) to write \(\bar{U}_{\perp }^{\alpha }=\gamma (\bar{U},P)D\Delta x^{\alpha }/d\tau \), computes \(D\bar{S}^{\alpha \beta }/d\bar{\tau }\) from (43) using (15) to obtain \(\bar{\tau }^{\alpha \beta }=\gamma (\bar{U},P)\tau ^{\alpha \beta }\), and finally uses the assumption above \(\tau ^{\alpha \beta }P_{\beta }=0\Rightarrow (h^{P})_{\ \sigma }^{\alpha }\bar{\tau }^{\sigma \beta }=\bar{\tau }^{\alpha \beta }\).

  13. 13.

    For example, in the case of the Schwarzschild spacetime, \(\Vert \mathbf {R}\Vert \sim m_{\mathrm{S}}/r^{3}\), \(\lambda =(m_{\mathrm{S}}/r)(a^{2}/r^{2})\); since \(m_{\mathrm{S}}/r<1\) for any point outside the horizon, \(\lambda \ll 1\) is guaranteed just by taking the size of the body much smaller than its the distance to the source, \(r^{2}\gg a^{2}\).

  14. 14.

    The inequality (69) means not only that the components of \(\Delta \mathbf{P}\) in the system \(\{x^{\hat{\alpha }}\}\) (where \(P^{\hat{i}} = 0\)) are much smaller than M, but also much smaller than the typical spatial momentum in other frames. For instance, in normal coordinates \(\{x^{\alpha '}\}\) comoving with \(u'^{\alpha }\), one has \(|P^{i'}| \sim \gamma (u', u)v(u', u)M\) and \(|\Delta {P}^{i'}| \lesssim |\Delta {P}^{\varvec{\xi }}|\gamma (u', u)\), thus \(|\Delta {P}^{i'}| \ll |P^{i'}|\) when \(\lambda \ll 1\).

References

  1. J. Frenkel, Die Elektrodynamik des rotierenden Elektrons. Z. Phys. 37, 243 (1926)

    Article  ADS  Google Scholar 

  2. J. Frenkel, Spinning electrons. Nature 117, 514 (1926)

    Article  Google Scholar 

  3. H.J. Bhabha, H.C. Corben, General classical theory of spinning particles in a Maxwell field. Proc. R. Soc. Lond. A 178, 273 (1940)

    ADS  MathSciNet  MATH  Google Scholar 

  4. H.C. Corben, Spin in classical and quantum theory. Phys. Rev. 121, 1833 (1961)

    Article  ADS  Google Scholar 

  5. W.G. Dixon, Description of extended bodies by multipole moments in special relativity. J. Math. Phys. 8, 1591 (1967)

    Google Scholar 

  6. C. Moller, On the definition of the centre of gravity in an arbitrary closed system in the theory of relativity. Commun. Dublin Inst. Adv. Stud. A 5, 3 (1949)

    Google Scholar 

  7. C. Moller, Sur la dynamique des systemes ayant un moment angulaire interne. Ann. Inst. Henri Poincaré 11, 251 (1949)

    MathSciNet  MATH  Google Scholar 

  8. M. Mathisson, Neue Mechanik materieller Systeme. Acta Phys. Pol. 6, 163 (1937)

    MATH  Google Scholar 

  9. E. Corinaldesi, A. Papapetrou, Spinning test-particles in general relativity II. Proc. R. Soc. Lond. A 209, 259–268 (1951)

    Article  ADS  MathSciNet  Google Scholar 

  10. W. Tulczyjew, Motion of multipole particles in general relativity theory. Acta Phys. Pol. 18, 393 (1959)

    MathSciNet  MATH  Google Scholar 

  11. W.G. Dixon, A covariant multipole formalism for extended test bodies in general relativity. II Nuovo Cimento 34, 317 (1964)

    Google Scholar 

  12. A. Ohashi, Multipole particle in relativity. Phys. Rev. D 68, 044009 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  13. K. Kyrian, O. Semerák, Spinning test particles in a Kerr field—II. Mon. Not. R. Astron. Soc. 382, 1922 (2007)

    Article  ADS  Google Scholar 

  14. O. Semerák, Spinning test particles in a Kerr field—I. Mon. Not. R. Astron. Soc. 308, 863 (1999)

    Article  ADS  Google Scholar 

  15. M. Mathisson, Das zitternde Elektron und seine Dynamik. Acta Phys. Pol. 6, 218 (1937)

    MATH  Google Scholar 

  16. L.F. Costa, C. Herdeiro, J. Natário, M. Zilhão, Mathisson’s helical motions for a spinning particle: are they unphysical? Phys. Rev. D 85, 024001 (2012)

    Article  ADS  Google Scholar 

  17. S. Gralla, A. Harte, R. Wald, Bobbing and Kicks in electromagnetism and gravity. Phys. Rev. D 81, 104012 (2010)

    Article  ADS  Google Scholar 

  18. W. Shockley, R.P. James, Try simplest cases discovery of hidden momentum forces on magnetic currents. Phys. Rev. Lett. 18, 876 (1967)

    Article  ADS  Google Scholar 

  19. L. Vaidman, Torque and force on a magnetic dipole. Am. J. Phys. 58, 978 (1990)

    Article  ADS  Google Scholar 

  20. V. Hnizdo, Hidden momentum and the electromagnetic mass of a charge and current carrying body. Am. J. Phys. 65, 92 (1997)

    Article  ADS  Google Scholar 

  21. S. Coleman, J.H. Van Vleck, Origin of hidden momentum forces on magnets. Phys. Rev. 171, 1370 (1968)

    Article  ADS  Google Scholar 

  22. D. Babson, S.P. Reynolds, R. Bjorquist, D.J. Griffiths, Hidden momentum, field momentum, and electromagnetic impulse. Am. J. Phys. 77, 826 (2009)

    Article  ADS  Google Scholar 

  23. L.F. Costa, J. Natário, Gravito-electromagnetic analogies. Gen. Rel. Grav. 46, 1792 (2014)

    Google Scholar 

  24. L.F. Costa, J. Natário, M. Zilhão, Spacetime dynamics of spinning particles—exact gravito-electromagnetic analogies (2012) arXiv:1207.0470

  25. J. Madore, The equations of motion of an extended body in general relativity. Ann. Inst. Henri Poincaré 11, 221 (1969)

    MathSciNet  Google Scholar 

  26. W.G. Dixon, Dynamics of extended bodies in general relativity. I. Momentum and angular momentum. Proc. R. Soc. Lond. A 314, 499 (1970)

    Article  ADS  MathSciNet  Google Scholar 

  27. C.W. Misner, K.S. Thorne, J.A. Wheeler, Gravitation (W. H. Freeman and Company, San Francisco, 1973)

    Google Scholar 

  28. R.T. Jantzen, P. Carini, D. Bini, The many faces of gravitoelectromagnetism. Ann. Phys. 215, 1 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  29. J.L. Synge, Relativity: The Special Theory (North-Holland Publishing Company, Amsterdam, 1956)

    MATH  Google Scholar 

  30. S. Gralla, F. Herrmann, Hidden momentum and black hole kicks. Class. Quantum Gravity 30, 205009 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  31. V. Bolós, Intrinsic definitions of relative velocity in general relativity. Commun. Math. Phys. 273, 217 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  32. L. Brewin, Riemann normal coordinate expansions using Cadabra. Class. Quantum Gravity 26, 175017 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  33. W. Beiglbock, The center-of-mass in Einstein’s theory of gravitation. Commun. Math. Phys. 5, 106–130 (1967)

    Article  ADS  MathSciNet  Google Scholar 

  34. R. Schattner, The uniqueness of the center of mass in general relativity. Gen. Reletiv. Gravit. 10, 395–399 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  35. J. Weyssenhoff, Relativistic dynamics of spin-fluids and spin-particles. Nature 157, 766 (1946)

    ADS  Google Scholar 

  36. J. Weyssenhoff, A. Raabe, Relativistic dynamics of spin-fluids and spin particles. Acta Phys. Pol. 9, 7 (1947)

    MathSciNet  Google Scholar 

  37. W.G. Dixon, On a classical theory of charged particles with spin and the classical limit of the Dirac equation. Il Nuovo Cimento 38, 1616 (1965)

    Article  MathSciNet  Google Scholar 

  38. R. Plyatsko, O. Stephanyshin, Mathisson equations: non-oscillatory solutions in a Schwarzschild field. Acta Phys. Pol. B 39, 23 (2008)

    ADS  MathSciNet  MATH  Google Scholar 

  39. R. Plyatsko, O. Stephanyshin, M. Fenyk, Mathisson-Papapetrou-Dixon equations in the Schwarzschild and Kerr backgrounds. Class. Quantum Gravity 28, 195025 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  40. N. Kudryashova, YuN Obukhov, On the dynamics of classical particles with spin. Phys. Lett. A 374, 3801 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  41. Yu.N. Obukhov, D. Puetzfeld, Dynamics of test bodies with spin in de Sitter spacetime. Phys. Rev. D 83, 044024 (2011)

    Google Scholar 

  42. W.G. Dixon, The definition of multipole moments for extended test bodies. Gen. Reletiv. Gravit. 4, 199 (1973)

    Article  ADS  Google Scholar 

  43. D. Bini, F. de Felice, A. Geralico, Strains in general relativity. Class. Quantum Gravity 23, 7603 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  44. W.G. Dixon, Dynamics of extended bodies in general relativity. III. Equations of motion. Philos. Trans. R. Soc. Lond. A 277, 59 (1974)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgments

We thank the participants of the 524 WE-Heraeus-Seminar for the enlightening discussions that helped shape this work. We thank also O. Semerák for useful correspondence, and Rui Quaresma (quaresma.rui@gmail.com) for his assistance in the illustrations. L.F.C. is funded by FCT through grant SFRH/BDP/85664/2012.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to L. Filipe O. Costa .

Editor information

Editors and Affiliations

Momentum and Angular Momentum in Curved Spacetime

Momentum and Angular Momentum in Curved Spacetime

In rectangular coordinates in flat spacetime, the momenta \(P^{\alpha }\) and \(S^{\alpha \beta }\) of an extended body, as measured by some observer of 4-velocity \(u^{\alpha }\), are well defined by the integrals

$$ P^{\alpha }=\int _{\Sigma (z,u)}T^{\alpha \beta }d\Sigma _{\beta };\qquad S^{\alpha \beta }=2\int _{\Sigma (z,u)}r^{[\alpha }T^{\beta ]\gamma }d\Sigma _{\gamma }, $$

where \(\Sigma (z,u)\) is the hyperplane orthogonal to \(u^{\alpha }\) (the rest space of \(u^{\alpha }\)), and \(r^{\alpha }=x^{\alpha }-z^{\alpha }\) is the vector connecting the reference worldline \(z^{\alpha }\) to the point of integration \(x^{\alpha }\). In curved spacetime the situation is different, as these integrals amount to summing tensors defined at different points; different generalizations of the flat spacetime notions have been proposed in the literature (e.g. [11, 25, 26]), none of them seeming a priori more natural than the others. Herein we discuss the mathematical meaning of the definitions used in this work, and how they relate to the schemes by Dixon [11, 26].

All schemes agree on generalizing \(\Sigma (z,u)\) by the geodesic hypersurface orthogonal to \(u^{\alpha }\), and on replacing \(r^{\alpha }\) by the vector \(\mathbf {X}\in \mathcal {T}_{z}\) tangent to the geodesic connecting \(z^{\alpha }\) and \(x^{\alpha }\), and whose length equals that of the geodesic. That is, \(\mathbf {X}=\Phi (x)\), where \(\Phi \equiv \exp _{z}^{-1}\) is the inverse exponential map, mapping points in the spacetime manifold to vectors in the tangent space \(\mathcal {T}_{z}\), \(\Phi :\mathcal {M\rightarrow }\mathcal {T}_{z}\). Where the schemes differ is in the way the vector \(\mathcal {A}^{\alpha }\equiv T^{\alpha \beta }d\Sigma _{\beta }\) is integrated. We adhere to the scheme proposed in [25]: using the natural map for tensors induced by \(\exp _{z}\) to pull back the energy-momentum tensor and the volume element to \(\mathcal {T}_{z}\), and integrate therein, which is then a well defined tensor operation. Let \(\varvec{\Omega }^{\hat{\alpha }}\) denote an orthonormal co-frame on \(\mathcal {T}_{z}\); the moments can then be written in the manifestly covariant form

$$\begin{aligned} \mathbf {P}(\varvec{\Omega }^{\hat{\alpha }})= & {} \int _{\Sigma (z,u)}\mathbf {T}(\Phi ^{*}\varvec{\Omega }^{\hat{\alpha }},d\varvec{\Sigma })\ ;\end{aligned}$$
(66)
$$\begin{aligned} \mathbf {S}(\varvec{\Omega }^{\hat{\alpha }},\varvec{\Omega }^{\hat{\beta }})= & {} 2\int _{\Sigma (z,u)}\mathbf {X}(\varvec{\Omega }^{[\hat{\alpha }})\mathbf {T}(\Phi ^{*}\varvec{\Omega }^{\hat{\beta }]},d\varvec{\Sigma }). \end{aligned}$$
(67)

Note that since \(\mathbf {T}(\Phi ^{*}\varvec{\Omega }^{\hat{\alpha }},d\varvec{\Sigma })=(\exp _{z}^{*}\mathbf {T})(\varvec{\Omega }^{\hat{\alpha }},\exp _{z}^{*}d\varvec{\Sigma })\), one is indeed pulling back the integrands from \(\mathcal {M}\) to \(\mathcal {T}_{z}\). Note also that Eqs. (66) and (67) are equivalent to (1) and (2), i.e. they just amount to perform the integration in a system of Riemann normal coordinates \(\{x^{\hat{\alpha }}\}\) centered at \(z^{\alpha }\) (the coordinates naturally adapted to the exponential map). This is because such system is constructed from geodesics radiating out of \(z^{\alpha }\); thus the components of \(\mathbf {X}\), in global Lorentz coordinates in \(\mathcal {T}_{z}\), are equal to the coordinates \(x^{\hat{\alpha }}\) on \(\mathcal {M}\); also the basis 1-forms of such system are the pullbacks of \(\varvec{\Omega }^{\hat{\alpha }}\) to \(\mathcal {M}\), \(dx^{\hat{\alpha }}=\Phi ^{*}\varvec{\Omega }^{\hat{\alpha }}\); and, taking it comoving with \(u^{\alpha }\) (i.e., at \(z, \partial _{\hat{0}}=\mathbf {u}\)), \(\Sigma (z,u)\) coincides with the spatial hypersurface \(x^{\hat{0}}=0\).

Let us now compare these definitions with other schemes in the literature. In [11], \(P^{\alpha }\) and \(S^{\alpha \beta }\) are defined as

$$\begin{aligned} P_{\mathrm{Dix}}^{\kappa }=\int _{\Sigma (z,u)}\bar{g}_{\alpha }^{\ \kappa }T^{\alpha \beta }d\Sigma _{\beta };\qquad S_{\mathrm{Dix}}^{\kappa \lambda }=-2\int _{\Sigma (z,u)}\sigma ^{[\kappa }\bar{g}_{\alpha }^{\ \lambda ]}T^{\alpha \beta }d\Sigma _{\beta }, \end{aligned}$$
(68)

where \(\sigma ^{\kappa }(x,z)=-(\Phi (x))^{\kappa }=-X^{\kappa }\), cf. [44]. These definitions thus differ from (66) and (67) only in the way the vector \(\mathcal {A}^{\alpha }\equiv T^{\alpha \beta }d\Sigma _{\beta }\) is integrated: \(\bar{g}_{\alpha }^{\ \kappa }\) is a bitensor which parallel transports \(\mathcal {A}^{\alpha }\) at \(x^{\alpha }\) to \(z^{\kappa }\) along the geodesic connecting the two points, so that the integral is performed over vectors \(\mathcal {A}^{\kappa }|_{z}=\bar{g}_{\alpha }^{\ \kappa }\mathcal {A}^{\alpha }|_{x}\) defined at \(z^{\kappa }\) (in [26, 44] different propagators, \(K_{\alpha }^{\ \kappa }\), \(H_{\alpha }^{\ \kappa }\) in the notation therein, are employed; the two schemes are not equivalent though, as noted in [26]). Writing \(\bar{g}_{\hat{\beta }}^{\ \hat{\alpha }}\mathcal {A}^{\hat{\beta }}|_{x}=\mathcal {A}^{\hat{\alpha }}|_{x}+\Delta \mathcal {A}^{\hat{\alpha }}\), with \(\Delta \mathcal {A}^{\hat{\alpha }}=-\int _{x}^{z}\Gamma _{\hat{\beta }\hat{\gamma }}^{\hat{\alpha }}(x')\mathcal {A}^{\hat{\beta }}dx'^{\hat{\gamma }}\), expanding the integrand in Taylor series around \(z^{\alpha }\), and noting that, in the normal coordinates \(\{x^{\hat{\alpha }}\}\) (see e.g. [27]), we have \(\Gamma _{\hat{\beta }\hat{\gamma }}^{\hat{\alpha }}(z)=0\) and \(\Vert \Gamma _{\hat{\beta }\hat{\gamma },\hat{\delta }}^{\hat{\alpha }}(z)\Vert \sim \Vert \mathbf {R}\Vert \), where \(\Vert \mathbf {R}\Vert \equiv \sqrt{|R_{\alpha \beta \gamma \delta }R^{\alpha \beta \gamma \delta }|}\) denotes the magnitude of the curvature, we have \(\Delta \mathcal {A}^{\alpha }=\mathcal {O}(\Vert \varvec{\mathcal {A}}\Vert \Vert \mathbf {R}\Vert x^{2})\). Therefore \(P_{\mathrm{Dix}}^{\hat{\alpha }}=P^{\hat{\alpha }}+\mathcal {O}(\lambda \Vert \mathbf {P}\Vert )\), where

$$\begin{aligned} \lambda =\Vert \mathbf {R}\Vert a^{2}\ , \end{aligned}$$
(69)

and a is the largest dimension of the body. Thus, when \(\lambda \ll 1\), i.e. when the curvature is not too strong compared to the scale of the size of the body,Footnote 13 \(P_{\mathrm{Dix}}^{\hat{\alpha }}\simeq P^{\hat{\alpha }}\). The two schemes are actually indistinguishable in a pole-dipole approximation, where only terms to linear order in x are kept in the integrals defining the moments; the resulting equations of motion are the same (compare Eqs. (43), (49) of [25] with Eqs. (6.31) and (6.32) of [11], or Eqs. (7.1) and (7.2) of [26]), both schemes leading to the well known Mathisson-Papapetrou equations (the latter derived using less sophisticated formalisms). These conclusions are natural, for the metric in Riemann normal coordinates is (e.g. [27]) of the form \(g_{\hat{\alpha }\hat{\beta }}=\eta _{\hat{\alpha }\hat{\beta }}+\mathcal {O}(\Vert \mathbf {R}\Vert x^{2})\); hence the assumption \(\lambda \ll 1\) amounts to say that, for the computation of \(P^{\alpha }\) and \(S^{\alpha \beta }\), one may, to a good approximation, take the spacetime as nearly flat throughout the body.

1.1 The Dependence of the Particle’s Momenta on \(\Sigma \)

The momenta (1) and (2) depend, in general, on the spacelike hypersurface \(\Sigma (z,u)\equiv \Sigma (z(\tau ),u)\) on which the integration is performed, see e.g. [11, 25, 26]. This is so even in flat spacetime; when forces and torques act on the body, it is clear that \(P^{\alpha }(z,u)\), \(S^{\alpha \beta }(z,u)\) depend on \(z^{\alpha }(\tau )\), and also on the argument \(u^{\alpha }\) of \(\Sigma \). Curvature brings additional complications, as \(u^{\alpha }\) is no longer a “free vector”, and \(\Sigma \) itself is in principle point dependent. Herein we shall show that, in the absence of electromagnetic field (\(F^{\alpha \beta }=0\)), and under the assumption \(\lambda \ll 1\) made above, for hypersurfaces \(\Sigma (z,u)\)through a point \(z^{\alpha }\)within the body’s convex hull, the \(u^{\alpha }\) dependence of the momentum and angular momentum is negligible.

Denote by \(\varvec{\xi }=dx^{\hat{\alpha }}\) a particular basis 1-form of the Riemann normal coordinate system \(\{x^{\hat{\alpha }}\}\); \(P^{\varvec{\xi }}\equiv P^{\alpha }\xi _{\alpha }\) is thus the \(\varvec{\xi }\) component of \(P^{\alpha }\). From definition (1), and since \(\xi _{\alpha }\) has constant components, we may write the \(\varvec{\xi }\) component of the momentum as the integral of a vector \(A^{\alpha }\equiv T^{\alpha \beta }\xi _{\beta }\) on a 3-surface,

$$ P^{\varvec{\xi }}(z,u)=\xi _{\hat{\alpha }}\int _{\Sigma (z,u)}T^{\hat{\alpha }\hat{\beta }}d\Sigma _{\hat{\beta }}=\int _{\Sigma (z,u)}T^{\hat{\alpha }\hat{\beta }}\xi _{\hat{\alpha }}d\Sigma _{\hat{\beta }}=\int _{\Sigma (z,u)}A^{\beta }d\Sigma _{\beta }\ . $$

Take \(u^{\alpha }=P^{\alpha }/M\), and consider another vector \(u'^{\alpha }\) at the same point \(z^{\alpha }\); the \(\varvec{\xi }\) component of the difference between the momenta computed in the hypersurfaces \(\Sigma (z,u')\) and \(\Sigma (z,u)\), \(\Delta P^{\varvec{\xi }}\equiv P^{\varvec{\xi }}(z,u')-P^{\varvec{\xi }}(z,u)\) is, from an application of the Gauss theorem (see Fig. 7),

$$ \Delta P^{\varvec{\xi }}=\int _{\Sigma (z,u)}A^{\beta }d\Sigma _{\beta }-\int _{\Sigma (z,u')}A^{\beta }d\Sigma _{\beta }=\int _{V_{Left}}A_{\ ;\beta }^{\beta }dV-\int _{V_{Right}}A_{\ ;\beta }^{\beta }dV\ . $$
Fig. 7
figure 7

Shadowed regions \(V_{Left}\) and \(V_{Right}\) are the 4-volumes delimited by the hypersurfaces \(\Sigma (z,u')\), \(\Sigma (z,u)\), and the boundary of the body’s worldtube, of convex hull W. \(u^{\alpha }\) is chosen parallel to \(P^{\alpha }\). a Curved spacetime; b flat spacetime

Here \(V_{Left}\) and \(V_{Right}\) denote the shadowed regions of Fig. 7 (where \(A^{\alpha }\ne 0\)), i.e. the “left” and “right” 4-volumes delimited by \(\Sigma (z,u')\), \(\Sigma (z,u)\) and the boundary of the body’s worldtube. Now, using the conservation law \(T_{\ \ ;\beta }^{\alpha \beta }=0\), one notes that

$$ A_{\ ;\beta }^{\beta }=T_{\ \ ;\beta }^{\alpha \beta }\xi _{\alpha }+T^{\alpha \beta }\xi _{\alpha ;\beta }=T^{\alpha \beta }\xi _{\alpha ;\beta }; $$

thus

$$ \Delta P^{\varvec{\xi }}=\int _{V_{Left}}T^{\alpha \beta }\xi _{\alpha ;\beta }dV-\int _{V_{Right}}T^{\alpha \beta }\xi _{\alpha ;\beta }dV. $$

Since \(\varvec{\xi }\) is a basis 1-form, \(\xi _{\hat{\alpha },\hat{\beta }}=0\), and

$$ \xi _{\hat{\alpha };\hat{\beta }}=-\Gamma _{\hat{\alpha }\hat{\beta }}^{\hat{\gamma }}\xi _{\hat{\gamma }}=\mathcal {O}(\Vert \mathbf {R}\Vert x); $$

therefore

$$ |\Delta P^{\varvec{\xi }}|\lesssim \Vert \mathbf {R}\Vert \int _{V}T^{\hat{0}\hat{0}}|x|dV=\Vert \mathbf {R}\Vert V\left\langle T^{\hat{0}\hat{0}}|x|\right\rangle , $$

where \(V\equiv V_{Left}+V_{Right}\), \(\left\langle \,\cdot \,\right\rangle \) denotes the average on the shadowed region of Fig. 7a, and we noted that \(T^{\hat{0}\hat{0}}\) is the largest component of \(T^{\alpha \beta }\) and always positive. Since \(b<av(u',u)\) (see Fig. 7), with \(v^{\alpha }(u',u)\) defined by Eq. (9), and \(v(u',u)<1\), then \(\left\langle |x|\right\rangle <a\); moreover (assuming \({\partial _{\hat{0}} = \mathbf {u}\,\mathrm{at}\,z}\)), \(V\left\langle T^{\hat{0}\hat{0}}\right\rangle <Mav(u',u)\); hence we get

$$\begin{aligned} |\Delta P^{\varvec{\xi }}|\lesssim M\lambda v(u',u)=\Vert \mathbf {P}\Vert \lambda v(u',u)\ , \end{aligned}$$
(70)

showing that \(\Delta P^{\alpha }\) is negligibleFootnote 14 compared to \(P^{\alpha }\) under the restriction above on the strength of the gravitational field, \(\lambda \ll 1\) (the same under which the different multipole schemes become equivalent, and one can take local Lorentz coordinates as nearly rectangular throughout the extension of the body; see also footnote 13). In the application in Sect. 3.4—Schwarzschild spacetime, far field limit—we can write

$$ |\Delta P^{\varvec{\xi }}|\lesssim \frac{Mm_{\mathrm{S}}}{r^{3}}a^{2}v(u',u)\simeq \Vert P_{\mathrm{hidI}}\Vert \frac{a}{R_{\mathrm{Moller}}}\frac{a}{r}v(u',u) $$

where \(P_{\mathrm{hidI}}^{\alpha }\) is the inertial hidden momentum of the CP condition, Eq. (57) (\(P_{\mathrm{hidI}}^{\alpha }\) is zero or negligible for the other solutions). Thus \(\Delta P^{\alpha }\) is negligible compared to \(P_{\mathrm{hidI}}^{\alpha }\) under the condition \(\frac{R_{\mathrm{Moller}}}{a}\gg \frac{a}{r}v(u',u)\), which is reasonable in a problem where the particle’s spin is worth taking into account (e.g. in the problem of nearly circular motion in Sect. 3.4 this amounts to taking \(\omega _{\mathrm{body}}\gg \omega _{\mathrm{orbit}}\), where \(\omega _{\mathrm{body}}\) and \(\omega _{\mathrm{orbit}}\) are the body’s rotation and orbital angular velocities).

Through an analogous procedure, one can show that the dependence of \(S^{\alpha \beta }\) on \(u^{\alpha }\) is negligible in this regime. Let \(\fancyscript{J}^{\hat{\alpha }\hat{\beta }\hat{\gamma }}\equiv 2x^{[\hat{\alpha }}T^{\hat{\beta }]\hat{\gamma }}\), so that \(S^{\hat{\alpha }\hat{\beta }}=\int _{\Sigma (z,u)}\fancyscript{J}^{\hat{\alpha }\hat{\beta }\hat{\gamma }}d\Sigma _{\hat{\gamma }}\); and consider the two basis spatial 1-forms \(\varvec{\xi }\) and \(\varvec{\eta }\). Constructing the vector \(\fancyscript{J}^{\gamma }\equiv \fancyscript{J}^{\alpha \beta \gamma }\xi _{\alpha }\eta _{\beta }\), we can write the \(\varvec{\xi }\otimes \varvec{\eta }\) component of \(S^{\alpha \beta }\) as \(S^{\varvec{\xi \eta }}(z,u)=\int _{\Sigma (z,u)}\fancyscript{J}^{\beta }d\Sigma _{\beta }\). By the Gauss theorem,

$$\begin{aligned} \Delta S^{\varvec{\xi \eta }}=&\int _{V_{Left}}\fancyscript{J}_{\ ;\beta }^{\beta }dV-\int _{V_{Right}}\fancyscript{J}_{\ ;\beta }^{\beta }dV\ \sim \ \Vert \mathbf {R}\Vert \int _{V}x^{2}|\vec {J}|dV=\Vert \mathbf {R}\Vert V\langle |\vec {J}|x^{2}\rangle \\&<\Vert \mathbf {R}\Vert a^{2}V\langle |\vec {J}|\rangle =\lambda V\langle |\vec {J}|\rangle \lesssim \lambda a^{4}v(u',u)\langle |\vec {J}|\rangle \ , \end{aligned}$$

where \(J^{\hat{i}}=T^{\hat{0}\hat{i}}\). In the second relation again we used \(\Gamma _{\hat{\alpha }\hat{\beta }}^{\hat{\gamma }}=\mathcal {O}(\Vert \mathbf {R}\Vert x)\). Since \(S=\mathcal {O}(a^{4}\langle |\vec {J}|\rangle )\), cf. Eq. (13), we see that indeed when \(\lambda \ll 1\), \(\Vert \Delta S^{\varvec{\xi \eta }}\Vert \ll S\).

The Case with Electromagnetic Field

When \(F^{\alpha \beta }\ne 0\), the conservation law is \(T_{\ \ ;\beta }^{\alpha \beta }=F^{\alpha \beta }j_{\beta }\) (denoting by \(T^{\alpha \beta }\)the particle’s energy momentum tensor). Consider for simplicity flat spacetime, and let \(\varvec{\xi }\) be a basis 1-form of a global Lorentz system; then \(T_{\ \ ;\beta }^{\alpha \beta }\xi _{\alpha }=(T^{\alpha \beta }\xi _{\alpha })_{;\beta }\equiv A_{\ ;\beta }^{\beta }=F^{\alpha \beta }j_{\beta }\xi _{\alpha }\). Note that \(f^{\alpha }\equiv F^{\alpha \beta }j_{\beta }\) is the Lorentz force density. It follows (see Fig. 7b)

$$\begin{aligned} \Delta P^{\varvec{\xi }}=&\int _{V_{Left}}A_{\ ;\beta }^{\beta }dV-\int _{V_{Right}}A_{\ ;\beta }^{\beta }dV=\xi _{\alpha }\left( V_{Left}\left\langle f^{\alpha }\right\rangle _{Left}-V_{Right}\left\langle f^{\alpha }\right\rangle _{Right}\right) . \end{aligned}$$

We have \(V_{Left}=V_{Left}^{(3)}b_{Left}/2\), \(V_{Right}=V_{Right}^{(3)}b_{Right}/2\) (where \(V^{(3)}\) denote 3-volumes orthogonal to \(u^{\alpha }\)). Herein we allow \(z^{\alpha }\) to be any point within the worldtube of centroids; it follows that

$$\begin{aligned} b_{Left}\ge v(u,u')\left( \frac{a}{2}-R_{\mathrm{Moller}}\right) ;&\qquad b_{Right}\le v(u,u')\left( \frac{a}{2}+R_{\mathrm{Moller}}\right) \\ V_{Left}^{(3)}\sim a^{2}\left( \frac{a}{2}-R_{\mathrm{Moller}}\right) ;&\qquad V_{Right}^{(3)}\sim a^{2}\left( \frac{a}{2}+R_{\mathrm{Moller}}\right) . \end{aligned}$$

Let \(\left\langle f^{\alpha }\right\rangle _{Left}=\left\langle f^{\alpha }\right\rangle _{Right}+\Delta f^{\alpha }\), with \(\Vert \Delta f^{\alpha }\Vert \lesssim \Vert \nabla _{\beta }f^{\alpha }\Vert a\); we obtain

$$\begin{aligned} |\Delta P^{\varvec{\xi }}|\lesssim \Vert F_{\mathrm{L}}^{\alpha }\Vert R_{\mathrm{Moller}}v(u',u)+\Vert \nabla _{j}F_{\mathrm{L}}^{\alpha }\Vert v(u',u)a^{2}. \end{aligned}$$
(71)

Hence \(\Delta P^{\varvec{\xi }}\) has, as upper bound, the sum of two terms: the impulse of the Lorentz force \(F_{L}^{\alpha }\) in the time interval \(R_{\mathrm{Moller}}v(u',u)\) (as measured in the \(u^{i}=0\) frame) between the two points where the hyperplane \(\Sigma (z,u')\) crosses the worldtube of centroids, plus a term analogous to the gravitational one (70). For the field of a Coulomb charge, discussed in Sect. 3.4.1, they read

$$\begin{aligned} \Vert F_{\mathrm{L}}^{\alpha }\Vert R_{\mathrm{Moller}}v(u',u)&=|E_{\mathrm{p}}|v(u',u)\frac{R_{\mathrm{Moller}}}{r}\sim \Vert P_{\mathrm{hidI}}\Vert v(u',u)\\ \Vert \nabla _{j}F_{\mathrm{L}}^{\alpha }\Vert v(u',u)a^{2}&=|E_{\mathrm{p}}|v(u',u)\frac{a^{2}}{r^{2}}\sim \Vert P_{\mathrm{hidI}}\Vert \frac{a}{R_{\mathrm{Moller}}}\frac{a}{r}v(u',u) \end{aligned}$$

where \(E_{\mathrm{p}}=qQ/r\) is the electric potential energy, and \(P_{\mathrm{hidI}}^{\alpha }\) is the inertial hidden momentum of the TD/FMP (non-helical) solutions, Eq. (52) (for the CP/OKS conditions, \(P_{\mathrm{hidI}}^{\alpha }=0\)). Assuming \(|E_{\mathrm{p}}|<M\), if \(R_{\mathrm{Moller}}/r\ll 1\) and \(a^{2}/r^{2}\ll 1\) (as is the case in the far-field regime), then \(|\Delta P^{\varvec{\xi }}| \ll M = \parallel \!\!\mathbf{P}\!\! \parallel \), and \(\Delta {P}^{\alpha }\) is negligible compared to \(P^{\alpha }\) by arguments analogous the ones given in footnote 14. It is also negligible compared to \(\Vert P_{\mathrm{hidI}}\Vert \) under the following conditions: (i) \(v(u',u)\ll 1\) so that the first term of (71) can be neglected (this is guaranteed by the slow motion assumption in Sect. 3.4); (ii) that \(\frac{R_{\mathrm{Moller}}}{a}\gg \frac{a}{r}v(u',u)\), a condition analogous to the one we obtained gravitational case above, which is reasonable whenever the particle’s spin is worth taking into account.

Note that the argument above can equally be used to show that \(P^{\alpha }\) does not depend on the spin condition. Start with the TD centroid: \(z^{\alpha }=x_{\mathrm{CM}}^{\alpha }(u)\), with \(u^{\alpha }=P^{\alpha }/M\); the centroids \(x_{\mathrm{CM}}^{\alpha }(u')\) of other spin conditions are reached by \(x_{\mathrm{CM}}^{\alpha }(u')=x_{\mathrm{CM}}^{\alpha }(u)+\Delta x^{\alpha }\), with \(\Delta x^{\alpha }\in \Sigma (u,z)\), cf. Eq. (10). Since the argument above applies to any spacelike hyperplane \(\Sigma (u',z')\) through any arbitrary centroid \(z'^{\alpha }\) on \(\Sigma (u,z)\), it effectively means that, to the accuracy at hand, \(P^{\alpha }\) does not depend on the particular centroid chosen.

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Costa, L.F.O., Natário, J. (2015). Center of Mass, Spin Supplementary Conditions, and the Momentum of Spinning Particles. In: Puetzfeld, D., Lämmerzahl, C., Schutz, B. (eds) Equations of Motion in Relativistic Gravity. Fundamental Theories of Physics, vol 179. Springer, Cham. https://doi.org/10.1007/978-3-319-18335-0_6

Download citation

Publish with us

Policies and ethics