This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

General theory of swimming in curved spacetime

Rodrigo Andrade e Silva [email protected] University of Maryland College Park, College Park, MD 20742, USA
Abstract

Swimming in curved spacetime is a phenomenon in which free bodies in curved spacetimes are able to propel themselves by performing cyclic internal motions. Most surprisingly, it has been suggested that, in the limit of fast cycles, the effective motion would not depend on the rapidity of the internal motions but only on the sequence of shapes assumed by the body, just like a swimmer in a non-turbulent viscous fluid. Since a rigorous treatment of this problem was still missing, in this paper we present a general, covariant theory of swimming in curved spacetime. We show that the effect can indeed take place, but it only displays this geometric-phase character in special circumstances (even within the limit of fast cycles). Our methods are fairly general, providing a technique to study the motion of free small articulated bodies in general relativity by mapping the problem to an analogue in special relativity.

I Introduction

Swimming in curved spacetime is a phenomenon where small bodies evolving freely in curved spacetimes appear to be able to propel themselves, by performing fast cyclic internal motions, in a “non-dynamical” manner [1]. More precisely, the resulting motion would depend only on the sequence of conformational changes of the body and not on how fast they are performed, similarly to the motion of a swimmer in a fluid at low Reynolds numbers [2]. But a rigorous general treatment of this problem, consistent with general relativity, was still missing. The main goal of this paper is then to fill this gap.

The problem of interest consists of describing the trajectory of a light, small, articulated body, performing a given sequence of internal motions, in a curved spacetime. The body is light so that it can be regarded as a test body in a fixed background spacetime, it is small in comparison to the length scale over which the curvature varies, and it is articulated in the sense that it can adjust its shape in a controlled manner (say, it has a couple of arms which can be willfully stretched or waved). In order to solve this problem we employ the theory of dynamics developed by Dixon [4, 5, 6]. Dixon’s formalism has the valuable property of being automatically consistent with the fundamental conservation equation aTab=0\nabla_{a}T^{ab}=0 of general relativity, which is directly implemented into the formalism from the very beginning. In addition, the equations of motion are given in a convenient multipole-expansion form which makes clear how the overall motion is affected by different orders in l/rl/r, where ll is the (spatial) length scale of the body and rr is the characteristic (spatial) length scale that the gravitational field varies.

We will divide this paper into four sections, beginning with a succinct review of Dixon’s formalism, which will be the basis for the subsequent developments. Next, we deduce from Dixon’s equations an approximate equation for the trajectory of the center of mass of a general “small” body, in terms of the net gravitational force and torque acting on it. Basically we shall find expressions for the velocity of the center of mass of the body, correct up to order (l/r)3(l/r)^{3}, so that we may keep only terms up to octupole order in the expressions for the force and torque. Then, we proceed to discuss a convenient manner to prescribe the quadrupole and octupole moments, based on a local correspondence to a flat spacetime. Finally, considering the equations for the trajectory of the body, we are able to formally study the effective dynamics induced by fast cyclic internal motions. With the general formalism established, we proceed to analyze a regime suitable for “swimming at low Reynolds number”, showing that although the effect can take place, it is limited to specific scenarios (which does not mean that bodies are generally incapable of propelling themselves by performing swimming-like internal motions, but only that the resulting motion is typically not described by a geometric phase). In Appendix B we show how to apply the formalism to a concrete example, putting our theory to the test. This paper is supplementary to [3], where a summary of these results are presented.

We use Latin letters for abstract tensor indices and Greek letters for tensor components in a basis. The curvature tensor is defined by (abba)ωc=Rabcdωd\left(\nabla_{a}\nabla_{b}-\nabla_{b}\nabla_{a}\right)\omega_{c}=-R_{abc}^{\;\;\;\;\,\,d}\,\omega_{d}, and the Ricci tensor and scalar curvature are respectively defined by Rac=RabcbR_{ac}=-R_{abc}^{\;\;\;\;\,\,b} and R=RaaR=R_{a}^{\;\;a}. The signature of the metric is (+,,,)(+,-,-,-) and the system of units is such that c=G=1c=G=1.

II Dixon’s formalism, a short summary

In a series of papers published in the 70’s [4, 5, 6], W. G. Dixon developed a formal theory to describe the dynamics of extended bodies in general relativity, in an explicitly covariant way. This section is dedicated to review (in a rather compressed way) the main results of his theory111Dixon’s theory is actually more general than this, allowing for the presence of external electromagnetic forces..

Let TabT^{ab} be the stress-energy tensor of the body under consideration, assumed to be conserved (i.e., aTab=0\nabla_{a}T^{ab}=0), and define the world-tube WW of the body as the support of TabT^{ab}. Then, under reasonable assumptions about TabT^{ab} 222It is assumed that the spatial extension of the body is finite and such that the convex hull {\cal B} of the intersection of its world-tube WW with any space-like hypersurface Σ\Sigma is contained in a normal neighborhood of any point of {\cal B}. This means, in particular, that there is only one geodesic in {\cal B} connecting any two spatially-separated points of the body. Moreover, wherever Tab0T^{ab}\neq 0, the momentum density is assumed to be time-like and future-directed according to all observers; i.e., Tabnanb>0T_{ab}n^{a}n^{b}>0 and TabTcbnanc>0T_{ab}T^{b}_{\,\;c}n^{a}n^{c}>0 for all time-like vectors nan^{a}. Finally, these conditions are supposed to hold also for the momentum density “propagated” from any point xx to any point xx^{\prime}, both in {\cal B}, where the “propagation” is performed along the geodesic connecting xx and xx^{\prime}, according to maps KbaK^{a^{\prime}}_{\;\;b} and HbaH^{a^{\prime}}_{\;\;b} which are properly defined in Appendix A. and somewhat weak conditions on the “strength” of the gravitational field [9], the following general results hold:

  • (i)

    At any point xWx\in W, there is a unique time-like, future-pointing, unit vector na=na(x)n^{a}=n^{a}(x) such that the linear momentum pa=pa(x,n)p^{a}=p^{a}(x,n) of the body (properly defined [4] with respect to xx and nan^{a}) is entirely in the direction of nan^{a}. In essence, this means that there is a (unique) family of observers (those with four-velocity ua=nau^{a}=n^{a}) according to whom the total spatial momentum of the body is zero. In addition, at each point xWx\in W, there is a natural definition for the angular momentum Skl=S[kl]=Skl(x,n)S^{kl}=S^{[kl]}=S^{kl}(x,n) of the body;

  • (ii)

    There is a unique, covariantly defined, time-like curve z(s)z(s), contained in the convex hull of WW, which can be consistently identified as the world-line of the center of mass of the body (here ss is the proper time along this world-line). More specifically, it is defined as the set of points zz which satisfy pk(z,n(z))Skl(z,n(z))=0p_{k}(z,n(z))S^{kl}(z,n(z))=0. It is worth mentioning that the tangent vector to this curve at z(s)z(s), va(s)v^{a}(s), is not, in general, parallel to ua(s)=na(z(s))u^{a}(s)=n^{a}(z(s)). We refer to uau^{a} and vav^{a} respectively as the dynamical and kinematical velocities of the body;

  • (iii)

    At each point z(s)z(s) on the world-line of the center of mass, there are also natural definitions for the (higher) multipole moments Jk1knpqlr(s)J^{k_{1}\cdots k_{n}pqlr}(s), n0n\geq 0, of the body 333We are referring to these as higher multipole moments because the linear and angular momenta, pkp^{k} and SklS^{kl}, can also be regarded as multipole moments – respectively, the monopole and dipole moments.. The set of tensors {pk,Skl}{Jk1knpqlr;n0}\{p^{k},S^{kl}\}\cup\{J^{k_{1}\cdots k_{n}pqlr};n\geq 0\} depends, at each ss, only on the restriction of TabT^{ab} to WΣ(s)W\cap\Sigma(s), where Σ(s)\Sigma(s) is the space-like surface generated by all geodesics starting at z(s)z(s) orthogonal to uk(s)u^{k}(s). Conversely, if the tensors in this set are given for all ss, then TabT^{ab} can be completely reconstructed;

  • (iv)

    The conservation equation aTab=0\nabla_{a}T^{ab}=0 can be translated into only two dynamical equations for the linear and angular momenta

    δpkds\displaystyle\frac{\delta p^{k}}{ds} =\displaystyle= Fk\displaystyle F^{k}
    δSklds\displaystyle\frac{\delta S^{kl}}{ds} =\displaystyle= 2p[kvl]+Qkl,\displaystyle 2p^{[k}v^{l]}+Q^{kl}\,, (1)

    where δ/ds:=vkk\delta/ds:=v^{k}\nabla_{k} is the covariant derivative operator along z(s)z(s), and FkF^{k} and QklQ^{kl} are respectively called the gravitational force and torque acting on the body (and they can be expressed in terms of couplings between the Riemann curvature tensor or its derivatives and the multipole moments);

  • (v)

    The proper mass of the body, MM, is defined as the scalar in the expression pk=Mukp^{k}=Mu^{k}. If ϑklqr\vartheta_{klqr} is the volume element of the spacetime (associated with the metric), then we define the spin vector of the body as Sk=12ϑklqrulSqrS_{k}=\frac{1}{2}\vartheta_{klqr}u^{l}S^{qr}, and it satisfies

    δSkds=uku˙lSl+Qk,\displaystyle\frac{\delta S^{k}}{ds}=-u^{k}\dot{u}_{l}S^{l}+Q^{k}\,, (2)

    where u˙k=δuk/ds{\dot{u}}^{k}=\delta u^{k}/ds and the torque vector QkQ^{k} is defined from QklQ^{kl} similarly to the spin vector. Then, the misalignment between uku^{k} and vkv^{k} is given by

    u[kvl]=12M2ϑklqrF¯qSr+1Mu[kQl]quq,\displaystyle u^{[k}v^{l]}=\frac{1}{2M^{2}}\vartheta^{klqr}\bar{F}_{q}S_{r}+\frac{1}{M}u^{[k}Q^{l]q}u_{q}\,\,, (3)

    where F¯k=(δlkukul)Fl\bar{F}^{k}=({\delta}^{k}_{\>\>{l}}-u^{k}u_{l})F^{l}. Note that this provides an equation for vkv^{k} in terms of the linear momentum, spin, force and torque.

  • (vi)

    For each (if any) symmetry of the spacetime with generator ξa\xi^{a}, there is a conserved dynamical quantity χξ\chi_{\xi} given by

    χξ=pkξk+12Sklkξl,\displaystyle\chi_{\xi}=p^{k}\xi_{k}+\frac{1}{2}S^{kl}\nabla_{k}\xi_{l}\,\,, (4)

    which can be seen as a version of Noether’s theorem.

  • (vii)

    For a spatially small test-body, one can keep only a few multipole terms in the expressions for the force and torque. Up to octupole order, they are

    Fk=\displaystyle F^{k}= 12vlSqrRklqr+16JqrlwkRqrlw+\displaystyle\frac{1}{2}v^{l}S^{qr}{R^{k}}_{lqr}+\frac{1}{6}J^{qrlw}\nabla^{k}R_{qrlw}+
    +118Jqrwtlk(qRrwtl2tRqrwl),\displaystyle+\frac{1}{18}J^{qrwtl}\nabla^{k}\left(\nabla_{q}R_{rwtl}-2\nabla_{t}R_{qrwl}\right)\,,
    Qkl=\displaystyle Q^{kl}= 43Jqrw[kRqrwl]+\displaystyle\frac{4}{3}J^{qrw[k}{R_{qrw}}^{l]}+
    +13Jqrwt[k(3qRrwtl]+2tRqrwl]),\displaystyle+\frac{1}{3}J^{qrwt[k}\left(3\nabla_{q}{R_{rwt}}^{l]}+2\nabla_{t}{R_{qrw}}^{l]}\right)\,, (5)

    and we shall refer to the three terms in the expression for the force as, respectively, FspinkF^{k}_{spin}, FquadkF^{k}_{quad} and FoctukF^{k}_{octu}; and to the two terms in the expression for the torque as, respectively, QquadklQ^{kl}_{quad} and QoctuklQ^{kl}_{octu}.

III Octupole-order equations of motion

The problem of solving for the dynamical evolution of a finite body in general relativity can be formulated as an initial-value problem with Dixon’s equations. Hence, we shall assume that the initial position of the center of mass of the body z(s0)z(s_{0}) is known, as well as the initial linear momentum pk(s0)p^{k}(s_{0}), the initial spin Sk(s0)S^{k}(s_{0}), and the initial multipole moments J(s0)J(s_{0}) (the indices are omitted). The worldline of the center of mass \ell is parameterized as z(s)z(s), where ss is the proper time along \ell, so that the tangent vector vkv^{k} has unit modulus.

In a coordinate system {xμ}\{x^{\mu}\} around z(s0)z(s_{0}), the trajectory zμ(s)z^{\mu}(s) of the center of mass of the body satisfies

vμ(s)=dzμds,v^{\mu}(s)=\frac{dz^{\mu}}{ds}\,, (6)

where vμv^{\mu} can be obtained algebraically from (3), in terms of other dynamical quantities like pμp^{\mu}, SμS^{\mu} and the JJ’s. This equation can then be solved in conjunction with (1), as long as we know how to evolve the JJ’s. The most direct approach for determining the moments is to translate the internal laws governing the substances constituting the body into differential equations for the JJ’s. In this case, depending on the order of the system of differential equations, the values of the JJ’s and the appropriate set of their time derivatives dJ/dsdJ/ds, d2J/ds2d^{2}\!J/ds^{2}, etc must be provided at s=s0s=s_{0}. We shall, nonetheless, adopt a different approach to specify the moments (see next section), significantly more appropriate for our purposes. In our description, we will find a way to evaluate the multipole moments as functions of the internal motions performed by the body. Thus, the internal motions (and the JJ’s) should be seen as part of the input of the problem, not the output.

Our task now is to find an explicit formula for the kinematical velocity vkv^{k} in terms of dynamical quantities like uku^{k}, the spin and the force and torque. This formula does not need to be exact, but it must be accurate up to order (l/r)3(l/r)^{3}. The motivation for assuming this degree of accuracy comes from (i)(i) the fact that the swimming effect is expected to be observed at this order444The Newtonian analysis suggests that the swimming effect should be observed at order (l/r)2(l/r)^{2}. The analysis of the motion of a bipod moving freely on a spacetime with topology ×S3\mathbb{R}\times S^{3} shows that the swimming propulsion can indeed be at this same order. The study of a tripod falling aligned with the radial direction of Schwarzschild spacetime, in a specific setup, yields that the propulsion caused by the internal motions takes place at order (l/r)3(l/r)^{3} only. and that (ii)(ii) the equations are still sufficiently simple and manageable (for instance, we can only keep terms up to octupole order in the equations of motion).555Dixon’s equations have been used to study the motion of extended bodies in curved spacetimes, for example in [7, 8], but here we focus on the application to the swimming effect.

Let us briefly study the magnitudes, in orders of l/rl/r, of the various terms in the equations of motion. Typically, in a (normal) coordinate system centered at z(s)z(s), aligned with uk(s)u^{k}(s), the components of the energy-momentum tensor TμνT^{\mu\nu} have comparable magnitudes. The components of the multipole moments, in the same coordinate system, are usually 666In the low “internal” velocities regime, some of these components can be much smaller than implied, so we can think of these magnitudes as upper bounds in the subsequent analysis. of the following orders of magnitude

pμM,SμνMl,JμνλσMl2,JμνλσρMl3,\displaystyle p^{\mu}\sim M,\,\,S^{\mu\nu}\sim Ml,\,\,J^{\mu\nu\lambda\sigma}\sim Ml^{2},\,\,J^{\mu\nu\lambda\sigma\rho}\sim Ml^{3}\,\,,

where M=pkpkM=\sqrt{p^{k}p_{k}} is the total mass of the body and ll is its characteristic (spatial) length scale (as measured over the hypersurfaces Σ(s)\Sigma(s) orthogonal to uku^{k}). If it is possible to write the components of the curvature tensor and its first derivative as

Rμνλσkr2,ρRμνλσkr3,\displaystyle R_{\mu\nu\lambda\sigma}\sim\frac{k}{r^{2}}\,,\quad\nabla_{\rho}R_{\mu\nu\lambda\sigma}\sim\frac{k}{r^{3}}\,, (7)

where kk is some dimensionless factor (not necessarily constant) and rr is some parameter with dimension of length (to be interpreted as the local radius of curvature), then each term in (5) has the respective order of magnitude

FμkMr(lr)+kMr(lr)2+kMr(lr)3,\displaystyle F^{\mu}\sim\frac{kM}{r}{\left(\frac{l}{r}\right)}+\frac{kM}{r}{\left(\frac{l}{r}\right)\!}^{2}+\frac{kM}{r}{\left(\frac{l}{r}\right)\!}^{3}\,\,,
QμνkM(lr)2+kM(lr)3,\displaystyle Q^{\mu\nu}\sim kM{\left(\frac{l}{r}\right)\!}^{2}+kM{\left(\frac{l}{r}\right)\!}^{3}\,\,, (8)

where the order-to-order separation, in the parameter l/rl/r, is evident.

In order to find the desired formula for vkv^{k}, we first write

u[kvl]=ϵkl,u^{[k}v^{l]}=\epsilon^{kl}\,, (9)

where ϵkl\epsilon^{kl} is an anti-symmetric tensor corresponding to the right-hand side of equation (3). We see that it has order

ϵμνk[(lr)2+(lr)3+(lr)4]+k[(lr)2+(lr)3],\epsilon^{\mu\nu}\sim k\left[{\left(\frac{l}{r}\right)\!}^{2}+{\left(\frac{l}{r}\right)\!}^{3}+{\left(\frac{l}{r}\right)\!}^{4}\right]+k\left[{\left(\frac{l}{r}\right)\!}^{2}+{\left(\frac{l}{r}\right)\!}^{3}\right]\,\,, (10)

where the first three terms come from the contraction of the force and spin (inside the square brackets we have, respectively, the spin-order, the quadrupole-order and the octupole-order force terms); and the last two terms come from the torque contracted with the dynamical velocity (inside the square brackets we have, respectively, the quadrupole-order and the octupole-order torque terms). Thus, the tensor ϵμν\epsilon^{\mu\nu} is of order (l/r)2(l/r)^{2} or higher.

Contracting u[kvl]u^{[k}v^{l]} with uku_{k}, and doing some approximations, we get

vkuk=12ϵklϵrlukur+𝒪(ϵ4),v^{k}u_{k}=1-2\epsilon^{kl}\epsilon_{rl}u_{k}u^{r}+\mathcal{O}(\epsilon^{4})\,\,, (11)

where 𝒪(ϵ4)\mathcal{O}(\epsilon^{4}) represent terms that contain at least four factors of ϵkl\epsilon^{kl}. Since we are only keeping terms up to order (l/r)3(l/r)^{3}, we can even neglect terms with two factors of ϵkl\epsilon^{kl}, i.e., 𝒪(ϵ2)\mathcal{O}(\epsilon^{2}), obtaining

vkuk+2ϵlkul=uk+1M2SklFl+1MQklul,v^{k}\approx u^{k}+2\epsilon^{lk}u_{l}=u^{k}+\frac{1}{M^{2}}S^{kl}F_{l}+\frac{1}{M}Q^{kl}u_{l}\,, (12)

which is correct up to order (l/r)3(l/r)^{3}. It is interesting to note that if we evaluate u[kvl]u^{[k}v^{l]} from this approximate formula for vlv^{l}, we obtain ϵkl\epsilon^{kl} exactly.

If the body were point-like and free, its worldline would be a geodesic of the spacetime and thus satisfy the geodesic equation vrrvk=0v^{r}\nabla_{r}v^{k}=0. However, since the body is not point-like, its worldline \ell should satisfy a different equation, namely

vrrvk=vrruk+vrr[1M2SklFl+1MQklul].v^{r}\nabla_{r}v^{k}=v^{r}\nabla_{r}u^{k}+v^{r}\nabla_{r}\left[\frac{1}{M^{2}}S^{kl}F_{l}+\frac{1}{M}Q^{kl}u_{l}\right]\,\,. (13)

From equation (1) we obtain

vllvk\displaystyle v^{l}\nabla_{l}v^{k} =1M(FkFluluk)3M3FrurSklFl\displaystyle=\frac{1}{M}(F^{k}-F^{l}u_{l}u^{k})-\frac{3}{M^{3}}F^{r}u_{r}S^{kl}F_{l}
+1M2Qkl(2Fl3Frurul)+1M2ukQlrFlur\displaystyle+\frac{1}{M^{2}}Q^{kl}(2F_{l}-3F_{r}u^{r}u_{l})+\frac{1}{M^{2}}u^{k}Q^{lr}F_{l}u_{r}
+1M2SklvrrFl+1MvrrQklul,\displaystyle+\frac{1}{M^{2}}S^{kl}v^{r}\nabla_{r}F_{l}+\frac{1}{M}v^{r}\nabla_{r}Q^{kl}u_{l}\,, (14)

which is the desired equation for the worldline of the body. Note that the terms involving two factors of force or a product of a force and a torque implicitly contain two factors of curvature, which contributes with the small factor of order k2/r4k^{2}/r^{4}. Since we are dropping terms of this order, this equation can be simplified to

vllvk=1M(FkFluluk)+1M2SklvrrFl+1MvrrQklulv^{l}\nabla_{l}v^{k}=\frac{1}{M}(F^{k}-F^{l}u_{l}u^{k})+\frac{1}{M^{2}}S^{kl}v^{r}\nabla_{r}F_{l}+\frac{1}{M}v^{r}\nabla_{r}Q^{kl}u_{l} (15)

which contains a force term, a term with the spin and the time derivative of the force, and a term involving the time derivative of the torque. The order of magnitude of the last two terms, containing the time derivatives of the force and the torque, cannot be precisely evaluated yet as we still do not have a prescription to explicitly calculate the multipole moments. Therefore, before going further with the analysis of this equation, let us discuss a proposal to specify the multipole moments.

IV Prescription of the multipole moments

This section is devoted to introduce the core element of the covariant theory of swimming: the special manner to prescribe the multipole moments of the body. Considering equation (15) for the worldline of the center of mass of the body, we can see that all the information concerning its geometric shape and internal motions have to be fully encoded in the multipole moments.

The most straightforward manner to prescribe the multipole moments, as noted by Dixon in his papers, is to translate the internal differential equations governing the substances compounding the body into equations relating the multipole moments. For instance, Dixon discusses a reasonable definition for a rigid body in general relativity, where the components of the moments are assumed to be constant in a tetrad rotating in a way determined by the total spin of the body. In the general case, one could use the correspondence between the stress-energy tensor and the set of multipole moments, mentioned in item (iii)(\text{iii}) of section (II), to translate the dynamical differential equations from one description to the other. Since we are going to this correspondence in this section, let us write it explicitly below. The (higher) multipole moments are given, for n0n\geq 0, by

Jk1knlpqr=Υk1kn[l[qp]r]+1n+1Πk1kn[l[qp]r]tvt,J^{k_{1}\cdots k_{n}lpqr}=\Upsilon^{k_{1}\cdots k_{n}[l[qp]r]}+\frac{1}{n+1}\Pi^{k_{1}\cdots k_{n}[l[qp]r]t}v_{t}\,, (16)

where Υ\Upsilon and Π\Pi are respectively given, for n2n\geq 2, by

Υk1knlq(s)=(1)nΣ(s)σk1σknσalσbqTabwc𝑑Σc\displaystyle\Upsilon^{k_{1}\cdots k_{n}lq}(s)=(-1)^{n}\!\int_{\Sigma(s)}\sigma^{k_{1}}\ldots\sigma^{k_{n}}{\sigma}^{l}_{\>\>{a}}{\sigma}^{q}_{\>\>{b}}T^{ab}w^{c}d\Sigma_{c}
Πk1knlqr(s)=2(1)nΣ(s)σk1σknnprlqHapTab𝑑Σb\displaystyle\Pi^{k_{1}\cdots k_{n}lqr}(s)=2(-1)^{n}\!\int_{\Sigma(s)}\sigma^{k_{1}}\ldots\sigma^{k_{n}}{\langle n\rangle}^{prlq}H_{ap}T^{ab}d\Sigma_{b} (17)

where Υk1knlq(s)=Υk1knlq(z(s))\Upsilon^{k_{1}\cdots k_{n}lq}(s)=\Upsilon^{k_{1}\cdots k_{n}lq}(z(s)) and Πk1knlqr(s)\Pi^{k_{1}\cdots k_{n}lqr}(s) are tensors at z(s)z(s). The integral run over xΣ(s)x\in\Sigma(s), with dΣcd\Sigma^{c} being the vector-valued element of volume induced on Σ(s)\Sigma(s). The stress-energy tensor TabT^{ab} is based at xx, and waw^{a} is any vector field such that waδsw^{a}\delta s carries Σ(s)\Sigma(s) to Σ(s+δs)\Sigma(s+\delta s). The quantities σk\sigma^{k}, σal{\sigma}^{l}_{\>\>{a}}, σbq{\sigma}^{q}_{\>\>{b}}, HapH_{ap} and nprlq{\langle n\rangle}^{prlq} are bitensors, based at (x,z(s))(x,z(s)), defined in Appendix A. Note that we use a similar index convention for bitensors as for tensors, with the difference that for bitensors based at (x,z)(x,z) the first half of the alphabet (“a,b,c,a,b,c,\ldots” or “α,β,γ,\alpha,\beta,\gamma,\ldots”) is reserved for indices at xx (first entry) and the second half of the alphabet (“k,l,m,k,l,m,\ldots” or “κ,λ,μ,\kappa,\lambda,\mu,\ldots”) for indices at zz (second entry).

So, how exactly one should proceed to specify these moments in a particular situation? In the case of the swimming bipod, for example, we could consider that a certain physical device is responsible for producing the desired internal movements. A simple realization is to consider a set of springs interconnecting the legs, so as to make them open or close as the springs oscillate; also, the legs could be considered to be springs themselves, so that they would be stretching and shrinking in an oscillatory fashion. We would then have to provide a local, covariant description for these springs. However, even the simplest model (of locally Hookean springs777By locally Hookean spring we mean a relativistic model for a spring that obeys the Hooke’s law in the local instantaneous rest frame of any infinitesimal piece of it. See [10].) is described by highly involved differential equations. Moreover, an important drawback of this type of approach is that the internal motions are not easily controllable. That is, it would be a considerable effort to find which structural modifications (such as changing the position or stiffness of the springs) would have to be implemented, at each time, in order to make the body perform the desired sequence of internal motions.

Therefore, we shall look for an alternative approach to the prescription of the multipole moments, particularly one that allows us to directly provide the sequence of internal motions rather than the underlying mechanisms necessary to produce them. As proposed by Wisdom, we might assume that there is some ingeniously-designed device, continuously coupled along the whole extension of the body, that applies the exact amount of tension to each of its points such as to produce a formerly specified sequence of internal motions. One way to conceive this is to imagine that the problem is first solved theoretically to determine the required internal forces that should be applied to each point, then all the information is stored in the memory of microcomputers spread all over the body, which will tell the local mechanisms of the device how to act in order to apply the correct tensions once the body is actually released to swim. We shall not be too concerned with the engineering details of the swimming mechanisms, as we are only interested in the resulting dynamics of a body going through a sequence of shape changes.

The basis of the covariant theory of swimming is the realization that, if only (l/r)3(l/r)^{3} order of precision is required in the equations of motion, then the quadrupole and octupole moments can be evaluated as if the neighboring spacetime were flat. This follows from the fact that curvature only affects the quadrupole and octupole moments at order (l/r)2(l/r)^{2}. Thus, since the lowest-order contribution to the equations of motion from these moments (that would come from their evaluation in a flat background) are already at order (l/r)2(l/r)^{2}, then curvature would only produce terms of order (l/r)4(l/r)^{4} or higher in the equations of motion. Basically, we will treat the equations of motions perturbatively in the curvature, so that we first study the internal motions as if the spacetime were flat, computing the multipole moments accordingly, and then insert them into the Dixon’s equations of motion. It should be stressed that this procedure plays a role only in the translation from the description of the internal states of a body in terms of an energy-momentum tensor to the corresponding description in terms of the multipole moments.

Before going through the justification of this claim, let us consider the case where the spacetime is actually flat. In Minkowski spacetime the equations of motion reduce to

δpkds=0\displaystyle\frac{\delta p^{k}}{ds}=0
δSklds=0\displaystyle\frac{\delta S^{kl}}{ds}=0
u[kvl]=0,\displaystyle u^{[k}v^{l]}=0\,, (18)

which imply that the mass, the dynamical velocity and the spin are conserved. Also, since vk=ukv^{k}=u^{k}, the worldline of the center of mass \ell is a straight line (i.e., geodesic). The internal motions have only to satisfy the conservation equation

aTab=0,\partial_{a}T^{ab}=0\,, (19)

where a\partial_{a} is the derivative operator associated with any inertial (Cartesian) coordinate system.

In inertial (Cartesian) coordinates, the bitensors appearing in (17) become

σκ=(xκzκ)\displaystyle\sigma^{\kappa}=-(x^{\kappa}-z^{\kappa})
σα=(zαxα)\displaystyle\sigma^{\alpha}=-(z^{\alpha}-x^{\alpha}) (20)
σβα=δβα\displaystyle{\sigma}^{\alpha}_{\>\>{\beta}}={\delta}^{\alpha}_{\>\>{\beta}}
σλα=δλα\displaystyle{\sigma}^{\alpha}_{\>\>{\lambda}}=-{\delta}^{\alpha}_{\>\>{\lambda}}
σλκ=δλκ\displaystyle{\sigma}^{\kappa}_{\>\>{\lambda}}={\delta}^{\kappa}_{\>\>{\lambda}}
σβκ=δβκ\displaystyle{\sigma}^{\kappa}_{\>\>{\beta}}=-{\delta}^{\kappa}_{\>\>{\beta}} (21)
Hκα=δκα\displaystyle{H}^{\alpha}_{\>\>{\kappa}}={\delta}^{\alpha}_{\>\>{\kappa}}
Kκα=δκα\displaystyle{K}^{\alpha}_{\>\>{\kappa}}={\delta}^{\alpha}_{\>\>{\kappa}} (22)
nμνκλ=ημ(κηλ)νfor n ≥2,\langle n\rangle^{\mu\nu\kappa\lambda}=\eta^{\mu(\kappa}\eta^{\lambda)\nu}\quad\text{for {\hbox{n \geq 2}}}\,, (23)

where δ\delta is the Kronecker delta and η\eta is the flat (Minkowski) metric.

Since the worldline of the center of mass is straight, the spatial hypersurfaces Σ(s)\Sigma(s) are all parallel to each other, so that we may simply take the ww-vector as

wμ=uμ=vμ.w^{\mu}=u^{\mu}=v^{\mu}\,\,. (24)

Consider the particular inertial frame that is at rest with respect to the center of mass. Further, let it be located at the spatial origin of the coordinates. Thus, the worldline of the center of mass \ell is described by the vertical line

zμ(s)=(s,0,0,0),z^{\mu}(s)=(s,0,0,0)\,\,, (25)

where ss is the proper time along it. The momentum-energy tensor can be decomposed888In this section, the indices i,j,k,l,mi,j,k,l,m may be used to indicate spatial indices of tensor components, so that they assume values from 11 to 33. Hopefully, it will be sufficiently clear where they are being used with this meaning and where the standard convention (abstract indices) is being used. as

Tμν=(ρg1g2g3g1h11h12h13g2h12h22h23g3h31h32h33),T^{\mu\nu}=\left(\begin{array}[]{cccc}\rho&g^{1}&g^{2}&g^{3}\\ g^{1}&h^{11}&h^{12}&h^{13}\\ g^{2}&h^{12}&h^{22}&h^{23}\\ g^{3}&h^{31}&h^{32}&h^{33}\end{array}\right)\,\,, (26)

where ρ\rho can be interpreted as the mass density of the matter distribution, gi=(g1,g2,g3)g^{i}=(g^{1},g^{2},g^{3}) can be seen as the density of linear 3-momentum, and the 3-tensor hijh^{ij} is called the total stress 3-tensor. If uiu^{i} is the 3-velocity of a piece of matter at a given point, then the total stress 3-tensor hijh^{ij} is related with the stress 3-tensor τij\tau^{ij} by hij=τij+giujh^{ij}=\tau^{ij}+g^{i}u^{j}, and τij\tau^{ij} is such that, for an infinitesimal oriented surface of area δA\delta A and a direction defined by the unit 3-vector n^=(n1,n2,n3)\hat{n}=(n^{1},n^{2},n^{3}), then the 3-vector τijnjδA\tau^{ij}n_{j}\delta A yields the force that the matter on the negative side of the surface applies to the matter on the positive side of it.

Inserting all these results in the definitions of Υ\Upsilon and Π\Pi, we get the following expressions for the quadrupole and octupole moments in flat spacetime

Jμνρσ=Σ𝑑Σr[μr[ρ(Tν]σ]+vν]Tσ]0+vσ]Tν]0)\displaystyle J^{\mu\nu\rho\sigma}=\int_{\Sigma}d\Sigma\,r^{[\mu}r^{[\rho}\left(T^{\nu]\sigma]}+v^{\nu]}T^{\sigma]0}+v^{\sigma]}T^{\nu]0}\right)
Jλμνρσ=Σ𝑑Σrλr[μr[ρ(Tν]σ]+12vν]Tσ]0+12vσ]Tν]0).\displaystyle J^{\lambda\mu\nu\rho\sigma}=\int_{\Sigma}d\Sigma\,r^{\lambda}r^{[\mu}r^{[\rho}\left(T^{\nu]\sigma]}+\frac{1}{2}v^{\nu]}T^{\sigma]0}+\frac{1}{2}v^{\sigma]}T^{\nu]0}\right)\,\,. (27)

where rμ=xμzμ(s)=(0,x1,x2,x3)r^{\mu}=x^{\mu}-z^{\mu}(s)=(0,x^{1},x^{2},x^{3}), with xΣ(s)x\in\Sigma(s), are the “position vectors” across the spatial surfaces Σ(s)\Sigma(s). It will be useful to write down explicitly, in terms of ρ\rho, gig^{i} and hijh^{ij}, each non-vanishing component of these moments. For the quadrupole,

J0i0j=34d3xxixjρ\displaystyle J^{0i0j}=\frac{3}{4}\int d^{3}\!x\,x^{i}x^{j}\rho
J0ijk=14d3x(3xixkgj2xixjgkxjxkgi)\displaystyle J^{0ijk}=\frac{1}{4}\int d^{3}\!x\left(3x^{i}x^{k}g^{j}-2x^{i}x^{j}g^{k}-x^{j}x^{k}g^{i}\right)
Jijkl=d3xx[ix[khj]l],\displaystyle J^{ijkl}=\int d^{3}\!x\,x^{[i}x^{[k}h^{j]l]}\,, (28)

and for the octupole,

Jm0i0j=12d3xxmxixjρ\displaystyle J^{m0i0j}=\frac{1}{2}\int d^{3}\!x\,x^{m}x^{i}x^{j}\rho
Jm0ijk=18d3xxm(4xixkgj3xixjgkxjxkgi)\displaystyle J^{m0ijk}=\frac{1}{8}\int d^{3}\!x\,x^{m}\left(4x^{i}x^{k}g^{j}-3x^{i}x^{j}g^{k}-x^{j}x^{k}g^{i}\right)
Jmijkl=d3xxmx[ix[khj]l],\displaystyle J^{mijkl}=\int d^{3}\!x\,x^{m}x^{[i}x^{[k}h^{j]l]}\,, (29)

where the functions ρ\rho, gig^{i} and hijh^{ij} are all evaluated at the point (s,x1,x2,x3)(s,x^{1},x^{2},x^{3}). This makes clear how the different components of the moments depend on the dynamical quantities associated with the matter distribution: when there are two time-indices (i.e., μ=0\mu=0) among the last four indices, then the component depends on the mass distribution of the matter; when there is only one time-index among the last four, then the component depends on the linear 3-momentum distribution; and when there is no time-index among the last four, then the component depends on the total stress distribution (which includes the 3-velocity, linear 3-momentum and the stress 3-tensor distributions).

Now we show that the “correct” expressions for the multipole moments, when curvature effects are included, differ from the flat spacetime expressions (27) only at order (l/r)2(l/r)^{2}. First let us argue that, in the suitable sense, the geometry of a bounded object999An object here may be seen as a subset of the space defined by a set of intrinsic geometric rules. of size (length) ll is affected by curvature only at order Rl2×lk(l/r)2×lRl^{2}\times l\sim k(l/r)^{2}\times l, if Rl21Rl^{2}\ll 1. It is clear that we need to define precisely what it means to “turn on the curvature” in the neighborhood of the object. Although there is no absolute way to define this operation, the exponential map provides a very convenient manner to relate the curved with the flat space – we just let the tangent space (say, at the center of mass) be the “flattened” projection of the real (curved) space.

Assume that there is some point p𝒱p\in\mathcal{V}\subset\mathcal{M}, where 𝒱\mathcal{V} is any open normal neighborhood in \mathcal{M} which contains the convex envelope of the geometrical object. Since 𝒱\mathcal{V} is a normal neighborhood of pp, we have that, for each point x𝒱x\in\mathcal{V}, there is a unique vector vv in TpT_{p}\mathcal{M} such that exppv=x\exp_{p}v=x. In other words, there is a diffeomorphism of a neighborhood of TpT_{p}\mathcal{M} (containing the 0 vector) onto 𝒱\mathcal{V} given by the exponential map at pp. Also, there is a natural coordinate system on 𝒱\mathcal{V} induced from a Cartesian coordinate system on TpT_{p}\mathcal{M}: if vv is decomposed in an orthonormal vector basis {eμ}\{e_{\mu}\} at pp as v=vμeμv=v^{\mu}e_{\mu}, then the coordinates vμv^{\mu} can be assigned to the point xx, i.e., we define xμ=vμx^{\mu}=v^{\mu} (called Riemann normal coordinates). On top of being a geometrically natural coordinate system, it has the interesting property that the coefficients of the Taylor expansion of the metric gμν(x)g_{\mu\nu}(x), with respect to xμx^{\mu} about 0, are determined by the curvature and its (covariant) derivatives. For instance, up to second order we have

gμν(x)=δμν+13Rμρνσ(0)xρxσ+,g_{\mu\nu}(x)=\delta_{\mu\nu}+\frac{1}{3}R_{\mu\rho\nu\sigma}(0)x^{\rho}x^{\sigma}+\ldots\,, (30)

which suggests a very natural way to implement the “turning on” of the curvature. We simply define a one-parameter family of metrics gλg_{\lambda} by

[gλ]μν(x)=δμν+13λRμρνσ(0)xρxσ+,[g_{\lambda}]_{\mu\nu}(x)=\delta_{\mu\nu}+\frac{1}{3}\lambda R_{\mu\rho\nu\sigma}(0)x^{\rho}x^{\sigma}+\ldots\,\,, (31)

so that λ=0\lambda=0 corresponds to a flat metric and λ=1\lambda=1 corresponds to the actual metric of the space.

Let us now try to convince ourselves that curvature affects the basic laws of flat geometry at order Rl2Rl^{2} only (note that this is not intended to be a rigorous argument, in part because the proposition itself lacks the necessary precision). We shall consider a few basic geometrical constructions and try to identify the effect of curvature. For instance, note that the trajectory of a geodesic starting at x0μx_{0}^{\mu} with (unit) tangent vector vμv^{\mu} is described by

xμ(s)=x0μ+vμs13s2Rνρσμx0νvρvσ+,x^{\mu}(s)=x_{0}^{\mu}+v^{\mu}s-\frac{1}{3}s^{2}{R_{\nu\rho\sigma}}^{\mu}x_{0}^{\nu}v^{\rho}v^{\sigma}+\ldots\,\,, (32)

where ss is the proper length along it and the curvature tensor is evaluated at the origin, pp. If a geodesic defined by the same prescription lived in a flat space, then its trajectory x~μ(s)\widetilde{x}^{\mu}(s) would be given by

x~μ(s)=x0μ+vμs,\widetilde{x}^{\mu}(s)=x_{0}^{\mu}+v^{\mu}s\,\,, (33)

which indicates that they deviate from each other by

Δxμ(s)=xμ(s)x~μ(s)=13s2Rνρσμx0νvρvσ+,\Delta x^{\mu}(s)=x^{\mu}(s)-\widetilde{x}^{\mu}(s)=-\frac{1}{3}s^{2}{R_{\nu\rho\sigma}}^{\mu}x_{0}^{\nu}v^{\rho}v^{\sigma}+\ldots\,\,, (34)

We then see that the effect of the curvature Rk/r2R\sim k/r^{2}, in a geodesic of length ll within 𝒱\mathcal{V} (so that x0μlx_{0}^{\mu}\lesssim l), is of order l×k(l/r)2l\times k(l/r)^{2}, supporting the claim.

Refer to caption
Figure 1: Pictorial representation of an equilateral triangle being deformed as the curvature is turned on - the dotted-line triangle corresponds to the flat-space object and the full-line triangle corresponds to the curved-space object. The triangle has edges of length ll and the deformation takes place at the scale l×Rl2l\times Rl^{2}.

For a few more examples [12], consider the (i)(i) geodesic distance L(x1,x2)L(x_{1},x_{2}) between two points x1μx_{1}^{\mu} and x2μx_{2}^{\mu} in 𝒱\mathcal{V},

L(x1,x2)2=δμνΔx12μΔx12ν+13Rμρνσx1ρx1σx2μx2ν+,L(x_{1},x_{2})^{2}=\delta_{\mu\nu}\Delta x_{12}^{\mu}\Delta x_{12}^{\nu}+\frac{1}{3}R_{\mu\rho\nu\sigma}x_{1}^{\rho}x_{1}^{\sigma}x_{2}^{\mu}x_{2}^{\nu}+\ldots\,\,, (35)

where Δxijμ=xjμxiμ\Delta x_{ij}^{\mu}=x_{j}^{\mu}-x_{i}^{\mu}; the (ii)(ii) cosine law for a geodesic triangle with vertices x0x_{0}, x1x_{1} and x2x_{2},

2L01L02cosθ0\displaystyle 2L_{01}L_{02}\cos\!\theta_{0} =L012+L022L122\displaystyle=L_{01}^{2}+L_{02}^{2}-L_{12}^{2}
+13RμρνσΔx01ρΔx01σΔx02μΔx02ν+,\displaystyle+\frac{1}{3}R_{\mu\rho\nu\sigma}\Delta x_{01}^{\rho}\Delta x_{01}^{\sigma}\Delta x_{02}^{\mu}\Delta x_{02}^{\nu}+\ldots\,, (36)

where θ0\theta_{0} is the angle subtended at the vertex x0x_{0} and Lij=L(xi,xj)L_{ij}=L(x_{i},x_{j}); and the (iii)(iii) parallel transport of a vector v0μv_{0}^{\mu} at x0x_{0}, along an arbitrary curve xμ(s)x^{\mu}(s) starting at x0x_{0},

vμ(s)=v0μ+13(Rνρσμ+Rνσρμ)v0ρa1σ(sa0ν+12s2a1ν)+v^{\mu}(s)=v_{0}^{\mu}+\frac{1}{3}\left({R_{\nu\rho\sigma}}^{\mu}+{R_{\nu\sigma\rho}}^{\mu}\right)v_{0}^{\rho}a_{1}^{\sigma}\left(sa_{0}^{\nu}+\frac{1}{2}s^{2}a_{1}^{\nu}\right)+\ldots (37)

where a0μ=x0μa_{0}^{\mu}=x_{0}^{\mu} and a1μ=dxμ/ds|s=0a_{1}^{\mu}=\left.dx^{\mu}/ds\right|_{s=0}. This suggests that the geometry of the body can be well-represented in a flat space, given our approximations.

Let us now examine other aspects of this approximation. Observe that the fundamental conservation equation, aTab=0\nabla_{a}T^{ab}=0, is only affected by the curvature at an order assumed to be negligible in Dixon’s equations

0\displaystyle 0 =μTμν(x)=μTμν+ΓμσμTσν+ΓμσνTμσ\displaystyle=\nabla_{\mu}T^{\mu\nu}(x)=\partial_{\mu}T^{\mu\nu}+\Gamma^{\mu}_{\,\,\mu\sigma}T^{\sigma\nu}+\Gamma^{\nu}_{\,\,\mu\sigma}T^{\mu\sigma}\approx
μTμν13RρσxρTσν+23RρμσνxρTμσ,\displaystyle\approx\partial_{\mu}T^{\mu\nu}-\frac{1}{3}R_{\rho\sigma}x^{\rho}T^{\sigma\nu}+\frac{2}{3}{R_{\rho\mu\sigma}}^{\nu}x^{\rho}T^{\mu\sigma}\,, (38)

which implies that μTμν0\partial_{\mu}T^{\mu\nu}\approx 0 as the terms containing the curvature can be dropped (for they contain a factor of order k/r2k/r^{2}). Hence, the dynamics of the internal motions of the body can be described by special relativity.

The first curvature-related corrections to the flat spacetime expressions for the bitensors in (17), if we particularize to z=pz=p, are

σα=xα\displaystyle\sigma^{\alpha}=x^{\alpha}
σκ=xκ,\displaystyle\sigma^{\kappa}=-x^{\kappa}\,\,, (39)
σκα=δκα+13Rβκγαxβxγ\displaystyle{\sigma}^{\alpha}_{\>\>{\kappa}}=-\delta^{\alpha}_{\,\,\kappa}+\frac{1}{3}{R_{\beta\kappa\gamma}}^{\alpha}x^{\beta}x^{\gamma}
σβα=δβα+13Rϵβγαxϵxγ\displaystyle{\sigma}^{\alpha}_{\>\>{\beta}}=\delta^{\alpha}_{\,\,\beta}+\frac{1}{3}{R_{\epsilon\beta\gamma}}^{\alpha}x^{\epsilon}x^{\gamma}
σακ=δακ\displaystyle{\sigma}^{\kappa}_{\>\>{\alpha}}=-\delta^{\kappa}_{\,\,\alpha}
σλκ=δλκ+13Rαλβκxαxβ,\displaystyle{\sigma}^{\kappa}_{\>\>{\lambda}}=\delta^{\kappa}_{\,\,\lambda}+\frac{1}{3}{R_{\alpha\lambda\beta}}^{\kappa}x^{\alpha}x^{\beta}\,\,, (40)
Hκα=δκα\displaystyle{H}^{\alpha}_{\>\>{\kappa}}=\delta^{\alpha}_{\,\,\kappa}
Kκα=δκα+13Rβκγαxβxγ,\displaystyle{K}^{\alpha}_{\>\>{\kappa}}=\delta^{\alpha}_{\,\,\kappa}+\frac{1}{3}{R_{\beta\kappa\gamma}}^{\alpha}x^{\beta}x^{\gamma}\,, (41)

showing that all these quantities only deviate from their flat spacetime counterparts, (20-22), at order k(l/r)2k(l/r)^{2}. Other geometrical quantities in (17) are the nn-potential function n\langle n\rangle, the volume element and the ww-vector. The first two can be similarly shown to deviate from their flat spacetime counterparts by terms proportional to the curvature tensor.

The ww-vector, on the other hand, deserves a special consideration for it is not exclusively related to the curvature. For a straight line \ell in the flat spacetime, it is clear that waw^{a} can be chosen as the covariant extension of uk=vku^{k}=v^{k}, that is, the vector field that coincides with uku^{k} on \ell and satisfies awb=0\nabla_{a}w^{b}=0 everywhere. If \ell is taken as a general (non-straight) line, so that the orthogonal hypersurfaces Σ(s)\Sigma(s) are not parallel to each other, then waw^{a} will not be a constant vector field. Let zμ(s)z^{\mu}(s) be the trajectory of \ell in the Riemann normal coordinate system centered at zμ(0)=0z^{\mu}(0)=0, with temporal vector (e0)μ=vμ(0)(e_{0})^{\mu}=v^{\mu}(0). Using the Taylor expansion in the proper length ss, we can write

zμ(s)=(e0)μs+12aμ(0)s2,z^{\mu}(s)=(e_{0})^{\mu}s+\frac{1}{2}a^{\mu}(0)s^{2}\,\,, (42)

where aμa^{\mu} denotes the covariant acceleration vννvμv^{\nu}\nabla_{\nu}v^{\mu}. The hypersurface Σδs\Sigma_{\delta s} at the infinitesimal proper time δs\delta s is formed by all geodesics starting at zμ(δs)z^{\mu}(\delta s) orthogonal to uμ(δs)u^{\mu}(\delta s). Since uμu^{\mu} only differs from vμv^{\mu} by terms involving the curvature tensor, we may consider that Σδs\Sigma_{\delta s} is actually formed by geodesics orthogonal to vμ(δs)v^{\mu}(\delta s) instead. Let qμ(s)q^{\mu}(s) be a general vector at zμ(s)z^{\mu}(s) orthogonal to vμ(s)v^{\mu}(s), so that the set of all qμ(s)q^{\mu}(s) form a 3-dimensional vector subspace at zμ(s)z^{\mu}(s). At δs\delta s, we have 0=qμ(δs)vμ(δs)ημνqμvν=q0qaδs0=q^{\mu}(\delta s)v_{\mu}(\delta s)\approx\eta_{\mu\nu}q^{\mu}v_{\nu}=q^{0}-\vec{q}\cdot\vec{a}\delta s, where the curvature term in gμν(z(δs))ημνg_{\mu\nu}(z(\delta s))\approx\eta_{\mu\nu} was dropped. In this equation we are writing qμ=(q0,q)q^{\mu}=(q^{0},\vec{q}), in which q=(q1,q2,q3)\vec{q}=(q^{1},q^{2},q^{3}), and aμ=aμ(0)=(0,a)a^{\mu}=a^{\mu}(0)=(0,\vec{a}). The dot denotes the trivial Euclidean scalar product, qa=i=13qiai\vec{q}\cdot\vec{a}=\sum_{i=1}^{3}q^{i}a^{i}. Thus, we have qμ(δs)=(qaδs,q)q^{\mu}(\delta s)=(\vec{q}\cdot\vec{a}\delta s,\vec{q}). A geodesic starting at zμ(δs)z^{\mu}(\delta s) and tangent to qμ(δs)q^{\mu}(\delta s), denoted by xμ(λ)x^{\mu}(\lambda), where λ\lambda is any affine parameter, can be given approximately by

xμ(λ)=zμ(δs)+qμ(δs)λ13λ2Rνρσμzν(δs)qρ(δs)qσ(δs),x^{\mu}(\lambda)=z^{\mu}(\delta s)+q^{\mu}(\delta s)\lambda-\frac{1}{3}\lambda^{2}{R_{\nu\rho\sigma}}^{\mu}z^{\nu}(\delta s)q^{\rho}(\delta s)q^{\sigma}(\delta s)\,, (43)

in which the last term, containing the curvature tensor, can be dropped. The ww-vector can then be chosen at Σ0\Sigma_{0} as the vector field satisfying

wα(0,x(λ))δs=(e0)μxμ(λ)(e0)αw^{\alpha}(0,\vec{x}(\lambda))\delta s=(e_{0})_{\mu}x^{\mu}(\lambda)(e_{0})^{\alpha} (44)

for infinitesimal δs\delta s. The xμ(λ)x^{\mu}(\lambda) in the right-hand side of the equation has to be evaluated up to first order in δs\delta s, yielding (e0)μxμ(λ)(e0)α=(δs+λqaδs)(e0)α(e_{0})_{\mu}x^{\mu}(\lambda)(e_{0})^{\alpha}=(\delta s+\lambda\vec{q}\cdot\vec{a}\delta s)(e_{0})^{\alpha}; while the argument of wαw^{\alpha} in the left-hand side can be evaluated to the lowest order, x(λ)=λq\vec{x}(\lambda)=\lambda\vec{q}. It then follows that

wα(0,x)=(1+xa)(e0)α,w^{\alpha}(0,\vec{x})=(1+\vec{x}\cdot\vec{a})(e_{0})^{\alpha}\,\,, (45)

showing that, if \ell could be prescribed arbitrarily, then the ww-factor would not necessarily be a constant vector field, even if the spacetime were flat. In our case, nonetheless, \ell is the worldline of the center of mass and its 4-acceleration is given by equation (15), where it is clear that a\vec{a} necessarily contains a curvature factor (from the gravitational force and torque), that implies that its order of magnitude has a factor k/r2\sim k/r^{2}. Therefore the ww-factor can be approximated as

wα(e0)α=δ  0α,w^{\alpha}\approx(e_{0})^{\alpha}=\delta^{\alpha}_{\,\,0}\,\,, (46)

as if they were evaluated in a exactly flat spacetime (where \ell would be exactly straight).

Now that we have justified our claim that the multipole moments can be evaluated as if the spacetime were flat (under our approximations), let us make one assumption on the kind of bodies that shall be considered – from now on, we shall focus on articulated bodies. More precisely, the assumption is that, at any instant, the internal configuration state of the body can be described by a finite number of continuous parameters. In other words, there should exist an nn-dimensional manifold \mathcal{I}, called the internal configuration space of the body, such that there is a one-to-one correspondence between any point qq\in\mathcal{I} and a particular internal configuration state of the body (i.e., a specification of the instantaneous geometry of the body). Note that \mathcal{I} is entirely external to the spacetime given that the instantaneous internal state of the body can be described in flat spacetime, which is a maximally symmetric space and contains no information about the structure of the real spacetime.

As an illustration, consider Wisdom’s tripod in a 3-dimensional Euclidean space [1]. It is defined by a “pivot” body, of proper mass m0m_{0}, connected to three other bodies, of equal proper masses m1m_{1}, by straight massless rods of length ll. These three rods form equal angles to each other, defining a symmetrical tripod. Define the direction of the symmetry axis by a unit vector n^1\hat{n}_{1}, specify the angle α\alpha that any of the rods makes with this axis, and let n^2\hat{n}_{2} (orthogonal to n^1\hat{n}_{1}) indicate the orientation of the tripod around the symmetry axis. The parameters that can be controlled are ll and α\alpha, while n^1\hat{n}_{1} and n^2\hat{n}_{2} are determined by the initial conditions and the dynamics of the tripod. Hence, if l>0l>0 and 0<α<π0<\alpha<\pi, the internal space \mathcal{I} of this body is diffeomorphic to +×(0,π)\mathbb{R}^{+}\times(0,\pi). If we specify the routine of internal motions by giving the functions l(s)l(s) and α(s)\alpha(s), then the energy-momentum tensor of the body is determined on each hypersurface Σs\Sigma_{s}, which can be used to evaluate the quadrupole and octupole moments at each “time” ss. In Newtonian mechanics, the mass distribution will depend solely on the instantaneous configuration state (l,α)(l,\alpha), the momentum distribution will also depend on the “velocities” (l˙,α˙)(\dot{l},\dot{\alpha}), where the dot denotes differentiation with respect to ss, and the tensions may also depend on the “accelerations” (l¨,α¨)(\ddot{l},\ddot{\alpha}); however, in relativity all these dynamical quantities will be mixed up in the momentum-energy tensor, so that we can only say that TμνT^{\mu\nu} will typically depend101010If physics were still Newtonian in the local rest frame of any infinitesimal portion of the matter, this conclusion would follow naturally. Otherwise, this proposed functional dependence for the energy-momentum tensor may be seen as an additional assumption of the theory. on (l,α,l˙,α˙,l¨,α¨)(l,\alpha,\dot{l},\dot{\alpha},\ddot{l},\ddot{\alpha}).

Generally, in order to fully characterize the prescribed routine of shapes assumed by the body, we may specify a corresponding curve in the configuration space, γ:\gamma:\mathbb{R}\rightarrow\mathcal{I}. In an arbitrary coordinate system {βμ}\{\beta^{\mu}\} for \mathcal{I}, the curve can be described by the functions βμ(s)\beta^{\mu}(s). The quadrupole and octupole moments are, supposedly, functions of the β\beta’s, β˙\dot{\beta}’s and β¨\ddot{\beta}’s,

J=J(β1,βn,β˙1,β˙n,β¨1,β¨n),J=J(\beta^{1},\ldots\beta^{n},\dot{\beta}^{1},\ldots\dot{\beta}^{n},\ddot{\beta}^{1},\ldots\ddot{\beta}^{n})\,\,, (47)

where JJ may represent either JμνρσJ^{\mu\nu\rho\sigma} or JμνρσλJ^{\mu\nu\rho\sigma\lambda}. For a given curve γ\gamma, the JJ’s become111111Note that we are slightly abusing of the notation by using the same name for objects of different functional dependence, i.e., the JJ’s as functions of the point z(s)z(s) (in the worldline \ell) or as functions of the configuration coordinates βi(s)\beta^{i}(s) and their derivatives. Since the arguments of the functions will be explicitly indicated, it should be clear which functional form is being used. implicit functions of the parameter ss,

J(z(s))=J(β1(s),β˙1(s),β¨1(s),),J(z(s))=J(\beta^{1}(s)\ldots,\dot{\beta}^{1}(s)\ldots,\ddot{\beta}^{1}(s),\ldots)\,, (48)

which can be used to solve for the worldline of the center of mass, \ell, according to the previous section. It is interesting to note that we may generalize this by considering that the JJ’s can also have an explicit parameter dependence

J(z(s))=J(β(s),β˙(s),β¨(s),s),J(z(s))=J(\beta(s),\dot{\beta}(s),\ddot{\beta}(s),s)\,\,, (49)

where β\beta denotes the full string of βi\beta^{i}, for i=1ni=1\ldots n, and analogously for β˙\dot{\beta} and β¨\ddot{\beta}. The introduction of this explicit ss dependence can be useful to deal with the case in which the correspondence between points in \mathcal{I} and the basic geometrical shape of the body change with time – for example, consider a body initially shaped like a three-dimensional ellipsoid, described by the lengths of its semi-axis β1\beta^{1}, β2\beta^{2} and β3\beta^{3}, eventually morphing into a three-dimensional parallelepiped, with these same three parameters now representing the length of its edges (i.e., the parameter ss can be viewed as the homotopy parameter of the continuous transformation that turns an ellipsoid into a rectangular parallelepiped).

Let us make a quick point on the parameterization of γ\gamma. In principle, we can take any physically meaningful parameter that can be diffeomorphically mapped to ss, such as the time τ\tau as measured by an observer moving instantaneously parallel to the dynamical velocity uku^{k}. Note that the difference, in the case of τ\tau, does not need to be accounted since it affects the equations of motion only at a negligible order of magnitude, as can be seen by the chain rule

δds=dτdsδdτ=ukvkδdτ(12ϵklϵrlukur+)δdτ,\frac{\delta}{ds}=\frac{d\tau}{ds}\frac{\delta}{d\tau}=u^{k}v_{k}\frac{\delta}{d\tau}\approx\left(1-2\epsilon^{kl}\epsilon_{rl}u_{k}u^{r}+\cdots\right)\frac{\delta}{d\tau}\,\,, (50)

where ϵkl\epsilon^{kl} is the right-hand side of equation (3). Since it includes the curvature tensor, in the evaluation of the JJ’s one can simply approximate

δdsδdτ,\frac{\delta}{ds}\approx\frac{\delta}{d\tau}\,\,, (51)

showing that it is meaningless to make a distinction between derivatives in ss or in τ\tau. Also, if the total time for which one is interested in evolving the dynamical quantities is not too large, these parameterizations are really indistinguishable,

τ(s)\displaystyle\tau(s) =0suk(s¯)vk(s¯)𝑑s¯=s20sϵklϵrlukur𝑑s¯+\displaystyle=\int_{0}^{s}u^{k}(\bar{s})v_{k}(\bar{s})d\bar{s}=s-2\int_{0}^{s}\epsilon^{kl}\epsilon_{rl}u_{k}u^{r}d\bar{s}+\cdots
s2s𝒪(ϵ2),\displaystyle\approx s-2s\mathcal{O}(\epsilon^{2})\,, (52)

that is, if s𝒪(ϵ2)s\mathcal{O}(\epsilon^{2}) is sufficiently small, then τ(s)s\tau(s)\approx s. Thus, for simplicity, we will use ss as the standard parameter of γ\gamma.

Finally, let us consider an important detail that has not been emphasized in the discussion above. By using this local mapping to a flat spacetime in a neighborhood of each point z(s)z(s), we have seen that it was possible to evaluate the energy-momentum tensor on Σs\Sigma_{s} which, consequently, determined the quadrupole and octupole moments at z(s)z(s). But how to connect this mapping for different points z(s)z(s)? That is, how the components JμνρσJ^{\mu\nu\rho\sigma} obtained in the Minkowski spacetime translate into a tensor JklqrJ^{klqr} in the true spacetime? Let us discuss a very natural proposal for establishing this connection. Take the initial position of the center of mass to be z(s0)z(s_{0}) and the initial dynamical velocity to be uk(s0)u^{k}(s_{0}). Let {(eμ)k}\{(e_{\mu})^{k}\} be any tetrad at z(s0)z(s_{0}) with (e0)k=uk(s0)(e_{0})^{k}=u^{k}(s_{0}). At this point, z(s0)z(s_{0}), consider the correspondence between curved and flat spacetimes induced by the inverse exponential mapping from some (open) neighborhood Σ~s0\widetilde{\Sigma}_{s_{0}} of Σs0W\Sigma_{s_{0}}\cap W to Tz(s0)T_{z(s_{0})}\mathcal{M}, and set global inertial (Cartesian) coordinates in Tz(s0)T_{z(s_{0})}\mathcal{M} such that each base vector /xμ\partial/\partial x^{\mu} is mapped to the corresponding (eμ)k(e_{\mu})^{k}. It is then natural to assume that, at this point, one should have

Jklqr(z(s0))=J¯μνρσ(z(s0))(eμ)k(eν)l(eρ)q(eσ)r,J^{klqr}(z(s_{0}))=\bar{J}^{\mu\nu\rho\sigma}(z(s_{0}))(e_{\mu})^{k}(e_{\nu})^{l}(e_{\rho})^{q}(e_{\sigma})^{r}\,, (53)

and analogously for Jklqrm(z(s0))J^{klqrm}(z(s_{0})). The bar is only to make clear that the J¯\bar{J}’s are evaluated in the flat spacetime. As the time passes, we can solve the dynamics of the body in the flat spacetime so as to get the value of its energy-momentum tensor at any future time t>s0t>s_{0}, which allows one to evaluate the multipole moments, J¯μνρσ(β(t),β˙(t),β¨(t),t)\bar{J}^{\mu\nu\rho\sigma}(\beta(t),\dot{\beta}(t),\ddot{\beta}(t),t) and J¯μνρσλ(β(t),β˙(t),β¨(t),t)\bar{J}^{\mu\nu\rho\sigma\lambda}(\beta(t),\dot{\beta}(t),\ddot{\beta}(t),t), in the inertial coordinate frame by using the information of the energy-momentum on the section (x0=t,x1,x2,x3)(x^{0}=t,x^{1},x^{2},x^{3}). The assumption here is that, as long as tt is not too large (see formula (52)), then expression (53) can be extended without modifications to an arbitrary point z(s)z(s) by evaluating the J¯\bar{J}’s at the time t=τ(s)st=\tau(s)\approx s (where τ(s0)=s0\tau(s_{0})=s_{0}),

Jklqr(z(s))=\displaystyle J^{klqr}(z(s))=
J¯μνρσ(β(τ(s)),β˙(τ(s)),β¨(τ(s)),τ(s))(e¯μ)k(e¯ν)l(e¯ρ)q(e¯σ)r\displaystyle\bar{J}^{\mu\nu\rho\sigma}(\beta(\tau(s)),\dot{\beta}(\tau(s)),\ddot{\beta}(\tau(s)),\tau(s))(\bar{e}_{\mu})^{k}(\bar{e}_{\nu})^{l}(\bar{e}_{\rho})^{q}(\bar{e}_{\sigma})^{r} (54)

and analogously for Jklqrm(z(s))J^{klqrm}(z(s)). The tetrad {(e¯μ)k}\{(\bar{e}_{\mu})^{k}\} at z(s)z(s) is obtained by any reasonable transportation of {(eμ)k}\{(e_{\mu})^{k}\} from z(s0)z(s_{0}) along \ell. For instance, we can consider the M-transport

δ(e¯μ)kds=(u˙kuluku˙l)(e¯μ)l,\frac{\delta(\bar{e}_{\mu})^{k}}{ds}=(\dot{u}^{k}u_{l}-u^{k}\dot{u}_{l})(\bar{e}_{\mu})^{l}\,\,, (55)

or the Fermi transport or the parallel transport, and they should all agree within our approximations. This completes the discussion on the basic idea behind the prescription of the quadrupole and octupole moments, which is going to be central to the study of the swimming phenomenon.

V Swimming deviation formula

In the this section we will consider the theory of swimming in spacetime, investigating the possibility that a body could move through the space in a way resembling the motion of swimmers in non-turbulent viscous fluids (i.e., low Reynolds number). More precisely, we must find a regime where the effective displacement of the center of mass does not depend on how fast the internal motions happen, but only on the geometric phase associated with the sequence of shapes assumed by the body. As we are going to see, this is not very typical of the motion of free bodies in general relativity, but can take place in particular scenarios and settings. Our formulas, nevertheless, are not limited to this special regime and can describe the dynamics of general small articulated bodies, free falling in curved spacetimes, performing internal motions, as long as the observation time is not too long.

Let the internal configuration space \mathcal{I} be nn-dimensional, covered by coordinates (β1,β2,βn)(\beta^{1},\beta^{2},\ldots\beta^{n}). Let there be dd dynamical variables (q1,q2,qd)(q^{1},q^{2},\ldots q^{d}) describing the external configuration of the body – they may correspond, for instance, to the three spatial coordinates (x1,x2,x3)(x^{1},x^{2},x^{3}) of the center of mass, the three spatial components of the dynamical velocity (u1,u2,u3)(u^{1},u^{2},u^{3}), the spatial components of the spin (S1,S2,S3)(S^{1},S^{2},S^{3}), the proper mass MM, or any other dynamical attribute associated with the body (these components being associated with a suitable coordinate system, such as a Riemann normal coordinate system around the initial position, aligned with the initial velocity of the center of mass). Let q0(t)q_{0}(t) be the solution for the dynamical quantities when the body is kept “rigid”, i.e., with fixed configuration coordinates β0\beta_{0}. Define the deviation of the dynamical quantity qiq^{i}, denoted by ζi\zeta^{i}, as the difference between the general solution and the rigid body solution,

ζi(t)=qi(t)q0i(t),\zeta^{i}(t)=q^{i}(t)-q_{0}^{i}(t)\,\,, (56)

which is supposed to measure how the internal motions affect the motion of the center of mass.

It is clear that when the body performs a certain sequence of internal motions, represented by a curve γ:[0,t]\gamma:[0,t]\rightarrow\mathcal{I} in the internal configuration space, the dynamical equations can be solved to determine ζ(t)\zeta(t). In other words, keeping all other parameters fixed (spacetime curvature, initial conditions, etc), the variations ζi\zeta^{i} are functions of the curve γ\gamma, ζi=fi(γ)\zeta^{i}=f^{i}(\gamma). The swimming at “low Reynolds number” happens when the deviations ζi\zeta^{i} display a geometric phase behavior, depending only on the static sequence of internal motions. That is, the deviations ζi\zeta^{i} become functions of the image of the curve only, ζi=fi(Im(γ))\zeta^{i}=f^{i}(\mathrm{Im}(\gamma)).

Inspired by the Newtonian version of this problem, we shall search for a regime in which the deviation variables satisfy a set of differential equations of the form

ζ˙i=λai(β,ζ)β˙a,\dot{\zeta}^{i}=\lambda^{i}_{\,\,a^{\prime}}(\beta,\zeta)\dot{\beta}^{a^{\prime}}\,\,, (57)

where the dot continues to denote derivatives in the time tt, there is an implicit summation over a=1na^{\prime}=1\ldots n, and β\beta and ζ\zeta are short notations for the strings of variables (β1,β2,βn)(\beta^{1},\beta^{2},\ldots\beta^{n}) and (ζ1,ζ2,ζd)(\zeta^{1},\zeta^{2},\ldots\zeta^{d}), respectively. In this section, we shall use a different convention for the indices, where Latin letters will correspond to components of the dynamical (qiq^{i}) or deviation (ζi\zeta^{i}) variables (assuming values from 11 to dd), and primed Latin letters will correspond to configuration coordinates βa\beta^{a^{\prime}} (assuming values from 11 to nn); the Einstein summation convention for repeated indices will be preserved.

Let us show that, in such regime, the so described geometric-phase behavior is observed. Note that the space of the deviations ζ\zeta can be identified with a manifold 𝒵\mathcal{Z} that is diffeomorphic to d\mathbb{R}^{d}, which in turn has a natural vector space structure. It is then convenient to define the state space 𝒫\mathcal{P} of the problem as the Cartesian product 𝒫=×𝒵\mathcal{P}=\mathcal{I}\times\mathcal{Z}, so that each point of it corresponds to a certain internal configuration and dynamical state of the body. Since 𝒵\mathcal{Z} has a vector space structure, 𝒫\mathcal{P} can be seen as the vector bundle (𝒫,,ρ1)(\mathcal{P},\mathcal{I},\rho_{1}), where \mathcal{I} is the base space and the projection of the bundle is the projection on the first component of the product ×𝒵\mathcal{I}\times\mathcal{Z}; this means that for each internal configuration state in \mathcal{I} there is a fiber attached to it representing the dynamical space of the body. We shall use the natural isomorphism Tp𝒫TβTζ𝒵T_{p}\mathcal{P}\sim T_{\beta}\mathcal{I}\oplus T_{\zeta}\mathcal{Z}, where p=(β,ζ)p=(\beta,\zeta) is a point in 𝒫\mathcal{P}, to represent vectors in the state space; in the component form it will generally read V=(v1,vn;η1,ηd)V=(v^{1},\ldots v^{n};\eta^{1},\ldots\eta^{d}), succinctly written as V=(va;ηi)V=(v^{a^{\prime}};\eta^{i}). For an arbitrary vector vv at TβT_{\beta}\mathcal{I}, define the lifted vector VV in T(β,ζ)𝒫T_{(\beta,\zeta)}\mathcal{P} as

V=(va;λai(β,ζ)va),V=(v^{a^{\prime}};\lambda^{i}_{\,\,{a^{\prime}}}(\beta,\zeta)v^{a^{\prime}})\,\,, (58)

which can be alternatively expressed as a linear map Λ:TβT(β,ζ)𝒫\Lambda:T_{\beta}\mathcal{I}\rightarrow T_{(\beta,\zeta)}\mathcal{P}

V=Λv.V=\Lambda v\,\,. (59)

Note that this map Λ\Lambda can be used to lift a vector field vv in \mathcal{I} to a vector field VV in 𝒫\mathcal{P}. Also, any curve γ:\gamma:\mathbb{R}\rightarrow\mathcal{I} can be lifted to a curve Γ:𝒫\Gamma:\mathbb{R}\rightarrow\mathcal{P} such that, given a point p𝒫p\in\mathcal{P} in the fiber at γ(0)\gamma(0), then Γ(t)\Gamma(t) is the unique curve in 𝒫\mathcal{P} satisfying Γ(0)=p\Gamma(0)=p and

Γ(t)=Λγ(t).\Gamma^{\prime}(t)=\Lambda\gamma^{\prime}(t)\,\,. (60)

Observe that if we write Γ(t)\Gamma(t) in the coordinates {β,ζ}\{\beta,\zeta\} as Γ(t)=(βa(t),ζi(t))\Gamma(t)=(\beta^{a^{\prime}}(t),\zeta^{i}(t)), then the equation above translates to the system of differential equations

β˙a(t)\displaystyle\dot{\beta}^{a^{\prime}}(t) =γ˙a(t)\displaystyle=\dot{\gamma}^{a^{\prime}}(t) (61)
ζ˙i(t)\displaystyle\dot{\zeta}^{i}(t) =λai(β(t),ζ(t))γ˙a(t)\displaystyle=\lambda^{i}_{\,\,{a^{\prime}}}(\beta(t),\zeta(t))\dot{\gamma}^{a^{\prime}}(t) (62)

that can be immediately identified as the differential equation (57). In other words, the curve Γ(t)\Gamma(t) represents the precise solution of equation (57) when the curve γ(t)\gamma(t) is traversed in the configuration space. Therefore, the projection of Γ(t)\Gamma(t) by ρ2\rho_{2} (i.e., the projection on 𝒵\mathcal{Z}) gives the exact values of the dynamical deviations ζi(t)\zeta^{i}(t), with initial conditions ζ(0)=ρ2(Γ(0))=ρ2(p)\zeta(0)=\rho_{2}(\Gamma(0))=\rho_{2}(p), when the internal motions run along γ(t)\gamma(t).

Note that the total deviation of the dynamical quantities is given by the difference of ρ2Γ\rho_{2}\circ\Gamma at the final point and the initial point Γ(0)=p\Gamma(0)=p [see figure (2)]. Therefore, it is a function of the image of Γ\Gamma between the initial and final points (thence independent of its parametrization). In addition, it is possible to show that the image of Γ\Gamma is a function of the image of γ\gamma, so that the net deviation of the dynamical quantities is a function of the image of γ\gamma. Hence, the dynamics of the problem is indeed of the geometrical phase kind. In order to show this result, consider a general reparameterization of γ\gamma given by γ¯=γχ\bar{\gamma}=\gamma\circ\chi, where χ:\chi:\mathbb{R}\rightarrow\mathbb{R} is a smooth function. The chain rule implies that

γ¯=γχ=dχdtγ,\bar{\gamma}^{\prime}=\gamma^{\prime}\circ\chi^{\prime}=\frac{d\chi}{dt}\gamma^{\prime}\,\,, (63)

where, in the last equality, it was used the fact that vectors in \mathbb{R} can be naturally identified with real numbers. Since Λ\Lambda is a linear map, we have

Γ¯(t)=Λγ¯(t)=dχdtΛγ(t)=dχdtΓ(t),\bar{\Gamma}^{\prime}(t)=\Lambda\bar{\gamma}^{\prime}(t)=\frac{d\chi}{dt}\Lambda\gamma^{\prime}(t)=\frac{d\chi}{dt}\Gamma^{\prime}(t)\,\,, (64)

showing that the curve Γ¯\bar{\Gamma} lifted from γ¯\bar{\gamma} is just a reparametrization of the curve Γ\Gamma lifted from γ\gamma,

Γ¯=Γχ,\bar{\Gamma}=\Gamma\circ\chi\,\,, (65)

which implies that Γ¯\bar{\Gamma} has the same image as Γ\Gamma, as we intended to prove.

Refer to caption
Figure 2: The state space 𝒫\mathcal{P}, the configuration space \mathcal{I} and the dynamical space 𝒵\mathcal{Z} are shown (the dotted lines represent fibers of 𝒫\mathcal{P}). A curve γ\gamma in \mathcal{I} is lifted to a curve Γ\Gamma in 𝒫\mathcal{P} starting at the point p𝒫p\in\mathcal{P}. The projection ρ2\rho_{2} maps Γ\Gamma to 𝒵\mathcal{Z}, where the total deviation of the dynamical quantities Δζ\Delta\zeta is the vector connecting the initial and final points of the projected curve.

Let us now derive a formula for the net deviation of the dynamical quantities when γ\gamma is a small closed curve. More precisely, we shall limit ourselves to the case of a small rectangular cycle defined by two transversal smooth vector fields uu and vv in the configuration space \mathcal{I}. Let the parameters of the integral curves of the vector fields uu and vv be xx and yy, respectively. Assume that γ\gamma starts at the point β0\beta_{0}, then goes along the integral curve of uu by an small variation in the parameter Δx\Delta x, then moves along the direction of vv by a parameter change Δy\Delta y, then goes back in the direction of uu by a parameter Δx-\Delta x, and finally moves along vv by an parameter Δy-\Delta y. After going though the four sides of this “parallelogram”, the curve will have returned to the initial point β0\beta_{0} if and only if these two vector fields commute

[u,v]=0,[u,v]=0\,\,, (66)

and this will be assumed to hold. The lifted curve, Γ\Gamma, will be precisely the curve obtained by going through the corresponding parallelogram defined by the lifted vector fields U=ΛuU=\Lambda u and V=ΛvV=\Lambda v. Since the lifted fields may not commute, Γ\Gamma may not be closed and there may be some non-vanishing net deviation of the dynamical quantities after the cycle is traversed. For the quantity ζi\zeta^{i}, we have that, at lowest approximation, its variation is given by

Δζi=.[U,V]ζi|(β0,ζ0)ΔxΔy,\Delta\zeta^{i}=\Big{.}[U,V]\zeta^{i}\Big{|}_{(\beta_{0},\zeta_{0})}\Delta x\Delta y\,\,, (67)

where Δζi=ζi(Γ(Δt))ζ0i\Delta\zeta^{i}=\zeta^{i}(\Gamma(\Delta t))-\zeta_{0}^{i}. For simplicity, let us redefine the vector fields so as to include Δx\Delta x and Δy\Delta y implicitly in the respective vector

uΔx\displaystyle u\Delta x u\displaystyle\mapsto u (68)
UΔx\displaystyle U\Delta x U\displaystyle\mapsto U (69)
vΔy\displaystyle v\Delta y v\displaystyle\mapsto v (70)
VΔy\displaystyle V\Delta y V,\displaystyle\mapsto V\,\,, (71)

so that the new fields can be seen as fields of infinitesimal vectors, and the formula above can be put in a slightly more appealing form

Δζi=[U,V]ζi.\Delta\zeta^{i}=[U,V]\zeta^{i}\,\,. (72)

If \partial is the derivative operator associated with the coordinates (β,ζ)(\beta,\zeta), we have

Δζi=[U,V]aaζi+[U,V]jjζi=[U,V]i,\Delta\zeta^{i}=[U,V]^{a^{\prime}}\partial_{a^{\prime}}\zeta^{i}+[U,V]^{j}\partial_{j}\zeta^{i}=[U,V]^{i}\,\,, (73)

where it was used that ζi/βa=0\partial\zeta^{i}/\partial\beta^{a^{\prime}}=0 and ζi/ζj=δji\partial\zeta^{i}/\partial\zeta^{j}=\delta^{i}_{\,\,j}. After some simple manipulations we get

Δζi=λai[u,v]a+2(aλbi+λajjλbi)u[avb].\Delta\zeta^{i}=\lambda^{i}_{\,\,{a^{\prime}}}[u,v]^{a^{\prime}}+2\left(\partial_{a^{\prime}}\lambda^{i}_{\,\,{b^{\prime}}}+\lambda^{j}_{\,\,{a^{\prime}}}\partial_{j}\lambda^{i}_{\,\,b^{\prime}}\right)u^{[{a^{\prime}}}v^{b^{\prime}]}\,\,. (74)

Thus, using [u,v]a=0[u,v]^{a^{\prime}}=0, we obtain

Δζi=2(aλbμ+λajjλbi)u[avb],\Delta\zeta^{i}=2\left(\partial_{a^{\prime}}\lambda^{\mu}_{\,\,b^{\prime}}+\lambda^{j}_{\,\,{a^{\prime}}}\partial_{j}\lambda^{i}_{\,\,b^{\prime}}\right)u^{[{a^{\prime}}}v^{b^{\prime}]}\,\,, (75)

where the right-hand side is evaluated at (β0,ζ0)(\beta_{0},\zeta_{0}).

It is interesting to point out that if λ\lambda in (57) depended explicitly on the time tt, then the motion would not display this geometric phase character, so that the net variation of the dynamical quantities would depend on γ(t)\gamma(t) and not only on Im(γ)\mathrm{Im}(\gamma). Notwithstanding, the geometric phase character of the dynamics is restored in the limit of small evolution times, for we can approximate

ζ˙i=λai(β,ζ,t)β˙aλai(β,ζ,t0)β˙a+λait|(β,ζ,t0)Δtβ˙a\dot{\zeta}^{i}=\lambda^{i}_{\,\,{a^{\prime}}}(\beta,\zeta,t)\dot{\beta}^{a^{\prime}}\approx\lambda^{i}_{\,\,{a^{\prime}}}(\beta,\zeta,t_{0})\dot{\beta}^{a^{\prime}}+\left.\frac{\partial\lambda^{i}_{\,\,{a^{\prime}}}}{\partial t}\right|_{(\beta,\zeta,t_{0})}\Delta t\dot{\beta}^{a^{\prime}} (76)

where Δt=tt0\Delta t=t-t_{0} is small during the whole integration (for instance, during the small time lapse of a fast cycle). Thus, one can obtain a fairly good approximation for the solution by considering the approximate differential equation

ζ˙iλai(β,ζ,t0)β˙a,\dot{\zeta}^{i}\approx\lambda^{i}_{\,\,{a^{\prime}}}(\beta,\zeta,t_{0})\dot{\beta}^{a^{\prime}}\,\,, (77)

which displays the desired geometrical-phase kind of dynamics we just studied. It is in this sense that we refer to “small time” in this problem.

Note that if ζ˙\dot{\zeta} had a different functional dependence on β˙\dot{\beta} or there were terms containing higher derivatives such as β¨\ddot{\beta}, then the motion would (most likely) not be of the geometrical-phase kind, even in the limit of fast cycles. Indeed, for very fast cycles, β˙\dot{\beta} and β¨\ddot{\beta} would increase accordingly so as to contribute sensibly to the dynamics of the problem. Thus, we shall look for a particular scenario in which small bodies in general relativity could be well described by equations of the form (57).

Note that if the functions λ\lambda are small in some sense, then

Δζi2aλbiu[avb],\Delta\zeta^{i}\approx 2\partial_{a^{\prime}}\lambda^{i}_{\,\,b^{\prime}}u^{[{a^{\prime}}}v^{b^{\prime}]}\,\,, (78)

where the term involving λ2\lambda^{2} were dropped. This implies that it is only the β\beta-dependence of λ\lambda that effectively matters in the final formula, so that λ(β,ζ)\lambda(\beta,\zeta) can be evaluated at the initial values of the dynamical quantities

λai(β,ζ)λai(β,ζ0).\lambda^{i}_{\,\,{a^{\prime}}}(\beta,\zeta)\mapsto\lambda^{i}_{\,\,{a^{\prime}}}(\beta,\zeta_{0})\,\,. (79)

This happens to be our case for the functions λ\lambda will contain factors of curvature and thence contribute with factors of order of magnitude k(l/r)2k(l/r)^{2} or higher, so that λ2\lambda^{2} would be of order k2(l/r)4k^{2}(l/r)^{4} or higher, negligible in our approximations.

The equation for the trajectory of the center of mass, given in (15), is of second order in zμ(s)z^{\mu}(s). Since we will be mostly interested in fast cycles of internal motions, we will only have to solve these equations for a small time interval ss (assuming that the initial conditions are given at s=0s=0). It is then convenient to take the local normal coordinate system centered at z(0)z(0) (so that zμ(0)=0z^{\mu}(0)=0) and oriented along the initial dynamical velocity uk(0)u^{k}(0) (so that uμ(0)=(1,0,0,0)u^{\mu}(0)=(1,0,0,0)). If the trajectory of the center of mass is described by zμ(s)z^{\mu}(s), then vllvkv^{l}\nabla_{l}v^{k} decomposes as

vννvμ=d2zμds2+Γνσμ(z)dzνdsdzσds,v^{\nu}\nabla_{\nu}v^{\mu}=\frac{d^{2}z^{\mu}}{ds^{2}}+\Gamma^{\mu}_{\,\,\nu\sigma}(z)\frac{dz^{\nu}}{ds}\frac{dz^{\sigma}}{ds}\,\,, (80)

where Γνσμ(z)\Gamma^{\mu}_{\,\,\nu\sigma}(z) can be approximated as

Γνσμ(z)=23Rρ(νσ)μ(0)zρ+12ρκΓνσμ(0)zρzκ,\Gamma^{\mu}_{\,\,\nu\sigma}(z)=\frac{2}{3}{R_{\rho(\nu\sigma)}}^{\mu}(0)z^{\rho}+\frac{1}{2}\partial_{\rho}\partial_{\kappa}\Gamma^{\mu}_{\,\,\nu\sigma}(0)z^{\rho}z^{\kappa}\,\,, (81)

in which ρκΓνσμ(0)\partial_{\rho}\partial_{\kappa}\Gamma^{\mu}_{\,\,\nu\sigma}(0) can be expressed in terms of the first derivative of the curvature at the origin, κRρσνμ(0)\nabla_{\kappa}{R_{\rho\sigma\nu}}^{\mu}(0). Thus, equation (15) gets

z¨μ\displaystyle\ddot{z}^{\mu} =Γνσμ(z)z˙νz˙σ+1M(FμFνuνuμ)\displaystyle=-\Gamma^{\mu}_{\,\,\nu\sigma}(z)\dot{z}^{\nu}\dot{z}^{\sigma}+\frac{1}{M}\left(F^{\mu}-F^{\nu}u_{\nu}u^{\mu}\right)
+1M2SμνF˙ν+1MQ˙μνuν,\displaystyle+\frac{1}{M^{2}}S^{\mu\nu}\dot{F}_{\nu}+\frac{1}{M}\dot{Q}^{\mu\nu}u_{\nu}\,, (82)

where the Christoffel symbols associated with the derivatives of the force and torque could be dropped due to our approximations. For the sake of completeness, consider the differential equations corresponding to the time evolution of other relevant dynamical quantities

vlluk=1M(FkFluluk)\displaystyle v^{l}\nabla_{l}u^{k}=\frac{1}{M}\left(F^{k}-F^{l}u_{l}u^{k}\right)
vllM=Flul\displaystyle v^{l}\nabla_{l}M=F^{l}u_{l}
vrrSkl=2Mu[kvl]+Qkl\displaystyle v^{r}\nabla_{r}S^{kl}=2Mu^{[k}v^{l]}+Q^{kl}
vll(e¯μ)k=2MF[kul](e¯μ)l.\displaystyle v^{l}\nabla_{l}(\bar{e}_{\mu})^{k}=\frac{2}{M}F^{[k}u^{l]}(\bar{e}_{\mu})_{l}\,\,. (83)

Note that, integrating these equations to the lowest order in time we obtain something of the form

ζ(s)ζ(0)+αs,\zeta(s)\approx\zeta(0)+\alpha s\,\,, (84)

where α\alpha is a quantity that necessarily contains a factor of the magnitude k/r2k/r^{2}. Hence, considering our approximations, all the dynamical quantities in the right-hand side of equation (82) can be approximated by their initial values

MM0\displaystyle M\mapsto M_{0}
uμδ  0μ\displaystyle u^{\mu}\mapsto\delta^{\mu}_{\,\,0}
SμνS0μν\displaystyle S^{\mu\nu}\mapsto S_{0}^{\mu\nu}
(e¯ν)μδνμ,\displaystyle(\bar{e}_{\nu})^{\mu}\mapsto\delta^{\mu}_{\,\,\nu}\,, (85)

as long as ss is not too big. Also, since the kinematical velocity z˙μ\dot{z}^{\mu} only differs from the dynamical velocity uμu^{\mu} by a quantity containing the curvature, we may also take

z˙μδ  0μ\dot{z}^{\mu}\mapsto\delta^{\mu}_{\,\,0} (86)

in the right-hand side. Thence, equation (82) can be simplified to

z¨μ=Γ  00μ(z)+1M0(FμF0δ  0μ)+1M02S0μνF˙ν+1M0Q˙μ0,\ddot{z}^{\mu}=-\Gamma^{\mu}_{\,\,00}(z)+\frac{1}{M_{0}}\left(F^{\mu}-F^{0}\delta^{\mu}_{\,\,0}\right)+\frac{1}{M_{0}^{2}}S_{0}^{\mu\nu}\dot{F}_{\nu}+\frac{1}{M_{0}}\dot{Q}^{\mu 0}\,, (87)

and, in addition, the force and torque can be given by

Fμ\displaystyle F^{\mu} =12S0νρRμ0νρ+16J¯νρσλμRνρσλ\displaystyle=\frac{1}{2}S_{0}^{\nu\rho}{R^{\mu}}_{0\nu\rho}+\frac{1}{6}\bar{J}^{\nu\rho\sigma\lambda}\nabla^{\mu}R_{\nu\rho\sigma\lambda}
Qμ0\displaystyle Q^{\mu 0} =43J¯νρσ[μRνρσ0]+\displaystyle=\frac{4}{3}\bar{J}^{\nu\rho\sigma[\mu}{R_{\nu\rho\sigma}}^{0]}\,+
+13J¯νρσλ[μ(3νRρσλ0]+2λRνρσ0]),\displaystyle+\frac{1}{3}\bar{J}^{\nu\rho\sigma\lambda[\mu}\left(3\nabla_{\nu}{R_{\rho\sigma\lambda}}^{0]}+2\nabla_{\lambda}{R_{\nu\rho\sigma}}^{0]}\right)\,, (88)

where it was used that Jμνρσ=J¯μνρσJ^{\mu\nu\rho\sigma}=\bar{J}^{\mu\nu\rho\sigma} and Jμνρσλ=J¯μνρσλJ^{\mu\nu\rho\sigma\lambda}=\bar{J}^{\mu\nu\rho\sigma\lambda} since the non-spinning basis e¯μ\bar{e}_{\mu} coincides with the coordinate basis μ\partial_{\mu}. Note also that the octupole contribution to the force was omitted, which is due to the fact that it depends on the second derivative of the curvature tensor, assumed to be of order k/r4k/r^{4}.

The zμz^{\mu} appearing in the right-hand side of equation (87) can be approximated using a similar argument as the one employed for the dynamical quantities. By equation (12), we conclude that

zμ(s)sδ  0μ+αμs,z^{\mu}(s)\approx s\delta^{\mu}_{\,\,0}+\alpha^{\mu}s\,\,, (89)

where αμ\alpha^{\mu} are quantities involving factors of curvature k/r2\sim k/r^{2}. Hence, the zμz^{\mu} that appears in the arguments of Γ  00μ(z)\Gamma^{\mu}_{\,\,00}(z) and the curvature (and its derivatives) can be approximated by zμ(s)=(s,0,0,0)z^{\mu}(s)=(s,0,0,0).

The equation for the spatial position of the center of mass, zi(s)z^{i}(s), is then given by

z¨i=Γ  00i(z)+1M0Fi+1M02S0ijF˙j+1M0Q˙i0\ddot{z}^{i}=-\Gamma^{i}_{\,\,00}(z)+\frac{1}{M_{0}}F^{i}+\frac{1}{M_{0}^{2}}S_{0}^{ij}\dot{F}_{j}+\frac{1}{M_{0}}\dot{Q}^{i0} (90)

or, alternatively,

dds[z˙i1M02S0ijFj1M0Qi0]=Γ  00i(z)+1M0Fi.\frac{d}{ds}\left[\dot{z}^{i}-\frac{1}{M_{0}^{2}}S_{0}^{ij}F_{j}-\frac{1}{M_{0}}Q^{i0}\right]=-\Gamma^{i}_{\,\,00}(z)+\frac{1}{M_{0}}F^{i}\,\,. (91)

But this equation still does not have the desired form (57). At this point it is convenient to assume that J¯\bar{J} is linear in β¨\ddot{\beta}, that is,

J¯(β,β˙,β¨,s)=J¯(β,β˙,0,s)+J¯β¨|(β,β˙,0,s)β¨.\bar{J}(\beta,\dot{\beta},\ddot{\beta},s)=\bar{J}(\beta,\dot{\beta},0,s)+\left.\frac{\partial\bar{J}}{\partial\ddot{\beta}}\right|_{(\beta,\dot{\beta},0,s)}\ddot{\beta}\,. (92)

This is motivated by the fact that, in Newtonian mechanics, the β¨\ddot{\beta}’s in the J¯\bar{J}’s typically come from the tension contributions in the energy-momentum tensor, which usually appear linearly in the stress tensor. Since a boost by β˙\dot{\beta} cannot introduce extra β¨\ddot{\beta} in the stress-energy tensor, this should still be true in special relativity, as long as the local physics in the instantaneous rest frame of each infinitesimal piece of matter is essentially Newtonian – if this is not true, then we shall take this special form of J¯\bar{J} as an additional assumption. Accordingly, the following quantities can be written as

1M0Fi=ϕi(β,β˙,s)+ϕ~ai(β,β˙,s)β¨a\displaystyle\frac{1}{M_{0}}F^{i}=\phi^{i}(\beta,\dot{\beta},s)+\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},s)\ddot{\beta}^{a^{\prime}}
1M02S0ijFj+1M0Qi0=ψi(β,β˙,s)+ψ~ai(β,β˙,s)β¨a,\displaystyle\frac{1}{M_{0}^{2}}S_{0}^{ij}F_{j}+\frac{1}{M_{0}}Q^{i0}=\psi^{i}(\beta,\dot{\beta},s)+\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},s)\ddot{\beta}^{a^{\prime}}\,\,, (93)

where the functions ϕ\phi, ϕ~\widetilde{\phi}, ψ\psi and ψ~\widetilde{\psi} depend implicitly on the initial conditions. The components of the spatial acceleration z¨i\ddot{z}^{i} can be generally cast in the form

dds[z˙iϕ~ai(β,β˙,s)β˙aψi(β,β˙,s)ψ~ai(β,β˙,s)β¨a]\displaystyle\frac{d}{ds}\left[\dot{z}^{i}-\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},s)\dot{\beta}^{a^{\prime}}-\psi^{i}(\beta,\dot{\beta},s)-\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},s)\ddot{\beta}^{a^{\prime}}\right]
=ϵi(s)+ϕi(β,β˙,s)dϕ~aidsβ˙a,\displaystyle=\epsilon^{i}(s)+\phi^{i}(\beta,\dot{\beta},s)-\frac{d\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{ds}\dot{\beta}^{a^{\prime}}\,\,, (94)

where ϵi(s)=Γ  00i(z)\epsilon^{i}(s)=-\Gamma^{i}_{\,\,00}(z). Note that the quantity inside the square brackets in the left-hand side corresponds to a first order differential equation in ziz^{i} that differs from equation (57) only by the term with β¨\ddot{\beta} and the possible non-linearity in β˙\dot{\beta}. If we integrate equation (V) in time, it is conceivable that the right-hand side could be neglected for very short times of evolution,

z˙iϕ~ai(β,β˙,s)β˙aψi(β,β˙,s)ψ~ai(β,β˙,s)β¨a=\displaystyle\dot{z}^{i}-\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},s)\dot{\beta}^{a^{\prime}}-\psi^{i}(\beta,\dot{\beta},s)-\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},s)\ddot{\beta}^{a^{\prime}}=
=0s()𝑑s0,\displaystyle\qquad=\int_{0}^{s}(\cdots)ds^{\prime}\approx 0\,, (95)

which could be true if the right-hand side of (V) depended only on β\beta, β˙\dot{\beta} and ss, and not on superior derivatives of β\beta. However, the term involving the time derivative of ϕ~\widetilde{\phi} can be expressed as

dϕ~aidsβ˙a=(ϕ~aiβbβ˙b+ϕ~aiβ˙bβ¨b+ϕ~ais)β˙a,\frac{d\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{ds}\dot{\beta}^{a^{\prime}}=\left(\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial\beta^{b^{\prime}}}\dot{\beta}^{b^{\prime}}+\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial\dot{\beta}^{b^{\prime}}}\ddot{\beta}^{b^{\prime}}+\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial s}\right)\dot{\beta}^{a^{\prime}}\,\,, (96)

and we can clearly see that it may depend121212Note that one could have tried to integrate ϕ~aiβ¨a\widetilde{\phi}^{i}_{\,\,a^{\prime}}\ddot{\beta}^{a^{\prime}} into the right-hand side of equation (V) by finding a function ϕ^i(β,β˙,s)\widehat{\phi}^{i}(\beta,\dot{\beta},s) such that ϕ^i/β˙a=ϕ~ai\partial\widehat{\phi}^{i}/\dot{\beta}^{a^{\prime}}=\widetilde{\phi}^{i}_{\,\,a^{\prime}}. However, that would only be possible if ϕ~ai/β˙bϕ~bi/β˙a=0\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}/\partial\dot{\beta}^{b^{\prime}}-\partial\widetilde{\phi}^{i}_{\,\,b^{\prime}}/\partial\dot{\beta}^{a^{\prime}}=0, yet there is no physical reason for that being true. on β¨\ddot{\beta}. Nonetheless, this β¨\ddot{\beta}-dependent term is absent if we ask for the condition

β˙a(J¯β¨b)=0,\frac{\partial}{\partial\dot{\beta}^{a^{\prime}}}\left(\frac{\partial\bar{J}}{\partial\ddot{\beta}^{b^{\prime}}}\right)=0\,, (97)

which means that the terms in the energy-momentum tensor TμνT^{\mu\nu} involving the “accelerations” β¨\ddot{\beta} would not depend on the “velocities” β˙\dot{\beta}. This is fairly common in Newtonian problems, where the tensions tend to assume the form

τκ(β)+θ(β)β˙β˙+ω(β)β¨\tau\sim\kappa(\beta)+\theta(\beta)\dot{\beta}\dot{\beta}+\omega(\beta)\ddot{\beta} (98)

and it can be easily verified that this form is invariant by a (non velocity-dependent) change of coordinates ββ(β)\beta\rightarrow\beta^{\prime}(\beta). Under this assumption, we can write equation (V) as

dds[Δz˙iϕ~ai(β,s)β˙aψi(β,β˙,s)ψ~ai(β,s)β¨a]=\displaystyle\frac{d}{ds}\left[\Delta\dot{z}^{i}-\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,s)\dot{\beta}^{a^{\prime}}-\psi^{i}(\beta,\dot{\beta},s)-\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,s)\ddot{\beta}^{a^{\prime}}\right]=
=ϕi(β,β˙,s)(ϕ~aiβbβ˙b+ϕ~ais)β˙a,\displaystyle=\phi^{i}(\beta,\dot{\beta},s)-\left(\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial\beta^{b^{\prime}}}\dot{\beta}^{b^{\prime}}+\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial s}\right)\dot{\beta}^{a^{\prime}}\,\,, (99)

where the arguments of ϕ~\widetilde{\phi} and ψ~\widetilde{\psi} were accordingly reduced to (β,s)(\beta,s).

It seems difficult to proceed much further from this point if we do not make any assumptions on the β˙\dot{\beta}’s. Clearly, since they are mainly related with the velocities, it seems reasonable to assume that they are bounded. For example, if (l,α)(l,\alpha) are the radial coordinates of point moving on a 2-dimensional plane, then its velocity would be given by v=(l˙,lα˙)\vec{v}=(\dot{l},l\dot{\alpha}), which puts a restriction of the kind l˙1\dot{l}\lesssim 1 and α˙l1\dot{\alpha}\lesssim l^{-1}. We shall, nonetheless, go beyond and assume that the β˙\dot{\beta}’s are “small” quantities. This is supposed to ensure that the internal motions are slow enough so as to be essentially Newtonian, such that relativistic-related non-linear terms in the velocities could be neglected in the equations. Note that this is essential to the tenability of hypothesis (97), whose underlying argument is based in (98), that is only reasonable in the Newtonian regime (in fact, a boost would necessarily introduce non-linear velocity factors in the tensions). Thus, we shall assume that the velocities are small and that the tensions are much smaller that the mass density scale of the body, so that the dynamics of the internal motions can be treated in a purely Newtonian way.

Let z0(s)z_{0}(s) be the trajectory of the rigid body (i.e., when β(s)=β0\beta(s)=\beta_{0}), but starting with the same initial conditions z0μ(0)=0z_{0}^{\mu}(0)=0 and z˙0μ(0)=δ  0μ\dot{z}_{0}^{\mu}(0)=\delta^{\mu}_{\,\,0}. Its second derivative satisfies the equation

dds[Δz˙0iψi(β0,0,s)]=ϵi(s)+ϕi(β0,0,s),\frac{d}{ds}\left[\Delta\dot{z}_{0}^{i}-\psi^{i}(\beta_{0},0,s)\right]=\epsilon^{i}(s)+\phi^{i}(\beta_{0},0,s)\,\,, (100)

so that the deviation Δzi(s)=zi(s)z0i(s)\Delta z^{i}(s)=z^{i}(s)-z_{0}^{i}(s) satisfies

dds[Δz˙iϕ~ai(β,s)β˙aψi(β,β˙,s)+ψi(β0,0,s)ψ~ai(β,β˙,s)β¨a]=\displaystyle\frac{d}{ds}\left[\Delta\dot{z}^{i}-\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,s)\dot{\beta}^{a^{\prime}}-\psi^{i}(\beta,\dot{\beta},s)+\psi^{i}(\beta_{0},0,s)-\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},s)\ddot{\beta}^{a^{\prime}}\right]=
=ϕi(β,β˙,s)ϕi(β0,0,s)(ϕ~aiβbβ˙b+ϕ~ais)β˙a.\displaystyle=\phi^{i}(\beta,\dot{\beta},s)-\phi^{i}(\beta_{0},0,s)-\left(\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial\beta^{b^{\prime}}}\dot{\beta}^{b^{\prime}}+\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial s}\right)\dot{\beta}^{a^{\prime}}\,\,. (101)

Expanding ϕ\phi and ψ\psi in Taylor and keeping first order terms in Δβ\Delta\beta but second order terms in β˙\dot{\beta} we get

dds[Δz˙iϕ~ai(β,s)β˙aψiβa|χ0(s)Δβaψiβ˙a|χ(s)β˙a\displaystyle\frac{d}{ds}\left[\Delta\dot{z}^{i}-\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,s)\dot{\beta}^{a^{\prime}}-\left.\frac{\partial\psi^{i}}{\partial\beta^{a^{\prime}}}\right|_{\chi_{0}(s)}\!\Delta\beta^{a^{\prime}}-\left.\frac{\partial\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}}\right|_{\chi(s)}\!\dot{\beta}^{a^{\prime}}\right.
122ψiβ˙aβ˙b|χ(s)β˙aβ˙bψ~ai(β,s)β¨a]\displaystyle\left.-\frac{1}{2}\left.\frac{\partial^{2}\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\right|_{\chi(s)}\!\dot{\beta}^{a^{\prime}}\dot{\beta}^{b^{\prime}}-\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,s)\ddot{\beta}^{a^{\prime}}\right]
=ϕiβa|χ0(s)Δβa+ϕiβ˙a|χ(s)β˙a\displaystyle=\left.\frac{\partial\phi^{i}}{\partial\beta^{a^{\prime}}}\right|_{\chi_{0}(s)}\Delta\beta^{a^{\prime}}+\left.\frac{\partial\phi^{i}}{\partial\dot{\beta}^{a^{\prime}}}\right|_{\chi(s)}\dot{\beta}^{a^{\prime}}
+122ϕiβ˙aβ˙b|χ(s)β˙aβ˙b(ϕ~aiβbβ˙b+ϕ~ais)β˙a,\displaystyle+\frac{1}{2}\left.\frac{\partial^{2}\phi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\right|_{\chi(s)}\dot{\beta}^{a^{\prime}}\dot{\beta}^{b^{\prime}}-\left(\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial\beta^{b^{\prime}}}\dot{\beta}^{b^{\prime}}+\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial s}\right)\dot{\beta}^{a^{\prime}}\,\,, (102)

where χ(s)=(β,0,s)\chi(s)=(\beta,0,s) and χ0(s)=(β0,0,s)\chi_{0}(s)=(\beta_{0},0,s). For small ss (of the order of the cycle duration), we typically have that Δββ˙s\Delta\beta\sim\dot{\beta}s and β¨β˙/s\ddot{\beta}\sim\dot{\beta}/s. Therefore, the order of magnitude of each term in the equation above, after the time integration, can be estimated as

Δz˙iϕ~aiβ˙aψiββ˙sψiβ˙aβ˙a\displaystyle\Delta\dot{z}^{i}-\widetilde{\phi}^{i}_{\,\,a^{\prime}}\dot{\beta}^{a^{\prime}}-\frac{\partial\psi^{i}}{\partial\beta}\dot{\beta}s-\frac{\partial\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}}\dot{\beta}^{a^{\prime}}
122ψiβ˙aβ˙bβ˙aβ˙bψ~aiβ˙as1\displaystyle-\frac{1}{2}\frac{\partial^{2}\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\dot{\beta}^{a^{\prime}}\dot{\beta}^{b^{\prime}}-\widetilde{\psi}^{i}_{\,\,a^{\prime}}\dot{\beta}^{a^{\prime}}s^{-1}\,\sim
ϕiβaβ˙as2+ϕiβ˙aβ˙as+122ϕiβ˙aβ˙bβ˙aβ˙bs\displaystyle\sim\,\frac{\partial\phi^{i}}{\partial\beta^{a^{\prime}}}\dot{\beta}^{a^{\prime}}s^{2}+\frac{\partial\phi^{i}}{\partial\dot{\beta}^{a^{\prime}}}\dot{\beta}^{a^{\prime}}s+\frac{1}{2}\frac{\partial^{2}\phi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\dot{\beta}^{a^{\prime}}\dot{\beta}^{b^{\prime}}s
(ϕ~aiβbβ˙b+ϕ~ais)β˙as.\displaystyle-\left(\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial\beta^{b^{\prime}}}\dot{\beta}^{b^{\prime}}+\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial s}\right)\dot{\beta}^{a^{\prime}}s\,\,. (103)

Thus, assuming that any terms that go to zero when ss is taken to zero (while keeping all other parameters fixed) can be neglected, we obtain the following first order differential equation for Δz\Delta z,

Δz˙i=ψ~ai(β,s)β¨a+\displaystyle\Delta\dot{z}^{i}=\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,s)\ddot{\beta}^{a^{\prime}}\,+
(ϕ~ai(β,s)+ψiβ˙a(β,0,s)+122ψiβ˙aβ˙b(β,0,s)β˙b)β˙a.\displaystyle\left(\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,s)+\frac{\partial\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}}(\beta,0,s)+\frac{1}{2}\frac{\partial^{2}\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}(\beta,0,s)\,\dot{\beta}^{b^{\prime}}\right)\dot{\beta}^{a^{\prime}}. (104)

Under the same s0s\rightarrow 0 condition, we may consider a Taylor expansion in ss, so that, also assuming that the terms proportional to ss can be neglected, we get

Δz˙i=ψ~ai(β,0)β¨a+\displaystyle\Delta\dot{z}^{i}=\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,0)\ddot{\beta}^{a^{\prime}}\,+
(ϕ~ai(β,0)+ψiβ˙a(β,0,0)+122ψiβ˙aβ˙b(β,0,0)β˙b)β˙a,\displaystyle\left(\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,0)+\frac{\partial\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}}(\beta,0,0)+\frac{1}{2}\frac{\partial^{2}\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}(\beta,0,0)\,\dot{\beta}^{b^{\prime}}\right)\dot{\beta}^{a^{\prime}}, (105)

that has the form

Δz˙i=λai(β,β˙)β˙a+λ~ai(β)β¨a,\Delta\dot{z}^{i}=\lambda^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta})\dot{\beta}^{a^{\prime}}+\widetilde{\lambda}^{i}_{\,\,a^{\prime}}(\beta)\ddot{\beta}^{a^{\prime}}\,\,, (106)

where

λai(β,β˙)\displaystyle\lambda^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta}) =ϕ~ai(β,0)+ψiβ˙a(β,0,0)+\displaystyle=\widetilde{\phi}^{i}_{\,\,a^{\prime}}(\beta,0)+\frac{\partial\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}}(\beta,0,0)\,+
+122ψiβ˙aβ˙b(β,0,0)β˙b\displaystyle\qquad\qquad+\frac{1}{2}\frac{\partial^{2}\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}(\beta,0,0)\,\dot{\beta}^{b^{\prime}}
λ~ai(β)\displaystyle\widetilde{\lambda}^{i}_{\,\,a^{\prime}}(\beta) =ψ~ai(β,0).\displaystyle=\widetilde{\psi}^{i}_{\,\,a^{\prime}}(\beta,0)\,. (107)

Note that ϕ~\widetilde{\phi} is essentially related to tension contributions to the force, ψ/β˙\partial\psi/\partial\dot{\beta} is related to linear momentum contributions to the torque or spin-force coupling, 2ψ/β˙2\partial^{2}\psi/\partial\dot{\beta}^{2} is related to tension contributions to the torque or spin-force coupling, and ψ~\widetilde{\psi} to tension contributions to the torque or spin-force coupling.

This equation is not of the geometric-phase kind for it differs from equation (57). Notwithstanding, it still has relatively simple mathematical properties. The resulting motion depends on the image of the path traversed in the tangent bundle of the internal space, instead of in the internal space as desired. In order to be precise, let the extended state space 𝒫~\widetilde{\mathcal{P}} be defined as the vector bundle (T×𝒵,T,ρ1)(T\mathcal{I}\times\mathcal{Z},T\mathcal{I},\rho_{1}), where ρ1\rho_{1} is the Cartesian projection of T×𝒵T\mathcal{I}\times\mathcal{Z} on the first component TT\mathcal{I}. Given a curve γ:\gamma:\mathbb{R}\rightarrow\mathcal{I} in the configuration space, consider its canonical lift γ~:T\widetilde{\gamma}:\mathbb{R}\rightarrow T\mathcal{I} to the tangent bundle (which corresponds to the field of tangent vectors along γ\gamma). For a coordinate system {βa}\{\beta^{a^{\prime}}\} in \mathcal{I}, take the natural coordinate system {βa,β¯b}\{\beta^{a^{\prime}},\bar{\beta}^{b^{\prime}}\} associated with TT\mathcal{I}. If the coordinates of γ\gamma are γa(s)\gamma^{a^{\prime}}(s), so that (γ~)a=γ˙a=dγa/ds(\widetilde{\gamma})^{a^{\prime}}=\dot{\gamma}^{a^{\prime}}=d\gamma^{a^{\prime}}/ds, then the components of the canonical lift are given by (γ1(s),γd(s),γ˙1,γ˙d(s))(\gamma^{1}(s),\ldots\gamma^{d}(s),\dot{\gamma}^{1},\ldots\dot{\gamma}^{d}(s)). Consider a generalization of equation (106) given by

Δζ˙i=λai(β,β˙,ζ)β˙a+λ~ai(β,β˙,ζ)β¨a,\Delta\dot{\zeta}^{i}=\lambda^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},\zeta)\dot{\beta}^{a^{\prime}}+\widetilde{\lambda}^{i}_{\,\,a^{\prime}}(\beta,\dot{\beta},\zeta)\ddot{\beta}^{a^{\prime}}\,\,, (108)

which clearly reduces to (106) when λ\lambda and λ~\widetilde{\lambda} are independent of ζ\zeta. Based on this equation we define131313Consider that the notation is analogous to the one employed before, except that now vectors in the base space are represented by (va,v¯a)(v^{a^{\prime}},\bar{v}^{a^{\prime}}) instead of simply (va)(v^{a^{\prime}}). the Λ~\widetilde{\Lambda}-lift of a vector v~\widetilde{v} in TT\mathcal{I} to a vector V~\widetilde{V} in 𝒫~\widetilde{\mathcal{P}} as

V~=Λ~v~=(va,v¯a;λai(β,β¯,ζ)va+λ~ai(β,β¯,ζ)v¯a),\widetilde{V}=\widetilde{\Lambda}\widetilde{v}=\left(v^{a^{\prime}},\bar{v}^{a^{\prime}};\lambda^{i}_{\,\,a^{\prime}}(\beta,\bar{\beta},\zeta)v^{a^{\prime}}+\widetilde{\lambda}^{i}_{\,\,a^{\prime}}(\beta,\bar{\beta},\zeta){\bar{v}}^{a^{\prime}}\right)\,\,, (109)

where v~=(va,v¯a)\widetilde{v}=(v^{a^{\prime}},\bar{v}^{a^{\prime}}) is an arbitrary vector at T(β,β¯)TT_{(\beta,\bar{\beta})}T\mathcal{I} and V~\widetilde{V} is a vector at T(β,β¯;ζ)𝒫T_{(\beta,\bar{\beta};\zeta)}\mathcal{P}. Thus, a curve γ\gamma in \mathcal{I} can be canonically lifted to a curve γ~\widetilde{\gamma} in TT\mathcal{I}, which then can be lifted to a curve Γ~\widetilde{\Gamma} in 𝒫~\widetilde{\mathcal{P}} defined analogously to the lift from \mathcal{I} to 𝒫\mathcal{P} discussed before. It is therefore clear that the curve Γ~\widetilde{\Gamma} represents a solution of equation (108). It is also clear that the deviations of the dynamical quantities Δζi\Delta\zeta^{i} will be functions of the image of Γ~\widetilde{\Gamma}, and thence functions of the image of γ~\widetilde{\gamma}. However, we see that different parameterizations for the curve γ\gamma in the configuration space lift to different curves γ~\widetilde{\gamma} in the tangent space - pictorially, the faster the path in the configuration space is traversed, more “stretched” in the direction of the fibers is the lifted curve in the tangent bundle. Therefore, the deviations of the dynamical quantities will not be functions of the image of the curve in the configuration space, i.e.,

Δζ=F~(Im(Γ~))=f~(Im(γ~))f(Im(γ)),\Delta\zeta=\widetilde{F}(\mathrm{Im}(\widetilde{\Gamma}))=\widetilde{f}(\mathrm{Im}(\widetilde{\gamma}))\neq f(\mathrm{Im}(\gamma))\,\,, (110)

where F~\widetilde{F}, f~\widetilde{f} and ff are suitable functions. Thus, the overall motion of a body described by equation (106) does not depend solely on the sequence of geometrical shapes assumed by the body, but also on how fast the internal motions are performed.

This suggests that swimming (at low Reynolds number) in spacetime can only occur in very particular scenarios in which the interaction between the multipole moments and the curvature is such that λ~\widetilde{\lambda} is negligible and λ\lambda does not depend on β˙\dot{\beta}. This happens if there is no tension contributions to the torque and spin-force couple. More precisely, if the following vanish, or can be neglected in some regime,

λ~ai\displaystyle\widetilde{\lambda}^{i}_{\,\,a^{\prime}} =1M02S0ijFjβ¨a+1M0Qi0β¨a=0\displaystyle=\frac{1}{M_{0}^{2}}S_{0}^{ij}\frac{\partial F_{j}}{\partial\ddot{\beta}^{a^{\prime}}}+\frac{1}{M_{0}}\frac{\partial Q^{i0}}{\partial\ddot{\beta}^{a^{\prime}}}=0
2ψiβ˙aβ˙b\displaystyle\frac{\partial^{2}\psi^{i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}} =1M02S0ij2Fjβ˙aβ˙b+1M02Qi0β˙aβ˙b=0,\displaystyle=\frac{1}{M_{0}^{2}}S_{0}^{ij}\frac{\partial^{2}F_{j}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}+\frac{1}{M_{0}}\frac{\partial^{2}Q^{i0}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}=0\,\,, (111)

where the right-hand sides are evaluated at β˙=0\dot{\beta}=0, β¨=0\ddot{\beta}=0 and s=0s=0.

Consider first that we are only interested in observing swimming at order (l/r)2(l/r)^{2}, then we have to ask for the condition

J¯νρσ[iβ¨aRνρσ0]=2J¯νρσ[iβ˙aβ˙bRνρσ0]=0,\frac{\partial\bar{J}^{\nu\rho\sigma[i}}{\partial\ddot{\beta}^{a^{\prime}}}{R_{\nu\rho\sigma}}^{0]}=\frac{\partial^{2}\bar{J}^{\nu\rho\sigma[i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}{R_{\nu\rho\sigma}}^{0]}=0\,\,, (112)

and, since in the Newtonian regime the tensions only contribute to components of the multipole moments with the form JijklJ^{ijkl} (i.e., with all space-like indices), then this condition reduces to

J¯ijklβ¨aRijk0=J¯ijklβ˙aβ˙bRijk0=0,\frac{\partial\bar{J}^{ijkl}}{\partial\ddot{\beta}^{a^{\prime}}}{R_{ijk}}^{0}=\frac{\partial\bar{J}^{ijkl}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}{R_{ijk}}^{0}=0\,\,, (113)

which can be satisfied for all bodies if the spacetime is such that

Rijk0=0,{R_{ijk}}^{0}=0\,\,, (114)

recalling that the 0-index is defined by the initial dynamical velocity (and the spatial indices are orthogonal to it). If Rijk00{R_{ijk}}^{0}\neq 0, then only bodies with certain geometric shapes or performing specific internal motions could satisfy the condition above.

If one wants to observe swimming at (l/r)3(l/r)^{3} order, then one must ask for the additional conditions

2M0J¯νρσλ[iβ¨a(3νRρσλ0]+2λRνρσ0])+\displaystyle 2M_{0}\frac{\partial\bar{J}^{\nu\rho\sigma\lambda[i}}{\partial\ddot{\beta}^{a^{\prime}}}\left(3\nabla_{\nu}{R_{\rho\sigma\lambda}}^{0]}+2\nabla_{\lambda}{R_{\nu\rho\sigma}}^{0]}\right)\,+
+S0ijJ¯νρσλβ¨ajRνρσλ=0\displaystyle\qquad\qquad+S_{0}^{ij}\frac{\partial\bar{J}^{\nu\rho\sigma\lambda}}{\partial\ddot{\beta}^{a^{\prime}}}\nabla_{j}R_{\nu\rho\sigma\lambda}=0
2M02J¯νρσλ[iβ˙aβ˙b(3νRρσλ0]+2λRνρσ0])+\displaystyle 2M_{0}\frac{\partial^{2}\bar{J}^{\nu\rho\sigma\lambda[i}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\left(3\nabla_{\nu}{R_{\rho\sigma\lambda}}^{0]}+2\nabla_{\lambda}{R_{\nu\rho\sigma}}^{0]}\right)\,+
+S0ij2J¯νρσλβ˙aβ˙bjRνρσλ=0,\displaystyle\qquad\qquad+S_{0}^{ij}\frac{\partial^{2}\bar{J}^{\nu\rho\sigma\lambda}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\nabla_{j}R_{\nu\rho\sigma\lambda}=0\,, (115)

which, in the Newtonian regime for internal motions, reduce to

M0J¯klmriβ¨a(3kRlmr0+2rRklm0)+\displaystyle M_{0}\frac{\partial\bar{J}^{klmri}}{\partial\ddot{\beta}^{a^{\prime}}}\left(3\nabla_{k}{R_{lmr}}^{0}+2\nabla_{r}{R_{klm}}^{0}\right)\,+
+S0ijJ¯klmrβ¨ajRklmr=0\displaystyle\qquad\qquad+S_{0}^{ij}\frac{\partial\bar{J}^{klmr}}{\partial\ddot{\beta}^{a^{\prime}}}\nabla_{j}R_{klmr}=0
M02J¯klmriβ˙aβ˙b(3kRlmr0+2rRklm0)+\displaystyle M_{0}\frac{\partial^{2}\bar{J}^{klmri}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\left(3\nabla_{k}{R_{lmr}}^{0}+2\nabla_{r}{R_{klm}}^{0}\right)\,+
+S0ij2J¯klmrβ˙aβ˙bjRklmr=0,\displaystyle\qquad\qquad+S_{0}^{ij}\frac{\partial^{2}\bar{J}^{klmr}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}\nabla_{j}R_{klmr}=0\,, (116)

which can be satisfied for all bodies if

lRijkμ=0,\nabla_{l}{R_{ijk}}^{\mu}=0\,\,, (117)

otherwise only particular shapes and internal motions could satisfy them. Note that if there were no spin involved, then we could ask simply for lRijk0=0\nabla_{l}{R_{ijk}}^{0}=0.

Under these conditions, the equation for the position deviation is given by

Δz˙i=λai(β)β˙a,\Delta\dot{z}^{i}=\lambda^{i}_{\,\,a^{\prime}}(\beta)\dot{\beta}^{a^{\prime}}\,\,, (118)

manifesting the geometric-phase character of the dynamics. We define the differential displacement as the 1-form in the configuration space \mathcal{I} given by

δzi=λai(β)dβa,\delta z^{i}=\lambda^{i}_{\,\,a^{\prime}}(\beta)d\beta^{a^{\prime}}\,\,, (119)

where λ\lambda is defined in (107), but can be written explicitly as

λai(β)=1M0Fiβ¨a+1M02S0ijFjβ˙a+1M0Qi0β˙a,\lambda^{i}_{a^{\prime}}(\beta)=\frac{1}{M_{0}}\frac{\partial F^{i}}{\partial\ddot{\beta}^{a^{\prime}}}+\frac{1}{M_{0}^{2}}S_{0}^{ij}\frac{\partial F_{j}}{\partial\dot{\beta}^{a^{\prime}}}+\frac{1}{M_{0}}\frac{\partial Q^{i0}}{\partial\dot{\beta}^{a^{\prime}}}\,\,, (120)

in which the right-hand side is being evaluated at (β,β˙,β¨,s)=(β,0,0,s0)(\beta,\dot{\beta},\ddot{\beta},s)=(\beta,0,0,s_{0}). It is then clear that linear momenta and tensions are “thrusting” the swimming body.

When the body goes along a curve γ\gamma in configuration space, the total displacement Δzi\Delta z^{i} is given by

Δzi=γδzi=γλai(β)𝑑βa.\Delta z^{i}=\int_{\gamma}\delta z^{i}=\int_{\gamma}\lambda^{i}_{\,\,a^{\prime}}(\beta)d\beta^{a^{\prime}}\,\,. (121)

For a cyclic sequence of internal motions, we get

Δzneti=𝒟d(λai(β)dβa)=𝒟λbiβa𝑑βadβb,\Delta z^{i}_{net}=\int_{\mathcal{D}}d\left(\lambda^{i}_{\,\,a^{\prime}}(\beta)d\beta^{a^{\prime}}\right)=\int_{\mathcal{D}}\frac{\partial\lambda^{i}_{\,\,b^{\prime}}}{\partial\beta^{a^{\prime}}}\,d\beta^{a^{\prime}}\!\!\wedge d\beta^{b^{\prime}}\,\,, (122)

where 𝒟\mathcal{D} is a 2-dimensional surface in \mathcal{I} enclosed by γ\gamma. If the cycle is a small parallelogram defined by the infinitesimal vectors uau^{a^{\prime}} and vav^{a^{\prime}} in \mathcal{I}, then formula (78) can be used,

Δzneti2aλbiu[avb],\Delta z^{i}_{net}\approx 2\partial_{a^{\prime}}\lambda^{i}_{\,\,b^{\prime}}u^{[{a^{\prime}}}v^{b^{\prime}]}\,\,, (123)

which is the ultimate formula of the theory of swimming (at low Reynolds number) in curved spacetime!

One should keep in mind that the approximations in (103) are not obviously valid in general, so one must verify if they are indeed satisfactory in the particular case of analysis. In order to show how certain terms dropped in (102) could be problematic, let us consider a hypothetical case in which one of the configuration coordinates is ll, representing the overall spatial size of the body. Assume that the mass distribution of the body contributes to ψ\psi according to

ψmass=Al2r2+Bl3r3+,\psi_{mass}=A\frac{l^{2}}{r^{2}}+B\frac{l^{3}}{r^{3}}+\cdots\,\,, (124)

where AA and BB are functions of the other configuration coordinates, and they are assumed to have the same order of magnitude kM\sim\!kM. The triple dots represent mass-like terms that do not depend on ll. Similarly, let the contribution from the linear momentum distribution to ψ\psi be

ψmomentum=Al2r2l˙+Bl3r3l˙+,\psi_{momentum}=A^{\prime}\frac{l^{2}}{r^{2}}\dot{l}+B^{\prime}\frac{l^{3}}{r^{3}}\dot{l}+\cdots\,\,, (125)

where AA^{\prime} and BB^{\prime} are other functions of the other configuration parameters, also of order of magnitude kM\sim\!kM. The triple dots here denote terms that are proportional to β˙\dot{\beta}’s other than l˙\dot{l}. The total ψ\psi is given by the sum of these two contributions

ψ=ψmass+ψmomentum.\psi=\psi_{mass}+\psi_{momentum}\,\,. (126)

Observe that in the approximations (103), for very short integration times ss, the term (ψ/β)χ0Δβ(\partial\psi/\partial\beta)_{\chi_{0}}\Delta\beta was supposedly negligible compared to (ψ/β˙)χβ˙(\partial\psi/\partial\dot{\beta})_{\chi}\dot{\beta}. In this model they become

ψβ|χ0Δβ=2Al0r2Δl+2Bl02r3Δl+\displaystyle\left.\frac{\partial\psi}{\partial\beta}\right|_{\chi_{0}}\Delta\beta=2A\frac{l_{0}}{r^{2}}\Delta l+2B\frac{l_{0}^{2}}{r^{3}}\Delta l+\cdots
ψβ˙|χβ˙=Al2r2l˙+Bl3r3l˙+,\displaystyle\left.\frac{\partial\psi}{\partial\dot{\beta}}\right|_{\chi}\dot{\beta}=A\frac{l^{2}}{r^{2}}\dot{l}+B\frac{l^{3}}{r^{3}}\dot{l}+\cdots\,\,, (127)

where l=l0+Δll=l_{0}+\Delta l. Hence, to the first order of approximation, (ψ/β)χ0Δβ(\partial\psi/\partial\beta)_{\chi_{0}}\Delta\beta differs from (ψ/β˙)χβ˙(\partial\psi/\partial\dot{\beta})_{\chi}\dot{\beta} by a factor Δl/ll˙s/l\Delta l/l\dot{l}\sim s/l. For small times ss, the first quantity is indeed negligible compared to the second; in general, we have only to ask for the condition

sl,s\ll l\,, (128)

which means that the time of the cycle should be much smaller than the spatial dimensions of the body. This is a case where the body cannot be centrally controlled, in “real time”, for that would require faster-than-light signals propagating across its extension - thus, that ingeniously-designed device proposed by Wisdom, coupled along the body, seems to be a good fit here (see beginning of section IV). Note, however, that this condition only ensures that (ψ/β)χ0Δβ(\partial\psi/\partial\beta)_{\chi_{0}}\Delta\beta is negligible compared to the lowest order approximation of (ψ/β˙)χβ˙(\partial\psi/\partial\dot{\beta})_{\chi}\dot{\beta}. Consider, for instance, the first order expansion of (ψ/β˙)χβ˙(\partial\psi/\partial\dot{\beta})_{\chi}\dot{\beta} in ll, around l0l_{0},

ψβ˙|χβ˙ψβ˙|χ0β˙+β(ψβ˙|χ)|χ0β˙Δβ=\displaystyle\left.\frac{\partial\psi}{\partial\dot{\beta}}\right|_{\chi}\dot{\beta}\approx\left.\frac{\partial\psi}{\partial\dot{\beta}}\right|_{\chi_{0}}\dot{\beta}+\left.\frac{\partial}{\partial\beta}\left(\left.\frac{\partial\psi}{\partial\dot{\beta}}\right|_{\chi}\right)\right|_{\chi_{0}}\dot{\beta}\Delta\beta=
Al02r2l˙+Bl03r3l˙+2Al0r2l˙Δl+3Bl02r3l˙Δl+.\displaystyle A\frac{l_{0}^{2}}{r^{2}}\dot{l}+B\frac{l_{0}^{3}}{r^{3}}\dot{l}+2A\frac{l_{0}}{r^{2}}\dot{l}\Delta l+3B\frac{l_{0}^{2}}{r^{3}}\dot{l}\Delta l+\cdots\,\,. (129)

Although (ψ/β)χ0Δβ(\partial\psi/\partial\beta)_{\chi_{0}}\Delta\beta is much smaller than the first two terms on the right-hand side of the expression above by a small factor s/ls/l, the next two terms are much smaller than it by a factor l˙Δl/s1\dot{l}\sim\Delta l/s\ll 1 (if Newtonian conditions are assumed). Therefore, assuming no other term in (102) is problematic, the resulting motion would be of the geometric-phase kind only up to first order in Δβ\Delta\beta. In particular, the swimming formula (123) for closed cycles, being of order Δβ2\Delta\beta^{2}, would become unreliable – unless, of course, (ψ/β)χ0(\partial\psi/\partial\beta)_{\chi_{0}} vanishes. At order (l/r)2(l/r)^{2}, (ψ/β)χ0(\partial\psi/\partial\beta)_{\chi_{0}} is zero if

J¯νρσ[iβaRνρσ0]=0,\frac{\partial\bar{J}^{\nu\rho\sigma[i}}{\partial\beta^{a^{\prime}}}{R_{\nu\rho\sigma}}^{0]}=0\,\,, (130)

which is evaluated at β=β0\beta=\beta_{0} and β˙=0\dot{\beta}=0. Since only the mass distribution contributes to this, we may require simply

J¯j0k0βaRj0ki=0,\frac{\partial\bar{J}^{j0k0}}{\partial\beta^{a^{\prime}}}{R_{j0k}}^{i}=0\,\,, (131)

and, at order (l/r)3(l/r)^{3}, we must also ask for

2M0J¯νρσλ[iβa(3νRρσλ0]+2λRνρσ0])+\displaystyle 2M_{0}\frac{\partial\bar{J}^{\nu\rho\sigma\lambda[i}}{\partial\beta^{a^{\prime}}}\left(3\nabla_{\nu}{R_{\rho\sigma\lambda}}^{0]}+2\nabla_{\lambda}{R_{\nu\rho\sigma}}^{0]}\right)\,+
+S0ijJ¯νρσλβajRνρσλ=0,\displaystyle\qquad\qquad+S_{0}^{ij}\frac{\partial\bar{J}^{\nu\rho\sigma\lambda}}{\partial\beta^{a^{\prime}}}\nabla_{j}R_{\nu\rho\sigma\lambda}=0\,, (132)

which reduces to

S0ijJ¯k0l0βajRk0l0+2M0J¯m0j0iβa0Rm0ji\displaystyle S_{0}^{ij}\frac{\partial\bar{J}^{k0l0}}{\partial\beta^{a^{\prime}}}\nabla_{j}R_{k0l0}+2M_{0}\frac{\partial\bar{J}^{m0j0i}}{\partial\beta^{a^{\prime}}}\nabla_{0}{R_{m0j}}^{i}\,-
M0J¯mj0k0βa(3mRj0ki+2kRm(0j)i)=0.\displaystyle-\,M_{0}\frac{\partial\bar{J}^{mj0k0}}{\partial\beta^{a^{\prime}}}\left(3\nabla_{m}{R_{j0k}}^{i}+2\nabla_{k}{R_{m(0j)}}^{i}\right)=0\,\,. (133)

In order to have these conditions automatically fulfilled regardless of the body shape and internal motions, the curvature tensor and its first derivative have to satisfy Rijk0=0R_{ijk0}=0 (which is the only condition imposed at (l/r)2(l/r)^{2} order), μRijk0=0\nabla_{\mu}R_{ijk0}=0 and, if there is spin, jRk0l0=0\nabla_{j}R_{k0l0}=0.

In the general case, it seems that most of the terms in the right-hand side of (102) can also impair the validity of formula (123) at (l/r)3(l/r)^{3} order. Notwithstanding, even if that happens, the overall motion could still be swimming-like up to first order in the amplitudes Δβ\Delta\beta of the internal motions. Thus, formula (121) would still be reliable up to first order in the amplitudes, i.e.,

Δziλai(β0)Δβa\Delta z^{i}\approx\lambda^{i}_{\,\,a^{\prime}}(\beta_{0})\Delta\beta^{a^{\prime}} (134)

even at (l/r)3(l/r)^{3} order.

Let us briefly consider these other terms. The term (ϕi/βa)χ0Δβa(\partial\phi^{i}/\partial\beta^{a^{\prime}})_{\chi_{0}}\Delta\beta^{a^{\prime}} cannot cause problems since it only affects the swimming formulas at order Δβ3\Delta\beta^{3}. The term (ϕ~ai/s)β˙a(\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}/\partial s)\dot{\beta}^{a^{\prime}} could only cause trouble if the multipole moments were explicitly dependent on time, as can be seen by

ϕ~aisβ˙a=16M0[s(J¯jklmβ¨a)iRjklm+\displaystyle\frac{\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}}{\partial s}\dot{\beta}^{a^{\prime}}=\frac{1}{6M_{0}}\left[\frac{\partial}{\partial s}\left(\frac{\partial{\bar{J}}^{jklm}}{\partial\ddot{\beta}^{a^{\prime}}}\right)\nabla^{i}R_{jklm}+\right.
+J¯jklmβ¨as(iRjklm)]β˙a,\displaystyle\left.+\frac{\partial{\bar{J}}^{jklm}}{\partial\ddot{\beta}^{a^{\prime}}}\frac{\partial}{\partial s}\left(\nabla^{i}R_{jklm}\right)\right]\dot{\beta}^{a^{\prime}}\,\,, (135)

so that the first term above may affect the equations at the order k/r3k/r^{3} while the second can be discarded for it is proportional to a second derivative of the curvature tensor (supposedly of order k/r4k/r^{4}). If tension-related moments were explicitly time-dependent, then we would have to require

s(J¯jklmβ¨a)iRjklm=0,\frac{\partial}{\partial s}\left(\frac{\partial{\bar{J}}^{jklm}}{\partial\ddot{\beta}^{a^{\prime}}}\right)\nabla^{i}R_{jklm}=0\,\,, (136)

which is satisfied for arbitrary bodies only if iRjklm=0\nabla_{i}R_{jklm}=0.

The quantity (ϕi/β˙a)χβ˙a(\partial\phi^{i}/\partial\dot{\beta}^{a^{\prime}})_{\chi}\dot{\beta}^{a^{\prime}} tends to be non-vanishing in most cases. The condition for it to vanish is basically that there should be no linear momentum contribution to the quadrupole-order force, which translates as

J¯μνρσβ˙aiRμνρσ=0,\frac{\partial{\bar{J}}^{\mu\nu\rho\sigma}}{\partial\dot{\beta}^{a^{\prime}}}\nabla^{i}R_{\mu\nu\rho\sigma}=0\,\,, (137)

that reduces to

J¯0jklβ˙aiR0jkl=0,\frac{\partial{\bar{J}}^{0jkl}}{\partial\dot{\beta}^{a^{\prime}}}\nabla^{i}R_{0jkl}=0\,\,, (138)

which can be generally satisfied if iR0jkl=0\nabla_{i}R_{0jkl}=0. Note that it does not cause problems at order (l/r)2(l/r)^{2}.

The quantities (1/2)(2ϕi/β˙aβ˙b)χβ˙aβ˙b(1/2)(\partial^{2}\phi^{i}/\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}})_{\chi}\dot{\beta}^{a^{\prime}}\dot{\beta}^{b^{\prime}} and (ϕ~ai/βb)β˙aβ˙b(\partial\widetilde{\phi}^{i}_{\,\,a^{\prime}}/\partial\beta^{b^{\prime}})\dot{\beta}^{a^{\prime}}\dot{\beta}^{b^{\prime}} are associated with the tension contributions to the force. They vanish if

(122J¯νρσλβ˙aβ˙b2J¯νρσλβaβ¨b)iRνρσλ=0,\left(\frac{1}{2}\frac{\partial^{2}\bar{J}^{\nu\rho\sigma\lambda}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}-\frac{\partial^{2}\bar{J}^{\nu\rho\sigma\lambda}}{\partial\beta^{a^{\prime}}\partial\ddot{\beta}^{b^{\prime}}}\right)\nabla^{i}R_{\nu\rho\sigma\lambda}=0\,\,, (139)

which reduces to

(122J¯jklmβ˙aβ˙b2J¯jklmβaβ¨b)iRjklm=0.\left(\frac{1}{2}\frac{\partial^{2}\bar{J}^{jklm}}{\partial\dot{\beta}^{a^{\prime}}\partial\dot{\beta}^{b^{\prime}}}-\frac{\partial^{2}\bar{J}^{jklm}}{\partial\beta^{a^{\prime}}\partial\ddot{\beta}^{b^{\prime}}}\right)\nabla^{i}R_{jklm}=0\,\,. (140)

Thus, a general body can satisfy this if iRjklm=0\nabla_{i}R_{jklm}=0. Again, no problem is posed at (l/r)2(l/r)^{2} order.

Due to the many different conditions that we have obtained, it is appropriate to summarize them in a more organized manner. Note that we are mostly interested in the conditions that allow a general body to swim, despite of its geometrical shape or internal motions. So, we have found that swimming can be fully observed, up to order (l/r)2(l/r)^{2}, if only Rijk0=0R_{ijk0}=0. When investigating the dynamics at order (l/r)3(l/r)^{3}, however, it is possible that formula (121) could be correct only up to first order in the amplitudes Δβ\Delta\beta, so that formula (123) would be unreliable - in this case, we shall say that the body can quasi-swim. When there is no spin, lRijk0=0\nabla_{l}R_{ijk0}=0 is enough to ensure that the body can quasi-swim, while 0Rijk0=lRijkm=0\nabla_{0}R_{ijk0}=\nabla_{l}R_{ijkm}=0 is further required for a legitimate swimming. When there is spin, lRijk0=lRijkm=0\nabla_{l}R_{ijk0}=\nabla_{l}R_{ijkm}=0 is required for quasi-swimming, while jRk0l0=0\nabla_{j}R_{k0l0}=0 is additionally demanded for legitimate swimming (i.e., only the components 0Ri0j0\nabla_{0}R_{i0j0} could be non-vanishing). Recall that these components refer to the coordinate system associated with the initial conditions; covariantly, the condition Rijk0=0R_{ijk0}=0, for instance, would be translated as “the restriction of R(u,,,)R(u,\cdot,\cdot,\cdot) to the vector subspace orthogonal to uu at z(s0)z(s_{0}) must vanish”.

We reinforce that all these conditions are associated with the hypotheses on the configuration space and the form of the multipole moments mentioned right below equation (123). In any particular case one must carefully go through similar steps to verify if the approximations in (103) are indeed reasonable. Still, this model seems to apply to a handful of interesting scenarios such as, to name a few, a bipod on a sphere (appendix B) and a tripod in Schwarzschild [3] or FRW spacetime.

VI Discussion

The main purpose of this paper was to provide a rigorous and comprehensive treatment of the phenomenon of swimming (at low Reynolds number) in curved spacetime. We found that this effect can indeed take place, although only in very special circumstances where the spacetime abounds with symmetries and the initial conditions and body configurations are favorable. The framework here developed, deeply rooted in Dixon’s multipole theory of dynamics, is not limited to the “low Reynolds number” regime, being suitable to describe the dynamics of any free small articulated body performing fast cyclic internal motions in general relativity.

The choice of Dixon’s formalism as the starting point was due to some of its important features. Namely, the fact that (i)(i) the fundamental conservation equation for matter in general relativity, aTab=0\nabla_{a}T^{ab}=0, is automatically implemented in the theory, (ii)(ii) the multipole structure of the equations allow a controlled separation of the dynamical effects at different orders in l/rl/r and (iii)(iii) it provides a reasonable definition for the worldline of the center of mass, which can be used to track the “effective” motion of the body.

One important aspect of our approach was the special manner proposed to prescribe the quadrupole and octupole moments. More precisely, we have shown that, when solving the multipole equations up to (l/r)3(l/r)^{3} order of precision, the relevant moments could be evaluated as if the background spacetime were flat, using the exponential map to establish the connection to the real (curved) spacetime. Moreover, the dynamics of the body could be described by the equations of special relativity, namely the energy-momentum tensor would satisfy the conservation equation aTab=0\partial_{a}T^{ab}=0, where \partial is the derivative operator associated with any inertial frame of the flat spacetime. This means that, under the (l/r)3(l/r)^{3} approximation, the problem of solving the dynamics in general relativity can be formally reduced to solving an analogous problem in special relativity. Obviously this amounts to a great simplification.

Considerable attention was devoted to the special case of articulated bodies, with finite-dimensional internal configuration spaces, \mathcal{I}, performing a predetermined routine of internal motions, γ:\gamma:\mathbb{R}\rightarrow\mathcal{I}. We found that, for small integration times and in the Newtonian regime for the internal movements, the effective motion can be described by a generalized concept of swimming. In this generalization, the deviation of dynamical quantities, with respect to the rigid body solution, are functions of the image of the (canonical) lift of γ\gamma to the tangent space of the configuration space, TT\mathcal{I}. Although this kind of motion is relatively simple, it does not display the defining feature of the traditional swimming at low Reynolds numbers. Indeed, the overall motion depends on the rapidity of the internal motions. It is interesting to recall that the Newtonian regime condition comes to ensure that the β¨\ddot{\beta}-dependent part of the force does not depend on the “velocities” β˙\dot{\beta}; therefore, this condition is not required at order (l/r)2(l/r)^{2}, possibly making this form of swimming a fairly general effect at this order of approximation (as long as certain terms can be neglected in the limit of fast cycles).

When there are no tension contributions to the torque, then it might be possible to swim (or quasi-swim) in spacetime. This is a very strong condition, not often satisfied. Moreover, this condition depends on initial values such as the initial dynamical velocity and spin. This can be understood by noting that swimming is not possible in Newtonian gravitation for only the mass content of the body interacts with gravity, so that speeding up the cycles of internal motions implies that the gravitational impulse becomes smaller, influencing the resulting motion. In general relativity the linear momentum and tension densities also “feel” the gravitational field and, since they increase for faster cycles, they can compensate for the shorter cycle duration in such a way that the gravitational impulse becomes independent of its duration – this is swimming at low Reynolds number. However, tension contributions to the torque tend to “overcompensate” for these shorter cycle times, in the sense that, at some point the impulse starts to increase for faster cycles and the resulting motion becomes dependent again on the rapidity of the internal motions.

Although swimming (at low Reynolds number) in curved spacetime is indeed a phenomenon of limited occurrence, there are some important spacetimes where swimming (or quasi-swimming) can take place. Namely, it happens in Friedmann-Robertson-Walker spacetimes, if the body starts at rest with respect to the family of isotropic observers. It is worth to point out that this scenario, where the spatial sections are homogeneous and isotropic, represents the closest analogue to the Newtonian swimmer in a non-turbulent viscous fluid. Indeed, the fluid mechanics theory of swimming at low Reynolds number, in which the effective motion is described by geometric phases associated with the sequence of shapes assumed by the body, is not expected to apply to a non-homogeneous (or non-isotropic) fluid, such as a fluid with non-constant density or viscosity.

In conclusion, the dynamics of small articulated bodies performing fast internal motions in curved spacetimes is much richer than originally expected, as the effective motion can respond differently to different rapidities of the internal movements. Accordingly, we propose that the term swimming in curved spacetime should not be restricted to this limited sense of “low Reynolds number” regime, in the same way that the term swimming in the context of fluids is not. Finally, we pose a possibly interesting question: what can a spacetime swimmer do to become more efficient? That is, what is the regime, shape and internal motions required to achieve the maximum displacement per cycle? This is just one example of stimulating questions that may arise in this field of swimming in curved spacetime.

Acknowledgements.
I would like to thank George Matsas and Daniel Vanzella for helpful discussions, and also for comments on the first version of this paper. I acknowledge full financial support from São Paulo Research Foundation (FAPESP) under Grant No. 2015/10373-4 during the time this work was produced, and full financial support from the University of Maryland, College Park during the time this paper was written.

Appendix A Definition of relevant bitensors

In section (IV) we presented formula (16) giving the multipole moments in terms of the stress-energy tensor. In this appendix we provide the definitions for the bitensors appearing in that formula.

We define the world-function biscalar σ\sigma by

σ(x1,x2)=12(u2u1)u1u2gabvavb𝑑u,\sigma(x_{1},x_{2})=\frac{1}{2}(u_{2}-u_{1})\int^{u_{2}}_{u_{1}}g_{ab}v^{a}v^{b}du\,\,, (141)

where x(u)x(u) is an affine-parametrized geodesic linking x1=x(u1)x_{1}=x(u_{1}) and x2=x(u2)x_{2}=x(u_{2}), and vav^{a} is the tangent vector field to it. Note that, in the hypothesis of normal neighborhood, there is a single geodesic linking each two points, making this bi-scalar well-defined.

We represent higher-rank bitensors, based on the pair of points (x,z)(x,z), using a convention for the indices similar to the abstract index notation for tensors. The difference is that we reserve the first half of the Latin alphabet (“a,b,c,a,b,c,\ldots”) to denote slots associated with the point xx (i.e., the first entry) and the second half of the alphabet (“k,l,m,k,l,m,\ldots”) to denote slots associated with the point zz (i.e., the second entry). Components associated with xx will be denoted by letters in the first half of the Greek alphabet (“α,β,γ,\alpha,\beta,\gamma,\ldots”) and those associated with zz by letters in the second half (“κ,λ,μ,\kappa,\lambda,\mu,\ldots”). We can act with covariant derivatives on bitensors by keeping the non-involved point fixed (and also ignoring the tensor indices associated with that point). Whenever the argument of bitensors are omitted, it is implicit that they are (x,z)(x,z).

When covariant derivatives act on σ\sigma we denote it as, for example, abkσ=σkba\nabla_{a}\nabla_{b}\nabla_{k}\sigma=\sigma_{kba}. In particular, we have σa\sigma^{a} and σk\sigma^{k}. Also, when a second derivative acts on σk\sigma^{k}, now with respect to the point xx, we get σak{\sigma}^{k}_{\>\>{a}}. This bitensor can be seen as a (linear) mapping of vectors at xx to vectors at zz, i.e., a map of the kind TxTzT_{x}\mathcal{M}\rightarrow T_{z}\mathcal{M} (if vaTxv^{a}\in T_{x}\mathcal{M}, then σakvaTz{\sigma}^{k}_{\>\>{a}}v^{a}\in T_{z}\mathcal{M}). In a region such that every point is in the normal neighborhood of the others, this is an invertible map, and we define the inverse map σ^ka{\widehat{\sigma}}^{a}_{\>\>k} by

σ^kaσbk=δba,{\widehat{\sigma}}^{a}_{\>\>k}\sigma^{k}_{\>\>b}=\delta^{a}_{\>\>b}\,\,, (142)

where δba\delta^{a}_{\>\>b} is the identity tensor at xx. We also define the transport bitensors KK and HH by

Kka=σ^laσkl\displaystyle K^{a}_{\>\>k}=-{\widehat{\sigma}}^{a}_{\>\>l}\sigma^{l}_{\>\>k}
Hka=σ^ka,\displaystyle H^{a}_{\>\>k}=-{\widehat{\sigma}}^{a}_{\>\>k}\,, (143)

which can be thought as (linear) mappings of vectors at zz to vectors at xx (i.e., maps of the kind TzTxT_{z}\mathcal{M}\rightarrow T_{x}\mathcal{M}). Finally, define the nn-potential function, for n2n\geq 2, as

nklqr(x,z)=(n1)01σkaσa(qσbr)σlbun2𝑑u,{\langle n\rangle}^{klqr}(x,z)=(n-1)\int_{0}^{1}\sigma^{ka}{\sigma}^{(q}_{\>\>{\,\,a}}{\sigma}^{r)}_{\>\>{\,b}}\sigma^{lb}u^{n-2}du\,, (144)

where the arguments of the σ\sigma’s are (x¯(u),z)(\overline{x}(u),z), with x¯(u)\overline{x}(u) being a geodesic with affine parameter uu such that x¯(0)=z\overline{x}(0)=z and x¯(1)=x\overline{x}(1)=x.

Appendix B A bipod on a sphere

In this appendix we analyze the problem of a bipod constrained to move on a 2-sphere. This problem was one of the major motivations behind Wisdom’s realization that swimming (at low Reynolds number) in the curved spacetimes of general relativity could be possible. We first review the Newtonian results, and then we apply the covariant theory of swimming to the equivalent problem in general relativity, showing that the answers match. Hence, this serves as a check of consistency for our theory.

A bipod is set to move freely (i.e., without friction) on a 2-sphere of radius rr, which will be covered with spherical coordinates {ϕ,ψ}\{\phi,\psi\}, where ϕ=[0,π]\phi=[0,\pi] and ψ=[0,2π)\psi=[0,2\pi). The bipod is formed by a central mass m0m_{0} placed (without loss of generality) at the point (π/2,0)(\pi/2,0), connected by two straight (geodesic) massless rods of length ll to two other identical masses m1=m2m_{1}=m_{2}. The bipod is allowed to stretch and shrink its legs or open and close the angle 2α2\alpha between the legs at m0m_{0}; hence, the associated internal configuration space \mathcal{I} is spanned by the coordinates {α,β}\{\alpha,\beta\}, where β=l/r\beta=l/r.

Let the axis of symmetry (bisectrix of the bipod) be initially aligned with the ψ\psi direction; thence, because of the symmetry of the bipod and the sphere, the effective motion of the bipod will necessarily be along the ψ\psi direction, and the bisectrix direction will not tilt. Therefore, the only meaningful dynamical variable is the ψ\psi coordinate of, say, the central mass. The (Newtonian) differential displacement141414Recall that the differential displacement 1-form is defined from the equation satisfied by the trajectory deviation, Δψ(t)=ψ(t)ψ0(t)\Delta\psi(t)=\psi(t)-\psi_{0}(t), where ψ0(t)\psi_{0}(t) is the trajectory of the rigid bipod. Namely, if Δψ˙=Al˙+Bα˙\Delta\dot{\psi}=A\,\dot{l}+B\,\dot{\alpha}, then δψ=Adl+Bdα\delta\psi=A\,dl+B\,d\alpha. is given by

δψ=2m1sinβcosβsinαm0+2m1(cos2β+sin2βcos2α)dα\displaystyle\delta\psi=\frac{2m_{1}\sin\!\beta\cos\!\beta\sin\!\alpha}{m_{0}+2m_{1}(\cos^{2}\!\beta+\sin^{2}\!\beta\cos^{2}\!\alpha)}\,d\alpha
2m1cosαm0+2m1(cos2β+sin2βcos2α)dβ.\displaystyle-\frac{2m_{1}\cos\!\alpha}{m_{0}+2m_{1}(\cos^{2}\!\beta+\sin^{2}\!\beta\cos^{2}\!\alpha)}\,d\beta\,. (145)

After a closed cycle of internal motions, the net additional displacement is

Δψ=𝒟[4m0m1sinαsin2β(m0+2m1(cos2α+cos2βsin2α))2]𝑑βdα.\Delta\psi=\int_{\mathcal{D}}\left[-\frac{4m_{0}m_{1}\sin\!\alpha\sin^{2}\!\beta}{(m_{0}+2m_{1}(\cos^{2}\!\alpha+\cos^{2}\!\beta\sin^{2}\!\alpha))^{2}}\right]\,d\beta\wedge d\alpha\,\,. (146)

where 𝒟\mathcal{D} is the region in \mathcal{I} enclosed by the cycle. Thus, considering the rectangular path (β,α)(β+Δβ,α)(β+Δβ,α+Δα)(β,α+Δα)(β,α)(\beta,\alpha)\rightarrow(\beta+\Delta\beta,\alpha)\rightarrow(\beta+\Delta\beta,\alpha+\Delta\alpha)\rightarrow(\beta,\alpha+\Delta\alpha)\rightarrow(\beta,\alpha) in the configuration space, with Δβ\Delta\beta and Δα\Delta\alpha small amplitudes, we can approximate

Δψ4m0m1sinαsin2β(m0+2m1(cos2α+cos2βsin2α))2ΔβΔα.\Delta\psi\approx-\frac{4m_{0}m_{1}\sin\!\alpha\sin^{2}\!\beta}{(m_{0}+2m_{1}(\cos^{2}\!\alpha+\cos^{2}\!\beta\sin^{2}\!\alpha))^{2}}\,\Delta\beta\Delta\alpha\,\,. (147)

For a small bipod (l/r1l/r\ll 1), the Taylor expansion in β\beta yields

Δψ4m0m1sinα(m0+2m1)2β2ΔβΔα+\displaystyle\Delta\psi\approx-\frac{4m_{0}m_{1}\sin\!\alpha}{(m_{0}+2m_{1})^{2}}\,\beta^{2}\Delta\beta\Delta\alpha+
4m0m1(m0+12m1cos2α10m1)sinα3(m0+2m1)3β4ΔβΔα+,\displaystyle\frac{4m_{0}m_{1}\left(m_{0}+12m_{1}\cos^{2}\!\alpha-10m_{1}\right)\sin\!\alpha}{3(m_{0}+2m_{1})^{3}}\,\beta^{4}\Delta\beta\Delta\alpha+\cdots\,\,, (148)

so that the lowest order effect is at order (l/r)2(l/r)^{2} and there is no correction at order (l/r)3(l/r)^{3}. Hence, we expect the result from the covariant theory of swimming to be

Δψ=4m0m1sinα(m0+2m1)2β2ΔβΔα.\Delta\psi=-\frac{4m_{0}m_{1}\sin\!\alpha}{(m_{0}+2m_{1})^{2}}\,\beta^{2}\Delta\beta\Delta\alpha\,. (149)

Fortunately, this is precisely what we will obtain.

Let us now look at this problem as the limiting case of a slowly moving bipod in the context of general relativity. The simplest thing that one can think of is to consider a spacetime with topology ×S2\mathbb{R}\times S^{2} and metric ds2=dt2r2(dϕ2+sin2ϕdψ2)ds^{2}=dt^{2}-r^{2}(d\phi^{2}+\sin^{2}\!\phi\,d\psi^{2}), where tt\in\mathbb{R}, ϕ[0,π]\phi\in[0,\pi] and ψ[0,2π)\psi\in[0,2\pi). However, in order to avoid going through the necessary modifications in our formulas associated with the different spacetime dimension, we shall have the bipod constrained in a spatial 2-sphere inside a 1+3 spacetime.

In the Newtonian case, there was no issue in considering the bipod to be held on the sphere by normal forces of constraint which could, for instance, be of contact nature: the Lagrangian formalism deals well with this. The Dixon’s formalism, on the other hand, provides no simple way of addressing constraints like this. The solution to this situation is to realize that the swimming of the bipod on the sphere should not depend on the nature of the forces of constraint. Thus, we might search for a spacetime in which this force is of gravitational nature, allowing us to directly apply Dixon’s formalism (as the body would be essentially free in this case). The natural candidate for the spacetime is one with topology ×S3\mathbb{R}\times S^{3} and metric

ds2=dt2r2(dθ2+sin2θdϕ2+sin2θsin2ϕdψ2),ds^{2}=dt^{2}-r^{2}(d\theta^{2}+\sin^{2}\!\theta\,d\phi^{2}+\sin^{2}\!\theta\sin^{2}\!\phi\,d\psi^{2})\,\,, (150)

where θ[0,π]\theta\in[0,\pi], ϕ[0,π]\phi\in[0,\pi] and ψ[0,2π)\psi\in[0,2\pi). Note that the submanifold θ=π/2\theta=\pi/2 is clearly diffeomorphic to the spacetime ×S2\mathbb{R}\times S^{2} mentioned above. Moreover, it has the convenient property that any geodesic tangent to it, at any point, lies entirely within it (intuitively, this submanifold is a sheaf of geodesics). This suggests that if the bipod is initially placed on this submanifold, and all the internal motions are tangent to it, then the body will not “take off” and leave the submanifold as it evolves (this can be confirmed by analysis of the equations of motion).

The only non-vanishing components of the curvature tensor on ×S3\mathbb{R}\times S^{3}, in the orthonormal basis eμe_{\mu} aligned with the coordinates (t,θ,ϕ,ψ)(t,\theta,\phi,\psi), restricted to the submanifold ×S2\mathbb{R}\times S^{2}, are given by

R1212=R1313=R2323=1r2,R_{1212}=R_{1313}=R_{2323}=\frac{1}{r^{2}}\,, (151)

and all the derivatives of the curvature tensor vanish everywhere, aRbcde=0\nabla_{a}R_{bcde}=0. This means that the Dixon’s equations become exact, with the force and torque given by

Fk=12RlqrkvlSqr\displaystyle F^{k}=\frac{1}{2}R^{k}_{\>\>lqr}v^{l}S^{qr}
Qkl=43Jpqr[kRpqrl],\displaystyle Q^{kl}=\frac{4}{3}J^{pqr[k}{R_{pqr}}^{l]}\,\,, (152)

Although the equations of motions are simple, the exact evaluation of the quadrupole moment can still be very complicated. Hence, let us assume that the body is sufficiently small compared to the radius (i.e., lrl\ll r), so that we can use the framework developed in this paper to obtain final results reliable up to (l/r)3(l/r)^{3} order.

According to formula (152), the force is orthogonal to the dynamical velocity, implying that the proper mass MM is always conserved. From an analysis of the conserved quantities (see item (vi)(\text{vi}) in section II) we find that, if the bipod is released from rest and with no spin, then the linear and angular momentum will not vary with time, that is,

pk(s)=M(e0)k\displaystyle p^{k}(s)=M(e_{0})^{k}
Skl(s)=0\displaystyle S^{kl}(s)=0 (153)

for all ss. Therefore the equation for the linear momentum becomes trivial, the one for the angular momentum yields

e0[avb]=12MQabe_{0}^{[a}v^{b]}=-\frac{1}{2M}Q^{ab} (154)

and equation (3) is automatically satisfied. In components associated with the eμe_{\mu} basis, we get

vi=1MQi0,v^{i}=\frac{1}{M}Q^{i0}\,, (155)

where ii denotes a “spatial” index. This can be integrated to yield the trajectory of the center of mass of the body.

According to (152), the non-vanishing components of the torque read

Q10=4v03r2(J2120+J3130)\displaystyle Q^{10}=\frac{4v^{0}}{3r^{2}}\left(J^{2120}+J^{3130}\right)
Q20=4v03r2(J1210+J3230)\displaystyle Q^{20}=\frac{4v^{0}}{3r^{2}}\left(J^{1210}+J^{3230}\right)
Q30=4v03r2(J1310+J2320),\displaystyle Q^{30}=\frac{4v^{0}}{3r^{2}}\left(J^{1310}+J^{2320}\right)\,\,, (156)

recalling that, in our approximations, v01v^{0}\approx 1. Thus, only the momentum-related components of the quadrupole moment, Jiji0=d3xxix[igj]J^{iji0}=\int\!d^{3}\!x\,x^{i}x^{[i}g^{j]}, are relevant. Since the bipod is contained to the 2-3 “plane” (i.e., the 2-sphere), then J1j10=0J^{1j10}=0; also, since there is no linear momentum in a direction orthogonal to the 2-sphere, then Ji1i0=0J^{i1i0}=0. Therefore, the torque components simplify to

Q10=0\displaystyle Q^{10}=0
Q20=4v03r2J3230\displaystyle Q^{20}=\frac{4v^{0}}{3r^{2}}J^{3230}
Q30=4v03r2J2320.\displaystyle Q^{30}=\frac{4v^{0}}{3r^{2}}J^{2320}\,\,. (157)

Furthermore, because of the inversion symmetry with respect to the bisectrix of the bipod, one should have J3230=Jklmn(e3)k(e2)l(e3)m(e0)n=Jklmn(e3)k(e2)l(e3)m(e0)n=0J^{3230}=J^{klmn}(e_{3})_{k}(e_{2})_{l}(e_{3})_{m}(e_{0})_{n}=J^{klmn}(e_{3})_{k}(-e_{2})_{l}(e_{3})_{m}(e_{0})_{n}=0, so that the only non-vanishing components of the torque are Q30=Q03Q^{30}=-Q^{03}. This implies that the bipod moves along the great circle of the 2-sphere (i.e., the worldline of the center of mass lies in the tt-ψ\psi submanifold), as expected. Parameterizing the worldline by the coordinate time tt, we obtain from (155) the simple formula

dψdt=43J2320Mr3.\frac{d\psi}{dt}=\frac{4}{3}\frac{J^{2320}}{Mr^{3}}\,\,. (158)

Note that this result is correct up to (l/r)3(l/r)^{3} order, and all the relativistic effects (e.g., those coming from very fast internal movements) must be encoded in J2320J^{2320}.

Let us now compute this quadrupole moment. We shall assume that the internal motions are sufficiently slow so that Newtonian mechanics apply. This is motivated by the general theory, which indicates that swimming (at low Reynolds number) is mainly limited to this regime.

Refer to caption
Figure 3: The local correspondence to a flat spacetime puts the bipod on an Euclidean plane. The coordinate x2x^{2} is aligned with the actual direction e2e_{2}, in the real spacetime, and x3x^{3} is aligned with the actual direction e3e_{3}.

The total mass is given by M=m0+2m1M=m_{0}+2m_{1}, and the momentum gjg^{j} is

gj(x)=ρ(x)ϑj(x),g^{j}(x)=\rho(x)\vartheta^{j}(x)\,, (159)

where ϑj(x)\vartheta^{j}(x) is the 3-velocity of the element of mass at the point xx. To evaluate the component J2320J^{2320}, we consider that the bipod is in an Euclidean 2-dimensional plane, covered with coordinates {x2,x3}\{x^{2},x^{3}\}, such that x2x^{2} is in e2e_{2} direction (i.e., ϕ\phi direction) and x3x^{3} is in e3e_{3} direction (i.e., ψ\psi direction) [see figure (3)]. Let the coordinates of the central mass, m0m_{0}, be (x02,x03)=(0,x0)(x_{0}^{2},x_{0}^{3})=(0,x_{0}), where x0x_{0} is some number. The coordinates of the mass m1m_{1}, at an arbitrary configuration, are given by (x12,x13)=(lsinα,x0+lcosα)(x_{1}^{2},x_{1}^{3})=(l\sin\!\alpha,x_{0}+l\cos\!\alpha) and those of the mass m2m_{2} are (x22,x23)=(lsinα,x0+lcosα)(x_{2}^{2},x_{2}^{3})=(-l\sin\!\alpha,x_{0}+l\cos\!\alpha). In order to have the center of mass at the origin of the coordinate system (which is essential to the definition of the quadrupole moment), we must set x0x_{0} as

x0=2m1m0+2m1lcosα.x_{0}=-\frac{2m_{1}}{m_{0}+2m_{1}}l\cos\!\alpha\,. (160)

The velocities x˙ki\dot{x}_{k}^{i} can be easily computed and the density of momentum can be expressed as

gi(x)=kmkδ(xxk)x˙ki,g^{i}(x)=\sum_{k}m_{k}\delta(x-x_{k})\dot{x}_{k}^{i}\,\,, (161)

where δ\delta is the Dirac delta function. We find the component J2320J^{2320} to be

J2320=m0m1m0+2m1l3sinαα˙,J^{2320}=-\frac{m_{0}m_{1}}{m_{0}+2m_{1}}l^{3}\sin\!\alpha\,\dot{\alpha}\,, (162)

so that

dψdt=43r3m0m1(m0+2m1)2l3sinαdαdt.\frac{d\psi}{dt}=-\frac{4}{3r^{3}}\frac{m_{0}m_{1}}{(m_{0}+2m_{1})^{2}}l^{3}\sin\!\alpha\,\frac{d\alpha}{dt}\,\,. (163)

When the bipod is rigid (i.e., l˙=0\dot{l}=0 and α˙=0\dot{\alpha}=0), the center of mass does not move, ψ0(t)=ψ(0)\psi_{0}(t)=\psi(0). Thus the differential displacement δψ\delta\psi is

δψ=43r3m0m1(m0+2m1)2l3sinαdα.\delta\psi=-\frac{4}{3r^{3}}\frac{m_{0}m_{1}}{(m_{0}+2m_{1})^{2}}l^{3}\sin\!\alpha\,d\alpha\,\,. (164)

Note that it differs from the corresponding 1-form (given in equation (145)) of the Newtonian approach, and this is because ψ\psi there denoted the position of the pivot mass while here it denotes the position of the center of mass. We expect, nonetheless, that after a complete cycle of internal motions the two distinct definitions for the net deviations should match – after all, the distance between the pivot and the center of mass depends only on the shape of the bipod, which is the same before or after a cycle. After a complete cycle (of infinitesimal internal motions) the total deviation is

Δψ=4m0m1(m0+2m1)2β2sinαΔβΔα,\Delta\psi=-\frac{4m_{0}m_{1}}{(m_{0}+2m_{1})^{2}}\beta^{2}\sin\!\alpha\,\Delta\beta\Delta\alpha\,\,, (165)

which is precisely the correct result, up to (l/r)3(l/r)^{3} order of approximation!

This result can be quickly re-obtained if we apply the swimming formula (123). The differential displacement is given by δψ=λa3dβa\delta\psi=\lambda^{3}_{\,\,a^{\prime}}\,d\beta^{a^{\prime}}, where (β1,β2)=(α,l)(\beta^{1},\beta^{2})=(\alpha,l). The function λa3\lambda^{3}_{\,\,a^{\prime}} is explicitly defined in (120). Since there is no force, it becomes simply

λa3=1M0Q30β˙a=43M0r2J2320β˙a,\lambda^{3}_{\,\,a^{\prime}}=\frac{1}{M_{0}}\frac{\partial Q^{30}}{\partial\dot{\beta}^{a^{\prime}}}=\frac{4}{3M_{0}r^{2}}\frac{\partial J^{2320}}{\partial\dot{\beta}^{a^{\prime}}}\,\,, (166)

where M0=m0+2m1M_{0}=m_{0}+2m_{1}. Using expression (162) for J2320J^{2320} in terms of ll and α\alpha, we get

λα3=43r3m0m1(m0+2m1)2l3sinα\displaystyle\lambda^{3}_{\,\,\alpha}=-\frac{4}{3r^{3}}\frac{m_{0}m_{1}}{(m_{0}+2m_{1})^{2}}l^{3}\sin\!\alpha (167)
λl3= 0,\displaystyle\lambda^{3}_{\,\,l}\,=\,0\,\,, (168)

which gives precisely the δϕ\delta\phi in (164).

Finally, it interesting to note that this swimming behavior can only be observed when the bipod is released from rest (with respect to the time symmetry). When uu has an initial component in, say, direction ψ\psi, the torque components become

Q10=0\displaystyle Q^{10}=0
Q20=ukvk4u03r2[u0J¯0323u3J¯0203]\displaystyle Q^{20}=u^{k}v_{k}\frac{4u^{0}}{3r^{2}}\Big{[}u^{0}\bar{J}^{0323}-u^{3}\bar{J}^{0203}\Big{]}
Q30=ukvk43r2[[(u0)2+(u3)2]J¯0232+.\displaystyle Q^{30}=u^{k}v_{k}\frac{4}{3r^{2}}\Big{[}\left[(u^{0})^{2}+(u^{3})^{2}\right]\bar{J}^{0232}+\Big{.}
.+u0u3(J¯0202+J¯2323)].\displaystyle\Big{.}\qquad\qquad+u^{0}u^{3}\left(\bar{J}^{0202}+\bar{J}^{2323}\right)\Big{]}\,. (169)

Since there is no torque propulsion in direction e1e_{1}, the body still does not leave the great sphere, as expected. Particularizing to the case of the symmetric bipod, the torque component Q20Q^{20} vanishes, showing that the bipod only moves along its symmetry axis (i.e., in ψ\psi direction). Clearly these torque components reduce to those in (157) when u=e0u=e_{0}. If uu has a non-negligible u3u^{3} projection, then we see that the torque component Q30Q^{30} contains a tension contribution coming form J¯2323\bar{J}^{2323}. Thus, if a bipod is set to move relativistically fast on a frictionless 2-sphere, it will not swim “at low Reynolds number”, unlike what happens in the Newtonian regime.

References

  • [1] J. Wisdom, Swimming in Spacetime: Motion by Cyclic Changes in Body Shape, Science 299, 1865 (2003).
  • [2] A. Shapere and F. Wilczek, Geometry of self-propulsion at low Reynolds number, J. Fluid Mech. 198, 557 (1989).
  • [3] Rodrigo Andrade e Silva, George E. A. Matsas and Daniel A. T. Vanzella, Rescuing the concept of swimming in curved spacetime, Phys. Rev. D 94, 121502 (2016).
  • [4] W. G. Dixon, Dynamics of extended bodies in general relativity I, Momentum and angular momentum, Proc. R. Soc. London, Ser. A 314, 499 (1970).
  • [5] W. G. Dixon, Dynamics of extended bodies in general relativity II, Moments of the charge-current vector, Proc. R. Soc. London, Ser. A 319, 509 (1970).
  • [6] W. G. Dixon, Dynamics of extended bodies in general relativity III, Equations of Motion, Philos. Trans. R. Soc. London, Ser. A 277, 59 (1974).
  • [7] A. I. Harte, Extended-body motion in black hole spacetimes: What is possible?, Physical Review D 102.12, 124075 (2020)
  • [8] A. I. Harte and M. T. Gaffney, Extended-body effects and rocket-free orbital maneuvering, Acta Astronautica 178, 625-633 (2021)
  • [9] W. Beiglböck, The center-of-mass in Einsteins Theory of gravitation, Commun. Math. Phys. 5, 106 (1967).
  • [10] R. Andrade e Silva, A. G. S. Landulfo, G. E. A. Matsas, D. A. T. Vanzella, Relativistic spring-mass system, arxiv:1810.13365
  • [11] W. G. Dixon, The definition of multipole moments for extended bodies, Gen. Rel. Grav. 4, 199 (1973).
  • [12] L. C. Brewin, Riemann normal coordinates, Monash U. Publishing (1997).