<<

arXiv:2012.03192v2 [gr-qc] 25 Jan 2021 us-peia ae n,i h us-lnradquasi-h be. and to quasi-planar happens the equivalent in and, case, quasi-spherical ilytesm hsclrls u hr r he oearbi coordinate (hyperboloid more the pseudo-spherical three on are or pend there planar, but spherical, roles, from physical same the Lemaˆıtre-Tolman-Ellis metric tially the of functions arbitrary unders physical proper descr a fuller to a indispensable scales, is of range for this a potential on much mos it structures the gives cosmological of which one equations, is field it Einstein’s because precisely Still, understanding. h td fti erchsbe iie yisrltv comp relative being its still by yet limited with and been models, has hand, hyperboloidal metric one and this the of on study vectors the cosmologi Killing inhomogeneous no Szekeres having the discovery, its Since Work Previous 1 ∗ et fPyisadAtooy nvriyo ea tSnAnton San at Texas of University Astronomy, and Physics of Dept. † zkrsmdl a evee sdsotoso Lemaˆıtre-T of distortions as viewed be can models Szekeres [email protected] [email protected] et fMts n ple ah,Uiest fCp on Ro Town, Cape of University Maths, Applied and Maths. of Dept. oaineet elb hwdta h angular the that showed Hellaby effect. rotation meddtrstplg.Svrlepii oesilsrt t agre illustrate complete effects. models direc in tilt explicit ‘radial’ are and the Several results rotation in of topology. closed sets torus are two that embedded the models q that Szekeres the construct show for to to em embedding, is it this 3-space use of the when properties and tilt some of investigate rate here the as We well ex as explicit shell, provided to shell Schlegel and Buckley while orientation, oain medn n oooyfrteSzekeres the for Topology and Embedding Rotation, eetwr nteSeee nooeeu omlgclmodel cosmological inhomogeneous Szekeres the on work Recent r hc sasr f‘ail oriaelblfrcmvn 3-s comoving for label coordinate ‘radial’ of sort a is which , ǫ +1 6= oetG Buckley G. Robert ako etna nlg rmwiht eiephysical derive to which from analogy Newtonian a of lack , hre Hellaby Charles Geometry Abstract 1 ( peri h zkrsmti lyn essen- playing metric Szekeres the in appear s ,φ θ, l ymty l rirr ucin de- functions arbitrary 6 All symmetry. al) proodlcss htvrteappropriate the whatever cases, yperboloidal tanding. of solutions exact inhomogeneous realistic t pino h emtyadteeouinof evolution the and geometry the of iption ) plcto nmdligrltvl complex relatively modelling in application a oe a lasitiudrelativists, intrigued always has model cal rr ucin htcnrltedeviation the control that functions trary ∗ rsin o h aeo oainfrom rotation of rate the for pressions oriae onthv constant a have not do coordinates iain n,i h ae fteplanar the of cases the in and, lication, eddi a - ulda space. Euclidean 4-d flat a in bedded † eebdiga ela h shell the as well as embedding he ieto h te ad However, hand. other the on silent la n li oes h three The models. Ellis and olman o a noi,Txs729 USA 78249, Texas Antonio, San io, in n ec aea‘natural’ a have hence and tion, aishrclrclasn case, recollapsing uasi-spherical dbsh 71 ot Africa South 7701, ndebosch, mn.W loso how show also We ement. noee surprising a uncovered s rae nthe in urfaces It was obvious from the beginning that the constant r 2-surfaces (on each time slice) were not arranged ‘concentrically’, and in Szekeres’ original paper [25] he described the 3-spaces as “a set of displaced or ‘non-concentric’ spheres, planes and pseudospheres.” This non-concentric arrangement was understood and taken into account when slices through Szekeres models were plotted, for example [1, 2, 5, 21, 23, 24, 4, 3, 10]. Again, because of the relative complication of doing this, it hasn’t been attempted very often. Since the native Szekeres coordinates are stereographic, the calculation of slices typically involved converting to angular r-θ-φ coordinates, and then projecting into Cartesian-style X-Y -Z coordinates. However, this method depends on the unstated assumption that the θ-φ coordinates maintain a ‘constant’ orientation in some sense. So it came as a considerable surprise when it was discovered, only recently, that, in the θ-φ representation, the constant r shells are also rotated relative to each other, except in highly specialised cases. In 2013, Buckley & Schlegel [6] stated that the constant r shells are rotated relative to each other by specific amounts, but did not explain how this was obtained. Unaware of this, Hellaby, in 2017 [15], questioned whether the θ & φ coordinates represent a constant orientation, and by studying the variation of a suitable orthonormal tetrad, showed that they don’t. We refer to this paper as “FR”. Two years later, Buckley & Schlegel [7] provided rigorous support for their relative rotation of adjacent shells by developing a local embedding, which required not only the previously stated rotations, but also higher dimensional rotations (or ‘tilts’) in the embedding space. In particular, it demonstrated that these rotations straighten up geodesics. Clearly, this effect must be incorporated into graphical routines for generating slices through Szekeres models, and methods for doing this were presented. We refer to this paper as “PG”. Confirming that the PG shell rotation and embedding result does indeed explain the origin of the frame rotations found in FR, would be a confirmation of both papers, and a useful validation of the new understanding of the Szekeres geometry, as well as the correct way to plot it graphically. Further, the consideration of embeddings opens up the question of whether less obvious topologies are possible.

2 Background

2.1 The Lemaˆıtre-Tolman Spacetime

The Lemaˆıtre-Tolman (LT) spacetime [22, 27] represents a spherically symmetric cloud of dust particles that is inhomogeneous in the radial direction; both the density and the rate of expansion or contraction can vary with radius. Its metric is

R′2 dr2 ds2 = dt2 + + R2 dθ2 + sin2 θdφ2 , (1) − 1+ f  where R = R(t, r) is the areal radius, and f = f(r) is a geometry-energy factor. From the Einstein field equations (EFEs), the evolution equation and the density are

2M ΛR2 R˙ 2 = + f + , (2) R 3 2M ′ κρ = , (3) R2R′

2 where M = M(r) is the total gravitational mass interior to each constant r shell, and Λ is the cosmological constant. When Λ=0, the evolution equation (2) has parametric solutions, for example when f < 0, M M R = 1 cos η , t = a + η sin η . (4) ( f) − ( f)3/2 − − −   and a = a(r) is the local ‘bang time’; the time on each constant r worldline at which R(t, r)=0 is t = a. The factors M M L = and T = (5) ( f) ( f)3/2 − − can be thought of as a scale length and a scale time for the worldline at r; 2L is the maximum areal radius reached, and 2πT is the duration from bang to crunch. The R′ that appears in the metric can be expressed parametrically as M ′ f ′M 3 R′ = (1 φ )(1 cos η) φ 1 (1 cos η) ( f) − 1 − − f 2 2 1 − − −   ( f)1/2a′φ (1 cos η) , (6) − − 2 − sin η(η sin η) sin η where φ = − , φ = . (7) 1 (1 cos η)2 2 (1 cos η)2 − − The Ellis [9] metrics are equivalents of the LT case, having planar and hyperboloidal (pseudo- spherical) symmetry.

2.2 The Szekeres Spacetime

The Szekeres (S) [25, 26] can be thought of as distortions of the LT and Ellis ones. In addition to the free functions f(r), M(r) and a(r) they have 3 more functions S(r), P (r) and Q(r) that specify the deviation from symmetry — spherical, planar, or hyperboloidal. The metric is RE′ 2 R′ dr2 − E R2 ds2 = dt2 +   + dp2 + dq2 , (8) − ǫ + f E2 where ǫ = +1, 0, 1, and  − S (p P )2 (q Q)2 E = − + − + ǫ . (9) 2 S2 S2   Since S = 0 is not possible for a regular metric, we assume S > 0. In fact the last term in (8) is R2 times the 2-metric for a unit sphere, pseudo-sphere, or plane, depending on whether ǫ is +1, 1, or 0. Thus the 3-spaces are foliated by a collection of symmetric 2-spaces, but they are not arranged− symmetrically, as we shall see. By the EFEs, R(t, r) obeys exactly the same evolution equation (2), while the density ρ is more complicated, 3ME′ 2 M ′ − E κρ =   . (10) RE′ R2 R′ − E   For more information about the Szekeres metric, see for example [20, 16, 17, 14, 28, 6, 7].

3 2.3 Angular Form of the Szekeres Metric

The standard stereographic mapping, for the ǫ = +1 case,

θ θ p = P + S cot cos φ , q = Q + S cot sin φ , (11) 2 2     transforms the metric into a more complicated, non-diagonal form,

1 R 2 ds2 = R′ + S′ cos θ + sin θ P ′ cos φ + Q′ sin φ ǫ + f S "   R2   + S′ sin θ + (1 cos θ) P ′ cos φ + Q′ sin φ S2 − 2   R 2 ′ ′ 2 + 2 (1 cos θ) P sin φ Q cos φ dr S − − # R2   S′ sin θ + (1 cos θ) P ′ cos φ + Q′ sin φ dr dθ − S − 2 R sin θ ′ ′  (1 cos θ)2 P sin φ Q cos φ dr dφ − S − − + R2(dθ2 + sin2 θ dφ2) .  (12)

For each (t, r) 2-sphere, it is sometimes convenient to define a local cartesian frame by

x = R sin θ cos φ y = R sin θ sin φ (13) z = R cos θ .

2.4 The Szekeres Dipole

The factor = E′/E that appears in both the metric (8) and the density (10) controls the deviation D from spherical, hyperboloidal or planar symmetry, and for ǫ = 0 it behaves like a dipole. The dipole has maximum value and orientation 6 ′ ′ ′ ′ E′ J P S ǫS J Q S ǫS J = = , p P = − , q Q = − , Dm E S m − H2 m − H2 m   (14)

2 ′2 ′2 ′2 ′2 ′2 where J = ǫ S + ǫ P + Q , H = P + Q . ′ q p The locus E =0 lies on the (p, q) circle 

′ ′ (p P )S ′ (q Q)S ′ ′ ′ ′ − + P + − + Q = P 2 + Q 2 + ǫS 2 . (15) S S     For the quasi-spherical case, ǫ = +1, the dipole function can be written as

′ ′ ′ E′ S cos θ + sin θ P cos φ + Q sin φ = = , (16) D E − S 

4 and it is evident that E′/E ranges between opposite extremes, passing through zero on an ‘equatorial’ circle. Note that here E′ still represents the r derivative at constant p & q. The angular of the dipole maximum is found from H S′ P ′ Q′ sin θ = − , cos θ = − , cos φ = , sin φ = , (17) m J m J m H m H or, expressed in local Cartesian coordinates, the position of the maximum on the (x, y, z) unit sphere is P ′ Q′ S′ x = sin θ cos φ = − , y = sin θ sin φ = − , z = cos θ = − , (18) m m m J m m m J m m J while the E′ =0 locus is in the plane P ′x + Q′y + S′z =0 . (19)

The dipole has two obvious effects — in the grr component of (8) it creates a non-uniform separation between adjacent 2-spheres of constant r, and in (10), it creates a variation of the density distribution around each sphere. These effects are in addition to the r-dependent inhomogeneity of the underlying LT model.

2.5 The Rotations

Relative to the angular form of the Szekeres metric (12), there is another more subtle effect of S, P & Q. The angular coordinates θ & φ of (12) do not in fact represent a constant orientation, and their cardinal directions do not parallel transport from one shell to the next. Adjacent 2-spheres have a relative rotation: the sphere at r + δr is rotated Q′ by δr about the x axis (20a) S P ′ and by − δr about the y axis , (20b) S relative to the one at r [6]. Four justifications for this were given in [7]. Firstly there is an argument about nearest points on the two spheres, explained in PG, section V.B and fig 4. Secondly, it was shown that geodesics look much straighter once these rotations are incorporated into plots. Thirdly was the calculation in PG appendix C, perhaps not entirely rigorous, that added displacements and rotations to the LT metric, ending up with the S metric. Fourthly, an embedding of any given constant t 3-space of the angular S metric for ǫ = +1, into a 4-d space that is flat or has constant curvature, turned out to require the above-stated rotations.

3 Embeddings

3.1 Global Embedding of Positively Curved LT 3-Spaces

Let E4 be a 4-d with Cartesian coordinates, X,Y,Z,W , and let the LT 3-spaces have positive curvature, 1 f 0. At a fixed time t, R becomes a function of r only. We define a 3-surface Σ by − ≤ ≤ X = R(r) sin θ cos φ , (21a)

5 Y = R(r) sin θ sin φ , (21b) Z = R(r) cos θ , (21c) r W = R′(r)α(r) dr , (21d) Z0 so then

dX = R′ sin θ cos φ dr + R cos θ cos φ dθ R sin θ sin φ dφ , (22a) − dY = R′ sin θ sin φ dr + R cos θ sin φ dθ + R sin θ cos φ dφ , (22b) dZ = R′ cos θ dr R sin θ dθ , (22c) − dW = R′α(r) . (22d)

Here R(r) is the LT areal radius R(t, r) at a particular time, and α is

f(r) α = − . (23) ±s1+ f(r)

Now in [13] the function f was interpreted as f = cos2 ψ where ψ is the angle of the tangent cone − to the embedded surface at r, and in this notation, then, α = cot ψ:

dR R′ 1+ f 1 tan ψ = = = = (24) dW ′ f ±s f α R − − 1+ f r

cos ψ = f , sin ψ = 1+ f , cot ψ = α . (25) → − p p Since f 0, closed models are quite likely, in which case there will be at least one rm that is a maximum≤ (or minimum) in R where R′ =0, f = 1, but R′/√1+ f is finite [18, 13]. As a spatial extremum is approached and traversed, R′, ψ and −α change sign, R′ & ψ passing through zero and α diverging. This ensures dW/dr retains a constant sign1. Using this embedding, (21) and (22) show that the metric of the 3-surface becomes

R′2 dr2 dX2 + dY 2 + dZ2 + dW 2 = + R2 dθ2 + sin2 θ dφ2 (26) 1+ f  which is the spatial part of the LT metric, dt =0. In the case that

R = K sin r , R′ = K cos r , f = sin2 r, K constant, (27) − we find

R′2 dr2 + R2 dθ2 + sin2 θ dφ2 = K2 dr2 + sin2 r dθ2 + sin2 θ dφ2 (28) 1+ f    and Σ becomes the 3-sphere.

1This is not essential, and other choices could lead to different valid embeddings.

6 3.2 Buckley & Schlegel’s Local Szekeres Embedding

For quasi-spherical S models with f < 0, their constant t 3-spaces can be embedded in a 4-d flat space, but for general f, one must embed in a 4-space of constant curvature [7]. We here consider the former case. Let E4 be a 4-d Euclidean space with Cartesian coordinates, X,Y,Z,W . In this 4-space, a 3- surface is constructed from a sequence of 2-spheres, by expanding, shifting and rotating a unit sphere, as a function of parameter r. Suitable choices of the expansion, shift and rotation functions ensure the intrinsic metric of the 3-surface is identical with the positively curved, quasi-spherical Szekeres 3-spaces of constant t. While an embedding is primarily a visualisation tool, in this case it also provides a clear confirmation of Buckley and Schlegel’s rotations, given in equation (20). It is also convenient to define local Cartesian coordinates, x,y,z,w near each constant r shell. In these coordinates, the 2-spherical shell lies in the w =0 3-space, according to (13), so the accumulated displacements and rotations are ignored, and the focus is on the local rate of displacement and rotation. Buckley & Schlegel’s embedding equation is

V (r,θ,φ)= R(r)AT (r)U(θ,φ)+∆(r) , (29) where V is a point in E4 and U is a unit sphere,

X sin θ cos φ Y sin θ sin φ V = , U = , (30)  Z   cos θ  W   0          while the rotation A, and the displacement vector ∆, are functions of r that satisfy the following differential equations (DEs)

P ′ P ′ 0 0 α S S  Q′ Q′  0 0 α ′ ′  S S  A (r) = Ω (r)A(r)=  ′ ′ ′  A(r) , (31)  P Q S   0 α  − S − S S   P ′ Q′ S′     α α α 0  − S − S − S   P ′  R S  Q′  ′ T ′ T R ∆ (r)= A (r)D (r)= A (r)  S  , (32)  S′  R   S   ′   R α      where α is given by (23), and here too one may specify that it have the same sign as R′. The intrinsic 3-metric of the embedded 3-surface is derived from the differential of V , using (29)-(32),

ds2 = g(dV, dV ) = dV T dV (33) ′ T T ′ T ′ dV =(R dr)A U + R((A ) dr)U + RA (Uθ dθ + Uφ dφ) + ∆ dr (34) ′ T T ′ T T T ′ = R A U dr + R(A (Ω ) )U dr + RA (Uθ dθ + Uφ dφ)+ A D dr (35)

7 T ′ ′ T ′ = A (R U + R(Ω ) U + D )dr + RUθ dθ + RUφ dφ (36) 2 ′ T T ′ ′ T T T ds = (RU + RU Ω +(D ) )dr + RUθ dθ + RUφ dφ ′ ′ T ′  (R U + R(Ω ) U + D )dr + RUθ dθ + RUφ dφ  (37) 2 ′2 T ′ T ′ T T ′ ′ T ′ ′ T = dr R U U + RR U (Ω ) U + U Ω U + R U D +(D ) U 2 T ′ ′ T T ′ ′ ′ T ′ T ′ T ′ + R U Ω (Ω ) U + R U Ω D +(D) (Ω ) U +(D ) D  ′ T T 2 T ′ T ′ T + drdθ RR U Uθ + Uθ U + R U Ω Uθ + Uθ (Ω ) U ′ T T ′ + R (D) Uθ + Uθ D    ′ T T 2 T ′ T ′ T + drdφ RR U Uφ + Uφ U + R U Ω Uφ + Uφ (Ω ) U ′ T T ′ + R (D ) Uφ + Uφ D    2 2 T 2 T T 2 2 T + dθ RUθ Uθ + dθdφ R Uθ Uφ + Uφ Uθ + dφ R Uφ Uφ (38)    o  where cos θ cos φ sin θ sin φ − cos θ sin φ sin θ cos φ U = , U = (39) θ  sin θ  φ  0  −  0   0          Evaluating the various matrix products, we find

T ′ T ′ T T T T T T T 0= U Ω U = U (Ω ) U = U Uθ = Uθ U = U Uφ = Uφ U = Uθ Uφ = Uφ Uθ (40a) T T 1= U U = Uθ Uθ (40b) T 2 Uφ Uφ = sin θ (40c) R (D′)T U = U T D′ = sin θ P ′ cos φ + Q′ sin φ + S′ cos θ (40d) S R (D′)T (Ω′)T U = U T Ω′D ′ = S′ sin θ P ′ cos φ+ Q′ sin φ P ′2 + Q′2 cos θ S2 − ′ 2 R α ′  ′ ′   + sin θ P cos φ + Q sin φ + cos θS (40e) S 2 ′ ′ R  ′ ′ ′ ′  (D )T D = P 2 + Q 2 + S 2 + R 2α2 (40f) S2 2 ′ ′ α ′ ′ ′ 2 U T Ω (Ω )T U = S cos θ + sin θ P cos φ + Q sin φ S2 2 sin θ ′ ′ 2 ′ ′ + P cos φ + Q sin φ + cos2 θ P 2 + Q 2 (40g) S2  P ′ cos φ + Q′ sin φ  U T Ω′U = U T (Ω′)T U = (40h) θ θ − S  ′ ′ T ′ T ′ T cos θ sin θ P sin φ Q cos φ U Ω Uφ = Uφ (Ω ) U = − (40i)  S R (D′)T U = U T D′ = cos θ P ′ cos φ + Q′ sin φ S′ sin θ (40j) θ θ S − R sin θ (D′)T U = U T D′ =  P ′ sin φ Q′ cos φ (40k) φ φ − S −  and we recover the metric (12) of the angular form of the quasi-spherical Szekeres metric.

8 3.3 Visualising the Local Embedding Geometry

Although there is no physical significance to the embedding of a given spacetime into one of higher , it can be very helpful for visualising the spacetime geometry, and that is what we explore here. We note that the constant w projection of D′, (32), is anti-parallel to the (x, y, z) direction of the dipole maximum, (18). For the local rate-of-rotation matrix Ω′, the fixed point locus is a 2-plane,

S′λ S′µ ′  ′ ′  Ω V =0 Vfp = (P λ + Q µ) , λ,µ independent parameters, (41) → −   (P ′λ + Q′µ)       α    which makes Ω′ the derivative of a simple rotation (not a double or isoclinic rotation). The constant ′ w projections of D and Vfp are orthogonal. This last fact ensures the maximum tilt-displacements occur along the dipole axis (where the max & min are located) — see fig 1 — and thus enables the following argument. In the local (x,y,z,w) frame, the constant r 2-sphere lies in the x-y-z 3-space, and the direction o = (0, 0, 0, 1) is orthogonal to that space. In going from the 2-sphere at r to the 2-sphere at r + δr, the displacement of the sphere centres is D′ δr; the rate of perpendicular displacement is R′α, and the rates of sideways displacement are RP ′/S, RQ′/S, & RS′/S, towards the +ve x, y, & z directions2. The slant angle, between o and the line of centres, is D′ R′α cos γ = o = (42) · D′ R2 2 + R′2α2 | | Dm which we could re-write as ‘component’ angles3 p RP ′ RQ′ RS′ tan γ = , tan γ = , tan γ = . (43) x R′αS y R′αS z R′αS Since the total rotation used in (29) is AT , we see from (31) that the local rotation-rate matrix to examine is (Ω′)T . The components of the local rate-of-shell-rotation are thus ′ ′ ′ P ′ Q ′ ω δr = − δr, ω δr = − δr, ω δr =0 , (44) xz S yz S xy with senses as noted above, and the components of rate-of-tilt are ′ ′ ′ ′ P α ′ Q α ′ S α ζ δr = − δr, ζ δr = − δr, ζ δr = − δr . (45) xw S yw S zw S Therefore the rates of tilt are α/R times the rates of sideways displacement, and a 3-plane parallel to w =0 is tilted down on the x, y, & z sides by ζ , ζ , & ζ , respectively. − − − xw yw zw Both these effects, the centre displacement and the tilt, decrease the shell separation where has ′ D the same sign as R (the ’closing edge’) and increase it where has the opposite sign (the ’opening D 2We here assume that P ′, Q′, & S′ are positive. Where the opposite sign occurs, the relevant direction is changed in the obvious way. 3 2 2 These ‘components’ are the projections of γ onto the x-w, y-w, and z-w planes, such that tan γ = tan γx + 2 2 tan γy + tan γz.

9 edge’). The no-shell-crossing conditions ensure the separation stays positive all round each sphere. Fig 1 illustrates the arrangement for the case when R′ > 0. Although finite displacements are shown, one should think of δr as infinitesimal, so that only first order effects are relevant. The four displacements shown are:

′ the w-separation of shell centres = δw0 = R α δr , (46a) ′ ′ the increase in shell radius = δd0 = R α tan ψδr = R δr , (46b) the tilt down displacement where is max = δw = Rζ′ δr = Rα δr , (46c) D 1 Dm the dipole displacement of the shell centre = δd = R δr . (46d) 1 Dm

Interestingly, δd0 & δw0 make the same angle as δd1 & δw1. This angle coincidence is possibly the reason why the calculation in appendix C of PG actually works — because the displaced & tilted shell ar r + δr more or less lies on the LT tangent cone to the 3-surface at r, to first order.

δw1 δd0 δd1 δd0

δd1 δw δw0 1 γ δw0 δw0 ψ δd1 max min Fig 1. Sketch of the displacement and tilt effects in the embedding, with one dimension suppressed. Because the sketch shows finite displacements, instead of infinitesimal ones, the distances shown are not exact. The green shows the embedding of the underlying LT model — the shells at r (lower) and r + δr (upper) and the local tangent cone. The blue vectors show the w displacement between the two shells and the radial expansion of the second shell (assumed positive here). The cyan shows the displaced and tilted shell at r + δr of the Szekeres model. The red vectors show the sideways displacement of the shell centre, and the down or up displacement due to the tilt. The dipole maximum is located at the left side, and the minimum at the right. The E′ = 0 locus, the ‘widest’ part of the 2-sphere at r + δr is shifted by δd1; this is consistent — a tilted slice through the LT cone has a displaced ‘greatest width’. The tilt axis is the magenta dashed line, and it is perpendicular to the red displacement δd1. The θ-φ rotation is not shown. Looking at the next shell pair, at r + δr and r +2δr, the local LT tangent cone is now tilted by ζ, and the r +2δr shell is tilted further. The slant angle γ is also tilted by ζ. A reflection in the bottom plane gives an idea of the R′ < 0 case. Where R′ < 0, α also flips sign and goes to in (46). Dm −Dm

3.4 Origins, Extrema & Self-Intersections

Using the expressions in (46), we examine the limiting values of these embedding quantities near origins and spatial extrema. We assume there are no shell crossings. If there are, then R′ & α do not always flip signs together. We also assume a well behaved r coordinate, with R finite & non-zero everywhere ′ except at an origin, r = ro, and R finite & non-zero everywhere except at a spatial extremum, r = re. We further assume ‘generic’ arbitrary functions, so that, for example, R′ & go linearly through zero D

10 at an extremum4. The results are gathered in table 1.

flips sign with flips sign behaviour at behaviour with R′ R′ 0 1/α at R 0 Dm → −Dm → ← → R R const 0 as (r ro) ′ ′ → → − R R 0 as (r re) const =0 α α → as (r −r )−1 →0 as (r 6 r ) → ∞ − e → − o m m 0 as (r re) const (or 0) D D →′ − → δw0 No No R α const 0 as (r ro) ′ → → − δd0 No Yes R 0 as (r re) const =0 δw Yes Yes Rα→ const− →0 as (r 6 r )2 1 Dm → → − o δd1 Yes No R m 0 as (r re) 0 as (r ro) D ′→ − → − dipole slant δd1/δw0 Yes No (R m)/(R α) 0 as (r re) const tilt rate δw /R Yes Yes D α → const − →0 as (r r ) 1 Dm → → − o Table 1. The behaviour of embedding displacements near origins and spatial extrema. See (46) and the illustration in fig. 1. Now, at an origin, where R(t, r ) 0 t, we see that the dipole slant does not disappear, but the o → ∀ tilt-rate must disappear. In contrast, at a spatial extremum, where R′(t, r ) 0 t, the dipole slant must disappear, e → ∀ whereas the tilt rate does not disappear. The condition for no local self-intersection (of adjacent shells), referring to fig 1 and (46), is just:

δw < δw , | 1| | 0| Rα δr < R′αδr , | Dm | | | R′ < . (47) Dm R

Interestingly, this is just the no-shell-crossing conditions for S, P & Q [16, 14].

3.5 Centre-Line Curvature

We now consider the locus of 2-sphere centres in the embedding space; that is, in the (X,Y,Z,W ) frame. (The previous section mostly used the local (x,y,z,w) frame.) For simplicity, we will consider a model in which P ′ = Q′ =0 everywhere. Between r and r + δr, ′ the slant angle between o and the line of centres changes by γz δr. Part of this change is due to the fact that in the same span, the 2-spheres of constant r undergo a tilt in the same direction by an angle ′ ζzw δr, and o’s angle changes along with them by the same amount. The rest must be due to the change in the angle of the line of centres itself. See fig 2. If we denote this change of the line-of-centres angle by δξ, we can write

′ ′ δξ = γz δr + ζzw δr (48)

4One may intentionally choose functions that give other behaviour at specific locations, such as R′ or going quadratically to zero and not changing sign. D

11 Fig 2. A diagram of the contributions to the centre line curvature. Solid lines represent values at r, and dashed lines represent values at r + δr. The green lines show the shells’ orientations. The blue lines are the vectors o orthogonal to the shells, and the change between them is the rate of tilt, ′ ζzw. The red lines are the tangent vectors of the line of centres, which changes by δξ. The angles between the two, written in purple, are the slant angles γz. Looking at the solid blue line and the dashed red line, we can see that the total angle between them is equal to γ + δξ, but also ′ ′ ′ ′ equal to ζzwδr + γz + γzδr, giving the result δξ = γzδr + ζzwδr.

′ ′ Eq (45) gives ζzw directly. We can calculate γz from eq (43) as follows: ′ 1 S′ R R′′ S′ Rα′ S′ R S′ γ′ sec2 γ = γ′ (1 + tan2 γ )= + (49) z z z z α S − R′2α S − R′α2 S R′α S ′ ′ ′ ′ ′   ′2 S ′′ S ′ ′ S ′ S ′ R α S R R α S R R α S + R R α S γz = − −′ , (50) 2 S 2 ′2 2 R S + R α  ′  and this can be re-expressed in terms of ζzw, ζ′ R R′′ Rα′ R γ′ sec2 γ = γ′ (1 + tan2 γ )= zw + ζ′ +2 ζ′ ζ′′ (51) z z z z − α2 R′2α2 zw R′α3 zw − R′α2 zw ′2 2 ′ ′′ 2 ′ ′ ′ ′ ′ 2 ′′ ′ R α ζzw + R R α ζzw +2R R αα ζzw R R α ζzw γz = − 2 ′2 ′2 4 − . (52) R ζzw + R α Locally about any point, the line of centres can be approximated as following a section of a circle. The radius of this circle is determined by dividing the length traversed in a small span by the corresponding change in angle δξ. We therefore need to express δξ in terms of path length along the line, rather than in terms of δr. The distance δl between the centre of shell r and that of shell r + δr follows from the displacement components given in eqs (46a) and (46d):

S′ 2 δl = δr R2 + R′2α2 (53) S s   The local radius of curvature of the line of centres is then δl ρ = zw δξ

′ 3/2 2 S 2 ′2 2 R S + R α = ′ ′ ′ ′ ′ ′ 3 (54) R′2(α α3) S R R′′hα S R R′α′ S +iR R′α S R2α S − S − S − S S − S ′   The relationship between the two contributions, the rate of tilt γz and the rate of change of slant ′ ζzw, is somewhat complicated, but becomes much simpler near an extremum in R. Starting from equation (51), replace R′ by R (r r ) (where R is the value of R′′ at r ), α by αe , and α′ by 2 e 2 e r−re αe − − 2 . This gives − (r re) ′ ′ 2 ζzw 2 R ′ R ′′ γz sec γz = 2 (r re) 2 ζzw 2 (r re)ζzw (55) − αe − − R2αe − R2αe −

12 ′ ′′ Since by table 1 ζzw approaches a constant at the spatial extremum, ζzw is small there, so we can drop the last term, as well as the first. The middle term dominates, and its sign depends only on R2 and ′ ′ ′ ′′ ζzw. Therefore, γz and ζzw have the same sign near maxima, where R is negative, and opposite signs near minima, where R′′ is positive. Consequently, the slant and tilt effects both increase the curvature of the line of centres near a spatial maximum, but oppose each other near a spatial minimum5.

4 Comparison of Rotations in PG & FR

In order to connect the shell rotations found in PG with the frame rotation effect found in FR, let us now set up the orthonormal of FR, within the E4 of section 3.2. Since the basis orientation is time-independent, we will only look at spatial components. Thus there will be 3 vectors intrinsic to the embedded surface, and one perpendicular to it. We shall need to refer to 3 different fames; the Cartesian coordinates of E4, the Szekeres angular coordinates, and the orthonormal tetrad of the embedded surface. Our index convention is

i, j, k, Szekeres coordinates r,θ,φ ··· − s, u, v, 4-d flat coordinates X,Y,Z,W (56) ··· − (a), (b), (c), ortho-normal basis indices (θ), (φ), (n), (N) . ··· −

4.1

From (36), the mapping between (r,θ,φ) and V =(X,Y,Z,W ) consists of

∂V ′ ′ ′ ∂V ∂V = AT (R U + R(Ω )T U + D ) , = AT RU , = AT RU , (57) ∂r ∂θ θ ∂φ φ ∂(X,Y,Z,W ) ∂V ∂V ∂V Λs = = (58) → i ∂(r,θ,φ) ∂r ∂θ ∂φ   which also constitute vectors in the 3-surface along the r, θ & φ directions. Thus one may easily find a vector orthogonal to that 3-surface,

α sin θ cos φ α sin θ sin φ N = AT , (59)  α cos θ   1   −    and consequently a third surface vector orthogonal to the θ and φ directions (as well as N) is

sin θ cos φ sin θ sin φ n = AT   . (60) cos θ  α      5This was first observed in numerical output, which also indicates the tilt-rate is the larger contribution in determining the centre-line curvature.

13 Normalising these, we obtain an orthonormal basis and its dual, with components in E4; the components of e(c) are in the columns of (61), and the basis order is (θ), (φ), (n), (N),

cos θ cos φ sin φ √1+ f sin θ cos φ √ f sin θ cos φ − − cos θ sin φ cos φ √1+ f sin θ sin φ √ f sin θ sin φ e (c) = AT = AT e (c) , (61) u  sin θ 0 √1+ f cos θ −√ f cos θ  u − −  0 0 √ f √1+ f   − −   cos θ cos φ cos θ sin φ sin θ 0 − sin φ cos φ 0 0 e s = A = e sA. (62) (b) √1+−f sin θ cos φ √1+ f sin θ sin φ √1+ f cos θ √ f  (b) −  √ f sin θ cos φ √ f sin θ sin φ √ f cos θ √1+ f   − − − −    4 Equation (62) gives the flat-space components of e(b) in E .

4.2 Variation of the Embedded Basis

Consider a path parametrised by λ, within the embedded 3-surface, which is also a 3-space of constant time in the Szekeres metric, so that r = r(λ), θ = θ(λ), φ = φ(λ). The tangent vector is vj = dxj/dλ, and the path in E4 is V = V (r(λ), θ(λ),φ(λ)). Along this path, the variation of the flat space basis 4 vector e(b) within E is

s u s u u s u ve = v e e e = v e e + e e e = v e e , (63) ∇ (b) ∇ s (b) u (b) ,s u (b) ∇ s u (b) ,s u so that the frame rotation matrix of FR is (see eq (21) of that paper)

(c) (c) s u (c) v s u (c) = v e (e )= v e e e e = v e e V (b) ∇ (b) (b) ,s u v (b) ,s u i s u (c) i u (c) = v Λi e(b) ,s eu =v e(b) ,i eu  .  (64) We have u u ′ u u ′ u e(b) ,r = e(b) A + e(b) ,r A = e(b) Ω + e(b) ,r A u u u u (65) e(b) ,θ = e(b) ,θ A , e(b) ,φ = e(b) ,φ A,  so that

′ (c) = vi e u e (c) = r˙ e u Ω + e u + θ˙ e u + φ˙ e u A AT e (c) V (b) (b) ,i u (b) (b) ,r (b) ,θ (b) ,φ u u ′ u ˙ u ˙ u (c) = r˙e(b) Ω + e(b) ,r + θ e(b) ,θ + φ e(b) ,φ eu . (66) The basis with respect to Szekeres angular coor dinates are

0 0 00 0 0 00 f ′  sin θ cos φ sin θ sin φ cos θ 1  e s = − , (67a) (b) ,r 2  √1+ f √1+ f √1+ f √ f     sin θ cos φ sin θ sin φ cos θ −1  − − − −   √ f √ f √ f √1+ f   − − −   sin θ cos φ sin θ sin φ cos θ 0 − − − s 0 0 00 e(b) ,θ =   , (67b) √1+ f cos θ cos φ √1+ f cos θ sin φ √1+ f sin θ 0 −  √ f cos θ cos φ √ f cos θ sin φ √ f sin θ 0  − − − −    14 cos θ sin φ cos θ cos φ 0 0 − s cos φ sin φ 0 0 e(b) ,φ =  − −  . (67c) √1+ f sin θ sin φ √1+ f sin θ cos φ 0 0 −  √ f sin θ sin φ √ f sin θ cos φ 0 0  − − −    Using eqs (67a)-(67c) and (61)-(62) in (66) above, we find that fS′ sin θ [(1 + f) f cos θ] P ′ cos φ + Q′ sin φ vr (θ) = − − − { } + 1+ f vθ , (68a) V (n) √1+ f S [1 (1 + f)(1 cos θ)] P ′ sin φ Q′ cos φ vr p (φ) = − − { − } + 1+ f sin θvφ , (68b) V (n) √1+ f S (P ′ sin φ Q′ cos φ) sin θ p (φ) = − − vr + cos θvφ . (68c) V (θ) S Eqs (68a)-(68c) agree with FR(25)-(27), confirming that the frame rotation effects found in [15] are fully consistent with, and therefore explained by, the shell rotations and tilts uncovered in [6] & [7].

5 Toroidal & Rotating Embeddings

As an illustration of the rotations and tilts described in PG & FR, it would be interesting to see if one can define a Szekeres model whose embedding ‘naturally’ bends round and closes on itself, i.e. the embedded surface has the topology of a torus in the 4-d flat space, without any arbitrary identifications. It doesn’t have to be a Datt-Kantowski-Sachs (DKS) type model [8, 19], it could be a quasi-spherical model that has both a spatial maximum and a minimum; or multiple spatial maxima & minima. Of course there is no physical significance to the embedding of a spacetime6, and the shape of the embedded surface depends on the space it is embedded into. The point here is to illustrate that the tilt can be continuously in the same sense. Where spatial extrema occur, the conditions for no shell crossings or surface layers [16, 14] require that f = 1, all 6 arbitrary functions have zero derivative, 0 = M ′ = f ′ = a′ = S′ = P ′ = Q′, and R′, M ′, α−must change sign together.

In order to achieve this, we require 3 things as r runs from ri to rf : (i) the tilt/rotation matrix A(r) must run round 2π, relative to some ‘axis’, and return to the identity; (ii) the locus of centres ∆(r) must form a loop; (iii) the areal radius R(r) and its derivative R′(r) must return to their starting values, so that the join is smooth — for example R could be (multiply) periodic in r around this loop — and this should hold at each constant t; (iv) ideally the 3-surface should not intersect itself in the embedding, and there should be no shell crossings.

5.1 Embedding DEs

The matrix DEs of (31) actually separate out into 4 identical sets of 4 linked DEs, but with different initial conditions,

ν1(r) ν2(r) ν3(r) ν4(r) χ (r) χ (r) χ (r) χ (r) A(r)=  1 2 3 4  , A(0) = I, (69a) λ1(r) λ2(r) λ3(r) λ4(r) σ (r) σ (r) σ (r) σ (r)  1 2 3 4    6... unless it’s a brane.

15 P ′ Q′ ν′ = (λ + ασ ) , χ′ = (λ + ασ ) , i S i i i S i i ′ ′ ′ ′ ′ ′ ′ P ν Q χ + S ασ ′ α(P ν + Q χ + S λ ) λ = − i − i i , σ = i i i , i =1, 2, 3, 4 ; (69b) i S i − S ν1(0) = 1 , χ1(0) = 0 , λ1(0) = 0 , σ1(0) = 0 , (69c)

ν2(0) = 0 , χ2(0) = 1 , λ2(0) = 0 , σ2(0) = 0 , (69d)

ν3(0) = 0 , χ3(0) = 0 , λ3(0) = 1 , σ3(0) = 0 , (69e)

ν4(0) = 0 , χ4(0) = 0 , λ4(0) = 0 , σ4(0) = 1 . (69f)

Similarly, from (32), the initial value problem (IVP) for the line of shell centres separates out into the same groups,

XC Y ∆(r)=  C  , ∆(0) = 0 , (70a) ZC W   C  R  (∆i)′ = (P ′ν + Q′χ + S′λ )+ R′ασ , i =1, 2, 3, 4 , (70b) S i i i i making 4 identical sets of 5 DEs.

5.2 Case With Only S Varying

Let’s take the simple case of P ′ =0= Q′, while f, M, a, S are general. (i) Then by (31)

00 0 0 00 0 0  ′ ′ ′ A (r) = Ω (r)A(r)=  S  A(r) (71) 00 0 α  S   S′    0 0 α 0   − S    which integrates up to

10 0 0

r ′ 01 0 0  S α A = , ζ(r)= dr . (72) 0 0 cos ζ sin ζ r S   Z i   0 0 sin ζ cos ζ    −  Then from (72) and (23) we have

S2ζ′2 1 f = − = − , (S2ζ′2 + S′2) S′2 1+ S2ζ′2   (73) f = 1 S′2 =0 , ζ′ =0 , − → 6 f = f S′2 = (1 f )S2ζ′2 . − a → − a

16 One may thus obtain all of f(r), α(r), S(r) or ζ(r) from specifying just two of them. It should be fairly easy to make ζ run from 0 to 2π, because it does not depend on time through R or R′. We show below that f and hence α will be oscillatory, and we note that α diverges where f goes to 1, so S′ = 0 will be needed here. Thus S′α/S needs to be non-negative, which means S′ changes sign− with α. Possible functional forms for ζ(r), S(r), f(r), and α(r) will be considered below. (ii) Next by integrating (32), and requiring that the line of centres closes up, we find

0 0 ′ ′   ∆ = AT D = AT S′ , (74) R   S   ′  R α    0 0 rf rf   ′ ′ ∆ dr =0= S ′ dr . (75) cos ζR sin ζR α Zri Zri  S −   ′   S ′  sin ζR + cos ζR α  S    Generically, R & R′ don’t have the same time dependence, so, in order for this to be true at all times, we attempt to arrange that

rf ′ rf S ′ cos ζR dr =0= sin ζR α dr , (76a) S Zri Zri rf S′ rf sin ζR dr =0= cos ζR′α dr . (76b) S Zri Zri In particular, if within 0 ζ π, one can arrange that R goes max-to-min-to-max (or min-to-max-to- min), in a manner that’s≤ symmetric≤ about ζ = π/2, then, however the time evolution goes, 2 copies of this should join up nicely. Similarly, if within 0 ζ π/2, one can arrange that R goes max-to- min-to-max (or min-to-max-to-min), symmetrically≤ about≤ ζ = π/4, then 4 copies of this should also join up nicely. And so on. (iii) To make R & R′ join nicely, so that the torus ‘tube’ joins itself smoothly, we could choose f, M & a periodic; and of course f = 1 is needed at the extrema, so f must oscillate twice as fast. − (iv) Once a detailed model has been chosen, it can be checked for shell crossings and self intersec- tions, and adjustments can be made as needed.

5.3 Case With Only P Varying

Next we consider the case S′ =0= Q′ (& general f, M, a, P ). Then by (69) & (70), the IVP becomes P ′ P ′ P ′α ν′ = (λ + ασ ) , λ′ = ν , σ′ = ν , i =1, 2, 3, 4 ; (77a) i S i i i − S i i − S i ν1(0) = 1 , λ1(0) = 0 , σ1(0) = 0 , (77b)

ν2(0) = 0 , λ2(0) = 0 , σ2(0) = 0 , (77c)

17 ν3(0) = 0 , λ3(0) = 1 , σ3(0) = 0 , (77d)

ν4(0) = 0 , λ4(0) = 0 , σ4(0) = 1 , (77e) RP ′ (∆i)′ = ν + σ R′α , ∆(0) = 0 . (77f) i S i

Clearly χi =0= ν2 = λ2 = σ2 = YC for all r. Cyclic choices for the arbitrary functions similar to those above should again ensure closure.

5.4 Model 1, S Only

We choose ζ to be uniformly increasing, satisfying (i) by construction, we fix P =0= Q, and we let S oscillate,

ζ = µr , (78) S = S + S 1 cos(nµr) , n N , (79) 0 1 − ∈ in which case, (72), (23) and (73) give7  1 f = 2 −2 2 , (80) n S1 sin (nµr) 1+ 2 S + S 1 cos(nµr) 0 1 − S S 1 cos(nµr) α = 0 − 1 −  ,  (81) → nS1 sin(nµr) and the maxima of f are at S 1 cos(nµr)= 1 f = − . (82) S + S → m n2S2 0 1 1+ 1 S0(S0 +2S1) For the mass and bang time we choose

M = M + M 1 cos(nµr) , a = a a 1 cos(nµr) , (83) 0 1 − 0 − 1 − in accordance with requirement (iii).  The relevant no-shell-crossing requirements are that a′ and a′ +2πT ′ have the same sign as M ′, and that (E′/E) M ′/(3M) . The function f, which− appears in the denominator of T , defined by m ≤ eq (5), oscillates twice as fast as M does. Thus the amplitude of f’s variation in eq (80) needs to be small. For the model considered here, this means S1 should be small, which in turn means S does not vary much. Now the calculation of R(r) and R′(r) on a constant time slice is a numerical exercise; the t equation of (4) has to be solved to get η at each r. But, using the parametric solution (4) with zero Λ, we can instead integrate ∆′ on surfaces of constant η; if the result is zero for every constant-η surface, then it will be zero for every constant-t surface also. Using constant η surfaces has the additional

7We could choose α to be the negative of the r.h.s. of (81), which would merely reflect the embedding in the W =0 plane.

18 advantage that the bang and crunch singularities are avoided. The integrals in (76) are therefore modified to become ′ ′ ′ rf S rf r S α MS cos ζR dr = (1 cos η) cos dr dr , (84a) S − S ( f)S Zri Zri Zri  − rf rf r S′α M ′α cos ζR′α dr = (1 φ )(1 cos η) cos dr dr − 1 − S ( f) Zri Zri Zri  − 3 rf r S′α f ′Mα φ 1 (1 cos η) cos dr dr − 2 1 − − S f 2   Zri Zri  rf r S′α φ (1 cos η) cos dr ( f)1/2a′α dr , (84b) − 2 − S − Zri Zri  with obvious variations for the other two. By symmetry arguments, the 4 integrals of (76) clearly evaluate to zero when the above choices are made, ensuring that (ii) is satisfied. See fig 3 for a sample plot of one of the terms in (84).

0.002

0.001

±3 ±2 ±1 1 2 3 r

±0.001

±0.002

Fig 3. The integrand cos ζ(Mf ′α/f 2)(3φ /2 1)(1 cos η) of (84) at η = π/3, for the 1 − − functions and parameter values given in section 5.4 and also used in fig 4. A particular example satisfying all requirements is plotted in fig 4, for the parameter times η = π/3 and η =5π/3. The parameter values are n =2, µ = π, r = 1, r =1, S =1, S =0.1, M =0.1, i − f 0 1 0 M1 = 0.05, a0 = 0, a1 = 0.1. The no-shell-crossing conditions [18, 16] that are applicable here are ′ ′ ′ ′ ′ ′ ′ ( a )/M 0, (a +2πT )/M 0, (E /E)m M /(3M) ; and it has been checked that the chosen functions− and≥ parameter values≥ satisfy them, though≤ not by much. The construction used here puts strong limits on the variation of S.

19 2 1.5 4

1

W 1 W W 2 0.5

0 1 2 3 4 0.5 1 1.5 2 0 1 2 3 4 5 6 7

Z Z Z ±0.5 ±2 ±1

±1

±4 ±2 ±1.5 Fig 4. An example of an embedded Szekeres model with a ‘natural’ torus topology as described in section 5. The path of the sphere centres (red), and a selection of sphere diameters (blue), are shown in the (Z,W ) plane. Each diameter represents a complete 2-sphere, and the sequence of 2-spheres describes an embedded 3-surface in the flat 4-d (X,Y,Z,W ) space. The arbitrary functions and parameter values of this model are listed in section 5.4. The left plot is for η = π/3, the middle one for η = π, and the right one for η = 5π/3. These are not constant time plots, so they are indicative rather than precise.

5.5 Model 2, S Only

This is a much wobblier version of the toroidal model, that has 3 lobes instead of 2. The key defining functions are M(r)= M + M 1 cos(nµ r) , 0 1 − 0 f(r)= 1+ f 1 cos(2nµ r) , − 1  − 0 a(r)=0 ,  (85) ′ ζ = µ0 1+ µ1 cos(nµ0r) , ′ ′ S /S = ζ /α , from which we find

2 1 2f1 sin (nµ0r) µ1 sin(nµ0r) α = − , ζ = µ0 r + p √2f1 sin(nµ0r) nµ0 ′   S √2f1 µ0 sin(nµ0r) 1+ µ1 cos(nµ0r) = S 2 1 2f1 sin (nµ0r)  − (86) 2 pµ1 1 2f1 sin (nµ0r) S = S0 exp − − ( p √2f1 n ) 2 1/(2n) (1 2f1 +4f1 cos (nµ0r)) 2 − cos(nµ0r) 1 2f1 sin (nµ0r) 2√2f1 − −  q  We can choose the value of S0 to make S =1 at r =0. For the plots in fig 5, the parameters are

M0 =0.55 , M1 =0.7 , f1 =0.06 , n =3 , µ0 =1 , µ1 =0.8 , t =1 . (87)

20 Fig 5. An embedded Szekeres torus model with 3 lobes, showing just its intersection with the (Z,W ) plane. The red curve is the path of the sphere centres, and each black line is the diameter of a 2-sphere at a particular r value. The arbitrary functions and parameter values of this model are listed in section 5.5. The times of the plots are t = 0.02, t = 1, t = 3, and t = 5. What’s interesting here is that there are parts where the tilt is changing more rapidly and parts where it’s hardly changing. Although the effect found in section 3.5, that the dipole displacement and the tilt have opposing effects on the curvature of the path of centres near a spatial minimum, is evident here, in the models we tried, after shell crossings had been eliminated, the tilt curvature seems to dominate.

21 5.6 Model 3, P Only

For this model, the only non-sphericity function we vary is P . By [11] we expect the line of centres to be bent, and by section 2.5, we expect there to be shell rotation. It is defined by

M(r)= M + M 1 cos(nµr) , 0 1 − f(r)= 1+ f 1 cos(2nµr) = 1+2f sin2(2nµr) , − 1  − − 1 a(r)=0 ,  S(r)=1 , Q(r)=0 , ζ = µr , P ′/S = ζ′/α , 1 2f sin2(nµr) α = − 1 , → p √2f1 sin(nµr) (88) µ√2f sin(nµr) P ′ = 1 , 1 2f sin(nµr)2 − 1 2 p1 (1 2f1 1 2 cos (nµr) ) P (r)= ln − { − } 2n 2√2f1

cos(nµr) 1 2f sin2(nµr) , − − 1 q ! M0 =0.75 , M1 =0.25 , f1 =0.02 , µ =1 , n =6 , t =1 .

22 and is illustrated in fig 6.

Fig 6. Three views of a less symmetric embedded Szekeres torus model, in the (X,Z,W ) 3-space, at time t = 1. The red curve is the path of the sphere centres, each blue circle is the outline of a 2-sphere at a particular r value, and the green dots indicate the location of θ = 0 on the 2-sphere; the rotation of the (θ, φ) coordinates is not very large in this model. The arbitrary functions and parameter values of this model are listed in section 5.6. Apart from the toroidal topology, the line of centres no longer lies in a 2-plane, and the direction of the ‘north pole’ θ =0, rotates between shells; however these variations are fairly small.

5.7 Model 4, P Only

This model is not toroidal; it is a spatially closed model defined by

3 M(r) = sin (µr) M0 + M1 sin(µr) , f(r)= sin2(µr) , −  πM a(r)= − , ( f)3/2 − (89) S(r)=1 , Q(r)=0 ,

P (r) = sin(µr) P0 + P1 sin(µr) , α = tan(µr) , →  M0 =1 , M1 =5 , P0 =1 , P1 =0.3 , µ = π , η =8π/5 .

Fig 7 shows 3 views of the resulting embedding. The right hand view shows there is substantial rotation of the θ =0 direction.

23 15

W 10

5

0 ±10 ±5 ±10 ±15 0 ±5 ±20 0 X Z

15

15 W 10 0 10 W ±5 5 ±10 5 2 0 0 ±2 ±4 Z ±6 ±8 ±10 ±15 ±5 ±14 ±10 0 ±20 ±15 X ±14 0 ±20 ±10 ±8 ±25 ±6 ±4 ±25 Z ±2 0 X 2 Fig 7. Three views of the embedded Szekeres model 4, at time η = π which corresponds to t = 0. The blue circles represent 2-spheres of constant r; tilting relative to the line of sight makes some appear elliptical. The red curve is the centre path — the locus of sphere centres, and the green lines show the direction of θ = 0 for each plotted sphere/circle. The Y coordinate is suppressed, and the centre path lies in Y = 0.

5.8 Model 5, RW as Szekeres

The Szekeres models become homogeneous, despite the non-symmetric coordinates, if the LT functions f, M & a take the RW form, regardless of the S, P & Q functions8. Such a model is obtained with M = M sin3(µr), f = sin2(µr), a =0, S =1, Q =0 and P = P sin(µr)(P + P sin(µr)), along 0 − 0 0 1 with parameter values M0 =1, P0 =0.7, P1 =0.09, µ = π, and the result is shown in fig 8.

8See [12] around eq (2.24) and section 1.3.4 of [20].

24 4

3

W 2

1

±2 ±1 1 0 0 0 1 ±1 2 X Z ±2

4

4 3 2 2 W 2 1 3 1 Z 0 1 X ±1 0 W 2 ±2 ±1 2 1 ±2 1

0 0 Z ±2 ±1 ±1 0 X 1 2 ±2 Fig 8. Three views of the embedded RW-Szekeres model 5, showing it is actually a 3-sphere in very non-symmetric coordinates. The centre path (red line) is bent round, and the ‘north’ (green lines) show significant variation. Although the overall embedded shape is clearly a 3-sphere, the slicing is not only non-parallel, it also has shell rotation as shown by the variation of the ‘north’ direction.

6 Conclusions

A correct understanding of the geometry of the Szekeres metric is important both for physical interpre- tation, when it is used for models of inhomogeneous gravitating structures, and also for a more accurate graphical depiction of those models. Though the ‘non-concentric’ property of constant r shells was known from the start, it was only recently shown there is a hidden shell rotation effect, that appears when the usual angular coordinates are used. Hitherto, it was tacitly assumed these coordinates had a constant orientation, as is the case for Lemaˆıtre-Tolman (LT) models. In this paper we have shown that two independent results about Szekeres shell rotations are in full agreement. In FR [15] it was shown, using the rotation rate of an orthonormal tetrad, that the (θ,φ) coordinates of (12) do not in general retain any kind of constant orientation. In [6] the relative rotation of adjacent (θ,φ) shells was stated, and in PG [7] this was explained in several ways, notably an embedding that further introduced higher dimensional tilts. The forms of the results in FR & PG are sufficiently different that an alignment is called for. We have reviewed the embedding of a f < 0 Szekeres model in flat 4-d Euclidean space, and discussed its visualisation. It is striking that the embedding of a Szekeres model is locally the same to first order as that of the underlying LT model at the corresponding r value. By showing how to derive the FR results from the PG results, we have confirmed the reality of these rotations and tilts, and thereby their importance for graphing Szekeres

25 slices. Methods for this graphing are suggested in PG. To illustrate the higher dimensional tilts, we have constructed Szekeres models whose 3-spaces ‘naturally’ have the topology of a torus, when embedded in a 4-d Euclidean flat space, without arbitrary identifications. Explicit Szekeres models that are closed in the r direction, with or without an arbitrary identification, had not been been previously considered. Models 1 & 2 had just S varying, but nicely demonstrated the toroidal embedding as well as non-trivial radius and tilt structure. Models 3 & 4 had just P varying, and additionally demonstrated a non-planar curve of shell centres. Importantly, model 4 clearly exhibited the shell rotation described in PG. Clearly, models in which all 3 of the Szekeres functions S, P & Q vary could produce even more interesting embedded shapes. Some questions for future investigations are (i) Can one find a different (p, q) to (θ,φ) transformation that incorporates the rotation, (20), found by Buckley & Schlegel? Is it reasonably neat or too complicated? Does the resulting metric look at all useful or useable? (ii) Can one apply or extend the embedding to the DKS-type (“β′ =0”) Szekeres models? (iii) Is there a reasonably simple or elegant embedding of the ǫ = +1 Szekeres models? (iv) Is it possible to create a quasi-spherical Szekeres model in which6 the shell rotation turns the ‘north pole’ through 180 degrees? Can such a model be given a toroidal topology, of any kind? That would make the transformation to (θ,φ) coordinates clearly inconsistent.

References

[1] K. Bolejko, Phys. Rev. D 73, 123508 (2006), “Structure Formation in the Quasispherical Szekeres Model”.

[2] K. Bolejko, Phys. Rev. D 75, 043508 (2007), “Evolution of Cosmic Structures in Different Envi- ronments in the Quasispherical Szekeres Model”.

[3] K. Bolejko, J. Cosm. Astropart. Phys. 2017, 06:025 (2017), “Cosmological Backreaction Within the Szekeres Model and Emergence of Spatial Curvature”. arXiv:1704.02810 [astro-ph.CO].

[4] K. Bolejko, M.A. Nazer, D.L. Wiltshire, J. Cosm. Astropart. Phys. 2016, 06:035 (2016), “Differ- ential Cosmic Expansion and the Hubble Flow Anisotropy”. arXiv:1512.07364.

[5] K. Bolejko & R. Sussman, Phys. Lett. B 697, 265-70 (2011), “Cosmic Spherical Void Via Coarse- Graining and Averaging Non-Spherical Structures”. arXiv:1008.3420.

[6] R.G. Buckley & E.M. Schlegel, Phys. Rev. D 87, 023524, 1-13 (2013), “CMB Dipoles and Other Low-Order Multipoles in the Quasispherical Szekeres Model”. arXiv:1907.08684 [astro-ph.CO].

[7] R.G. Buckley & E.M. Schlegel, Phys. Rev. D 101, 023511, 1-26 (2020), “Physical Geometry of the Quasispherical Szekeres Models”. arXiv:1908.02697 [gr-qc].

[8] B. Datt, Zeit. Physik 108, 314 (1938), “Uber¨ eine Klasse von L¨osungen der Gravitationsgleichun- gen der Relativit¨at”. Reprinted in English with historical introduction in: Gen. Rel. Grav. 31, 1619-27 (1999).

[9] G.F.R. Ellis, J. Math. Phys. 8, 1171-94 (1967), “Dynamics of Pressure-Free Matter in General Relativity”.

26 [10] I.D. Gaspar, J.C. Hidalgo, R.A. Sussman & I. Quiros, Phys. Rev. D 97, 104029 (2018), “Black Hole Formation from the Gravitational Collapse of a Non-Spherical Network of Structures”. arXiv:1802.09123 [gr-qc]. [11] I. Georg & C. Hellaby, Phys. Rev. D 95, 124016, 1-16 (2017), “Symmetry and Equivalence in Szekeres Models”. arXiv:1702.05347. [12] S.W. Goode & J. Wainwright, Phys. Rev. D 26, 3315-26 (1982), “Singularities and Evolution of the Szekeres Cosmological Models”. [13] C. Hellaby, Class. Quantum Grav. 4, 635-650 (1987), “A Kruskal-Like Model with Finite Density”. [14] C. Hellaby, Proc. Sci. PoS(ISFTG), 005, 1-50 (2009), “Modelling Inhomogeneity in the Uni- verse”. https://pos.sissa.it/cgi-bin/reader/conf.cgi?confid=81. arXiv:0910.0350 [gr-qc]. [15] C. Hellaby, Class. Quantum Grav. 34, 145006, 1-20 (2017), “Frame Rotation in the Szekeres Spacetimes”. arXiv:1706.00622 [gr-qc]. [16] C. Hellaby & A. Krasi´nski, Phys. Rev. D 66, 084011, 1-27 (2002), “You Can’t Get Through Szekeres Wormholes: Regularity, Topology and Causality in Quasi-Spherical Szekeres Models”. arXiv:gr-qc/0206052. [17] C. Hellaby & A. Krasi´nski, Phys. Rev. D 77, 023529, 1-26 (2008), “Physical and Geometrical Interpretation of the ǫ 0 Szekeres Models”. arXiv:0710.2171 [gr-qc]. ≤ [18] C. Hellaby & K. Lake, Astrophys. J. 290, 381-387 (1985), “Shell Crossings and the Tolman Model”.; plus errata in Astrophys. J.300 461 1986. [19] R. Kantowski & R.K. Sachs, J. Math. Phys. 7, 443-6 (1966), “Some Spatially Homogeneous Anisotropic Relativistic Cosmological Models”. [20] A. Krasi´nski, Inhomogeneous Cosmological Models, Cambridge U P, 1997, ISBN 0 521 48180 5. [21] A. Krasi´nski & K. Bolejko, Phys. Rev. D 85, 124016 (2012), “Apparent Horizons in the Quasi- spherical Szekeres Models”. [22] G. Lemaˆıtre, Ann. Soc. Sci. Bruxelles A53, 51-85 (1933), “L’Universe en Expansion”. Reprinted in English with historical introduction in: Gen. Rel. Grav. 29, 641-80 (1997). [23] R.A. Sussman & I.D. Gaspar, Phys. Rev. D 92, 083533 (2015), “Multiple Non-Spherical Structures from the Extrema of Szekeres Scalars”. arXiv:1508.03127 [gr-qc]. [24] R.A. Sussman, I.D. Gaspar & J.C. Hidalgo, J. Cosm. Astropart. Phys. 2016, 03:012 (2016), “Coarse-Grained Description of Cosmic Structure from Szekeres Models”. Errata in: J. Cosm. Astropart. Phys. 2016, 06:E03 (2016). arXiv:1507.02306 [gr-qc]. [25] P. Szekeres, Comm. Math. Phys. 41, 55-64 (1975), “A Class of Inhomogeneous Cosmological Models”. [26] P. Szekeres, Phys. Rev. D 12, 2941-8 (1975), “Quasispherical Gravitational Collapse”. [27] R.C. Tolman, Proc. Nat. Acad. Sci. U.S.A. 20, 169-76 (1934), “Effect of Inhomogeneity on Cosmological Models”. Reprinted with historical introduction in: Gen. Rel. Grav. 29, 935-43 (1997).

27 [28] A. Walters and C. Hellaby, J. Cosm. Astropart. Phys. 2012, 12:001, 1-40 (2012), “Constructing Realistic Szekeres Models from Initial and Final Data”. arXiv:1211.2110 [gr-qc].

28