Differential Geometry – Is Time-Orientability a Condition on the Metric, Smooth, or Topological Structure of a Manifold

characteristic-classesdifferential-geometrygeneral-relativitymathematical physicssemi-riemannian-geometry

I recently asked a question on Physics Stack Exchange about orientability and time-orientability of a manifold in the language of fiber bundles. This new question is related to, but independent, of that one.

Orientability of an $n$-dimensional manifold (for example in the sense that there is an everywhere non-vanishing smooth $n$-form) is a topological property of manifolds. This can be seen through the theory of characteristic classes and the fact that it is equivalent to the first Stiefel–Whitney class of the manifold vanishing.

Time-orientability can be defined by requiring that there is an everywhere non-vanishing timelike vector on the manifold (which immediately makes use of the metric structure in order to define what "timelike" means). Alternatively, one can define a manifold is time-orientable if the bundle of orthonormal frames $\mathrm{O}(M)$ admits a reduction to an $\mathrm{O}^+(M)$ subbundle under the inclusion $\mathrm{O}^+(s,t) \subseteq \mathrm{O}(s,t)$.

My question is: when we say that a Lorentzian (or, more generally, pseudo-Riemannian) manifold $(M,g)$ is time-orientable, is this a restriction on the metric structure $(M,g)$, on the smooth structure of $M$, or on the topology of $M$?

Best Answer

Here’s a theorem from O’Neill’s Semi-Riemannian Geometry, With Applications to Relativity:

Theorem (O’Neill, Chapter 5, Lemma 37).

For a smooth manifold $M$, the following statements are equivalent:

  1. $M$ admits a Lorentzian metric
  2. $M$ admits a time-orientable Lorentzian metric
  3. $M$ admits a smooth nowhere-vanishing vector field
  4. $M$ is either non-compact or it is compact with vanishing Euler-characteristic: $\chi(M)=0$.

The proof is relatively short once you accept the equivalence of (3) and (4) by referring to some other source. So, the question of smooth manifolds admitting (time-orientable) Lorentzian metrics is essentially a question about the underlying topology.


Next, suppose $(M,g)$ is a Lorentzian manifold, and let $M_{\tau}$ be the set of all time cones in all the tangent spaces; then there is a canonical projection map $\pi:M_{\tau}\to M$, and it is possible to provide $M_{\tau}$ a topology and smooth structure such that $\pi$ becomes a smooth covering map, which is 2-1 (because there are exactly two time cones for each point). Then, $\pi$ is in particular a local diffeomorphism, so the pullback $g_{\tau}:=\pi^*g$ is a Lorentzian metric and with this, $\pi: (M_{\tau},g_{\tau})\to (M,g)$ becomes a Lorentzian covering called the time-orientation covering of $(M,g)$. Now, we have the relatively easy theorem:

Theorem (O’Neill, Chapter 7, Lemma 17).

Let $(M,g)$ be a Lorentzian manifold. Then,

  1. $(M_{\tau},g_{\tau})$ is time-orientable.
  2. $(M,g)$ is time-orientable if and only if $\pi:M_{\tau}\to M$ is trivializable.

Btw, you can also define a smooth action of $\Bbb{Z}_2$ (doesn’t matter if it’s right or left, since it’s Abelian) on $M_{\tau}$ where $-1$ acts by sending a timecone in $T_pM$ to the opposite timecone in $T_pM$. The orbit of a given point in $M_{\tau}$ (i.e of a given timecone) is simply the set of two timecones in a given tangent space, but this is precisely the fiber of $\pi$. Hence, by the universal property of quotient manifolds, we have $M\cong M_{\tau}/\Bbb{Z}_2$, and since the $\Bbb{Z}_2$ action on $M_{\tau}$ is free (obviously), it follows $(M_{\tau},\pi,M,\Bbb{Z}_2)$ is a principal bundle. Since $M_{\tau}$ is a principal bundle, the second statement is thus equivalent to $M_{\tau}$ admitting a global section. So, the theorem is saying that existence of a smooth timelike vector field in $(M,g)$ is equivalent to a smooth choice of timecones.

Also, there’s a relatively standard topological lemma (which you can google or try to prove yourself) that for a covering (say of manifolds) $\pi:X\to B$, if the base $B$ is simply connected, then $(X,\pi,B)$ is trivializable. Thus, we have the following corollary of (2) above:

Corollary.

Every simply-connected Lorentzian manifold is time-orientable.

Thus, for a given Lorentzian manifold, we have a sufficient condition on the topology which ensures time-orientability. But, I’m not aware of necessary and sufficient conditions (not that I’ve looked too hard).


Finally, it should also be noted that not every Lorentzian manifold is time-orientable, and indeed, once we consider manifolds which are not simply-connected, there are examples of Lorentzian metrics on the same manifold which are time-orientable, and also ones which are not time-orientable. In fact, even if we assume that the manifold is orientable (a topological assumption as you say), that still doesn’t really help us with time-orientability. So, time-orientability of a given Lorentzian metric is a combination of both topological and metric properties.

First let’s record the following easy lemmas (which are of inherent independent interest)

Lemma 1.

A product of smooth manifolds is orientable if and only if each factor is orientable.

By induction, it suffices to prove the case of two factors; call the manifolds $X,Y$, and let $m:=\dim Y$. If $X,Y$ are orientable, then letting $\mu_X,\mu_Y$ be respective volume forms, we have that $\mu:=\pi_X^*\mu_X\wedge\pi_Y^*\mu_Y$ will be a volume form. Conversely, suppose $X\times Y$ is orientable. Letting $Y_0$ be an open subset of $Y$ which is diffeomorphic to $\Bbb{R}^m$, we see that the restriction of any volume form on $X\times Y$ to $X\times Y_0$ is still a volume form. Hence, $X\times\Bbb{R}^m$ is orientable; let $\mu$ be a volume form here. Now, define $\mu_X:=\iota^*\left(\mu(\cdots, \eta_1,\dots, \eta_m)\right)$, where $\iota:X\to X\times\{0\}\subset X\times\Bbb{R}^m$ is the canonical embedding and for each $i\in\{1,\dots, m\}$, $\eta_i$ is the canonical lift of $\frac{\partial}{\partial y^i}$ on $\Bbb{R}^m$ to $X\times \Bbb{R}^m$. Then, $\mu_X$ is a volume-form on $X$, so $X$ is orientable. Reversing the roles of $X,Y$ in the proof above shows $Y$ is also orientable.

Lemma 2.

Let $(B,g_B)$ be a Lorentzian manifold, $(R,g_R)$ a Riemannian manifold, $\phi:B\to (0,\infty)$ any smooth function, and consider the warped Lorentzian manifold $M=B\times_{\phi}R$ (i.e $M=B\times R$ as a manifold, with metric $g=g_B+\phi^2 g_R$, or more precisely $g=\pi_B^*g_B+(\phi\circ \pi_B)^2 \pi_R^*g_R$). Then, $(B,g_B)$ is time-orientable if and only if $(M,g)$ is.

To prove the $(\implies)$ direction, take any timelike vector field $T$ on $B$, and define a vector field $\tau$ on $M=B\times R$ by \begin{align} \tau(b,r):=(T(b),0)\in T_bB\times T_rR\cong T_{(b,r)}(B\times R). \end{align} Then, obviously (suppressing the projections in the notation) we have $g(\tau,\tau)=g_B(T,T)<0$, so $\tau$ is timelike on $M$. Conversely, given any timelike vector field $\tau$ on $M$, we can separate it out into components $\tau(b,r)=(\tau_B(b,r),\tau_R(b,r))\in T_bB\times T_rR$ with $\tau_B,\tau_R$ being smooth vector fields on $M$. Observe that \begin{align} g_B(\tau_B,\tau_B)&=g(\tau,\tau)-\phi^2g_R(\tau_R,\tau_R)\leq g(\tau,\tau)<0, \end{align} where we have used that $g_R$ is Riemannian. So, if we fix $r\in R$ and define $T(b)=\tau_B(b,r)$, then the above calculation shows we have a smooth timelike vector field on $B$.

With these lemmas, it suffices to find 2-dimensional counterexamples, because to get one of higher dimension $n> 2$, we just take a product with an $(n-2)$-dimensional orientable Riemannian manifold (e.g $\Bbb{R}^{n-2}$ with usual positive-definite metric and orientation).

Counterexamples.

  • The open Mobius band $M$ (which is not an orientable manifold) admits a time-orientable Lorentzian metric, and also one which is not time-orientable.

  • The cylinder $S^1\times\Bbb{R}$ (which is an orientable manifold) admits a time-orientable Lorentzian metric, and also one which is not time-orientable.


Details for the Mobius band.

We claim the open Mobius band $M$ admits a time-orientable Lorentzian metric, and also one which is not time-orientable. Here are the details:

Let $\Bbb{Z}$ act on $\Bbb{R}^2$ by $n\cdot (x,y):=(x+n,(-1)^ny)$. This is a smooth, free and proper action (exercise), so by the orbit manifold theorem (see Lee’s Introduction to Smooth Manifolds, Theorem 21.10) the quotient $M:= \Bbb{R}^2/\Bbb{Z}$ equipped with the quotient topology carries a unique smooth structure such that the projection $\pi:\Bbb{R}^2\to M$ is a smooth submersion; this is one of the many possible definitions of the Mobius band (if you have some other definition, you can easily show it’s equivalent to this one). Now, let $X=\frac{\partial}{\partial x}$ on $\Bbb{R}^2$; this vector field is in fact $\Bbb{Z}$-equivariant as a map $X:\Bbb{R}^2\to T\Bbb{R}^2$ (relative to the given action on $\Bbb{R}^2$ and the induced tangent action on the tangent bundle), which implies that it descends uniquely to a smooth vector field $\widetilde{X}:M\to TM$ which is $\pi$-related to $X$. Intuitively, since the $x$-direction is simply being wrapped around, rather than being “twisted”, it follows that the $X$ vector field which points along the $x$-axis will have no problems descending to the quotient.

Now, consider the following Lorentzian metrics on $\Bbb{R}^2$: $g_1=-dx^2+dy^2$ and $g_2=dx^2-dy^2=-g_1$. Relative to both of these, the above action of $\Bbb{Z}$ on $\Bbb{R}^2$ is by isometries; and the fibers are discrete, so in particular semi-Riemannian manifolds in their own right, and so $g_1,g_2$ descend to Lorentzian metrics $\widetilde{g_1},\widetilde{g_2}$ on the quotient $M$, such that $g_i=\pi^*\widetilde{g_i}$ (Lee’s Introduction to Riemannian Manifolds, Corollary 2.29 covers the Riemannian case, but with minor modifications you can convince yourself the semi-Riemannian case holds as well). It is easily verified that \begin{align} \pi^*(\widetilde{g_i}(\widetilde{X},\widetilde{X}))=g_i(X,X)=(-1)^i. \end{align} So, for $i=1$, we see (by surjectivity of $\pi$) that $\widetilde{X}$ is a smooth timelike vector field on $(M,\widetilde{g}_1)$, and hence this Lorentzian metric on the Mobius band is time-orientable.

However, for $i=2$, the vector field $\widetilde{X}$ is spacelike, so supposing for the sake of contradiction there existed a smooth timelike vector field $T$ on $M$, it obviously can’t be a multiple of $\widetilde{X}$, and so $\{T,\widetilde{X}\}$ forms a smooth global frame for the tangent bundle. This implies $M$ is orientable as a manifold (since the bivector field $T\wedge\widetilde{X}$ is nowhere-vanishing, and thus by a musical isomorphism, we get a nowhere-vanishing 2-form on $M$), but this is absurd since we know the Mobius band is not an orientable manifold. Therefore, $(M,g_2)$ is not time-orientable.

For the sake of completeness, here’s a proof that $M$ is not an orientable manifold. Supposing it is, let $\mu$ be a volume form on $M$. Then $\pi^*\mu$ is a volume form on $\Bbb{R}^2$ since $\Bbb{Z}$ being discrete implies $\pi$ is a local diffeomorphism. Hence can be written as $f(x,y)\,dx\wedge dy$ for some smooth nowhere-vanishing function $f:\Bbb{R}^2\to\Bbb{R}$, which by the intermediate-value theorem must maintain aconstant sign. But now letting $\theta_n(x,y)=(x+n,(-1)^ny)$ be the action map of the integer $n$, we easily see by pulling back by $\theta_n$ that \begin{align} f\,dx\wedge dy&=\pi^*\mu=(\pi\circ\theta_n)^*\mu=\theta_n^*\pi^*\mu=\theta_n^*(f\,dx\wedge dy)=(-1)^n(f\circ \theta_n)\,dx\wedge dy, \end{align} which implies $f\circ \theta_n=(-1)^nf$, contradicting that $f$ maintains a constant sign (and in fact it says that if we move along the $x$-axis by $n$ units with $n$ odd, then $f$ flips sign, which coincides with our intuitive understanding of the ‘twist’ in the Mobius strip as well).


Details for the Cylinder.

The cylinder $M=S^1\times\Bbb{R}$ is obviously orientable and time-orientable with the metric $g=d\theta^2-dt^2$ (i.e the product of the usual positive-definite round metric on $S^1$ with the negative-definite $(\Bbb{R},-dt^2)$) because $\frac{\partial}{\partial t}$ is well-defined and timelike for $g$, so provides a time-orientation.

To give an example of a non-time-orientable Lorentzian metric on the cylinder, consider the following picture: enter image description here

The picture on the left is from O’Neill Chapter 5, figure 5 (page 145), and the picture drawn on the right is an example of a vector field in the covering space $\Bbb{R}^2$ of the cylinder. Intuitively, the red vector field rotates clockwise as we right along the $x$-axis, i.e the timecones are themselves rotating; and if we arrange the speed of rotation correctly, the timecones after moving $2\pi$ will be in the flipped order, so when we project to the quotient, we’ll get an inconsistency.

More precisely, on $\Bbb{R}^2$ with coordinates $(x,y)$, let us define the tensor field \begin{align} g&=\cos x\,dx^2-2\sin x\,dx\,dy-\cos x\,dy^2 = -(dx^2+dy^2)+ 2\left(\cos\frac{x}{2}\,dx-\sin\frac{x}{2}\,dy\right)^2, \end{align} where in the second equality I have used the half-angle formulae. Note that the first equality tells us $g$ is symmetric. Next, define the vector fields \begin{align} T=-\sin\left(\frac{x}{2}\right)\frac{\partial}{\partial x} -\cos\left(\frac{x}{2}\right)\frac{\partial}{\partial y},\quad\text{and}\quad X=\cos\left(\frac{x}{2}\right)\frac{\partial}{\partial x} -\sin\left(\frac{x}{2}\right)\frac{\partial}{\partial y}. \end{align} Then, using the second expression for $g$, it is easily seen that $g(T,T)=-1$ and $g(X,X)=1$, and with the first expression $g(T,X)=0$, proving that $g$ has Lorentzian signature and $\{T,X\}$ provides a global $g$-orthonormal frame with $T$ timelike and $X$ spacelike. Btw, the way I got these formulas is I first tried to write down a simple ‘rotating’ vector field like the red arrows above, and this easily gave me the formula for $T$. Then, I started with the general anszatz $g=\alpha\,dx^2+2\beta\,dx\,dy+\gamma\,dy^2$ and tried to choose $\alpha,\beta,\gamma$ to ensure $g(T,T)=-1$, and with a little guess-work I got the above formula, and lastly it was easy to see that $X$ as defined above satisfies $g(X,X)=1$ and $g(T,X)=0$, proving that $g$ is indeed Lorentzian.

Now, consider the action of $\Bbb{Z}$ on $\Bbb{R}^2$ defined as $n\cdot (x,y)\equiv \theta_n(x,y):= (x+2\pi n, y)$. The quotient is the cylinder: $M:=\Bbb{R}^2/_{\theta}\Bbb{Z}\cong (\Bbb{R}/2\pi \Bbb{Z})\times\Bbb{R}\cong S^1\times\Bbb{R}$. Also, the action consists of isometries of $g$: $\theta_n^*g=g$ (using the basic trigonometric facts that $\sin\left(\frac{x}{2}+n\pi\right)=(-1)^n\sin\left(\frac{x}{2}\right)$, and likewise for $\cos$, but this sign disappears because it is squared). Hence, $g$ descends to a Lorentzian metric $\widetilde{g}$ on the cylinder (if by abuse of notation you use local coordinates $(x,y)$ on $S^1\times \Bbb{R}$ with $x$ defined only mod $2\pi$, then it has the same formula as above). Also, it is easy to verify that $\theta_n^*T=(-1)^nT$ (the LHS being pullback of a vector field).

Suppose for the sake of contradiction $(M,\widetilde{g})$ is time-orientable, and let $\widetilde{T}$ be a smooth timelike vector field. Since $\Bbb{Z}$ is discrete, the projection $p:\Bbb{R}^2\to M$ is a covering map, and in particular a local diffeomorphism, so we can define a unique pullback vector field $\tau:=\pi^*\widetilde{T}$ (more precisely, there’s a unique vector field $\tau$ on $\Bbb{R}^2$ which is $p$-related to $\widetilde{T}$). Now, it is easy to verify using $g=p^*\widetilde{g}$ and that $\tau$ is $p$-related to $\widetilde{T}$ to prove that \begin{align} g(\tau,\tau)=p^*\left(\widetilde{g}(\widetilde{T},\widetilde{T})\right)<0, \end{align} and so $\tau$ is timelike. Also, it is straightforward to verify (keeping in mind $p\circ \theta_n=p$) that \begin{align} \theta_n^*\left(g(T,\tau)\right)&=g(\theta_n^*T,\tau)=(-1)^ng(T,\tau). \end{align} Now, since both $T$ an $\tau$ are timelike vector fields it follows $g(T,\tau)$ is nowhere-vanishing hence maintains a constant sign by the intermediate-value theorem. However, this contradicts the above equation which tells us that when $n$ is odd, $g(T,\tau)$ has opposite sign at $(x,y)$ and $\theta_n(x,y)=(x+2\pi n,y)$. Therefore, $(M,\widetilde{g})$ is not time-orientable.

This example specifically is nice in hindsight because the proof really does mimic the intuitive picture, and the contradiction reached via the intermediate-value theorem is a precise way of saying that when we travel $2\pi$ along the $x$-axis, the timecone has flipped.

Some final words about intuition: notice that in both of the examples above, the lack of time-orientability comes from the timecones in the quotient flipping direction as you travel around in a ‘single’ loop (which is only possible due to non simply-connectedness). But a key difference is the reason for the flip. In the Mobius band case, our timecones in the covering space $\Bbb{R}^2$ were the usual ‘upright’ ones, but in the quotient Mobius band, they flipped because the Mobius band itself is defined with a twist. In the cylinder case, the timecones in the covering space $\Bbb{R}^2$ themselves were the ones rotating, and since the quotient is obtained without a twist, it follows that the flip in the covering space still persists in the quotient. In both cases, formalizing the occurrence of the flip is done via the intermediate-value theorem.