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

Hilbert’s Theorem, via moving frames

William D. Dunbar
(September 18, 2025)
Abstract

We present a proof that the hyperbolic plane cannot be isometrically immersed in Euclidean 33-space by a CC^{\infty} map. Ideas from many topics in (essentially) undergraduate mathematics are applied; the use of moving frames and connection forms to express the geometry simplifies the outline of the proof, compared to, say, using coordinate patches and Christoffel symbols. The key transition is from expressions in terms of the principal directions on the immersed surface (which give access to the Gaussian curvature) to expressions in terms of the asymptotic directions (which yield a coordinate system and give access to surface area).

1 Outline of the proof

  • Assume for simplicity that ϕ:23\phi:\mathbb{H}^{2}\rightarrow\mathbb{R}^{3} is a CC^{\infty} isometric embedding (immersions are no more difficult). In this section, “the surface” will refer to the image of ϕ\phi.

  • Construct an orthonormal frame field {e1,e2}\{e_{1},e_{2}\} on 2\mathbb{H}^{2} which is mapped by ϕ\phi_{*} to unit vectors in the principal directions on the surface (maximizing and minimizing normal curvature).

  • Construct a frame field {E1,E2}\{E_{1},E_{2}\} on 2\mathbb{H}^{2} which is mapped by ϕ\phi_{*} to linearly independent unit vectors in asymptotic directions (where the normal curvature equals zero).

  • Express the connection 11-form ω12\omega_{1}^{2} (for {e1,e2}\{e_{1},e_{2}\}) first in terms of the 11-forms dual to {e1,e2}\{e_{1},e_{2}\} and the principal curvatures [Lemma 1], then in terms of the same forms and the angle α\alpha between either of the asymptotic directions and e1e_{1} [Lemma 2], and finally change basis to the 11-forms dual to {E1,E2}\{E_{1},E_{2}\} [equation (23)]. (The Gaussian curvature on 2\mathbb{H}^{2} being identically equal to 1-1 is used in the proof of Lemma 2.)

  • Show that the dual 11-forms {η1,η2}\{\eta^{1},\eta^{2}\} for {E1,E2}\{E_{1},E_{2}\} are closed forms, hence exact [Lemma 3]; furthermore, construct a (global!) coordinate chart mapping P2P\in\mathbb{H}^{2} to F(P)=(u1(P),u2(P)){(x1,x2)}=2F(P)=(u^{1}(P),u^{2}(P))\in\{(x^{1},x^{2})\}=\mathbb{R}^{2}, such that u1=E1\frac{\partial}{\partial u^{1}}=E_{1} and u2=E2\frac{\partial}{\partial u^{2}}=E_{2} [Proposition 3].

  • Define Θ\Theta to be the (oriented) angle between E1E_{1} and E2E_{2}; hence, sinΘ\sin\Theta is the area distortion factor when mapping from the x1x2x^{1}x^{2}-plane to 2\mathbb{H}^{2}. Show that sinΘ=2Θu1u2\sin\Theta=\frac{\partial^{2}\Theta}{\partial u^{1}\partial u^{2}} [equation (26)].

  • Calculate the area in 2\mathbb{H}^{2} corresponding to [a,a]×[a,a][-a,a]\times[-a,a] in the x1x2x^{1}x^{2}-plane, and show that as aa\to\infty, area is bounded above by 2π2\pi [equation (2.4)] …but also calculate the area of 2\mathbb{H}^{2} by other means to show that it is infinite [equation (2.4)]. This is a contradiction, so ϕ\phi cannot exist.

This approach to Hilbert’s theorem was inspired by the argument given by Rubens Leão in the Appendix of [dC83]. Citations will be supplied for equations that come from undergraduate differential geometry and/or from tensor analysis, but proofs will usually be omitted. For differential geometry, [O’N06] and [Shi18, Chapter 3, Section 3] use moving frames; [MP77] and [dC16] may also be helpful. Basic facts about differential forms in n\mathbb{R}^{n} (exterior differentiation, wedge product, pullback) can be found in [dC94, Chapter 1] and in [Spi79a, Chapter 7].

2 Proof of the main theorem

Theorem 1.

There does not exist a CC^{\infty} isometric immersion of the hyperbolic plane into Euclidean 33-space.

Before starting the proof, we make a foundational definition and a remark on notation.

Definition 1.

A moving frame on an open subset UU of n\mathbb{R}^{n} is an ordered nn-tuple of smooth (i.e., CC^{\infty}) vector fields {v1,,vn}\{v_{1},\dots,v_{n}\}, such that the vectors at each point pUp\in U form an orthonormal basis for the tangent space Tp(U)T_{p}(U).

The previous definition agrees with [dC83, page 118]; see also [dC94, page 77] and [O’N06, Chapter II, Section 1]. Elsewhere, such as in [Spi79b, page 285], the vectors in a moving frame need only be linearly independent (and span).

Remark.

We will use lower subscripts for contravariant objects such as vectors, and upper subscripts for their components with respect to a basis (as in the preceding definition). Similarly, we will use upper subscripts for covariant objects such as differential forms, and lower subscripts for their components with respect to a basis. Summation signs are needed only a few times, and they are not omitted, as they would be in “Einstein summation convention” [Spi79a, pages 50–51, pages 155–158].

\begin{overpic}[scale={0.36}]{AllMapsR9a.pdf} \put(12.0,4.0){$\mathbb{R}^{2}$} \put(62.0,4.0){$\mathbb{D}$} \put(88.0,4.0){$\mathbb{R}^{3}$} \put(26.5,18.0){\LARGE$\overset{F}{\longleftarrow}$} \put(69.5,18.0){\LARGE$\overset{\phi}{\longrightarrow}$} \put(19.0,20.5){$x^{1}$} \put(13.5,25.5){$x^{2}$} \put(46.0,14.0){$P$} \put(54.5,19.8){$e_{1}$} \put(44.5,22.5){$e_{2}$} \put(54.4,15.8){$E_{1}$} \put(50.3,22.3){$E_{2}$} \end{overpic}
Figure 1: Overview
Proof of Theorem 1.

The Poincaré disk 𝔻\mathbb{D} will represent the hyperbolic plane; define O:=(0,0)𝔻O:=(0,0)\in\mathbb{D} (see the Appendix for the Riemannian metric and other details). Suppose that ϕ:𝔻3\phi:\mathbb{D}\longrightarrow\mathbb{R}^{3} is a CC^{\infty} isometric immersion.

2.1 Construct vector fields and 11-forms on 𝔻\mathbb{D}

For all P𝔻P\in\mathbb{D}, there is a positive number ϵP\epsilon_{P}, such that ϕ\phi is an embedding, when restricted to the open disk UPU_{P} of radius ϵP\epsilon_{P} in the hyperbolic metric. The tangent space to the surface ϕ(UP)=:U¯P\phi(U_{P})=:\bar{U}_{P} at ϕ(P)=:P¯\phi(P)=:\bar{P} can inherit an orientation from TP(𝔻)T_{P}(\mathbb{D}), via (ϕ)P(\phi_{*})_{P}, so there is a preferred unit normal vector at P¯\bar{P}, which can be continuously (indeed, smoothly) extended to a unit normal vector field e¯3:U¯PS2\bar{e}_{3}:\bar{U}_{P}\rightarrow S^{2}.

At every point QUPQ\in U_{P}, the map (de¯3)Q¯:TQ¯(U¯P)Te¯3(Q¯)S2(-d\bar{e}_{3})_{\bar{Q}}:T_{\bar{Q}}(\bar{U}_{P})\rightarrow T_{\bar{e}_{3}(\bar{Q})}S^{2}, is a self-adjoint linear transformation of TQ¯(U¯P)=ϕ(TQ(𝔻))T_{\bar{Q}}(\bar{U}_{P})=\phi_{*}(T_{Q}(\mathbb{D})) [dC16, page 142]; note that Te¯3(Q¯)S2T_{\bar{e}_{3}(\bar{Q})}S^{2} and TQ¯(U¯P)T_{\bar{Q}}(\bar{U}_{P}) are both names for the orthogonal complement in 3\mathbb{R}^{3} of span(e¯3(Q¯))\operatorname{span}(\bar{e}_{3}(\bar{Q})). The eigenvalues are equal to the principal curvatures (let κ¯1(Q¯)\bar{\kappa}_{1}(\bar{Q}) denote the positive eigenvalue, and κ¯2(Q¯)\bar{\kappa}_{2}(\bar{Q}) the negative eigenvalue), and the corresponding (one-dimensional, orthogonal) eigenspaces L¯1(Q¯),L¯2(Q¯)\bar{L}_{1}(\bar{Q}),\bar{L}_{2}(\bar{Q}) give the principal directions [dC16, page 146]. At this stage, we are using the fact that the Gaussian curvature (product of the principal curvatures) is negative at every point of U¯P\bar{U}_{P}, but not (yet) the fact that it is constant and equal to 1-1.

Via (ϕ|UP)1(\phi|_{U_{P}})_{*}^{-1}, we can transfer these eigenspaces back to form two smooth line fields on each UPU_{P}. Whenever RUPUQR\in U_{P}\cap U_{Q}, the line fields for UPU_{P} and UQU_{Q} will agree at RR, since they both describe the extrinsic geometry of ϕ(UPUQ)\phi(U_{P}\cap U_{Q}) at ϕ(R)\phi(R). Hence we can merge all of the “local” line fields into two “global” line fields L1,L2L_{1},L_{2} on 𝔻\mathbb{D}. There is a two-sheeted covering space of 𝔻\mathbb{D}, consisting of all the unit vectors in L1L_{1}, and since 𝔻\mathbb{D} is simply-connected, this covering space is not connected. In other words, there are two unit vector fields ±e1\pm e_{1} on 𝔻\mathbb{D} which everywhere point along L1L_{1}. We define e2e_{2} to be the result of rotating e1e_{1} a quarter-turn counter-clockwise (i.e., +π/2+\pi/2 radians).

This allows us to define e¯1:=ϕ(e1)\bar{e}_{1}:=\phi_{*}(e_{1}) and e¯2:=ϕ(e2)\bar{e}_{2}:=\phi_{*}(e_{2}) on each U¯P\bar{U}_{P}, so we now have a (positively oriented) moving frame {e¯1,e¯2,e¯3}\{\bar{e}_{1},\bar{e}_{2},\bar{e}_{3}\} on U¯P\bar{U}_{P}. As in [dC94, page 82], we can extend this moving frame to a small open subset WPW_{P} of 3\mathbb{R}^{3} by displacing frames normally by a sufficiently small amount (choose νP>0\nu_{P}>0 such that UP×(νP,νP)WPU_{P}\times(-\nu_{P},\nu_{P})\rightarrow W_{P} by (Q,t)Q¯+t[e¯3]Q¯(Q,t)\mapsto\bar{Q}+t\cdot[\bar{e}_{3}]_{\bar{Q}} is a diffeomorphism). See Figure 2 for an illustration.

Since de¯1,de¯2,de¯3d\bar{e}_{1},d\bar{e}_{2},d\bar{e}_{3} are linear transformations on 3\mathbb{R}^{3} at every point of WPW_{P} (in other words, they are 3\mathbb{R}^{3}-valued 11-forms, as defined in [Spi79a, page 546], while e¯1,e¯2,e¯3\bar{e}_{1},\bar{e}_{2},\bar{e}_{3} are 3\mathbb{R}^{3}-valued 0-forms), we can define the connection 11-forms ω¯jk(1j,k3)\bar{\omega}^{k}_{j}\ (1\leq j,k\leq 3) by

de¯j(v)=k=13ω¯jk(v)e¯k,d\bar{e}_{j}(v)=\sum_{k=1}^{3}\bar{\omega}^{k}_{j}(v)\ \bar{e}_{k}, (1)

for all vector fields vv on WPW_{P} [dC94, page 78] [Spi79b, page 287]. The notation ω¯jk\bar{\omega}_{j}^{k} follows [Spi79b, page 287] and corresponds to ωjk\omega_{jk} in [dC94, page 78].

On each WPW_{P} we can also define 11-forms θ¯1,θ¯2,θ¯3\bar{\theta}^{1},\bar{\theta}^{2},\bar{\theta}^{3} dual to the frame, that is,

θ¯i(e¯j)=δji;1i,j3\bar{\theta}^{i}(\bar{e}_{j})=\delta^{i}_{j};\quad 1\leq i,j\leq 3 (2)

(using the Kronecker delta). These 11-forms on WPW_{P} are related to each other by the Cartan structure equations, which follow.

Proposition 1.

For all 1i,j31\leq i,j\leq 3, and for any moving frame on an open subset of 3\mathbb{R}^{3}, with connection forms and dual forms defined as in equations (1) and (2),

dθ¯i\displaystyle d\bar{\theta}^{i} =k=13ω¯kiθ¯k\displaystyle=-\sum_{k=1}^{3}\bar{\omega}^{i}_{k}\wedge\bar{\theta}^{k} (3)
dω¯ji\displaystyle d\bar{\omega}^{i}_{j} =k=13ω¯kiω¯jk\displaystyle=-\sum_{k=1}^{3}\bar{\omega}^{i}_{k}\wedge\bar{\omega}^{k}_{j} (4)
Proof.

See [Spi79b, page 287] or [dC94, page 78]. The former obtains (3) by expanding 0=d2I0=d^{2}I as a sum of scalar multiples of the {e¯i}\{\bar{e}_{i}\}, where I:33I:\mathbb{R}^{3}\longrightarrow\mathbb{R}^{3} is the identity map (and dI=i=13θ¯ie¯idI=\sum_{i=1}^{3}\bar{\theta}^{i}\wedge\bar{e}_{i}, here an equation of 3\mathbb{R}^{3}-valued 11-forms). Similarly, (4) comes from 0=d2e¯j0=d^{2}\bar{e}_{j}. ∎

\begin{overpic}[height=126.47249pt]{CloseupU1bT.pdf} \put(31.0,15.0){$\bar{e}_{1}$} \put(68.0,31.0){$\bar{e}_{2}$} \put(45.0,47.0){$\bar{e}_{3}$} \put(53.0,34.0){$\bar{P}$} \end{overpic}
Refer to caption
Figure 2: At left, U¯P=ϕ(UP)\bar{U}_{P}=\phi(U_{P}) and frame at P¯=ϕ(P)\bar{P}=\phi(P); at right, U¯PWP\bar{U}_{P}\subset W_{P}.

Now, on each UPU_{P}, we can define the 11-forms θi:=ϕθ¯i\theta^{i}:=\phi^{*}\bar{\theta}^{i}, 1i31\leq i\leq 3 (noting that θ¯3\bar{\theta}^{3} pulls back to the zero 11-form) and ωji:=ϕω¯ji\omega^{i}_{j}:=\phi^{*}\bar{\omega}^{i}_{j}, 1i,j31\leq i,j\leq 3. The reader can easily check that θ1,θ2\theta^{1},\theta^{2} are dual to e1,e2e_{1},e_{2}.

Remark.

Orthonormality of the moving frame {e¯1,e¯2,e¯3}\{\bar{e}_{1},\bar{e}_{2},\bar{e}_{3}\} is, in fact, not required for equations (3) and (4); linear independence of the frame field suffices. However, orthonormality implies the relations ω¯ij=ω¯ji\bar{\omega}^{j}_{i}=-\bar{\omega}^{i}_{j}, hence ωij=ωji\omega^{j}_{i}=-\omega^{i}_{j}, and in particular, ωii=0\omega^{i}_{i}=0 [Spi79b, page 292] [dC94, page 78].

Since {e1,e2}\{e_{1},e_{2}\} are globally defined on 𝔻\mathbb{D}, the moving frames {e¯1,e¯2,e¯3}\{\bar{e}_{1},\bar{e}_{2},\bar{e}_{3}\} at R¯U¯PU¯Q\bar{R}\in\bar{U}_{P}\cap\bar{U}_{Q}, arising respectively from ϕ|UP\phi|_{U_{P}} and ϕ|UQ\phi|_{U_{Q}}, will agree. Hence the 11-forms θi\theta^{i} and ωji\omega^{i}_{j}, defined above, will agree on UPUQU_{P}\cap U_{Q}. Consequently, as we did above with the line fields, we merge the “local” θ\theta’s and ω\omega’s into “global” smooth 11-forms that are defined on 𝔻\mathbb{D}.

Furthermore, since the “unbarred” versions of equations (3) and (4) can be justified by basic facts about pullbacks and wedge products in each UPU_{P}, Proposition 1 implies the following equations on 𝔻\mathbb{D} (or more precisely, in the exterior algebra of CC^{\infty} differential forms on 𝔻\mathbb{D}):

{dθ1=(ω11θ1+ω21θ2+ω31θ3)=0θ1ω21θ2ω310=ω12θ2dθ2=(ω12θ1+ω22θ2+ω32θ3)==ω12θ1\begin{cases}d\theta^{1}&=-(\omega^{1}_{1}\wedge\theta^{1}+\omega^{1}_{2}\wedge\theta^{2}+\omega^{1}_{3}\wedge\theta^{3})=-0\wedge\theta^{1}-\omega^{1}_{2}\wedge\theta^{2}-\omega^{1}_{3}\wedge 0=\omega_{1}^{2}\wedge\theta^{2}\\ d\theta^{2}&=-(\omega^{2}_{1}\wedge\theta^{1}+\omega^{2}_{2}\wedge\theta^{2}+\omega^{2}_{3}\wedge\theta^{3})=\dots=-\omega_{1}^{2}\wedge\theta^{1}\end{cases} (5)
{dω12=(ω12ω11+ω22ω12+ω32ω13)=ω23ω13=ω13ω23dω13=(ω13ω11+ω23ω12+ω33ω13)=ω23ω12=ω12ω23dω23=(ω13ω21+ω23ω22+ω33ω23)=ω13ω12=ω12ω13\begin{cases}d\omega_{1}^{2}&=-(\omega_{1}^{2}\wedge\omega_{1}^{1}+\omega_{2}^{2}\wedge\omega_{1}^{2}+\omega_{3}^{2}\wedge\omega_{1}^{3})=\omega_{2}^{3}\wedge\omega_{1}^{3}=-\omega_{1}^{3}\wedge\omega_{2}^{3}\\ d\omega_{1}^{3}&=-(\omega_{1}^{3}\wedge\omega_{1}^{1}+\omega_{2}^{3}\wedge\omega_{1}^{2}+\omega_{3}^{3}\wedge\omega_{1}^{3})=-\omega_{2}^{3}\wedge\omega_{1}^{2}=\omega_{1}^{2}\wedge\omega_{2}^{3}\\ d\omega_{2}^{3}&=-(\omega_{1}^{3}\wedge\omega_{2}^{1}+\omega_{2}^{3}\wedge\omega_{2}^{2}+\omega_{3}^{3}\wedge\omega_{2}^{3})=\omega_{1}^{3}\wedge\omega_{1}^{2}=-\omega_{1}^{2}\wedge\omega_{1}^{3}\end{cases} (6)

Also, by definition, in each U¯P\bar{U}_{P},

de¯3(e¯1)=κ¯1e¯1 and de¯3(e¯2)=κ¯2e¯2d\bar{e}_{3}(\bar{e}_{1})=-\bar{\kappa}_{1}\bar{e}_{1}\textrm{ and }d\bar{e}_{3}(\bar{e}_{2})=-\bar{\kappa}_{2}\bar{e}_{2} (7)

[BL10, page 179], [MP77, pages 125–129], [Spi79b, pages 103–106]. From the first equation we have, using (1), ω¯31(e¯1)=κ¯1\bar{\omega}^{1}_{3}(\bar{e}_{1})=-\bar{\kappa}_{1} and ω¯32(e¯1)=0\bar{\omega}^{2}_{3}(\bar{e}_{1})=0; similarly, the second equation implies ω¯31(e¯2)=0\bar{\omega}^{1}_{3}(\bar{e}_{2})=0 and ω¯32(e¯2)=κ¯2\bar{\omega}^{2}_{3}(\bar{e}_{2})=-\bar{\kappa}_{2}. We conclude, after pullback to 𝔻\mathbb{D}, with κi:=κ¯iϕ(i=1,2)\kappa_{i}:=\bar{\kappa}_{i}\circ\phi\ (i=1,2), that

ω31=κ1θ1 and ω32=κ2θ2, so ω13=κ1θ1 and ω23=κ2θ2.\omega_{3}^{1}=-\kappa_{1}\theta^{1}\textrm{ and }\omega_{3}^{2}=-\kappa_{2}\theta^{2},\textrm{ so }\omega_{1}^{3}=\kappa_{1}\theta^{1}\textrm{ and }\omega_{2}^{3}=\kappa_{2}\theta^{2}. (8)
Remark.

To keep the signs straight, it is useful to keep in mind the example of the unit sphere (with one point removed, so that there exists a moving frame), oriented with outward normal. Then e¯3\bar{e}_{3} equals pp at pS2p\in S^{2}, so de¯3d\bar{e}_{3} is the identity map. The normal curvatures at each point all equal 1-1 (negative since curves on the sphere bend away from the normal vector). So for this surface the principal curvatures both equal 1-1, which is consistent with equation (7).

2.2 Expressions for ω12\omega_{1}^{2} in terms of {θ1,θ2}\{\theta^{1},\theta^{2}\}

We now turn to the proofs of two key lemmas, which relate ω12\omega_{1}^{2} to the principal curvatures and to the angle between the asymptotic directions and e1e_{1}.

Lemma 1.

With {e1,e2}\{e_{1},e_{2}\}, {θ1,θ2}\{\theta^{1},\theta^{2}\}, and ω12,κ1,κ2\omega_{1}^{2},\kappa_{1},\kappa_{2} defined as above,

(κ1κ2)ω12=dκ1(e2)θ1+dκ2(e1)θ2.(\kappa_{1}-\kappa_{2})\omega_{1}^{2}=d\kappa_{1}(e_{2})\theta^{1}+d\kappa_{2}(e_{1})\theta^{2}. (9)
Proof.

To verify this equation of 11-forms, it suffices to show the following two equations:

(κ1κ2)ω12(e1)=?dκ1(e2) and (κ1κ2)ω12(e2)=?dκ2(e1)(\kappa_{1}-\kappa_{2})\omega_{1}^{2}(e_{1})\overset{?}{=}d\kappa_{1}(e_{2})\textrm{ and }(\kappa_{1}-\kappa_{2})\omega_{1}^{2}(e_{2})\overset{?}{=}d\kappa_{2}(e_{1}) (10)

We rewrite dω13d\omega_{1}^{3} in two ways, using equations (5), (6), and (8), obtaining

κ1ω12θ2+dκ1θ1=(5)κ1dθ1+dκ1θ1=(8)dω13=(6)ω12ω23=(8)κ2ω12θ2.\kappa_{1}\omega_{1}^{2}\wedge\theta^{2}+d\kappa_{1}\wedge\theta^{1}\overset{\eqref{Equ:ddual}}{=}\kappa_{1}d\theta^{1}+d\kappa_{1}\wedge\theta^{1}\overset{\eqref{Equ:princurv}}{=}d\omega_{1}^{3}\overset{\eqref{Equ:dconn}}{=}\omega_{1}^{2}\wedge\omega_{2}^{3}\overset{\eqref{Equ:princurv}}{=}\kappa_{2}\omega_{1}^{2}\wedge\theta^{2}.

In other words, (κ1κ2)ω12θ2=dκ1θ1(\kappa_{1}-\kappa_{2})\omega_{1}^{2}\wedge\theta^{2}=-d\kappa_{1}\wedge\theta^{1}. Evaluating both sides of this equation on the pair (e1e_{1},e2e_{2}), we obtain

(κ1κ2)ω12θ2(e1,e2)\displaystyle(\kappa_{1}-\kappa_{2})\omega_{1}^{2}\wedge\theta^{2}(e_{1},e_{2}) =(κ1κ2)(ω12(e1)θ2(e2)ω12(e2)θ2(e1))=(κ1κ2)ω12(e1)\displaystyle=(\kappa_{1}-\kappa_{2})(\omega_{1}^{2}(e_{1})\ \theta^{2}(e_{2})-\omega_{1}^{2}(e_{2})\ \theta^{2}(e_{1}))=(\kappa_{1}-\kappa_{2})\omega_{1}^{2}(e_{1})
and dκ1θ1(e1,e2)\displaystyle\textrm{ and }-d\kappa_{1}\wedge\theta^{1}(e_{1},e_{2}) =dκ1(e1)θ1(e2)+dκ1(e2)θ1(e1)=dκ1(e2),\displaystyle=-d\kappa_{1}(e_{1})\ \theta^{1}(e_{2})+d\kappa_{1}(e_{2})\ \theta^{1}(e_{1})=d\kappa_{1}(e_{2}),

which justifies the first half of (10). Similarly, rewriting dω23d\omega_{2}^{3},

κ2ω12θ1+dκ2θ2=(5)κ2dθ2+dκ2θ2=(8)dω23=(6)ω12ω13=(8)κ1ω12θ1.-\kappa_{2}\omega_{1}^{2}\wedge\theta^{1}+d\kappa_{2}\wedge\theta^{2}\overset{\eqref{Equ:ddual}}{=}\kappa_{2}d\theta^{2}+d\kappa_{2}\wedge\theta^{2}\overset{\eqref{Equ:princurv}}{=}d\omega_{2}^{3}\overset{\eqref{Equ:dconn}}{=}-\omega_{1}^{2}\wedge\omega_{1}^{3}\overset{\eqref{Equ:princurv}}{=}-\kappa_{1}\omega_{1}^{2}\wedge\theta^{1}.

In other words, (κ1κ2)ω12θ1=dκ2θ2-(\kappa_{1}-\kappa_{2})\omega_{1}^{2}\wedge\theta^{1}=d\kappa_{2}\wedge\theta^{2}. Again evaluating the left-hand and right-hand sides at (e1e_{1},e2e_{2}),

(κ1κ2)ω12θ1(e1,e2)\displaystyle-(\kappa_{1}-\kappa_{2})\omega_{1}^{2}\wedge\theta^{1}(e_{1},e_{2}) ==(κ1κ2)ω12(e2)\displaystyle=\dots=(\kappa_{1}-\kappa_{2})\omega_{1}^{2}(e_{2})
and dκ2θ2(e1,e2)\displaystyle\textrm{ and }d\kappa_{2}\wedge\theta^{2}(e_{1},e_{2}) ==dκ2(e1),\displaystyle=\dots=d\kappa_{2}(e_{1}),

which justifies the second half of (10). ∎

For all P𝔻P\in\mathbb{D}, the normal curvatures at ϕ(P)\phi(P) can be described by II(v):=de¯3(ϕ(v)),ϕ(v)\operatorname{II}(v):=\langle-d\bar{e}_{3}(\phi_{*}(v)),\phi_{*}(v)\rangle, defined in terms of the Euclidean inner product for vectors in 3\mathbb{R}^{3} [dC94, page 87]. In essence, this is the second fundamental form for the immersed surface near ϕ(P)\phi(P), pulled back to a quadratic form on TP(𝔻)T_{P}(\mathbb{D}). Using Euler’s theorem [dC16, page 147] [MP77, page 129], and (for the first time) making use of the fact that Gaussian curvature K:=κ1κ2K:=\kappa_{1}\kappa_{2} equals 1-1 everywhere on 𝔻\mathbb{D}, we have normal curvature zero when

0=II(v)=κ1cos2α+κ2sin2α=κ1cos2α(1/κ1)sin2α0=\operatorname{II}(v)=\kappa_{1}\cos^{2}\alpha+\kappa_{2}\sin^{2}\alpha=\kappa_{1}\cos^{2}\alpha-(1/\kappa_{1})\sin^{2}\alpha

for a unit vector vv that makes an angle ±α\pm\alpha with e1e_{1}. We conclude that tan2α=κ12\tan^{2}\alpha=\kappa_{1}^{2}, so set α:=arctan(κ1)(0,π/2)\alpha:=\arctan(\kappa_{1})\in(0,\pi/2), and define the asymptotic directions

E1:=cosαe1sinαe2 and E2:=cosαe1+sinαe2.E_{1}:=\cos\alpha\cdot e_{1}-\sin\alpha\cdot e_{2}\textrm{ and }E_{2}:=\cos\alpha\cdot e_{1}+\sin\alpha\cdot e_{2}. (11)

It follows that

e1=12secα(E1+E2) and e2=12cscα(E1+E2).e_{1}=\frac{1}{2}\sec\alpha(E_{1}+E_{2})\textrm{ and }e_{2}=\frac{1}{2}\csc\alpha(-E_{1}+E_{2}). (12)
Lemma 2.

With {e1,e2}\{e_{1},e_{2}\}, {θ1,θ2}\{\theta^{1},\theta^{2}\}, and ω12,α\omega_{1}^{2},\alpha defined as above,

ω12=tanαdα(e2)θ1+cotαdα(e1)θ2\omega_{1}^{2}=\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}+\cot\alpha\,d\alpha(e_{1})\cdot\theta^{2} (13)
Proof.

Since κ2=1/κ1=1/tanα=cotα\kappa_{2}=-1/\kappa_{1}=-1/\tan\alpha=-\cot\alpha,

κ1κ2=sinαcosα+cosαsinα=1cosαsinα.\kappa_{1}-\kappa_{2}=\frac{\sin\alpha}{\cos\alpha}+\frac{\cos\alpha}{\sin\alpha}=\frac{1}{\cos\alpha\,\sin\alpha}.

Therefore, using dκ1=d(tanα)=sec2αdαd\kappa_{1}=d(\tan\alpha)=\sec^{2}\alpha\,d\alpha and dκ2=d(cotα)=csc2αdαd\kappa_{2}=d(-\cot\alpha)=\csc^{2}\alpha\,d\alpha,

dκ1κ1κ2\displaystyle\frac{d\kappa_{1}}{\kappa_{1}-\kappa_{2}} =(sec2αdα)(cosαsinα)=tanαdα\displaystyle=(\sec^{2}\alpha\,d\alpha)(\cos\alpha\,\sin\alpha)=\tan\alpha\,d\alpha (14)
dκ2κ1κ2\displaystyle\frac{d\kappa_{2}}{\kappa_{1}-\kappa_{2}} =(csc2αdα)(cosαsinα)=cotαdα\displaystyle=(\csc^{2}\alpha\,d\alpha)(\cos\alpha\,\sin\alpha)=\cot\alpha\,d\alpha (15)

After evaluating (14) at e2e_{2} and (15) at e1e_{1}, the desired conclusion follows from Lemma 1. ∎

2.3 Use {E1,E2}\{E_{1},E_{2}\} to construct a global coordinate map

We now turn our attention to the properties of the asymptotic vector fields E1,E2E_{1},E_{2}.

Lemma 3.

Let {η1,η2}\{\eta^{1},\eta^{2}\} be the 11-forms which are dual to {E1,E2}\{E_{1},E_{2}\}. Then dη1=0d\eta^{1}=0 and dη2=0d\eta^{2}=0.

Proof.

It is not hard to use the definitions of E1E_{1} and E2E_{2} in (11) to verify that

η1=12[secαθ1cscαθ2] and η2=12[secαθ1+cscαθ2]\eta^{1}=\frac{1}{2}[\sec\alpha\cdot\theta^{1}-\csc\alpha\cdot\theta^{2}]\textrm{ and }\eta^{2}=\frac{1}{2}[\sec\alpha\cdot\theta^{1}+\csc\alpha\cdot\theta^{2}] (16)

Next, we see that

d(secαθ1)\displaystyle d(\sec\alpha\cdot\theta^{1}) =(5)secαtanαdαθ1+secα(ω12θ2)\displaystyle\overset{\eqref{Equ:ddual}}{=}\sec\alpha\,\tan\alpha\,d\alpha\wedge\theta^{1}+\sec\alpha(\omega_{1}^{2}\wedge\theta^{2})
=secαtanα(dα(e1)θ1+dα(e2)θ2)θ1+secα(ω12θ2)\displaystyle=\sec\alpha\,\tan\alpha\,(d\alpha(e_{1})\cdot\theta^{1}+d\alpha(e_{2})\cdot\theta^{2})\wedge\theta^{1}+\sec\alpha(\omega_{1}^{2}\wedge\theta^{2})
=secαtanαdα(e2)θ1θ2+secα(ω12θ2)\displaystyle=-\sec\alpha\,\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}\wedge\theta^{2}+\sec\alpha(\omega_{1}^{2}\wedge\theta^{2})
=(13)secαtanαdα(e2)θ1θ2\displaystyle\overset{\eqref{Equ:conntancot}}{=}-\sec\alpha\,\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}\wedge\theta^{2}
+secα((tanαdα(e2)θ1+cotαdα(e1)θ2)θ2)\displaystyle\qquad+\sec\alpha((\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}+\cot\alpha\,d\alpha(e_{1})\cdot\theta^{2})\wedge\theta^{2})
=secαtanαdα(e2)θ1θ2+secα(tanαdα(e2)θ1θ2)=0\displaystyle=-\sec\alpha\,\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}\wedge\theta^{2}+\sec\alpha(\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}\wedge\theta^{2})=0 (17)

and similarly

d(cscαθ2)\displaystyle d(\csc\alpha\cdot\theta^{2}) =(5)cscαcotαdαθ2+cscα(ω12θ1)\displaystyle\overset{\eqref{Equ:ddual}}{=}-\csc\alpha\,\cot\alpha\,d\alpha\wedge\theta^{2}+\csc\alpha(-\omega_{1}^{2}\wedge\theta^{1})
=cscαcotα(dα(e1)θ1+dα(e2)θ2)θ2+cscα(ω12θ1)\displaystyle=-\csc\alpha\,\cot\alpha\,(d\alpha(e_{1})\cdot\theta^{1}+d\alpha(e_{2})\cdot\theta^{2})\wedge\theta^{2}+\csc\alpha(-\omega_{1}^{2}\wedge\theta^{1})
=cscαcotαdα(e1)θ1θ2cscα(ω12θ1)\displaystyle=-\csc\alpha\,\cot\alpha\,d\alpha(e_{1})\cdot\theta^{1}\wedge\theta^{2}-\csc\alpha(\omega_{1}^{2}\wedge\theta^{1})
=(13)cscαcotαdα(e1)θ1θ2\displaystyle\overset{\eqref{Equ:conntancot}}{=}-\csc\alpha\,\cot\alpha\,d\alpha(e_{1})\cdot\theta^{1}\wedge\theta^{2}
cscα((tanαdα(e2)θ1+cotαdα(e1)θ2)θ1)\displaystyle\qquad-\csc\alpha((\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}+\cot\alpha\,d\alpha(e_{1})\cdot\theta^{2})\wedge\theta^{1})
=cscαcotαdα(e1)θ1θ2cscα(cotαdα(e1)(θ1θ2))=0\displaystyle=-\csc\alpha\,\cot\alpha\,d\alpha(e_{1})\cdot\theta^{1}\wedge\theta^{2}-\csc\alpha(\cot\alpha\,d\alpha(e_{1})\cdot(-\theta^{1}\wedge\theta^{2}))=0 (18)

By subtracting and adding equations (2.3) and (2.3), and then comparing to equation (16), we conclude that dη1=0d\eta^{1}=0 and dη2=0d\eta^{2}=0. ∎

We next show that the vector fields E1E_{1} and E2E_{2} on 𝔻\mathbb{D} each generate a global flow (i.e., each integral curve is defined for all tt\in\mathbb{R}). We will use the following special case of [Cod61, Chapter 6, §6, Theorem 2].

Theorem 2.

Let a,ba,b be positive numbers, let t0t_{0}\in\mathbb{R}, let 𝐲𝟎2\mathbf{y_{0}}\in\mathbb{R}^{2}, and let R:={(t,𝐲):|tt0|a,|𝐲𝐲𝟎|b}R:=\{(t,\mathbf{y}):|t-t_{0}|\leq a,|\mathbf{y}-\mathbf{y_{0}}|\leq b\}. If

  1. 1.

    f:R2f:R\rightarrow\mathbb{R}^{2} is a continuous function, and

  2. 2.

    there exists L>0L>0 such that for all (t,𝐲𝟏),(t,𝐲𝟐)R(t,\mathbf{y_{1}}),(t,\mathbf{y_{2}})\in R, |f(t,𝐲𝟏)f(t,𝐲𝟐)|L|𝐲𝟏𝐲𝟐||f(t,\mathbf{y_{1}})-f(t,\mathbf{y_{2}})|\leq L|\mathbf{y_{1}}-\mathbf{y_{2}}|, and

  3. 3.

    there exists M>0M>0 such that for all (t,𝐲)R(t,\mathbf{y})\in R, |f(t,𝐲)|M|f(t,\mathbf{y})|\leq M,

then the initial-value problem d𝐲/dt=f(t,𝐲),𝐲(t0)=𝐲𝟎d\mathbf{y}/dt=f(t,\mathbf{y}),\ \mathbf{y}(t_{0})=\mathbf{y_{0}} has a solution

𝐲=γ(t),γ:(t0ϵ,t0+ϵ){𝐲2:|𝐲𝐲𝟎|b},\mathbf{y}=\gamma(t),\ \gamma:(t_{0}-\epsilon,t_{0}+\epsilon)\rightarrow\{\mathbf{y}\in\mathbb{R}^{2}:|\mathbf{y}-\mathbf{y_{0}}|\leq b\},

where ϵ:=min(a,b/M)\epsilon:=\min(a,b/M). ∎

Note that ϵ\epsilon does not depend on LL, the Lipschitz constant.

Proposition 2.

If EE is a CC^{\infty} vector field on 𝔻\mathbb{D} which everywhere has unit length with respect to the hyperbolic metric, then every maximal integral curve of EE is defined for all tt\in\mathbb{R}.

Proof.

Preparing to apply Theorem 2, we will think of the unit-hyperbolic-length vector field EE as a vector field on the open unit disk in 2\mathbb{R}^{2}, and now measure length with respect to the Euclidean metric. Set t0:=0t_{0}:=0 and 𝐲𝟎:=(0,0)=O\mathbf{y_{0}}:=(0,0)=O. Next, set a:=1a:=1, and set b:=1/2b:=1/2, so that R={|t|1,|𝐲|1/2}R=\{|t|\leq 1,|\mathbf{y}|\leq 1/2\}, and define f(t,𝐲):=E|𝐲f(t,\mathbf{y}):=E|_{\mathbf{y}}. By the definition of the hyperbolic metric on 𝔻\mathbb{D} (see Appendix), we can take M=1/2M=1/2 (for any choice of b(0,1)b\in(0,1)). Hence, ϵ=min(1,12/12)=1\epsilon=\min(1,\frac{1}{2}/\frac{1}{2})=1.

The partial derivatives of ff with respect to the standard coordinates of 2\mathbb{R}^{2} are continuous on the compact set RR, and hence are bounded, so we conclude that the Lipschitz constant LL exists. By Theorem 2, we have a solution γ:(1,1)𝔻2\gamma:(-1,1)\rightarrow\mathbb{D}\subset\mathbb{R}^{2}.

Now fix an arbitrary point P𝔻P\in\mathbb{D}. There exists a hyperbolic isometry ψ:𝔻𝔻\psi:\mathbb{D}\rightarrow\mathbb{D} which maps PP to OO (see equation (30)). Use ψ\psi_{*} to map the vector field EE to a “recentered” vector field E~\tilde{E} on 𝔻\mathbb{D}, then apply the previous reasoning to E~\tilde{E}, obtaining an integral curve γ:(1,1)𝔻\gamma:(-1,1)\rightarrow\mathbb{D} such that γ(0)=O\gamma(0)=O. The curve ψ1γ:(1,1)𝔻\psi^{-1}\circ\gamma:(-1,1)\rightarrow\mathbb{D} will be the desired integral curve through PP.

Since EE is a smooth vector field on the smooth manifold 𝔻\mathbb{D}, and every point of 𝔻\mathbb{D} can flow forward and backward along EE for at least one unit of time, it follows from the Uniform Time Lemma [Lee13, Lemma 9.15, page 216] that every integral curve can be extended to an integral curve defined on (,)(-\infty,\infty). ∎

Corollary 1.

All integral curves of E1E_{1} and E2E_{2}, the unit asymptotic directions, are defined for all time. ∎

Proposition 3.

There is a CC^{\infty} diffeomorphism FF from 𝔻\mathbb{D} onto 2:={(x1,x2):x1,x2}\mathbb{R}^{2}:=\{(x^{1},x^{2}):x^{1},x^{2}\in\mathbb{R}\}, such that

  1. 1.

    FF is a (global) coordinate chart for 𝔻\mathbb{D}, and

  2. 2.

    F(Ei)=xiF_{*}(E_{i})=\frac{\partial}{\partial x^{i}}, i=1,2i=1,2.

Proof.

Working in 𝔻\mathbb{D}, with O=(0,0)O=(0,0), define F(P):=(OPη1,OPη2)F(P):=(\int_{O}^{P}\eta^{1},\int_{O}^{P}\eta^{2}). FF is well-defined, using Green’s theorem, since η1\eta^{1} and η2\eta^{2} are closed and any two paths from OO to PP bound a region in 𝔻\mathbb{D}. By definition, the differential of FF at PP will map (Ei)P(E_{i})_{P} to (/xi)F(P)(\partial/\partial x^{i})_{F(P)}, for i=1,2i=1,2. Since E1E_{1} and E2E_{2} are everywhere linearly independent, the differential is everywhere nonsingular, so FF is a local CC^{\infty} diffeomorphism (using a CC^{\infty} version of the Inverse Function theorem; see [Lee13, Theorem C.34] ).

In addition, FF is surjective, since in order to find a point in 𝔻\mathbb{D} which maps to (a,b)2(a,b)\in\mathbb{R}^{2}, it suffices to start at OO at time zero, follow an integral curve for E1E_{1} (forward if a>0a>0, backward if a<0a<0) for time |a||a|, then similarly follow an integral curve for E2E_{2} for time |b||b|, using Corollary 1. By the definition of FF, along the first part of the curve, the second coordinate of the output will not change, and vice versa.

With respect to the orthonormal bases {e1,e2}\{e_{1},e_{2}\} in 𝔻\mathbb{D} and {/x1,/x2}\{\partial/\partial x^{1},\partial/\partial x^{2}\} in 2\mathbb{R}^{2}, any local inverse for FF has Jacobian matrix

Mα:=[cosαcosαsinαsinα],M_{\alpha}:=\begin{bmatrix}\cos\alpha&\cos\alpha\\ -\sin\alpha&\sin\alpha\end{bmatrix}, (19)

(using equation (11)) and the singular values of this matrix are easily calculated to be 2cosα\sqrt{2}\cos\alpha and 2sinα\sqrt{2}\sin\alpha. From this we can conclude that all local inverses are Lipschitz maps with constant 2\sqrt{2}. Now let γ:[0,1]2\gamma:[0,1]\rightarrow\mathbb{R}^{2} be a continuous path, and let QQ be a point in 𝔻\mathbb{D} such that F(Q)=γ(0)F(Q)=\gamma(0). We claim that γ\gamma has a unique lift to QQ; that is, there is a unique path γ~:[0,1]𝔻\tilde{\gamma}:[0,1]\rightarrow\mathbb{D} such that γ~(0)=Q\tilde{\gamma}(0)=Q and such that Fγ~=γF\circ\tilde{\gamma}=\gamma.

Let 𝒯:={t[0,1]:γ|[0,t] has a unique lift}\mathcal{T}:=\{t\in[0,1]:\gamma|_{[0,t]}\textrm{ has a unique lift}\}. This is an open subset of [0,1][0,1], since when γ\gamma lifts as far as [0,s][0,s], a local diffeomorphism from an open subset of 𝔻\mathbb{D} to an open neighborhood of γ(s)\gamma(s) will allow the lift to be (uniquely) extended a little further. It is also true that 𝒯\mathcal{T} is a closed subset of [0,1][0,1], since if γ|[0,s)\gamma|_{[0,s)} lifts, then γ\gamma is uniformly continuous on [0,s)[0,1][0,s)\subset[0,1]. Using the Lipschitz bound in the previous paragraph, γ~|[0,s)\tilde{\gamma}|_{[0,s)} is also uniformly continuous, and hence (see Appendix for details) γ~\tilde{\gamma} has a unique extension to [0,s][0,s]. Finally, since [0,1][0,1] is connected, and 𝒯\mathcal{T} is a non-empty subset that is both open and closed, we conclude that 𝒯=[0,1]\mathcal{T}=[0,1], hence γ\gamma lifts uniquely.

Since 𝔻\mathbb{D} is path-connected, and 2\mathbb{R}^{2} is simply-connected, the following lemma applies to the map F:𝔻2F:\mathbb{D}\rightarrow\mathbb{R}^{2}.

Lemma 4 ([Lim03, Proposition 6.12]).

Let f:XYf:X\rightarrow Y be a local homeomorphism with the unique-path-lifting property. If XX is path-connected and YY is simply-connected, then ff is a homeomorphism.

Proof.

(Sketch) Construct a path between any two points in XX that ff maps to the same place. Use ff to map the path, continuously deform the mapped path to a constant path, and then lift the homotopy, to show that the initial and final points of the original path are equal. ∎

Since FF is now known to be a homeomorphism (and in particular is injective), and has already been shown to be a local CC^{\infty} diffeomorphism, it follows that FF is a CC^{\infty} diffeomorphism. ∎

Corollary 2.

Let G:2𝔻G:\mathbb{R}^{2}\rightarrow\mathbb{D} be defined by mapping the point (x1,x2)2(x^{1},x^{2})\in\mathbb{R}^{2} to the point in 𝔻\mathbb{D} obtained by starting at OO, following an integral path for the vector field E1E_{1} for time x1x^{1}, then following an integral path for the vector field E2E_{2} for time x2x^{2} (moving backward when these numbers are negative). Then G=F1G=F^{-1}.

Proof.

By construction, FGF\circ G is the identity map on 2\mathbb{R}^{2}. Proposition 3 implies that FF is injective, so FI𝔻=I2F=(FG)F=F(GF)F\circ I_{\mathbb{D}}=I_{\mathbb{R}^{2}}\circ F=(F\circ G)\circ F=F\circ(G\circ F) implies I𝔻=GFI_{\mathbb{D}}=G\circ F. ∎

Since Proposition 3 has shown that FF is a coordinate mapping for 𝔻\mathbb{D}, we can define coordinate functions u1,u2u^{1},u^{2} on 𝔻\mathbb{D} such that u1=x1F,u2=x2Fu^{1}=x^{1}\circ F,\ u^{2}=x^{2}\circ F and

u1=E1,u2=E2; hence du1=η1,du2=η2.\frac{\partial}{\partial u^{1}}=E_{1},\ \frac{\partial}{\partial u^{2}}=E_{2};\textrm{ hence }du^{1}=\eta^{1},\ du^{2}=\eta^{2}. (20)

(By a common abuse of notation, we let x1x^{1} and x2x^{2} denote the functions on 2\mathbb{R}^{2} that return, respectively, the first and second coordinates of a point.)

2.4 Area computations

Adding and subtracting the equations in (16), we have

θ1=cosα(η1+η2) and θ2=sinα(η1+η2).\theta^{1}=\cos\alpha(\eta^{1}+\eta^{2})\textrm{ and }\theta^{2}=\sin\alpha(-\eta^{1}+\eta^{2}). (21)

Therefore, using Lemma 2,

ω12\displaystyle\omega_{1}^{2} =tanαdα(e2)θ1+cotαdα(e1)θ2\displaystyle=\tan\alpha\,d\alpha(e_{2})\cdot\theta^{1}+\cot\alpha\,d\alpha(e_{1})\cdot\theta^{2}
=(21)tanαdα(e2)cosα(η1+η2)+cotαdα(e1)sinα(η1+η2)\displaystyle\overset{\eqref{Equ:thetafrometa}}{=}\tan\alpha\,d\alpha(e_{2})\cdot\cos\alpha(\eta^{1}+\eta^{2})+\cot\alpha\,d\alpha(e_{1})\cdot\sin\alpha(-\eta^{1}+\eta^{2})
=sinαdα(e2)(η1+η2)+cosαdα(e1)(η1+η2)\displaystyle=\sin\alpha\,d\alpha(e_{2})\cdot(\eta^{1}+\eta^{2})+\cos\alpha\,d\alpha(e_{1})\cdot(-\eta^{1}+\eta^{2})
=(sinαdα(e2)cosαdα(e1))η1+(sinαdα(e2)+cosαdα(e1))η2\displaystyle=(\sin\alpha\,d\alpha(e_{2})-\cos\alpha\,d\alpha(e_{1}))\eta^{1}+(\sin\alpha\,d\alpha(e_{2})+\cos\alpha\,d\alpha(e_{1}))\eta^{2} (22)

while also (rewrite (12) using (20), apply dαd\alpha to both sides of each equation, and simplify)

sinαdα(e2)=12(αu1+αu2) and cosαdα(e1)=12(αu1+αu2).\sin\alpha\,d\alpha(e_{2})=\frac{1}{2}\left(-\frac{\partial\alpha}{\partial u^{1}}+\frac{\partial\alpha}{\partial u^{2}}\right)\textrm{ and }\cos\alpha\,d\alpha(e_{1})=\frac{1}{2}\left(\frac{\partial\alpha}{\partial u^{1}}+\frac{\partial\alpha}{\partial u^{2}}\right).

We define Θ:=2α\Theta:=2\alpha, the angle between E1E_{1} and E2E_{2} (using the capital letter to avoid confusion with the 1-forms θ1,θ2\theta^{1},\theta^{2}). Now (22) becomes

ω12=αu1η1+αu2η2=12Θu1η1+12Θu2η2,\omega_{1}^{2}=-\frac{\partial\alpha}{\partial u^{1}}\eta^{1}+\frac{\partial\alpha}{\partial u^{2}}\eta^{2}=-\frac{1}{2}\frac{\partial\Theta}{\partial u^{1}}\eta^{1}+\frac{1}{2}\frac{\partial\Theta}{\partial u^{2}}\eta^{2}, (23)

which means that (using (16) and 12secαcscα=cscΘ\frac{1}{2}\sec\alpha\csc\alpha=\csc\Theta to show the final equality)

dω12\displaystyle d\omega_{1}^{2} =122Θu2u1η2η1+122Θu1u2η1η2=2Θu1u2(η1η2)\displaystyle=-\frac{1}{2}\frac{\partial^{2}\Theta}{\partial u^{2}\partial u^{1}}\eta^{2}\wedge\eta^{1}+\frac{1}{2}\frac{\partial^{2}\Theta}{\partial u^{1}\partial u^{2}}\eta^{1}\wedge\eta^{2}=\frac{\partial^{2}\Theta}{\partial u^{1}\partial u^{2}}(\eta^{1}\wedge\eta^{2})
=cscΘ2Θu1u2(θ1θ2),\displaystyle=\csc\Theta\frac{\partial^{2}\Theta}{\partial u^{1}\partial u^{2}}(\theta^{1}\wedge\theta^{2}), (24)

where θ1θ2\theta^{1}\wedge\theta^{2} represents the “area form” dAdA on 𝔻\mathbb{D} (since {e1,e2}\{e_{1},e_{2}\} is orthonormal). At the same time,

dω12=(6)ω13ω23=(8)(κ1θ1)(κ2θ2)=κ1κ2(θ1θ2)=θ1θ2.d\omega_{1}^{2}\overset{\eqref{Equ:dconn}}{=}-\omega_{1}^{3}\wedge\omega_{2}^{3}\overset{\eqref{Equ:princurv}}{=}-(\kappa_{1}\theta^{1})\wedge(\kappa_{2}\theta^{2})=-\kappa_{1}\kappa_{2}(\theta^{1}\wedge\theta^{2})=\theta^{1}\wedge\theta^{2}. (25)

Comparing equations (2.4) and (25), we have the PDE

2Θu1u2=sinΘ,\frac{\partial^{2}\Theta}{\partial u^{1}\partial u^{2}}=\sin\Theta, (26)

on 𝔻\mathbb{D}, which we can transfer to the x1x2x^{1}x^{2}-plane, after defining Θ^:=ΘF1\hat{\Theta}:=\Theta\circ F^{-1}, as

2Θ^x1x2=sinΘ^.\frac{\partial^{2}\hat{\Theta}}{\partial x^{1}\partial x^{2}}=\sin\hat{\Theta}. (27)

By Proposition 3 and the definition of Θ\Theta, det(Mα)=(19)sin(2α)=sinΘ\det(M_{\alpha})\overset{\eqref{Equ:GJacob}}{=}\sin(2\alpha)=\sin\Theta is the factor by which area is distorted, when F1F^{-1} maps the x1x2x^{1}x^{2}-plane to 𝔻\mathbb{D}. Hence, the area AA of the hyperbolic plane is given by

A\displaystyle A =limaaaaasinΘ^dx1dx2=limaaaaa2Θ^x1x2𝑑x1𝑑x2\displaystyle=\lim_{a\to\infty}\int_{-a}^{a}\int_{-a}^{a}\sin\hat{\Theta}\,dx^{1}\,dx^{2}=\lim_{a\to\infty}\int_{-a}^{a}\int_{-a}^{a}\frac{\partial^{2}\hat{\Theta}}{\partial x^{1}\partial x^{2}}\,dx^{1}\,dx^{2}
=aa[Θ^x2]x1=ax1=a𝑑x2=aa[Θ^x2(a,x2)Θ^x2(a,x2)]𝑑x2\displaystyle=\int_{-a}^{a}\left[\frac{\partial\hat{\Theta}}{\partial x^{2}}\right]_{x^{1}=-a}^{x^{1}=a}\,dx^{2}=\int_{-a}^{a}[\frac{\partial\hat{\Theta}}{\partial x^{2}}(a,x^{2})-\frac{\partial\hat{\Theta}}{\partial x^{2}}(-a,x^{2})]\,dx^{2}
=lima[Θ^(a,a)Θ^(a,a)Θ^(a,a)+Θ^(a,a)],\displaystyle=\lim_{a\to\infty}[\hat{\Theta}(a,a)-\hat{\Theta}(-a,a)-\hat{\Theta}(a,-a)+\hat{\Theta}(-a,-a)], (28)

and since 0<Θ<π0<\Theta<\pi, AA is bounded above by 2π2\pi. On the other hand, the hyperbolic plane has infinite area, as computed below (in the Poincaré model; see the Appendix for the Riemannian metric):

limt102π0t4(1r2)2r𝑑r𝑑θ\displaystyle\lim_{t\to 1^{-}}\int_{0}^{2\pi}\int_{0}^{t}\frac{4}{(1-r^{2})^{2}}\,r\,dr\,d\theta =limt102π[2(1r2)]0t𝑑θ\displaystyle=\lim_{t\to 1^{-}}\int_{0}^{2\pi}\left[\frac{2}{(1-r^{2})}\right]_{0}^{t}\,d\theta
=2πlimt1(21t22)=\displaystyle=2\pi\cdot\lim_{t\to 1^{-}}\left(\frac{2}{1-t^{2}}-2\right)=\infty (29)

Assuming the existence of ϕ\phi has led to a contradiction, hence there does not exist an isometric immersion of 𝔻\mathbb{D} into 3\mathbb{R}^{3}, concluding the proof of Theorem 1. ∎

Corollary 3.

If SS is a CC^{\infty} Riemannian 22-manifold for which the metric is complete and has constant curvature 1-1, then there does not exist a CC^{\infty} isometric immersion of SS into 3\mathbb{R}^{3}.

Proof.

By [Wol77, Corollary 2.3.17], the universal cover, S~\tilde{S}, of any such SS would be isometric to 2\mathbb{H}^{2}, and the composition S~S3\tilde{S}\rightarrow S\rightarrow\mathbb{R}^{3} would be a counterexample to Theorem 1. ∎

Corollary 4.

If SS is a CC^{\infty} Riemannian 22-manifold for which the metric is complete and has constant curvature k2-k^{2} (k>0k>0), then there does not exist a CC^{\infty} isometric immersion of SS into 3\mathbb{R}^{3}.

Proof.

Suppose that ψ:S3\psi:S\rightarrow\mathbb{R}^{3} is such an isometric immersion, and let h:33h:\mathbb{R}^{3}\rightarrow\mathbb{R}^{3} be the dilation (x,y,z)(kx,ky,kz)(x,y,z)\mapsto(kx,ky,kz). Since the effect of the dilation is to divide all normal curvatures by kk, the image of the composition hψh\circ\psi has constant curvature 1-1 everywhere, which contradicts Corollary 3 (if the original inner product at every point on SS is multiplied by k2k^{2}, the new Riemannian metric will still be complete, and hψh\circ\psi will be an isometric immersion). ∎

3 Brief History

Hilbert’s original proof [Hil01] assumed that the surface in 3\mathbb{R}^{3} was real-analytic (and had constant negative curvature 1-1). However, the outline of his argument will work if ϕ\phi is smooth enough to ensure that the PDE in equation (26) makes sense. Tracing back the definition of Θ\Theta, we can see that if ϕC4(𝔻)\phi\in C^{4}(\mathbb{D}), then the normal vector e¯3\bar{e}_{3} is C3C^{3}, and the eigenvalues κ1¯,κ¯2\bar{\kappa_{1}},\bar{\kappa}_{2} of de¯3-d\bar{e}_{3} are C2C^{2}, as is Θ=2arctan(κ1)\Theta=2\arctan(\kappa_{1}). So 2Θ/u1u2\partial^{2}\Theta/\partial u^{1}\partial u^{2} exists and is continuous on 𝔻\mathbb{D}. With more sophisticated techniques, Efimov proved that there is no C2C^{2} isometric immersion of the hyperbolic plane; see [Mil72] for an exposition and more history. On the other hand, by work of Nash and Kuiper, there does exist a C1C^{1} isometric embedding of the hyperbolic plane in 3\mathbb{R}^{3} [Nas54] [Kui55].

4 Appendix: Metric properties of 2\mathbb{H}^{2}

Let O:=(0,0)𝔻:={(x,y)2:x2+y2<1}O:=(0,0)\in\mathbb{D}:=\{(x,y)\in\mathbb{R}^{2}:x^{2}+y^{2}<1\}. 𝔻\mathbb{D} is the Poincaré disk model for 2\mathbb{H}^{2}, with Riemannian metric g11=g22=4/(1x2y2)2,g12=g21=0g_{11}=g_{22}=4/(1-x^{2}-y^{2})^{2},\ g_{12}=g_{21}=0 [MP77, page 179], [Rat94, Theorem 4.5.5]. 𝔻\mathbb{D} inherits an orientation from the usual orientation on 2\mathbb{R}^{2}. Furthermore, the corresponding distance function on 𝔻\mathbb{D} is [Rat94, Theorem 4.5.1]

d𝔻(P,Q):=arccosh(1+2|PQ|2(1|P|2)(1|Q|2))d_{\mathbb{D}}(P,Q):=\operatorname{arccosh}\left(1+\frac{2|P-Q|^{2}}{(1-|P|^{2})(1-|Q|^{2})}\right)

where P,Q𝔻P,Q\in\mathbb{D} and |P||P| denotes the Euclidean length of the vector PP.

Proofs of the following properties can be found in many textbooks, including [Rat94].

  1. 1.

    The metrics gg and d𝔻d_{\mathbb{D}} are invariant under Möbius transformations which leave 𝔻\mathbb{D} invariant.

  2. 2.

    The d𝔻d_{\mathbb{D}}-metric topology on 𝔻\mathbb{D} equals the topology that the 𝔻\mathbb{D} inherits as a subset of 2\mathbb{R}^{2} (with the usual topology).

  3. 3.

    Every open d𝔻d_{\mathbb{D}}-ball centered at a point in 𝔻\mathbb{D} equals (as a set) an open Euclidean disk whose closure is contained in 𝔻\mathbb{D}.

  4. 4.

    The metric space (𝔻,d𝔻)(\mathbb{D},d_{\mathbb{D}}) is complete [Rat94, Theorem 8.5.1].

  5. 5.

    After identifying (x,y)(x,y) with the complex number z=x+iyz=x+iy, the maps

    zzPP¯z+1,zz+PP¯z+1z\mapsto\frac{z-P}{-\bar{P}z+1},\ z\mapsto\frac{z+P}{\bar{P}z+1} (30)

    are isometries of the Riemannian manifold (𝔻,g)(\mathbb{D},g) and inverses of each other, taking P𝔻P\in\mathbb{D} to OO and OO to PP, respectively. [Rat94, Section 4.5, Exercise 10] [Nee97, 3 IX & 6 III 11].

Furthermore, if r>0r>0 and γ:[0,r)𝔻\gamma:[0,r)\rightarrow\mathbb{D} is uniformly continuous, then there is a unique extension of γ\gamma to a (uniformly) continuous map γ¯:[0,r]𝔻\bar{\gamma}:[0,r]\rightarrow\mathbb{D}. This is a very special case of an exercise in [Mun00, Chapter 7]. The proof hinges on the fact that every sequence in [0,r)[0,r) converging to rr is a Cauchy sequence and will be mapped by γ\gamma to a Cauchy sequence in 𝔻\mathbb{D}.

References

  • [BL10] Thomas Banchoff and Stephen Lovett, Differential geometry of curves and surfaces, A K Peters, Ltd., Natick, MA, 2010. MR 2674651
  • [Cod61] Earl A. Coddington, An introduction to ordinary differential equations, Prentice-Hall Mathematics Series, Prentice-Hall, Inc., Englewood Cliffs, N.J., 1961. MR 0126573
  • [dC83] Manfredo Perdigão do Carmo, Formas diferenciais e aplicações, Monografías de Matemática [Mathematical Monographs], vol. 37, Instituto de Matemática Pura e Aplicada (IMPA), Rio de Janeiro, 1983, With an appendix by Rubens Leão. MR 752287
  • [dC94] Manfredo P. do Carmo, Differential forms and applications, Universitext, Springer-Verlag, Berlin, 1994, Translated from the 1971 Portuguese original. MR 1301070
  • [dC16]   , Differential geometry of curves & surfaces, Dover Publications, Inc., Mineola, NY, 2016, Revised & updated second edition of [MR0394451]. MR 3837152
  • [Hil01] David Hilbert, Ueber Flächen von constanter Gaussscher Krümmung, Trans. Amer. Math. Soc. 2 (1901), no. 1, 87–99. MR 1500557
  • [Kui55] Nicolaas H. Kuiper, On C1C^{1}-isometric imbeddings. I, II, Nederl. Akad. Wetensch. Proc. Ser. A. 58 = Indag. Math. 17 (1955), 545–556, 683–689. MR 0075640
  • [Lee13] John M. Lee, Introduction to smooth manifolds, second ed., Graduate Texts in Mathematics, vol. 218, Springer, New York, 2013. MR 2954043
  • [Lim03] Elon Lages Lima, Fundamental groups and covering spaces, A K Peters, Ltd., Natick, MA, 2003, Translated from the Portuguese by Jonas Gomes. MR 2000701
  • [Mil72] Tilla Klotz Milnor, Efimov’s theorem about complete immersed surfaces of negative curvature, Advances in Math. 8 (1972), 474–543. MR 301679
  • [MP77] Richard S. Millman and George D. Parker, Elements of differential geometry, Prentice-Hall Inc., Englewood Cliffs, N. J., 1977. MR 0442832
  • [Mun00] James R. Munkres, Topology, Prentice Hall, Inc., Upper Saddle River, NJ, 2000, Second edition of [MR0464128]. MR 3728284
  • [Nas54] John Nash, C1C^{1} isometric imbeddings, Ann. of Math. (2) 60 (1954), 383–396. MR 65993
  • [Nee97] Tristan Needham, Visual complex analysis, The Clarendon Press, Oxford University Press, New York, 1997. MR 1446490
  • [O’N06] Barrett O’Neill, Elementary differential geometry, second ed., Elsevier/Academic Press, Amsterdam, 2006. MR 2351345
  • [Rat94] John G. Ratcliffe, Foundations of hyperbolic manifolds, Graduate Texts in Mathematics, vol. 149, Springer-Verlag, New York, 1994. MR 1299730
  • [Shi18] Theodore Shifrin, Differential geometry: a first course in curves and surfaces, https://math.franklin.uga.edu/sites/default/files/inline-files/ShifrinDiffGeo.pdf, 2018, preliminary version, accessed March 16, 2021.
  • [Spi79a] Michael Spivak, A comprehensive introduction to differential geometry. Vol. I, second ed., Publish or Perish, Inc., Wilmington, Del., 1979. MR 532830
  • [Spi79b]   , A comprehensive introduction to differential geometry. Vol. II, second ed., Publish or Perish, Inc., Wilmington, Del., 1979. MR 532831
  • [Wol77] Joseph A. Wolf, Spaces of constant curvature, fourth ed., Publish or Perish, Inc., Houston, TX, 1977. MR 928600