3

One way to state Gauss's theorema egregium is as follows.

Theorema Egregium. Suppose $(M,g)$ is a $2$-dimensional Riemannian submanifold of $\Bbb{R}^3$. For every $p\in M$, the Gaussian curvature $K$ of $M$ at $p$ is equal to one-half the scalar curvature $R_g$ of $g$ at $p$; thus, the Gaussian curvature is a local isometry invariant of $(M,g)$.

I'm wondering if the theorem still holds with $\Bbb{R}^3$ replaced by an abstract Riemannian $3$-manifold $(N,\bar{g})$. I asked this question because I don't know why $$R_g=g^{ij}R_{ij}=K g^{ij}g_{ij}=2K$$ has anything to do with $\Bbb{R}^3$. Here $R_{ij}$ is the local expression of the Ricci tensor.

Does anyone have an idea? Thank you.

Edit. I grabbed a proof of this theorem. It seems to exploit flatness of $\Bbb{R}^3$, which prohibits our freedom to replace the Euclidean space with a general Riemannian manifold. But then I began to wonder why $R_g=2K$ still holds for a general ambient manifold. Maybe the point is that we have to show that Gaussian curvature is in fact an intrinsic property.

Boar
  • 357
  • How do you define the Gaussian curvature? – Arctic Char Dec 11 '23 at 13:00
  • @ArcticChar Hello, I would define the Gaussian curvature as the determinant of the Weingarten map, also known as the shape operator. That way, I would obtain $K$ as the product of the two principal curvatures. – Boar Dec 11 '23 at 14:19
  • 3
    If $M$ is a submanifold of $N$ $(e_1,e_2)$ is an orthonormal basis of $T_pM$,, then the generalized Theorem Egregium states that $$ g(e_1,R_M(e_1,e_2)e_2) = \det W + g(e_1,R_N(e_1,e_2)e_2), $$ where $W$ is the Weingarten map, $R_M$ and $R_N$ are the Riemann curvature tensors of $M$ and $N$ respectively. – Deane Dec 11 '23 at 16:37
  • 4
    You are using an extrinsic definition of the Gaussian curvature that is misleading you. A good example to think through is a great $S^2\subset S^3$. It is totally geodesic, so the second fundamental form is zero. Yet its curvature is surely not $0$. – Ted Shifrin Dec 11 '23 at 18:38

2 Answers2

5

Edit: The calculations are right, but intermediate reasoning is not, if $n>3$. (The issue, as pointed out by @LuckyJollyMoments, is that although the sectional curvature $A$ is intrinsic to $M$, I am using the eigenvectors of the shape operator $S_{\nu}$ in the sectional curvature; this makes the quantity not intrinsic anymore. I think this is a valuable mistake to leave up, so I have striked out the incorrect portions, and slight edits are made in blue. The correct proof of the higher dimensional Theorema-Egregium, which can be found in Spivak Vol 4, is given below — also that portion is written in a self-contained manner, so one can skip directly to it).

Let’s go general to special. Suppose $(M,g)$ is a Riemannian submanifold of $(\widetilde{M},\widetilde{g})$, and let $\alpha:TM\oplus TM\to (TM)^{\perp}$ be the (vector) second fundamental form. Then, for each point $p\in M$ and linearly independent tangent vectors $x,y\in T_pM$, Gauss’ equation tells us \begin{align} A_p(x,y)&=\widetilde{A}_p(x,y)+\frac{\langle\alpha(x,x),\alpha(y,y)\rangle-\|\alpha(x,y)\|^2}{\|x\|^2\|y\|^2-\langle x,y\rangle^2}, \end{align} where $A_p(x,y)$ is the sectional curvature of the tangent plane in $T_pM$ spanned by $x,y$ as measured in $(M,g)$, and $\widetilde{A}_p(x,y)$ is the sectional curvature measured by $(\widetilde{M},\widetilde{g})$. So, this equation tells you that the sectional curvature of the submanifold (which is intrinsic to the submanifold) equals the sectional curvature of the ambient manifold (intrinsic to the ambient manifold) + some stuff involving the second fundamental form (i.e extrinsic to the submanifold).

Now, suppose $\dim \widetilde{M}=n\geq 3$ and $M$ is a hypersurface. Fixing a unit normal $\nu$ for $T_pM$, we have that the shape operator/Weingarten map $S_{\nu}: T_pM\to T_pM$ is related to the second fundamental form $\alpha$ by $\alpha(x,y)=\langle S_{\nu}(x),y\rangle\nu$. With this, the above equation becomes \begin{align} A_p(x,y)&=\widetilde{A}_p(x,y)+\frac{\langle S_{\nu}(x),x\rangle\langle S_{\nu}(y),y\rangle-\langle S_{\nu}(x),y\rangle^2}{\|x\|^2\|y\|^2-\langle x,y\rangle^2}, \end{align} Symmetry of $\alpha$ implies $S_{\nu}$ is self-adjoint, so by the spectral theorem, we can find an orthonormal basis of eigenvectors $\{e_1,\dots, e_{n-1}\}$ of $S_{\nu}$ for $T_pM$, say with corresponding eigenvalues $\lambda_1,\dots,\lambda_{n-1}$. Then, for each $1\leq i<j\leq n-1$, we have that \begin{align} A_p(e_i,e_j)&=\widetilde{A}_p(e_i,e_j)+\lambda_i\lambda_j, \end{align} or rearranging, $\lambda_i\lambda_j=A_p(e_i,e_j)-\widetilde{A}_p(e_i,e_j)$ . Now, multiply over all possible pairs $i,j$ such that $1\leq i<j\leq \color{blue}{n-1}$. Then, we get \begin{align} (\lambda_1\cdots\lambda_{n-1})^{n-2}&=\prod_{1\leq i<j\leq n-1}\left[A_p(e_i,e_j)-\widetilde{A}_p(e_i,e_j)\right]. \end{align} On the left this is precisely $(\det S_{\nu})^{n-2}$, and on the right, we have a bunch of products of the difference in sectional curvatures. Therefore, depending on the parity of the dimension $n$ of the ambient manifold, we can solve for $\det S_{\nu}$: \begin{align} \begin{cases} \det S_{\nu}=\left(\prod_{1\leq i<j\leq n-1}\left[A_p(e_i,e_j)-\widetilde{A}_p(e_i,e_j)\right]\right)^{\frac{1}{n-2}}&,\quad\text{if $n$ odd}\\ |\det S_{\nu}|= \left(\prod_{1\leq i<j\leq n-1}\left[A_p(e_i,e_j)-\widetilde{A}_p(e_i,e_j)\right]\right)^{\frac{1}{n-2}}&,\quad\text{if $n$ even} \end{cases} \end{align} In particular,

  • if $n\geq 3$ is odd, and the ambient space $\widetilde{M}$ is flat then its sectional curvatures all vanish, and so we have expressed $\det S_{\nu}$ completely in terms of the sectional curvatures $A$ of the submanifold (which is completely intrinsic to the submanifold), thereby proving that $\det S_{\nu}$ (which is calculated in an extrinsic fashion) actually depends only on the geometry of the submanifold. Also, reading the formula right-to-left, we see that it does not matter which orthonormal basis $\{e_1,\dots, e_{n-1}\}$ one chooses to compute the sectional curvatures; the resulting product is independent of this choice.
  • if $n\geq 4$ is even, and the ambient space is flat, then it is $|\det S_{\nu}|$ which is intrinsic to the

That this sign ambiguity must be present is obvious because the choice of normal $\nu$ is only determined up to sign, so replacing $\nu$ by $-\nu$ changes $\det S_{\nu}$ to $\det S_{-\nu}=\det(-S_{\nu})=(-1)^{n-1}\det S_{\nu}$, i.e it stays the same if $n$ is odd and flips sign if $n$ is even.

Since $\det S_{\nu}$ is originally how Gauss defined his curvature, we get the result that the Gaussian curvature is intrinsic to the submanifold In particular, specializing to $n=3$ with flat ambient space, we see that there is only one term on the right: \begin{align} \det S_{\nu}=A_p(e_1,e_2)\equiv A_p,\tag{$*$} \end{align} where $A_p\in\Bbb{R}$ denotes the sectional curvature of the plane $T_pM$ (there is only one since $M$ is 2-dimensional). $\color{blue}{\textrm{This proves Gauss’ theorema Egregium for a $3$-dimensional ambient space}}$.

A completely separate result is the following: for any $(n-1)$-dimensional Riemannian manifold $(M,g)$, the scalar curvature of $(M,g)$ is equal to the sum of the distinct sectional curvatures in any orthonormal basis $\{e_1,\dots, e_{n-1}\}$ for $T_pM$: \begin{align} R_p&=2\sum_{1\leq i<j\leq n-1}A_p(e_i,e_j). \end{align} This follows just by definition of the various things, and using orthonormality. In particular, specializing to $n-1=2$-dimensional manifold $M$, there is only one term on the right, so \begin{align} R_p&=2A_p.\tag{$**$} \end{align}

So, if you now combine the two situations $(*)$ and $(**)$, we get that for a Riemannian 2-manifold $M$ embedded in a flat 3-dimensional space, \begin{align} \det S_{\nu}&=A_p=\frac{R_p}{2}.\tag{$***$} \end{align} It is this equation, $(***)$, which motivates the abstract intrinsic definition of Gaussian curvature, $K$, of a Riemannian 2-manifold $(M,g)$: we define $K:=\frac{R}{2}$, where $R$ is the scalar curvature. So, with this definition, the Gaussian curvature of a Riemannian 2-manifold coincides with the single sectional curvature $A$ of $M$, and when embedded into $\Bbb{R}^3$ it coincides with $\det S_{\nu}$, the product of the principal curvatures. So, I think you’re having an issue with what’s a definition vs what’s being proved.


Corrected proof of Theorema-Egregium.

We shall give a slightly more general result than Gauss’ Theorema-Egregium. We shall start with some generalities, and then naturally discover the statement and proof.

Let $(M,g)$ be a semi-Riemannian manifold, and let $R:\bigwedge^2(TM)\to\text{End}(TM)$ be the curvature of the Levi-Civita connection. First thing’s first, $R$ is actually skew-adjoint-valued (in terms of the $(0,4)$ version of the tensor field, defined by $R(W,Z,X,Y)=\langle R(X\wedge Y)Z,W\rangle$, this is the skew-symmetry $R_{abcd}=-R_{bacd}$). Hence, there is a unique vector bundle morphism $\rho:\bigwedge^2(TM)\to \bigwedge^2(TM)$ such that for all $p\in M$ and all $x,y,\xi,\eta\in T_pM$, we have \begin{align} \left\langle\rho(x\wedge y), \xi\wedge \eta\right\rangle&= \left\langle R(x\wedge y)\xi,\eta\right\rangle. \end{align} Here, the inner product on the left is the induced one on $\bigwedge^2(TM)$ (in general if $(V,g)$ is a pseudo-inner product space then there is one on $\bigwedge^k(V)$ defined such that for all pure wedges, we have $\langle v_1\wedge\cdots\wedge v_k, w_1\wedge\cdots\wedge w_k\rangle= \det(\langle v_i,w_j\rangle)$).

Now, let $(\widetilde{M},\widetilde{g})$ be a Riemannian manifold, and $(M,g)$ a Riemannian submanifold (i.e $g=\iota^*\widetilde{g}$ in particular, and we shall denote them all as $\langle\cdot,\cdot\rangle$) —- at this stage, $M$ can have any codimension. Let their Riemann curvatures be denoted $\widetilde{R},R$. Let $\mathrm{I\!I}:TM\oplus TM\to (TM)^{\perp}$ be the vector-valued second fundamental form (you could also phrase things in terms of the shape tensor $\alpha: TM\to \mathrm{Hom}(TM,(TM)^{\perp})$, which is related by $\alpha(x):= \mathrm{I\!I}(x,\cdot)$). Then, Gauss’ equation states that for all $x,y,\xi,\eta\in TM$ over the same base point, \begin{align} \langle R(x\wedge y)\xi,\eta\rangle&=\langle\widetilde{R}(x\wedge y)\xi,\eta\rangle- [\langle \mathrm{I\!I}(x,\xi), \mathrm{I\!I}(y,\eta)\rangle -\langle \mathrm{I\!I}(x,\eta), \mathrm{I\!I}(y,\xi)\rangle]. \end{align} In terms of the curvature endomorphisms $\widetilde{\rho}, \rho$, this can be written as \begin{align} \langle \rho(x\wedge y), \xi\wedge\eta\rangle&= \langle \widetilde{\rho}(x\wedge y), \xi\wedge\eta\rangle- [\langle \mathrm{I\!I}(x,\xi), \mathrm{I\!I}(y,\eta)\rangle -\langle \mathrm{I\!I}(x,\eta), \mathrm{I\!I}(y,\xi)\rangle].\tag{!} \end{align} The equation (!) is very useful.

Now, specialize to the case of $M$ having codimension $1$ in $\widetilde{M}$, fix a unit normal $\nu\in T_pM$ and let $S_{\nu}:T_pM\to T_pM$ be the shape operator along $\nu$. Recall this is related to $\mathrm{I\!I}$ as follows: for all $x,y\in T_pM$, \begin{align} \mathrm{I\!I}(x,y)&=\langle S_{\nu}(x),y\rangle \nu,\quad\text{or equivalently,}\quad \langle\mathrm{I\!I}(x,y),\nu\rangle=\langle S_{\nu}(x),y\rangle. \end{align} As a result, Gauss’ equation (!) can be written in terms of the shape operator as follows: \begin{align} \langle \rho(x\wedge y), \xi\wedge\eta\rangle&= \langle \widetilde{\rho}(x\wedge y), \xi\wedge\eta\rangle- [\langle S_{\nu}(x),\xi\rangle\langle S_{\nu}(y),\eta\rangle -\langle S_{\nu}(x),\eta\rangle\langle S_{\nu}(y),\xi\rangle].\tag{!!} \end{align}

Since $\mathrm{I\!I}$ is symmetric, it follows $S_{\nu}$ is self-adjoint, so by the spectral theorem, there is an orthonormal basis $\{e_1,\dots, e_{n-1}\}$ consisting of eigenvectors of $S_{\nu}:T_pM\to T_pM$; let $\lambda_1,\dots,\lambda_{n-1}$ be the corresponding eigenvalues. Then, (!!) applied to $x=e_i$ and $y=e_j$ with $1\leq i<j\leq n-1$ shows that \begin{align} \langle \rho(e_i\wedge e_j), \xi\wedge\eta\rangle&= \langle \widetilde{\rho}(e_i\wedge e_j), \xi\wedge\eta\rangle- \lambda_i\lambda_j[\langle e_i,\xi\rangle\langle e_j,\eta\rangle -\langle e_i,\eta\rangle\langle e_j,\xi\rangle]\\ &= \langle \widetilde{\rho}(e_i\wedge e_j) - \lambda_i\lambda_j e_i\wedge e_j, \xi\wedge\eta\rangle. \end{align} Since $\xi,\eta$ are arbitrary, it follows that Gauss’ equation can be very nicely written as \begin{align} \rho(e_i\wedge e_j)&=\widetilde{\rho}(x_i\wedge e_j)-\lambda_i\lambda_j \,e_i\wedge e_j.\tag{!!!} \end{align} In particular, if the ambient space is flat, then $\widetilde{\rho}=0$, and so this computation shows that $e_i\wedge e_j$ is an eigenvector of $\rho$ with eigenvalue $-\lambda_i\lambda_j$.

Since the operator $\rho:\bigwedge^2(TM)\to \bigwedge^2(TM)$ is constructed purely from the metric $g$ (and its Levi-Civita connection and curvature $R$), it follows that its (fiberwise) eigenvalues are also intrinsic to $M$. Thus, we have shown the following:

Theorem.

With notation as above, the collection $(\lambda_i\lambda_j)_{1\leq i<j\leq n-1}$ is intrinsic to $M$.

Hence, every symmetric function $f$ of $\binom{n-1}{2}$ variables when evaluated on these eigenvalues is also intrinsic to $M$; so if you wanted to, you could define a new type of curvature $K_f$ for each such function $f$.

The most obvious thing to do is take the product of all these eigenvalues of $\rho$ (i.e compute the (fiberwise) determinant of $\rho$): \begin{align} \det (-\rho)= \prod_{1\leq i<j\leq n-1}\lambda_i\lambda_j=\left(\prod_{i=1}^{n-1}\lambda_i\right)^{n-2}=(\det S_{\nu})^{n-2}. \end{align} Since the quantity on the left is intrinsic to $M$, so is the quantity on the right. In particular,

  • if $n\geq 3$ is odd, then we can solve this to find that $\det S_{\nu}$ is intrinsic to $M$
  • if $n\geq 4$ is even, then $|\det S_{\nu}|$ is intrinsic to $M$.

This case distinction is to be expected because the choice of unit normal is only unique up to sign, and we have \begin{align} \det(S_{-\nu})=\det(-S_{\nu})=(-1)^{n-1}\det S_{\nu}, \end{align} and this equals $\det S_{\nu}$ if $n$ is odd, and equals $-\det S_{\nu}$ if $n$ is even. Since $K\equiv \det S_{\nu}$ is how Gauss originally defined his curvature, we see get the following result:

Higher dimensional Theorema-Egregium.

Let $M$ be a Riemannian hypersurface in a flat Riemannian manifold $\widetilde{M}$, with $n:=\dim \widetilde{M}\geq 3$. Then, the Gauss curvature, $K$, of $M$ is intrinsic to $M$ if $n$ is odd, while merely $|K|$ is intrinsic if $n$ is even.

peek-a-boo
  • 65,833
  • Hello, in the last paragraph where you define $K$, are you telling me that the Riemannian $2$-manifold $(M,g)$ is not necessarily a submanifold? Thank you. – Boar Dec 12 '23 at 05:27
  • 1
    @Boar yes exactly. The equation $(***)$ which we proved for $M$ embedded in $\Bbb{R}^3$ is now used as motivation to define $K$ for abstract 2-dimensional $M$. – peek-a-boo Dec 12 '23 at 05:30
  • Thank you. Please allow me to ask one more stupid question. In the context of an embedded hypersurface $\Phi:(M,g)\to(\widetilde{M},\widetilde{g})$, people say that $K$ is intrinsic because $K$ is found to be dependent only on $g$, but isn't $g$ induced by the embedding $\Phi$? That embedding determines how $M$ is put into $\widetilde{M}$, doesn't it? How could they claim that $K$ is intrinsic to the hypersurface $M$? – Boar Dec 12 '23 at 05:35
  • 1
    @Boar yes, $g$ is defined in terms of $\widetilde{g}$, but intrinsic here means roughly ‘depends only on $g$’. This is not the same as ‘depends only on $\widetilde{g}$’. Anyway, I really don’t see what the confusion is. Once you have $(M,g)$, then forget about everything else. Don’t ask ‘where did it come from’ or anything else. No more distractions. Define everything using $(M,g)$ only and you have ‘intrinsic’ stuff. – peek-a-boo Dec 12 '23 at 05:38
  • Thank you. I suppose you are trying to express that "depends only on $g$" does not exclude the option of "$g$ might be induced by some other metric $\widetilde{g}$". I just found a related discussion https://math.stackexchange.com/questions/2206328/intrinsic-vs-extrinsic-properties-of-surfaces on this topic. – Boar Dec 12 '23 at 06:11
  • 1
    yes, it doesn’t matter how $g$ came to be. All we care about is we have a smooth manifold $M$ and a Riemannian metric $g$ on it, and all subsequent quantities/properties are being discussed relative to $g$ only. – peek-a-boo Dec 12 '23 at 06:30
  • 1
    It is possible for a Riemannian manifold to be isometrically embedded into the same higher dimensional Riemannian manifold in more than one way. Intrinsic means the invariant depends on the Riemannian metric and not on the embedding. The second fundamental form is an extrinsic invariant. The theorem egregium and its generalizations identify invariants of the second fundamental form that are intrinsic. – Deane Dec 12 '23 at 12:59
  • When $n>3$, the product of $\binom{n-1}{2}$ sectional curvatures is not invariant under a change of orthonormal basis in general. In Spivak’s Comprehensive Introduction to Differential Geometry, the quantities ${-\lambda_i \lambda_j}$ (counting multiplicities) are identified as the eigenvalues of an endomorphism of $\Omega^2 T_p M$ defined by $X \wedge Y \mapsto R(X, Y)$. Hence the set ${\lambda_i \lambda_j}$ is invariant under isometries. However, in general, an element of the orthogonal group $O(T_p M)$ does not arise as the differential of an isometry. – LuckyJollyMoments May 03 '25 at 07:49
  • @LuckyJollyMoments i don’t follow. The determinant on the LHS is the determinant of an endomorphism so doesn’t depend on the basis used, so the corresponding quantity on the RHS either. (I’ll have to remind myself of what Spivak writes another time :) But also if you say that the products $\lambda_i\lambda_j$ are invariant under isometries then surely the full product over all indices is also invariant, consistent with what I wrote, no? – peek-a-boo May 03 '25 at 08:02
  • @peek-a-boo The RHS involves an orthonormal basis (up to $n$ sign choices) of $T_p M$ determined by $S_\nu$. To show that Gaussian curvature is intrinsic, we must demonstrate that for any two embeddings of $M$ into $\mathbb{R}^n$, each inducing an orthonormal basis of $T_p M$, the RHS yields the same value. Although it seems natural to expect that the RHS is invariant under arbitrary orthonormal changes, this is not generally the case. – LuckyJollyMoments May 04 '25 at 06:43
  • @peek-a-boo Consider $\mathbb{R}^3$ equipped with an algebraic curvature tensor $T$. Consider a rotation $e_1' = \cos\theta\cdot e_1- \sin\theta\cdot e_2 , e_2' = \sin\theta\cdot e_1 + \cos\theta\cdot e_2$. It is easy to verify that $T(e_1, e_2, e_2, e_1)$ remains invariant. Now consider the product $T(e_1', e_3, e_3, e_1') \cdot T(e_2', e_3, e_3, e_2')$. Expanding the rotated expression in terms of the unrotated basis, we observe that invariance of the product imposes constraints on these mixed terms: $T(e_1, e_3, e_3, e_1), T(e_1, e_3, e_3, e_2), T(e_2, e_3, e_3, e_2)$. – LuckyJollyMoments May 04 '25 at 06:44
  • @LuckyJollyMoments ok, both those points sort of make sense, but something still isn’t clicking for me. So, it seems like you’re objecting to my first bullet point where I said ”it does not matter which orthonormal basis… the resulting product is independent of this choice”. I guess all I have shown is that it doesn’t matter which orthonormal basis of eigenvectors of the shape operator we use - do you agree up to here? – peek-a-boo May 04 '25 at 07:50
  • and in that case, I guess I can somewhat where things are going wrong: although the sectional curvature mapping $A_p$ (acting on all $2$-planes in $T_pM$) is intrinsic to the submanifold (i.e invariant under isometries), the choice of eigenvectors are those of the shape operator (which clearly is extrinsic information). And it is only the case of ambient dimension $3$ and a 2-dimensional submanifold where there is a unique $2$-plane, so the result is still instrinsic, and hence why you claimed my argument breaks down for $n>3$ (I guess another way of saying it is $\binom{n-1}{2}>1$ if $n>3$)? – peek-a-boo May 04 '25 at 07:54
  • so, although the final statement is true (that the Gauss curvature of a hypersurface in an ambient flat manifold, e.g $\Bbb{R}^n$ is intrinsic (up to sign depending on parity of the dimension), i.e invariant under isometries), my “proof” for it is flawed because it is based on products of sectional curvatures; and instead, I should argue based on it being related to the eigenvaleus of the curvature as a (skew-adjoint-valued) 2-form on $M$ (or more properly, by metric duality, a fiberwise linear map $\bigwedge^2(TM)\to \bigwedge^2(TM)$). Would you agree with this? – peek-a-boo May 04 '25 at 07:59
  • 1
    @peek-a-boo Yes, I completely agree — your summary is spot on. The issue is just that the shape operator and its eigenvectors come from the embedding. The curvature tensor as an operator on $\Lambda^2 T_p M$ avoids this and gives a truly intrinsic viewpoint. – LuckyJollyMoments May 04 '25 at 15:15
  • @LuckyJollyMoments alright perfect, I’ll rewrite the answer when i get the time (this is my only answer I remember where such a huge mistake has gone unnoticed for so long, so thanks for clarifying the issue:) – peek-a-boo May 04 '25 at 17:05
  • 1
    It’s worth noting that one can replace “a flat Riemannian manifold” with "a Riemannian manifold of constant sectional curvature $\kappa_0$", and define an alternative notion of Gaussian curvature relative to $\kappa_0$, which also turns out to be intrinsic. – LuckyJollyMoments May 05 '25 at 09:45
0

The Theorema Egregium asserts that the Gaussian curvature $K$ of a surface is independent of isometric embedding. This is proved by showing that $K$ can be expressed in terms of the first fundamental form (i.e., the metric) alone. The easiest way to show this is to use the existence of isothermic coordinates $(x,y)$ with respect to which the metric looks like $f^2(x,y)\big(dx^2+dy^2)$. Then $K=-\Delta_{LB}\ln f$, where $\Delta_{LB}$ is the Laplace-Beltrami operator. The latter can be expressed in terms of the ordinary (flat) Laplacian $\Delta_0$ by $\Delta_{LB}=\frac1{f^2}\Delta_0$.

Mikhail Katz
  • 47,573