24

Cauchy's theorem states that closed line integral of some holomorphic functions yields zero, in some good regions (i.e. simply connected domain). More explicitly,

$$ \oint_\gamma f(z) d z=0. $$

Many textbooks use Goursat's theorem or Green's theorem to prove this. I understand both proofs and can write it myself.

Example: in Stein and Shakarchi's Complex Analysis the proof that the integral over a triangle of a holomorphic function is zero is due to a limiting argument. While this proof is elegant and slick, it doesn't give me any intuition of why the integral is zero.

Are there any intuitive interpretations of this theorem that might help my understanding? Something that might explain why these integrals turn out to be zero while others are not?

random0620
  • 3,325
SunnyMath
  • 349
  • Please clarify your specific problem or provide additional details to highlight exactly what you need. As it's currently written, it's hard to tell exactly what you're asking. – Community May 11 '24 at 10:14
  • Hi, welcome to Math SE. Edit a proof you've seen into your question, then explain which step you think is dubious. You may have just encountered a poor explanation. – J.G. May 11 '24 at 10:15
  • 4
    Do you have intuition for Green's formula? – Trebor May 11 '24 at 10:15
  • 10
    From the perspective of multivariable calculus, the Cauchy-Riemann equations tell you that holomorphic functions on simply connected domains are just conservative vector fields, so it’s not that surprising that the line integral over any closed curve is zero. – kieransquared May 11 '24 at 10:18
  • 4
    One way to think about it is that the functions have antiderivatives, so when you evaluate at the endpoints (which are the same!) a la the Fundamental Theorem of Calculus, you get 0. (The existence of antiderivatives is where the conditions about being a closed curve on a simply connected domain with the function being holomorphic throughout come in.) – Cameron L. Williams May 11 '24 at 10:20
  • 2
    Cauchy's theorem is Green's theorem. So if you believe you understand Green's theorem, then you really do understand Cauchy's theorem. – Deane May 11 '24 at 16:37
  • 1
    Some things in math you don't get, you just get used to them. – Mitch May 13 '24 at 15:39
  • @Mitch I think you can “get” fundamental theorem of calculus if you understand that it’s telescoping sums + mean value theorem. Then Stokes/greens is just that idea but in higher dimensions. – random0620 May 13 '24 at 15:46
  • @Sam I was just paraphrasing von Neumann's saying, one of whose interpretations is that in general sometimes you can work with a tool just out of habit, without understanding it fully. – Mitch May 13 '24 at 16:00
  • @SamKirkiles As to Stokes theorem... wait... as to the FTC... wait... as to the mean value theorem... now -that- one I don't get. Feels like a lot of heavy weather over something I was never worried about (and then OK fine if you really want to go there, let's do it right with analysis). – Mitch May 13 '24 at 16:04
  • Fix a partition of $[a,b]$ $a=t_0<t_1,\dots,<t_n=b$. By MVT fix a $t_k^* \in [t_{k-1},t_k]$ such that $F(t_k)-F(t_{k-1})=f(t_k^)(t_k-t_{k-1})$. Now telescoping, we get $F(b)-F(a)=\sum_{k=0}^n F(t_k)-F(t_{k-1})=\sum_{k=0}^n f(t_k^)(t_{k}-t_{k-1})=\int_a^b f(x)dx$ as you send the max bin width to zero. – random0620 May 13 '24 at 16:12
  • @Mitch Same principle which underlies Stokes/greens. https://galileo-unbound.blog/2019/12/16/looking-under-the-hood-of-the-generalized-stokes-theorem/. Also the book Div Grad Curl and all that has a nice illustration of the principle. – random0620 May 13 '24 at 16:14
  • 1
    I write this out mainly because feeling you don't truly understand Cauchy's theorem might be related to feeling like you don't understand FTC/Green's. – random0620 May 13 '24 at 16:18
  • 1
    @Mitch That's a wonderful point, and I wish someone had told me that decades ago. That's actually been what has slowed me down in mathematics and quantum mechanics - the inability to move forward if there's even one thing I can't fully understand. – uhoh May 13 '24 at 21:44

3 Answers3

24

Here is a purely intuitive analogy that I was taught when I was first learning complex analysis.

Imagine the poles of a function as flag poles. We are a cowboy throwing out a lasso. We toss our lasso and we successfully get it around the pole. It gets snagged at the bottom and it gets pulled taut. In this case we get a residue contribution.

enter image description here

Now suppose the cowboy throws the lasso and it misses the flagpole. As we try to pull it tight, but the lasso gets yanked back to us. We got nothing! In this case, the integral is zero.

enter image description here


Now, let's justify this with a little calculation. Let's suppose $f(z)$ has a pole at zero of order 1, and we will expand it in a Laurent series around this pole

$$f(z)=\sum_{k=-\infty}^\infty a_kz^{k}$$

Let's parameterize our "lasso" contour $\gamma$ as $re^{i\theta}$ as a small circle of radius $r$ around $0$. Now, we just integrate the Laurent series, and exchange summation and integration.

$$\int_\gamma f(z)dz=\sum_{k=-\infty}^\infty a_k \int_\gamma z^k dz$$

If we explicitly compute these integrals, we see that

$$\int_\gamma z^k d z= \begin{cases}0 & k \neq-1 \\ 2 \pi i & k=-1 \end{cases}$$

meaning only the $a_{-1}$ residue term is non-zero, which is present in the Laurent series if and only if there is a pole of order 1. Therefore, we get

$$\int_\gamma f(z)dz = 2\pi i a_{-1}$$

and if there is no pole of order 1, $a_{-1}=0$.

random0620
  • 3,325
  • 2
    "meaning only the $a_{-1}$ residue term is non-zero, which is present in the Laurent series if and only if there is a pole." of order $1$. As in: $f(z)=1/z^2$ has perfectly vanishing loop integrals, and zero residue, and still has a pole at $0$. – Torsten Schoeneberg May 12 '24 at 04:00
  • @TorstenSchoeneberg Thanks, fixed! – random0620 May 12 '24 at 04:11
  • 3
    I think this is the one true answer to the question. As has been pointed out, the residue is exactly given by the coefficient of $1/z$ in the Laurent expansion. The entire point is that a complex-differentiable function has a Taylor expansion, and every Taylor series has an antiderivative, but this is not the whole story because the Laurent series tells you exactly what any given meromorphic function on a connected domain (which is a rational function of Taylor series) leaves as the residue when you integrate around the origin. – user21820 May 13 '24 at 14:18
14

This is not a proof of Cauchy's theorem, but a conceptual interpretation (and, as it happens, an expansion of Cameron Williams's comment): A holomorphic differential form $f(z)\, dz$ on a simply-connected region has a primitive.

On one hand this isn't formally surprising. As we learn in calculus, if $f$ is a continuous function of one variable on some simply-connected region—I mean, on some real interval $I$—then for each $a$ in $I$, the function $F$ defined by $$ F(x) = \int_{a}^{x} f(t)\, dt $$ is a primitive of $f$. Particularly, for every closed path $\gamma:[0, 1] \to I$, we have $$ \int_{\gamma} f(t)\, dt = F\bigl(\gamma(1)\bigr) - F\bigl(\gamma(0)\bigr) = 0 $$ (since $\gamma(0) = \gamma(1)$, cf. Cameron Williams's comment).

So, what's different in the complex setting?

First, continuity of the integrand is not enough; differentiability turns out to be necessary, and locally sufficient. That is, if $R$ is a simply-connected region in the complex plane and $f$ is holomorphic in $R$, then $f(\zeta)\, d\zeta$ has a primitive in $R$ defined by path integration: If we fix a point $z_{0}$ in $R$, then for each $z$ in $R$ and for any two paths $\gamma_{0}$ and $\gamma_{1}$ starting at $z_{0}$ and ending at $z$, Cauchy's theorem guarantees $$ \int_{\gamma_{0}} f(\zeta)\, d\zeta = \int_{\gamma_{1}} f(\zeta)\, d\zeta. $$ Since the value does not depend on the choice of path, it's reasonable to denote the common value by $$ \int_{z_{0}}^{z} f(\zeta)\, d\zeta, $$ and to define a holomorphic primitive $F$ in $R$ by $$ F(z) = \int_{z_{0}}^{z} f(\zeta)\, d\zeta. $$ To emphasize, this is not a proof of Cauchy's theorem, but an intuitive interpretation. It may be worth noting, however, that if we write $z = x + iy$ and $f(z) = u + iv$ with $x$, $y$, $u$, and $v$ real-valued, then $$ f(z)\, dz = (u + iv)(dx + i\, dy) = (u\, dx - v\, dy) + i(v\, dx + u\, dy). $$ When we integrate over the boundary of a region, Green's theorem leads us to consider the integrals of $-v_{x} - u_{y}$ and $u_{x} - v_{y}$. Both integrands vanish identically if and only if the Cauchy-Riemann equations hold in $R$. (In the language of differential forms, $f(z)\, dz$ is closed—has exterior derivative $0$—if and only if $f$ is holomorphic.)

Second, there is a topological aspect to defining primitives. On the real line a point is an obstacle (the complement of a point is disconnected), while in the complex plane a point is not an obstacle (the complement of a point is connected).

Particularly, if $a$ and $b$ are points of a real interval, there is essentially one path from $a$ to $b$. In that respect, it's not surprising that a continuous, one-variable (real) integrand $f(x)\, dx$ has a primitive!

By contrast, there are infinitely many paths between two points of a complex region. As noted above, local path-independence of integrals of $f(z)\, dz$ in $R$ is eqivalent to the Cauchy-Riemann equations in $R$, i.e., to $f$ being holomorphic in $R$.

One remarkable feature of complex analysis is, a holomorphic $1$-form $f(z)\, dz$ in an arbitrary connected region may or may not have a primitive. Let $R = \mathbf{C} \setminus\{0\}$ be the punctured plane and let $n$ be an integer. The power functions $p_{n}(z) = z^{n}$ in $R$ are prototypical: The holomorphic $1$-form $z^{n}\, dz$ has a primitive in $R$ if and only if $n \neq -1$, precisely because if $\gamma$ traces the unit circle once counterclockwise, then $$ \oint_{\gamma} z^{n}\, dz = \begin{cases} 0 & n \neq -1, \\ 2\pi i & n = -1. \end{cases} $$ In the first case, the integral over an arbitrary closed path is $0$, so the integral over an open path depends only on the endpoints. In the second case, integrating around the origin adds a constant so there is no primitive.

6

For any convex area, the Stokes formula says that its area integral can be transformed into a boundary integral, simply by integrating along parallel lines and summing the boundary values.

Perhaps the simple algebraic identities for complex functions independent on $x- i y$ helps:

with $$z=x+i y,\quad w = x-i y \quad dx \wedge dy =\frac{1}{2i}\ dz \wedge dw$$ reduce the contour integral over a simply conncted domain to the area integral

$$\int _ {\partial F} \ f (z, w) \ dz \ = \ \int _F \ d f (z, w)\wedge dz = \int _F \ \left( \partial_w f (z, w) d w + \partial_z f (z, w) d z \right) \wedge dz = \int _F \ \partial_w f (z, w) \ dw \wedge dz \text =0$$

Roland F
  • 5,122