3. Power series; Integration over curves  PDF TEX

Power series

Series of complex numbers

  • Given a sequence \((w_n)_{n\ge0}\subseteq \mathbb C\), consider the series \(\sum_{n=0}^{\infty} w_{n}\). If \[\lim _{n \rightarrow \infty} \sum_{k=0}^{n} w_{k}=w\] for some \(w\in\mathbb C\), then we say that the series converges to \(w\) and write \(w=\sum_{n=0}^{\infty} w_{n}\). Otherwise, the series is said to diverge.

  • A useful observation is that a series is convergent iff the partial sums \(\sum_{k=0}^{n} w_{k}\) form a Cauchy sequence, that is, \(\lim_{m, n \rightarrow \infty}\sum_{k=m}^{n} w_{k} =0\).

  • The series \(\sum_{n=0}^{\infty} w_{n}\) is said to converge absolutely if the series \(\sum_{n=0}^{\infty}\left|w_{n}\right|\) is convergent.

  • As in the real variables case, an absolutely convergent series is convergent.

  • A necessary and sufficient condition for absolute convergence is that the sequence of partial sums \(\sum_{k=0}^{n}\left|w_{k}\right|\) be bounded.

Ratio and root tests

Theorem.

Let \(\sum_{n\ge0} w_{n}\) be a series of nonzero terms.

  • If \(\limsup _{n \rightarrow \infty}\big|\frac{w_{n+1}}{w_{n}}\big|<1\), then the series converges absolutely.

  • If \(\big|\frac{w_{n}+1}{w_{n}}\big| \geq 1\) for all sufficiently large \(n\), the series diverges.

Proof: Exercise! $$\tag*{$\blacksquare$}$$

Theorem.

Let \(\sum_{n\ge0} w_{n}\) be any complex series.

  • If \(\limsup _{n \rightarrow \infty}|w_{n}|^{1 / n}<1\), the series converges absolutely.

  • If \(\limsup _{n \rightarrow \infty}|w_{n}|^{1 / n}>1\), the series diverges.

Proof: Exercise! $$\tag*{$\blacksquare$}$$

The Weierstrass \(M\)-Test

Fact Let \((f_{n})_{n\in\mathbb N}\) be sequence of complex-valued functions on a set \(S\).

  • Then \((f_{n})_{n\in\mathbb N}\) converges pointwise on \(S\) (that is, for each \(z \in S\), the sequence \((f_{n}(z))_{n\in\mathbb N}\) is convergent in \(\mathbb{C}\) ) iff \((f_{n})_{n\in\mathbb N}\) is pointwise Cauchy (that is, for each \(z \in S\), the sequence \((f_{n}(z))_{n\in\mathbb N}\) is a Cauchy sequence in \(\mathbb{C}\)).

  • Also, \((f_{n})_{n\in\mathbb N}\) converges uniformly iff \((f_{n})_{n\in\mathbb N}\) is uniformly Cauchy on \(S\), in other words, \(\lim_{m, n \rightarrow \infty}\left|f_{n}(z)-f_{m}(z)\right|= 0\), uniformly for \(z \in S\).

Theorem.

Let \((g_{n})_{n\in\mathbb N}\) be a sequence complex-valued functions on a set \(S\subseteq \mathbb C\), and assume that \(\left|g_{n}(z)\right| \leq M_{n}\) for all \(z \in S\). If \(\sum_{n=1}^{\infty} M_{n}<+\infty\), then the series \(\sum_{n=1}^{\infty} g_{n}(z)\) converges uniformly on \(S\).

Proof: Let \(f_{n}=\sum_{k=1}^{n} g_{k}\). Then \(|f_n-f_m|\le \sum_{k=m+1}^n|g_{k}|\le \sum_{k=m+1}^nM_k\) for \(n>m\). Thus \((f_{n})_{n\in\mathbb N}\) is uniformly Cauchy on \(S\) and we are done.$$\tag*{$\blacksquare$}$$

Power series

  • The series of the form \(\sum_{n=0}^{\infty} a_{n}\left(z-z_{0}\right)^{n}\), where \(z_{0}, a_{n}\in\mathbb C\) are called the power series. Thus we are dealing with series of functions \(\sum_{n=0}^{\infty} f_{n}\) of a very special type, namely \(f_{n}(z)=a_{n}\left(z-z_{0}\right)^{n}\).

Theorem.

If \(\sum_{n=0}^{\infty} a_{n}\left(z-z_{0}\right)^{n}\) converges at the point \(z\in\mathbb C\) with \(\left|z-z_{0}\right|=r\), then the series converges absolutely on \(D\left(z_{0}, r\right)\), uniformly on each closed subdisk of \(D\left(z_{0}, r\right)\), hence uniformly on each compact subset of \(D\left(z_{0}, r\right)\).

Proof: We have \(\left|a_{n}\left(z^{\prime}-z_{0}\right)^{n}\right|=\left|a_{n}\left(z-z_{0}\right)^{n}\right|\big|\frac{z^{\prime}-z_{0}}{z-z_{0}}\big|^{n}\).

  • The convergence at \(z\) implies that the sequence \((a_{n}\left(z-z_{0}\right)^{n})_{n\in\mathbb N}\) is bounded, since \(\lim_{n\to\infty}a_{n}\left(z-z_{0}\right)^{n} = 0\).

  • If \(\left|z^{\prime}-z_{0}\right| \leq r^{\prime}<r\), then \(\big|\frac{z^{\prime}-z_{0}}{z-z_{0}}\big| \leq \frac{r^{\prime}}{r}<1\) proving absolute convergence at \(z^{\prime}\) (by comparison with a geometric series). Thus the series converges uniformly on \(\overline{D}\left(z_{0}, r^{\prime}\right)\) by the Weierstrass \(M\)-test.$$\tag*{$\blacksquare$}$$

Radius of convergence

Theorem.

Let \(\sum_{n=0}^{\infty} a_{n}\left(z-z_{0}\right)^{n}\) be a power series. Let \(r=\Big[\lim \sup _{n \rightarrow \infty}\sqrt[n]{|a_{n}|}\Big]^{-1}\), be the radius of convergence of the series. (Adopt the convention that \(1 / 0=\infty, 1 / \infty=0\).) The series converges absolutely on \(D\left(z_{0}, r\right)\), and uniformly on its compact subsets. The series diverges for \(\left|z-z_{0}\right|>r\).

Proof: We have \(\limsup _{n \rightarrow \infty}\left|a_{n}\left(z-z_{0}\right)^{n}\right|^{1 / n}=\left|z-z_{0}\right| / r\), which will be less than \(1\) if \(\left|z-z_{0}\right|<r\).

  • By the root test, the series converges absolutely on \(D\left(z_{0}, r\right)\). Uniform convergence on compact subsets follows from the previous result. (We do not necessarily have convergence for \(\left|z-z_{0}\right|=r\), but we do have convergence for \(\left|z-z_{0}\right|=r^{\prime}\), for any \(r^{\prime}\in(0, r)\).

  • If the series converges at some point \(z\) with \(\left|z-z_{0}\right|>r\), then by the previous theorem it converges absolutely at points \(z^{\prime}\) so that \(r<\left|z^{\prime}-z_{0}\right|<\left|z-z_{0}\right|\). But then \(\left|z-z_{0}\right| / r>1\), contradicting the root test. $$\tag*{$\blacksquare$}$$

Power series are holomorphic

Definition.

Let \(\Omega\subseteq \mathbb C\) be an open set. We say that a function \(f:\Omega\to \mathbb C\) is representable by power series in \(\Omega\) if to every disc \(D(a,r) \subseteq \Omega\) we have \[f(z)=\sum_{n=0}^{\infty} c_{n}(z-a)^{n} \quad \text{ for all } \quad z \in D(a,r). \tag{*}\]

Theorem.

Let \(\Omega\subseteq \mathbb C\) be an open set. If \(f:\Omega\to \mathbb C\) is representable by power series in \(\Omega\), then \(f \in H(\Omega)\) and \(f^{\prime}\) is also representable by power series in \(\Omega\). In fact, if (*) holds, then we also have \[f^{\prime}(z)=\sum_{n=1}^{\infty} n c_{n}(z-a)^{n-1} \quad \text{ for all } \quad z \in D(a,r). \tag{**}\]

Proof: The key idea of the proof is to compare the differential quotient \[\frac{f(z) - f(w)}{z - w}\] with the power series (**) evaluated at \(w\). Then, we let \(z \to w\).

  • If the series (*) converges in \(D(a,r)\), then by the root test, the series (**) also converges in that domain.

  • Without loss of generality, set \(a = 0\) and let \[g(z)=\sum_{n=1}^{\infty} nc_{n}z^{n-1}.\]

  • Fix \(w \in D(0,r)\) and choose \(\rho\) such that \(|w| < \rho < r\). We have \[\frac{f(z)-f(w)}{z-w}-g(w)=\sum_{n=1}^{\infty} c_{n}\left[\frac{z^{n}-w^{n}}{z-w}-n w^{n-1}\right] \quad \text{ if } \quad z \neq w.\]

  • For \(n = 1\), the expression in brackets equals \(0\). For \(n \geq 2\), it becomes: \[\label{eq:15} \left[\frac{z^{n}-w^{n}}{z-w}-n w^{n-1}\right]=(z - w) \sum_{k=1}^{n-1} k w^{k-1} z^{n-k-1}.\]

  • This follows from the formula \[z^n - w^n = (z - w) \sum_{k=0}^{n-1} z^{n-1-k} w^k,\] and the telescoping identity \[\begin{aligned} \sum_{k=0}^{n-1} z^{n-1-k} w^k& = \sum_{k=0}^{n-1} ((k+1) - k) z^{n-1-k} w^k\\ &= \sum_{k=1}^{n} k z^{n-k} w^{k-1} - \sum_{k=1}^{n-1} k z^{n-1-k} w^k. \end{aligned}\]

  • If \(|z| < \rho\), then \[\Big|\sum_{k=1}^{n-1} k w^{k-1} z^{n-k-1}\Big|\le \frac{n(n-1)}{2} \rho^{n-2}.\]

  • Thus, we have \[\left|\frac{f(z) - f(w)}{z - w} - g(w)\right| \leq |z - w| \sum_{n=2}^{\infty} n^2 \left|c_n\right| \rho^{n-2}.\]

  • Since \(\rho < r\), the last series converges.

  • Hence, \[\begin{aligned} \lim_{z\to w}\left|\frac{f(z) - f(w)}{z - w} - g(w)\right|=0. \end{aligned}\]

  • This implies that \(f^{\prime}(w) = g(w)\), completing the proof.$$\tag*{$\blacksquare$}$$

Remark. Since \(f^{\prime}\) satisfies the same conditions as \(f\), the theorem can be applied to \(f^{\prime}\) as well. This implies that \(f\) has derivatives of all orders, each of which can be represented by a power series in \(\Omega\).

  • Specifically, if (*) holds, then \[f^{(k)}(z) = \sum_{n=k}^{\infty} n(n-1) \cdots (n-k+1) c_n (z-a)^{n-k}.\]

  • Consequently, (*) leads to \[k! c_k = f^{(k)}(a) \quad \text{for } \quad k = 0, 1, 2, \ldots,\] ensuring that for each \(a \in \Omega\), there exists a unique sequence \((c_n)_{n\ge0}\) satisfying (*).

Examples

  • The geometric series \[\sum_{n=0}^{\infty} z^{n}\] converges absolutely only within the disk \(|z| < 1\).

  • Its sum within this region is the function \(\frac{1}{1-z}\), which is holomorphic in the open set \(\mathbb{C}\setminus\{1\}\).

  • This identity is established similarly to the real case: \[\sum_{n=0}^{N} z^{n} = \frac{1 - z^{N+1}}{1 - z}.\] Then \(\lim_{N \to \infty} z^{N+1} = 0\) if \(|z| < 1\).

  • By the previous theorem, for \(z \in D(0,1)\), we have \[\frac{1}{(1-z)^2}=\bigg(\frac{1}{1-z}\bigg)'=\sum_{n=1}^\infty nz^{n-1}.\]

  • The most important example of a power series is the complex exponential function, which is defined for \(z \in \mathbb{C}\) by \[\exp(z) = \sum_{n=0}^{\infty} \frac{z^n}{n!}.\] When \(z\) is real, this definition coincides with the usual exponential function and in fact, the series above converges absolutely for every \(z \in \mathbb{C}\).

  • Further examples of power series that converge in the whole complex plane are given by the standard trigonometric functions; these are defined by \[\cos z = \sum_{n=0}^{\infty} (-1)^n \frac{z^{2n}}{(2n)!}, \quad \text{and} \quad \sin z = \sum_{n=0}^{\infty} (-1)^n \frac{z^{2n+1}}{(2n+1)!},\] and they agree with the usual cosine and sine of a real argument whenever \(z \in \mathbb{R}\).

More about \(\exp(z), \sin(z)\) and \(\cos(z)\)

  • In order to show that the series defining \(\exp(z)\) converges absolutely, observe that \[\left|\frac{z^{n}}{n!}\right| = \frac{|z|^n}{n!}.\] Thus, \(|\exp(z)|\) can be compared to the series \(\sum_{n=0}^\infty \frac{|z|^n}{n!} = e^{|z|} < \infty\). In fact, this estimate shows that the series defining \(\exp(z)\) is uniformly convergent in every disk in \(\mathbb{C}\).

  • A similar argument can be used to deduce the convergence of power series for \(\sin z\) and \(\cos z\).

  • By the previous theorem, for any \(z \in \mathbb{C}\), the complex derivative of \(\exp(z)\) exists and is given by \[\exp'(z) = \sum_{n=0}^{\infty} n \frac{z^{n-1}}{n!} = \sum_{m=0}^{\infty} \frac{z^{m}}{m!} = \exp(z),\] therefore \(\exp(z)\) is its own derivative.

  • A similar argument gives us that \(\cos'z = -\sin z\) and \(\sin'z = \cos z\). This shows that these are entire functions as well.

  • Since the series defining the exponential function is absolutely convergent, we may multiply it with itself to obtain that for \(z, w \in \mathbb{C}\), we have \[\begin{aligned} \left( \sum_{k=0}^{\infty} \frac{z^k}{k!} \right) \left( \sum_{m=0}^{\infty} \frac{w^m}{m!} \right) &= \sum_{n=0}^{\infty} \frac{1}{n!} \sum_{k=0}^{n} \frac{n!}{k!(n-k)!} z^k w^{n-k}\\ &= \sum_{n=0}^{\infty} \frac{(z+w)^n}{n!}, \end{aligned}\] which shows that \[\exp(z)\exp(w) = \exp(z+w). \tag{A}\]

  • A simple calculation shows that for \(y \in \mathbb{R}\) we have \[\exp(iy) = \cos(y) + i\sin(y).\tag{B}\]

  • We have \(\exp(z) = 1\) if and only if \(z = 2k\pi i\). Indeed, let \(z = x + iy\) and if \(\exp(z) = 1\), then \(|\exp(z)| = 1\).

  • The identities (A) and (B), together with the Pythagorean trigonometric identity, imply that \(x = 0\). Hence, \(z = iy\).

  • If \(\exp(iy) = 1\), this implies that \(\cos(y) = 1\) and \(\sin(y) = 0\). This implies that \(y = 2k\pi\), where \(k \in \mathbb{Z}\). Therefore, \(z = 2k\pi i\).

  • Consequently, by (A), we get that the complex exponential is a periodic function with period \(2\pi i\). This implies that \[\exp(z + 2k\pi i) = \exp(z), \quad \text{for all } \quad z \in \mathbb{C} \text{ and } k \in \mathbb{Z}.\]

Integration over paths

Continuous curves in topological space

Definition.

If \(X\) is a topological space, a curve in \(X\) is a continuous mapping \(\gamma\) of a compact interval \([\alpha, \beta] \subset \mathbb R\) into \(X\); here \(\alpha<\beta\). We call \([\alpha, \beta]\) the parameter interval of \(\gamma\) and denote the range of \(\gamma\) by \[\gamma^{*}=\gamma([\alpha, \beta])=\{\gamma(t): t\in[\alpha, \beta]\}.\]

  • Observe that \(\gamma^{*}\) is compact and connected.

  • If the initial point \(\gamma(\alpha)\) of \(\gamma\) coincides with its end point \(\gamma(\beta)\), we call \(\gamma\) a closed curve.

  • In the definition of a curve in \(\mathbb C\), we will distinguish between the one-dimensional geometric object in the plane (endowed with an orientation) \(\gamma^*\), and its parametrization \(\gamma\), which is a mapping from a closed interval to \(\mathbb{C}\). This parametrization is not uniquely determined.

Equivalent curves

  • Two parametrizations, \[\gamma:[\alpha, \beta] \rightarrow \mathbb{C} \quad \text{and} \quad \tilde{\gamma}:[\alpha_1, \beta_1] \rightarrow \mathbb{C},\] are equivalent if there exists a continuously differentiable bijection from \([\alpha_1, \beta_1] \ni s \mapsto \varphi(s)\in [\alpha, \beta]\) such that \(\varphi^{\prime}(s) > 0\) and \[\tilde{\gamma}(s) = \gamma(\varphi(s)).\]

  • The condition \(\varphi^{\prime}(s) > 0\) precisely ensures that the orientation is preserved: as \(s\) travels from \(\alpha_1\) to \(\beta_1\), \(\varphi(s)\) travels from \(\alpha\) to \(\beta\).

  • The family of all parametrizations that are equivalent to \(\gamma(t)\) determines a smooth curve \(\gamma^\ast \subset \mathbb{C}\), namely the image of \([\alpha, \beta]\) under \(\gamma\), with the orientation given by \(\gamma\) as \(t\) travels from \(\alpha\) to \(\beta\).

Path in topological space

Definition.

A path is a piecewise continuously differentiable curve in the plane. More precisely, a path with parameter interval \([\alpha, \beta]\) is a continuous complex function \(\gamma\) defined on \([\alpha, \beta]\), satisfying the following conditions:

  • There exist finitely many points \(s_{j}\) such that \[\alpha = s_{0} < s_{1} < \cdots < s_{n} = \beta,\] and on each interval \([s_{j-1}, s_{j}]\), the function \(\gamma\) has a continuous derivative.

  • However, at the points \(s_{1}, \ldots, s_{n-1}\), the left-hand and right-hand derivatives of \(\gamma\) may differ.

  • A closed path is a closed curve that is also a path.

Integration over paths

Definition.

Let \(\gamma\) be a path with parameter interval \([\alpha, \beta]\). Assume that \(f\) is continuous on \(\gamma^{*}\subset\mathbb C\). Then we define \[\int_{\gamma} f(z) d z=\int_{\alpha}^{\beta} f(\gamma(t)) \gamma^{\prime}(t) d t.\]

  • When \(\gamma\) is closed path, then integration over \(\gamma\) is understood to be in the anticlockwise direction, unless otherwise mentioned.

  • Here, the integral on the right-hand side is the Riemann integral since \(\gamma^{\prime}(t)\) is a bounded function of \(t\) in \([\alpha, \beta]\) with at most finitely many discontinuities.

Properties of the integrals over paths

  • Let \(\phi:[a, b] \rightarrow[\alpha, \beta]\) be continuous, strictly increasing and onto function. Further assume that \(\phi\) is continuously differentiable.

  • Then \(\phi(a)=\alpha, \phi(b)=\beta\) and \(\phi([a, b])=[\alpha, \beta]\).

  • Let \(\gamma\) be a path with parameter interval \([\alpha, \beta]\). Then \[\sigma=\gamma \circ \phi\] is a path with parameter interval \([a, b]\).

  • Let \(f\) be continuous on \(\gamma^{*}\). Then \(f\) is also continuous on \(\sigma^{*}\) and \[\int_{\sigma} f(z) d z=\int_{\gamma} f(z) d z.\]

  • We call \(\phi:[a, b] \rightarrow[\alpha, \beta]\) a change of parameter function.

  • Let \(\gamma_{1}\) and \(\gamma_{2}\) be paths such that the end point of \(\gamma_{1}\) coincides with the initial point of \(\gamma_{2}\). Then, after suitable re-parametrization, we obtain a path \(\gamma\) by first following \(\gamma_{1}\) and then \(\gamma_{2}\).

  • By (i), we have \[\int_{\gamma} f(z) d z=\int_{\gamma_{1}} f(z) d z+\int_{\gamma_{2}} f(z) d z,\] where \(f\) is continuous on \(\gamma_{1}^{*} \cup \gamma_{2}^{*}\). We write \(\gamma=\gamma_{1}+\gamma_{2}\).

  • For paths \(\gamma_{1}, \gamma_{2}, \ldots, \gamma_{n}\) such that the end points of \(\gamma_{j}\) coincides with the initial point of \(\gamma_{j+1}\) with \(1 \leq j<n\), the path \[\gamma=\gamma_{1}+\gamma_{2}+\cdots+\gamma_{n}\] is defined similarly.

  • Let \(\gamma:[0, 1]\to\mathbb C\) be a path. Define \(\gamma_{1}(t)=\gamma(1-t)\) for \(t\in[0, 1]\). Then \(\gamma_{1}\) is called a path opposite to \(\gamma\). We have \[\int_{\gamma} f(z) d z=-\int_{\gamma_{1}} f(z) d z,\] where \(f\) is continuous on \(\gamma^{*}\).

  • Let \(\gamma:[\alpha, \beta]\to\mathbb C\) be a path and \(f\) be continuous on \(\gamma^{*}\). Then \[\begin{aligned} \left|\int_{\gamma} f(z) d z\right|&=\left|\int_{\alpha}^{\beta} f(\gamma(t)) \gamma^{\prime}(t) d t\right|\\ & \leq \max _{z \in \gamma^{*}}|f(z)| \int_{\alpha}^{\beta}\left|\gamma^{\prime}(t)\right| d t \\ & =\ell(\gamma) \max _{z \in \gamma^{*}}|f(z)| \end{aligned}\] where \(\ell(\gamma)=\int_{\alpha}^{\beta}\left|z^{\prime}(t)\right| d t\) is the length of \(\gamma\).

Some remarks

Let \(\gamma\) be a closed path.

  • Then the complement of \(\gamma^{*}\) in metric space \(\mathbb{C}_{\infty}\) is open. Thus it is a disjoint union of regions, since every open set is a disjoint union of open and connected sets.

  • We say that these regions are determined by \(\gamma\) in \(\mathbb{C}_{\infty}\). There is only one region determined by \(\gamma\) which is unbounded and we call it the unbounded region determined by \(\gamma\). We observe that it contains \(\infty\).

  • The regions determined by \(\gamma\) in \(\mathbb{C}_{\infty}\) and the regions determined by \(\gamma\) in \(\mathbb{C}\) are identical except that the unbounded region determined by \(\gamma\) in \(\mathbb{C}\) does not contain \(\infty\).

Examples

  • If \(a\) is a complex number and \(r>0\), the path defined by \[\gamma(t):=a+r e^{i t}, \quad0 \leq t \leq 2 \pi,\] is called the positively oriented circle with center at \(a\) and radius \(r\) and then we have \[\int_{\gamma} f(z)\,{\rm d} z=\int_{0}^{2 \pi} f\left(a+r e^{i t}\right) ire^{i t}d t\] and the length of \(\gamma\) is \(2 \pi r\), as expected.

  • If \(a\) and \(b\) are complex numbers, the path \(\gamma\) given by

    \[\gamma(t) := a + (b-a) t, \quad 0 \leq t \leq 1,\] is the positively oriented interval \([a, b]\); its length is \(|b-a|\), and

    \[\int_{[a, b]} f(z) \, {\rm d} z = (b-a) \int_{0}^{1} f\big(a + (b-a)t\big) d t.\] Let \(\alpha < \beta\) be real numbers. If \[\gamma_{1}(t) := \frac{a(\beta - t) + b(t - \alpha)}{\beta - \alpha}, \quad \alpha \leq t \leq \beta,\] then we obtain an equivalent path, which we still denote by \([a, b]\).

  • The path opposite to \([a, b]\) is \([b, a]\).

  • Let \(\{a, b, c\}\) be an ordered triple of complex numbers, and let \[\Delta = \Delta(a, b, c)\] be the triangle with vertices at \(a\), \(b\), and \(c\). The set \(\Delta\) is the smallest convex set that contains \(a\), \(b\), and \(c\). Define \[\int_{\partial \Delta} f = \int_{[a, b]} f + \int_{[b, c]} f + \int_{[c, a]} f \tag{$\triangle$}\] for any \(f\) continuous on the boundary of \(\Delta\). We can regard (\(\triangle\)) as the definition of its left side. Alternatively, we can consider \(\partial \Delta\) as a path obtained by joining \([a, b]\) to \([b, c]\) to \([c, a]\), as outlined in definition of the path, in which case (\(\triangle\)) is easily proved to be true.

  • If \(\{a, b, c\}\) is permuted cyclically, we see from (\(\triangle\)) that the left side of (\(\triangle\)) is unaffected. If \(\{a, b, c\}\) is replaced by \(\{a, c, b\}\), then the left side of (\(\triangle\)) changes sign.

Top