166
$\begingroup$

Can someone give a simple explanation as to why the harmonic series

$$\sum_{n=1}^\infty\frac1n=\frac 1 1 + \frac 12 + \frac 13 + \cdots $$

doesn't converge, on the other hand it grows very slowly?

I'd prefer an easily comprehensible explanation rather than a rigorous proof regularly found in undergraduate textbooks.

$\endgroup$
7
  • 5
    $\begingroup$ This is not meant to be an answer but an interesting note. Suppose we denote $H(n) = 1/1 + 1/2 + ... + 1/n$ then $H(n!) - H((n-1)!) \approx log(n)$ for large n. Does this give a hint? ;) $\endgroup$ Commented Jul 11, 2011 at 4:14
  • 11
    $\begingroup$ Here is a weakly related question: What is a textbook, or even a popularization for the general public, that (1) discusses infinite series, but (2) does not have an explanation for the divergence of this exact series? $\endgroup$
    – GEdgar
    Commented Nov 3, 2013 at 19:50
  • $\begingroup$ to avoid defining the logarithm, use the Cauchy condensation test to show that $\sum 1/n$ converges iff $\sum 1$ converges $\endgroup$
    – reuns
    Commented Jan 30, 2016 at 23:31
  • 1
    $\begingroup$ If it converges, then it contradicts the dominated convergence theorem. This proof is easily comprehensible if you know the dominated convergence theorem, but that theorem is not the most comprehensible. $\endgroup$
    – Oiler
    Commented Oct 6, 2016 at 23:08
  • 1
    $\begingroup$ Here is the funny write up for what Oiler mentioned, given by terry tao: mathoverflow.net/q/44742 $\endgroup$ Commented Oct 14, 2016 at 14:27

26 Answers 26

188
$\begingroup$

Let's group the terms as follows:

Group $1$ : $\displaystyle\frac11\qquad$ ($1$ term)

Group $2$ : $\displaystyle\frac12+\frac13\qquad$($2$ terms)

Group $3$ : $\displaystyle\frac14+\frac15+\frac16+\frac17\qquad$($4$ terms)

Group $4$ : $\displaystyle\frac18+\frac19+\cdots+\frac1{15}\qquad$ ($8$ terms)

$\quad\vdots$

In general, group $n$ contains $2^{n-1}$ terms. But also, notice that the smallest element in group $n$ is larger than $\dfrac1{2^n}$. For example all elements in group $2$ are larger than $\dfrac1{2^2}$. So the sum of the terms in each group is larger than $2^{n-1} \cdot \dfrac1{2^n} = \dfrac1{2}$. Since there are infinitely many groups, and the sum in each group is larger than $\dfrac1{2}$, it follows that the total sum is infinite.

This proof is often attributed to Nicole Oresme.

$\endgroup$
10
  • 16
    $\begingroup$ +1: This is nice: it's easy to turn this into a rigorous proof, and it even gives you a lower bound for the order of growth! $\endgroup$ Commented Jul 21, 2010 at 5:19
  • 2
    $\begingroup$ I assume you mean that group 4 as 8 terms? Or do you mean to go all the way to 1/23? $\endgroup$ Commented Jul 21, 2010 at 7:37
  • 1
    $\begingroup$ Is there a closed-form function for this value? $\endgroup$ Commented Jul 21, 2010 at 18:29
  • 2
    $\begingroup$ Interestingly, this proof goes as far back as Nicole Oresme in the 14th century. Wikipedia has a nice display of this proof [en.wikipedia.org/wiki/Harmonic_series_%28mathematics%29] $\endgroup$ Commented Jul 22, 2010 at 13:20
  • 1
    $\begingroup$ @John: There's no explicit closed-form, but they're generally known as the Harmonic Numbers; there are a number of identities involving them (how to sum them or sum multiples of them, etc.) $\endgroup$ Commented Jul 10, 2011 at 21:23
54
$\begingroup$

There is a fantastic collection of $20$ different proofs that this series diverges. I recommend you read it (it can be found here). I especially like proof $14$, which appeals to triangular numbers for a sort of cameo role.


EDIT

It seems the original link is broken, due to the author moving to his own site. So I followed up and found the new link. In addition, the author has an extended addendum, bringing the total number of proofs to 42+.

$\endgroup$
3
  • 2
    $\begingroup$ Proof 6 is also nice. $\endgroup$
    – leonbloy
    Commented Mar 15, 2013 at 20:51
  • $\begingroup$ Apparently, the list has been updated. $\endgroup$ Commented Jun 19, 2013 at 16:15
  • $\begingroup$ In case the links go down again, the titles are: "The Harmonic Series Diverges Again and Again" by Steven J. Kifowit and Terra A. Stamps, and "More Proofs of Divergence of the Harmonic Series" by Steven J. Kifowit. $\endgroup$
    – Elle Najt
    Commented Oct 24, 2019 at 15:09
42
$\begingroup$

This was bumped, so I'll add a proof sweet proof I saw in this site. Exponentiate $H_n$ and get $$e^{H_n}=\prod_{k=1}^n e^{1/k}\gt\prod_{k=1}^n\left(1+\frac{1}{k}\right)=n+1.$$ Therefore, $H_n\gt\log(n+1)$, so we are done. We used $e^x\gt1+x$ and telescoped the resulting product.

$\endgroup$
1
  • 2
    $\begingroup$ Oh, that's unique. $\endgroup$ Commented Oct 7, 2016 at 1:12
35
$\begingroup$

The answer given by AgCl is a classic one. And possibly pedagogically best; I don't know.

I also like the following argument. I'm not sure what students who are new to the topic will think about it.

Suppose 1 + 1/2 + 1/3 + 1/4 + ... adds up to some finite total S. Now group terms in the following way:

$$1 + \frac{1}{2} > \frac{1}{2} + \frac{1}{2} = \frac{2}{2} = 1$$

$$\frac{1}{3} + \frac{1}{4} > \frac{1}{4} + \frac{1}{4} = \frac{2}{4} = \frac{1}{2}$$

$$\frac{1}{5} + \frac{1}{6} > \frac{1}{6} + \frac{1}{6} = \frac{2}{6} = \frac{1}{3}$$

Continuing in this way, we get $S > S$, a contradiction.

$\endgroup$
3
  • 2
    $\begingroup$ Not really. From $S_n > T_n$ you can only conclude that $\lim S_n \ge \lim T_n$. $\endgroup$
    – lhf
    Commented Jul 10, 2011 at 21:24
  • 9
    $\begingroup$ @lhf: That's right, but that can be easily fixed here (with $S_n = 1 + 1/2 + \dots + 1/2n$ and $T_n = 1 + 1/2 + \dots + 1/n$): we can use a better inequality, like say $S_n \ge T_n + 1/2$ (using just the first step) to conclude that $\lim S_n \ge \lim T_n + 1/2$, contradicting $S = \lim S_n = \lim T_n$. $\endgroup$ Commented Jul 11, 2011 at 4:18
  • 1
    $\begingroup$ I think "better" would be to use only $\frac{1}{4}$ in the first line, then $S - \frac{1}{4} \geq S$ $\endgroup$
    – dEmigOd
    Commented Feb 14, 2019 at 7:35
28
$\begingroup$

Let's group the terms as follows:$$A=\frac11+\frac12+\frac13+\frac14+\cdots\\ $$ $$ A=\underbrace{(\frac{1}{1}+\frac{1}{2}+\frac{1}{3}+\cdots+\frac{1}{9})}_{\color{red} {9- terms}} +\underbrace{(\frac{1}{10}+\frac{1}{11}+\frac{1}{12}+\cdots+\frac{1}{99})}_{\color{red} {90- terms}}\\+\underbrace{(\frac{1}{101}+\frac{1}{102}+\frac{1}{103}+\cdots+\frac{1}{999})}_{\color{red} {900- terms}}+\cdots \\ \to $$ $$\\A>9 \times(\frac{1}{10})+(99-10+1)\times \frac{1}{100}+(999-100+1)\times \frac{1}{1000}+... \\A>\frac{9}{10}+\frac{90}{100}+\frac{90}{100}+\frac{900}{1000}+...\\ \to A>\underbrace{\frac{9}{10}+\frac{9}{10}+\frac{9}{10}+\frac{9}{10}+\frac{9}{10}+\frac{9}{10}+...}_{\color{red} {\text{ m group} ,\text{ and} \space m\to \infty}} \to \infty $$

Showing that $A$ diverges by grouping numbers.

$\endgroup$
0
27
$\begingroup$

An alternative proof (translated and adapted from this comment by Filipe Oliveira, in Portuguese, posted also here). Let $ f(x)=\ln(1+x)$. Then $f'(x)=\dfrac {1}{1+x}$ and $ f'(0)=1$. Hence

$$\displaystyle\lim_{x\to 0}\dfrac{\ln(1+x)}{x}=\lim_{x\to 0}\dfrac{\ln(1+x)-\ln(1)}{x-0}=1,$$

and

$$ \displaystyle\lim_{n\to\infty} \dfrac{\ln\left(1+\dfrac{1}{n}\right)}{\dfrac {1}{n}}=1>0.$$

So, the series $\displaystyle\sum\dfrac{1}{n}$ and $\displaystyle\sum\ln\left(1+\dfrac {1}{n}\right)$ are both convergent or divergent. Since

$$\ln\left(1+\dfrac {1}{n}\right)=\ln\left(\dfrac{n+1}{n}\right)=\ln (n+1)-\ln(n),$$

we have

$$\displaystyle\sum_{n=1}^N\ln\left(1+\dfrac {1}{n}\right)=\ln(N+1)-\ln(1)=\ln(N+1).$$

Thus $\displaystyle\sum_{n=1}^{\infty}\ln\left(1+\dfrac {1}{n}\right)$ is divergent and so is $\displaystyle\sum_{n=1}^{\infty}\dfrac{1}{n}$.

$\endgroup$
1
  • $\begingroup$ Can you please help me to understand how you arrived at the conclusion that the series is divergent? What's the reasoning their...as I know that $log$ is kinda sluggish function so I am having trouble... $\endgroup$ Commented Jan 13, 2021 at 20:08
23
$\begingroup$

This is not as good an answer as AgCl's, nonetheless people may find it interesting.

If you're used to calculus then you might notice that the sum $$ 1+\frac{1}{2}+\frac{1}{3}+\dots+\frac{1}{n}$$ is very close to the integral from $1$ to $n$ of $\frac{1}{x}$. This definite integral is ln(n), so you should expect $1+\frac{1}{2}+\frac{1}{3}+\dots+ \frac{1}{n}$ to grow like $\ln(n)$.

Although this argument can be made rigorous, it's still unsatisfying because it depends on the fact that the derivative of $\ln(x)$ is $\frac{1}{x}$, which is probably harder than the original question. Nonetheless it does illustrate a good general heuristic for quickly determining how sums behave if you already know calculus.

$\endgroup$
3
  • 2
    $\begingroup$ If you look at a Riemann sum for intervals with width 1, you can pretty quickly see that the integral of 1/x from 1 to infinity must be less than the sum of the harmonic series. $\endgroup$
    – Isaac
    Commented Jul 21, 2010 at 5:51
  • $\begingroup$ Thank you for adding this answer. I was hoping to avoid an answer that involved integration, so I also prefer AgCl's answer. But I am happy to see more than one demonstration/proof. $\endgroup$
    – bryn
    Commented Jul 22, 2010 at 11:33
  • $\begingroup$ The sum is closer to the integral from $\frac{1}{2}$ to $n+\frac{1}{2}$ of $\frac{1}{x}$, which is $log(2n+1)$ math.stackexchange.com/a/1602945/134791 $\endgroup$ Commented Jan 25, 2016 at 23:02
17
$\begingroup$

Another interesting proof is based upon one of the consequences of the Lagrange's theorem applied on $\ln(x)$ function, namely:

$$\frac{1}{k+1} < \ln(k+1)-\ln(k)<\frac{1}{k} \space , \space k\in\mathbb{N} ,\space k>0$$

Taking $k=1,2,...,n$ values to the inequality and then summing all relations, we get the required result.

The proof is complete.

$\endgroup$
0
16
$\begingroup$

There also exists a proof for the divergence of the harmonic series that involves the Integral Test. It goes as follows.

It is possible to prove that the harmonic series diverges by comparing its sum with an improper integral. Specifically, consider the arrangement of rectangles shown in the figure of $ y = \dfrac {1}{x} $:

y=1/x

Each rectangle is $1$ unit wide and $\frac{1}{n}$ units high, so the total area of the rectangles is the sum of the harmonic series: $$ \displaystyle\sum \left( \text {enclosed rectangle are} \right) = \displaystyle\sum_{k=1}^{\infty} \dfrac {1}{k}. $$Now, the total area under the curve is given by $$ \displaystyle\int_{1}^{\infty} \dfrac {\mathrm{d}x}{x} = \infty. $$Since this area is entirely contained within the rectangles, the total area of the rectangles must be infinite as well. More precisely, this proves that $$ \displaystyle\sum_{n=1}^{k} \dfrac {1}{n} > \displaystyle\int_{1}^{k+1} \dfrac {\mathrm{d}x}{x} = \ln (k+1). $$This is the backbone of what we know today as the integral test.

Interestingly, the alternating harmonic series does converge: $$ \displaystyle\sum_{n=1}^{\infty} \dfrac {(-1)^n}{n} = \ln 2. $$And so does the $p$-harmonic series with $p>1$.

$\endgroup$
13
$\begingroup$

Let's assume that $\sum_{n=1}^{\infty}\frac1n=:H\in \mathbb{R}$, then $$H=\frac11+\frac12+\frac13+\frac14+\frac15+\frac16 +\ldots $$ $$H\geqslant \frac11+\frac12 +\frac14+\frac14+\frac16+\frac16+\ldots$$ $$H\geqslant \frac11+\frac12+\frac12+\frac13+\frac14+\frac15+\ldots$$ $$H\geqslant \frac12 +H \Rightarrow 0\geqslant \frac12$$ Since the last inequality doesn't hold, we can conclude that the sum doesn't converge.

$\endgroup$
12
$\begingroup$

Suppose to the contrary that converges.

Let $s_n$ denote the $n$-th partial sum. Since the serie converges, $(s_n)$ is a Cauchy sequence. Let $\varepsilon = 1/3$, then there is some $n_0$ such that $|s_q-s_p|< 1/3$ for all $q>p\ge n_0$. Let $q=2n_0$ and $p=n_0$. Then

$$\frac{1}{3}>\bigg|\sum_{n=n_0+1}^{2n_0} \frac{1}{n}\bigg|\ge\bigg|\sum_{n=n_0+1}^{2n_0} \frac{1}{2n_0}\bigg|=\frac{1}{2}$$

a contradiction. Then this contradiction shows that the series diverges.

$\endgroup$
11
$\begingroup$

Another (different) answer, by the Cauchy Condensation Test :

$$\sum_{n=1}^\infty \frac{1}{n} < \infty \iff \sum_{n=1}^\infty 2^n \frac{1}{2^n} = \sum_{n=1}^\infty 1< \infty $$

The latter is obviously divergent, therefore the former diverges. This is THE shortest proof there is.

$\endgroup$
11
$\begingroup$

$$\int_{0}^{\infty}e^{-nx}dx=\frac1n$$

$$\sum_{n=1}^{\infty}\int_{0}^{\infty}e^{-nx}dx=\lim_{ m \to \infty}\sum_{n=1}^{m}\frac1n$$

using the law of Geometric series

$$\int_{0}^{\infty}(\frac{1}{1-e^{-x}}-1)dx=\lim_{ m \to \infty}H_m$$

$$\lim_{ m \to \infty}H_m=\left [ \ln(e^x-1)-x \right ]_0^{\infty}\to\infty$$

$\endgroup$
3
  • $\begingroup$ Hm, the lower bound goes to $-\infty$ it appears. $\endgroup$ Commented Oct 6, 2016 at 16:40
  • $\begingroup$ @SimpleArt The upper bound goes to 0 and The Lower goes to $+\infty$ ,,,$-\ln 0^+$ $\endgroup$
    – mnsh
    Commented Oct 7, 2016 at 11:20
  • $\begingroup$ Oh right, duh, didn't quite use that FTOC correctly. $\endgroup$ Commented Oct 7, 2016 at 13:19
10
$\begingroup$

Another answer that's very similar to others. But it's prettier, and perhaps easier to understand for the 9-th grade student who asked the same question here.

The student's question was ... does the sum equal some number $S$. But, look:

enter image description here

So, whatever it is, $S$ is larger than the sum of the infinite string of $\tfrac12$'s shown in the last line. No number can be this large, so $S$ can't be equal to any number. Mathematicians say that the series "diverges to infinity".

$\endgroup$
9
$\begingroup$

A non-rigorous explanation I thought of once: consider a savings scheme where you put a dollar in your piggy bank every day. So after $n$ days, you have $n$ dollars; clearly, your savings approach infinity. On the other hand, each day you add an additional $1/n$ proportion of your existing savings, "so" (the non-rigorous step) the accumulated percentage after $n$ days is $1 + 1/2 + \cdots + 1/n$.

This can be made rigorous through the infinite product argument $$\prod_{n = 1}^\infty (1 + \tfrac{1}{n}) < \infty \iff \sum_{n = 1}^\infty \frac{1}{n} < \infty$$ which is obtained, essentially, by taking the logarithm of the left-hand side and using the power series for $\log (1 + x)$.

$\endgroup$
8
$\begingroup$

I think the integral test gives the most intuitive explanation. Observe that $$\int^n_1 \frac1x dx= \log n$$ The sum $\displaystyle\sum^n_{k=1}\frac1k$ can be viewed as the area of $n$ rectangles of height $\frac1k$, width $1$ (with the first one having it's left hand side on the y axis, and all having their bottom on the x axis). The graph of $x\mapsto \frac1x$ can be drawn under these, so the sum will grow with $n$ at (least) as fast as the integral - hence will grow (at least) logarithmically.

$\endgroup$
2
  • $\begingroup$ use \log to get nice formatting for $\log$ $\endgroup$
    – Tyler
    Commented Nov 1, 2013 at 17:24
  • $\begingroup$ And thus confirms the author's observation that the harmonic series diverges very slowly---as to grow logarithmically means to $\mathit{crawl \, off}$ to infinity, $\endgroup$
    – DDS
    Commented Jun 22, 2019 at 3:19
7
$\begingroup$

The proof I learned, in Rosenlicht's Introduction to Analysis, published by Dover, is essentially a variant of the most popular answers above.

Namely, we will show that the sequence $S_n=\sum_{k=1}^n \dfrac 1k$ is not Cauchy.

For given $N\gt0$, look at $\vert S_{2N}-S_N\vert=\dfrac1{2N}+\dots+\dfrac 1{N+1}\ge N\cdot \dfrac 1{2N}=\dfrac 12$.

$\endgroup$
6
$\begingroup$

Let be the partial sum $H_n = \frac11 + \frac12 + \frac13 + \cdots + \frac1n$. Using Cesàro-Stolz: $$ \lim_{n\to\infty}\frac{H_n}{\log n} = \lim_{n\to\infty}\frac{H_{n+1}-H_n}{\log(n+1)-\log n} = \lim_{n\to\infty}\frac{\frac1{n+1}}{\log(1+1/n)} = \lim_{n\to\infty}\frac{\frac1{n+1}}{\frac1n} = 1 $$ and $$\sum_{n=1}^\infty\frac1n = \lim_{n\to\infty}H_n = \infty.$$

$\endgroup$
6
$\begingroup$

Visualize the harmonic series as the area of a sequence of rectangles: enter image description here $$ S:=1+\frac12+\frac13+\frac14+\frac15+\cdots$$ Now compute this same area using horizontal rectangles: enter image description here $$ \begin{align} \textstyle S=\sum\text{width $\cdot$ height} &= 1\cdot\left(1-\frac12\right) + 2\cdot\left(\frac12-\frac13\right) + 3\cdot\left(\frac13-\frac14\right) + 4\cdot\left(\frac14-\frac15\right)+\cdots\\ &=\left(1-\frac12\right)+\left(1-\frac23\right)+\left(1-\frac34\right)+\left(1-\frac45\right)+\cdots\\ &=\frac12+\frac13+\frac14+\frac15+\cdots \end{align} $$ ... and the first term in the sum has disappeared! Since $S=S-1$, the sum cannot be finite.


Here's the same argument as a one-liner: $$ \sum_{n=1}^\infty\frac1n\stackrel{(a)}= \sum_{n=1}^\infty\sum_{k=n}^\infty\left[\frac1k-\frac1{k+1}\right] =\sum_{n=1}^\infty\sum_{k=n}^\infty\frac1{k(k+1)} \stackrel{(b)}= \sum_{k=1}^\infty\sum_{n=1}^k\frac1{k(k+1)}=\sum_{k=1}^\infty\frac1{k+1} $$ In (a) we have a telescoping sum; in (b) the interchange in the order of summation is legal since all terms are nonnegative.

$\endgroup$
3
$\begingroup$

We all know that $$\sum_{n=1}^\infty\frac1n=\frac 1 1 + \frac 12 + \frac 13 + \cdots $$ diverges and grows very slowly!! I have seen many proofs of the result but recently found the following: $$S =\frac 1 1 + \frac 12 + \frac 13 +\frac 14+ \frac 15+ \frac 16+ \cdots$$ $$> \frac 12+\frac 12+ \frac 14+ \frac 14+ \frac 16+ \frac 16+ \cdots =\frac 1 1 + \frac 12 + \frac 13 +\cdots = S.$$ In this way we see that $S > S$.

$\endgroup$
3
  • $\begingroup$ O.o This. Is. Amazing!! =) $\endgroup$
    – user378947
    Commented Dec 12, 2016 at 1:18
  • $\begingroup$ You can also see it here math.stackexchange.com/questions/1160527/… $\endgroup$
    – User8976
    Commented Dec 12, 2016 at 1:20
  • $\begingroup$ I have saved this to my personal The Book :) That being said... Come on! The last inequality itself is proof enough!! :P $\endgroup$
    – user378947
    Commented Dec 12, 2016 at 1:31
3
$\begingroup$

Using Euler's form of the Harmonic numbers,

$$\sum_{k=1}^n\frac1k=\int_0^1\frac{1-x^n}{1-x}dx$$

$$\begin{align} \lim_{n\to\infty}\sum_{k=1}^n\frac1k & =\lim_{n\to\infty}\int_0^1\frac{1-x^n}{1-x}dx \\ & =\int_0^1\frac1{1-x}dx \\ & =\left.\lim_{p\to1^+}-\ln(1-x)\right]_0^p \\ & \to+\infty \end{align}$$


Using the Taylor expansion of $\ln(1-x)$,

$$-\ln(1-x)=x+\frac{x^2}2+\frac{x^3}3+\frac{x^4}4+\dots$$

$$-\ln(1-1)=1+\frac12+\frac13+\frac14+\dots\quad\ $$


Using Euler's relationship between the Riemann zeta function and the Dirichlet eta function,

$$\begin{align} \sum_{k=1}^\infty\frac1{k^s} & =\frac1{1-2^{1-s}}\sum_{k=1}^\infty\frac{(-1)^{k+1}}{k^s} \\ \sum_{k=1}^\infty\frac1k & =\frac10\sum_{k=1}^\infty\frac{(-1)^{k+1}}k\tag{$s=1$} \\ & \to+\infty \end{align}$$

$\endgroup$
2
  • $\begingroup$ But isn't the series for $\ln(1-x)$ only valid for $-1\le x<1$? $\endgroup$
    – TheSimpliFire
    Commented Mar 17, 2018 at 9:31
  • $\begingroup$ Yes, but since the limit as $x\to1$ in $x^n$ is monotone, it equals the asked series, if they exist. $\endgroup$ Commented Mar 19, 2018 at 2:33
3
$\begingroup$

\begin{align} \sum_{n=1}^\infty\frac1n=\sum_{n=1}^\infty \int_0^1x^{n-1}\ dx=\int_0^1\sum_{n=1}^\infty x^{n-1}\ dx=\int_0^1\frac{dx}{1-x}=\int_0^1\frac{dx}{x}=\ln(1)-\ln(0)=\infty \end{align}

$\endgroup$
3
$\begingroup$

One of the possible answers.

Firstly:

$$ \sum_{n=1}^{\infty}\frac{(-1)^{n-1}}{n}=\sum_{n=1}^{\infty}\frac{1}{(2n-1)2n} = \frac1{2}+\frac1{12}+\sum_{n=3}^{\infty}\frac{1}{(2n-1)2n} > \frac{1}{2}$$

Since $\displaystyle H_m = \sum_{n=1}^{m}\frac{1}{n}$ is increasing, if it is bounded, it converges, otherwise diverges to infinity.

Assume it converges, then:

$$\sum_{n=1}^{\infty}\frac{1}{n} = 2\frac{1}{2}+2\frac{1}{4}+2\frac{1}{6}+...+\sum_{n=1}^{\infty}\frac{(-1)^{n-1}}{n} = \sum_{n=1}^{\infty}\frac{1}{n} + \sum_{n=1}^{\infty}\frac{(-1)^{n-1}}{n}$$

which would be giving

$$\sum_{n=1}^{\infty}\frac{(-1)^{n-1}}{n}=0$$

and that is not the case.

$\endgroup$
2
$\begingroup$

This is based on the same idea as several other answers, but the presentation, I think, is sufficiently different to make it worth adding. The key is to note that the inequalities $2\gt1$, $2/3\gt1/2$, and $2/5\gt1/3$ generalize to $2/(2n-1)\gt1/n$.

Let

$$S=1+{1\over2}+{1\over3}+{1\over4}+{1\over5}+{1\over6}+\cdots$$

If $S$ were finite then we would have

$$\begin{align} 2S&=2+{2\over2}+{2\over3}+{2\over4}+{2\over5}+{2\over6}+\cdots\\ &=\left(2+{2\over3}+{2\over5}+\cdots \right)+\left({2\over2}+{2\over4}+{2\over6}+\cdots \right)\\ &\gt\left(1+{1\over2}+{1\over3}+\cdots \right)+\left(1+{1\over2}+{1\over3}+\cdots \right)\\ &=2S \end{align}$$

But the strict inequality $2S\gt2S$ is impossible. So $S$ cannot be finite.

$\endgroup$
0
$\begingroup$

Let $r=\frac{m}{n}$ be any positive rational number. Beginning with the equation

$$ r=\underbrace{\frac{1}{n}+\frac{1}{n}+\dots+\frac{1}{n}}_{m\text{ times}} $$

and repeatedly applying the algebraic identity

$$ \frac{1}{p}=\frac{1}{p+1}+\frac{1}{p(p+1)} $$

we can eventually find an Egyptian fraction representation

$$r=\frac{1}{n_1}+\frac{1}{n_2}+\dots+\frac{1}{n_k}$$

of $r$ with $n_1<n_2<\dots<n_k$. So $r<H_{n_k}$.

Since $r$ was arbitrary, it follows that the sequence of harmonic numbers is unbounded. Thus the harmonic series diverges.

(I think the greedy algorithm for constructing Egyptian fractions doesn't work here, as you already need to know that the harmonic series diverges in order to prove that the greedy algorithm works for arbitrarily large rationals...)

$\endgroup$
0
$\begingroup$

If $x\ge 0$, then $\log (1+x)\le x~~~(*)$. Let $x=1/k$, then $$\frac{1}{k} \ge \log(k+1)-\log k$$ By telescopic summation, we get $$\sum_{k=1}^{n}\frac{1}{k}\ge \log(n+1)$$ $$\implies \sum_{k=1}^{\infty}\frac{1}{k}\to \infty ~~~~~~~ \text{Divergence}$$

A Proof for (*) Let $f(x)=\log(1+x)-x \implies f'(x)=\frac{-x}{1+x}<0$. Hence $f(x)$ is a decreasing function. Consequently $f(x)\le f(0)=0.$

$\endgroup$
4
  • $\begingroup$ That is essentially user 1591719's answer math.stackexchange.com/a/151157/42969, isn't it? $\endgroup$
    – Martin R
    Commented Jun 20 at 9:39
  • $\begingroup$ Here, I am not using LMVT. $\endgroup$
    – Z Ahmed
    Commented Jun 20 at 9:47
  • $\begingroup$ Please see my Edit. $\endgroup$
    – Z Ahmed
    Commented Jun 20 at 9:55
  • $\begingroup$ $f'(x) \le 0 \implies f$ is decreasing is the mean-value theorem. – This is not really new in my opinion, but others's may jugde differently. $\endgroup$
    – Martin R
    Commented Jun 20 at 11:03

You must log in to answer this question.

Not the answer you're looking for? Browse other questions tagged .