5

A friend of mine gave me the task to calculate the value of the following triple sum: $$ \sum_{a \in \mathbb{N}}\sum_{b \in \mathbb{N}}\sum_{c \in \mathbb{N}} \frac{(-1)^{a+b+c}}{1+a+b+c}. $$ I am not even sure whether or not it converges. I have tried to interchange the order of summation and just calculate the value of $$ \sum_{n \in \mathbb{N}} \frac{\phi(n) \cdot (-1)^n}{1 + n} $$ where $\phi(n)$ is the number of ways $n$ can be written as the sum of positive integers, but this sum clearly diverges.

So I wonder

  • Does this triple sum converge?
  • If so, what value does it converge to?
Blue
  • 83,939
  • 1
    While this doesn't converge it might still be possible to compute its standard divergent renormalization. To do requires considering the function $\Omega(x,y,z) = \sum_{a=1}^{\lfloor x \rfloor} \sum_{b=1}^{\lfloor y \rfloor} \sum_{c=1}^{\lfloor z \rfloor} \frac{(-1)^{a+b+c}}{1+a+b+c} $. Now $\Omega(x,y,z)$ should have an asymptotic decomposition in terms of polynomials + polylogarithmic functions of $x,y,z$. There is an $O(1)$ bounded term in this expansion $B(x,y,z)$. The average value of $B(x,y,z)$ as $x,y,z \rightarrow \infty $ (if its defined) then gives the renormalization to this. – Sidharth Ghoshal Sep 08 '24 at 19:25
  • The 1-D example is considerably easier, noticing that $\sum_{n=1}^{\lfloor x \rfloor} \frac{1}{n} = \ln(x) + \gamma + o(\frac{1}{x})$ The bounded function here is $\gamma$ and its average value is trivially $\gamma$ giving us the divergent renormalization of the harmonic series. See here for more info. And here for examples of manipulating divergent multi-index sums. – Sidharth Ghoshal Sep 08 '24 at 19:27
  • Techniques here and here would also be useful – Sidharth Ghoshal Sep 08 '24 at 19:33
  • I would really like to know why my question has been closed. Anyone can explain? – Bruno Krams Sep 18 '24 at 18:15

6 Answers6

8

$$\begin{split} S_3 &:=\sum_{a=0}^\infty\sum_{b=0}^\infty\sum_{c=0}^\infty \frac{(-1)^{a+b+c}}{1+a+b+c} \\ &=\sum_{a=0}^\infty\sum_{b=0}^\infty\sum_{c=0}^\infty(-1)^{a+b+c}\int_0^1x^{a+b+c}\,dx\\ &=\int_0^1\sum_{a=0}^\infty(-x)^a\cdot\sum_{b=0}^\infty(-x)^b\cdot\sum_{c=0}^\infty(-x)^c\, dx\\ &=\int_0^1 \frac{dx}{(1+x)^3} = \bbox[5px,border:2px solid red]{ \frac{3}{8}}\,. \end{split}$$ Can you fill in the details? :)


More generally, for integer $k\geqslant 2$, $$\begin{split} S_k&:=\sum_{a_1=0}^\infty\cdots\sum_{a_k=0}^\infty\frac{(-1)^{\sum_{j=1}^ka_j}}{1+\sum_{j=1}^ka_j}\\ &=\int_0^1\prod_{j=1}^k\sum_{a_j=0}^\infty (-x)^{a_j}dx\\ &=\int_0^1\frac{dx}{(1+x)^k}= \bbox[5px,border:2px solid red]{\frac{1-2^{1-k}}{k-1}}\,. \end{split}$$ Continuing the expression at $k=1$, we retrieve the well-known result $$S_1:=\sum_{n=0}^\infty \frac{(-1)^n}{1+n}=\log(2)\,. $$


Bonus I. Just for fun: $$\begin{split} \sum_{k=1}^n S_k &= \log(2) - \sum_{k=2}^n \frac{2^{1-k}}{k-1} + \sum_{k=2}^n \frac{1}{k-1} \\ &=\log(2) - \sum_{k=1}^n \frac{1}{k\cdot 2^k} + \sum_{k=1}^n \frac 1 k\,. \end{split}$$ However, $$-\log(1-x) = \sum_{k=1}^\infty \frac{x^k}{k}\,,\qquad |x|<1\,, \tag{$\ast$}$$ so that $$\sum_{k=1}^n \frac{1}{k\cdot 2^k} \xrightarrow{n\to\infty} -\log\left(1-\frac 1 2\right) = \log(2)\,, $$ that is, $$\log(2) - \sum_{k=1}^n \frac{1}{k\cdot 2^k} = \mathrm{o}(1)\qquad \text{as }n\to\infty\,. $$ As a consequence $$\begin{split} \sum_{k=1}^n S_k &= H_n + \mathrm{o}(1) \\ &=\log(n) + \gamma + \mathrm{o}(1)\qquad \text{as }n\to\infty\,, \end{split}$$ where $H_n$ is the $n$-th harmonic number, and $\gamma$ is the Euler–Mascheroni constant. These asymptotics can be sharpened by writing down Lagrange's remainder theorem for $(\ast)$ and by refining the expansion for $H_n$, for example via the Euler–Maclaurin formula or its equivalents (see answers here).


Bonus II. (Sorry, I'm bored on a long train journey.) Same thing but with doubled integers in the denominator: $$\begin{split} R_k &:= \sum_{a_1=0}^\infty \cdots \sum_{a_k=0}^\infty \frac{(-1)^{\sum_{j=1}^k a_j}}{1+\color{green}{2}\sum_{j=1}^k a_j} \\ &= \int_0^1 \prod_{j=1}^k \sum_{a_j=1}^\infty (-1)^{a_j} x^{2a_j}\,dx\\ &= \int_0^1 \frac 1{(x^2+1)^k}\, dx = \int_0^{\pi/4} \cos^{2k-2}(x)\,dx \end{split}$$ We immediately deduce $R_1 = \frac \pi 4$, in accordance with $$\sum_{a=0}^\infty \frac{(-1)^a}{1+2a} = \arctan(1) = \frac{\pi}4\,.$$ For integer $k\geqslant 2$, integrating by parts yields the recurrence relation $$\begin{split} R_k &=\int_0^{\pi/4} \cos^{2k-2}(x)\,dx = \left.\frac{\sin(x)\cdot\cos^{2k-3}(x)}{2k-2}\right|_0^{\pi/4}+ \frac{2k-3}{2k-2}\int_0^{\pi/4} \cos^{2k-4}(x)\,dx\\ &= \frac{1}{2^k\cdot(k-1)} + \frac{2k-3}{2k-2}\,R_{k-1}\,. \end{split}$$ Solving for $R_k$ is complicated (maybe Mathematica can give you a closed-form expression), but at the very least we can compute $$R_3 = \frac{1}{16} + \frac{3}{4}\,R_2 =\cdots= \frac{8+3\pi}{32}\,.$$ Not sure whether you can do this with numbers other than $2$ in that denominator.

giobrach
  • 7,852
4

I believe the sum does converge, as does the analogous $k$-fold sum for any $k\ge1$. Let me try to make a convincing argument for $k=2$. We can write $$ \sum_{a=0}^\infty \sum_{b=0}^\infty \frac{(-1)^{a+b}}{1+a+b} = \sum_{a=0}^\infty (-1)^a T(a), \quad\text{where}\quad T(a) = \sum_{b=0}^\infty \frac{(-1)^b}{1+a+b}. $$ Note that $T(a)$ converges by the alternating series test. We'd like to use the alternating series test on the remaining sum over $a$; but to do so, we need to show that $T(a)$ decreases to $0$. The fact that $\lim_{a\to\infty} T(a)\to0$ isn't hard, because $T(a)$ is bounded between its initial partial sum $0$ and its $0$th partial sum $\frac1{1+a}$. It's “decreasing” that we need to work for.

We want to say that $$ T'(a) = \sum_{b=0}^\infty \frac d{da} \frac{(-1)^b}{1+a+b} = - \sum_{b=0}^\infty \frac{(-1)^b}{(1+a+b)^2} $$ because this would show that $T'(a)<0$ by the same alternating series argument (it's bounded between $0$ and the $0$th partial sum $-\frac1{(1+a)^2}$). However, term-by-term differentiation of an infinite series needs to be justified: the series of derivatives must converge (locally) uniformly. Fortunately that is the case here since the series of derivatives converges absolutely and uniformly bounded by $\sum_{b=0}^\infty \frac1{(1+b)^2}$.

My instinct is that this argument can be iterated to the original triple sum (and beyond), but I haven't tried to write down the details.

Greg Martin
  • 92,241
4

THIS IS NOT A COMPLETE ANSWER!

I provide an approach that does work to provide the answer, although I do not have the knowledge of tools required to justify convergence issues. Let us concern ourselves with finding

$$S(x)=\sum_{a=0}^{\infty}\sum_{b=0}^{\infty}\sum_{c=0}^{\infty} \frac{x^{a+b+c}}{1+a+b+c}$$

Then the ordered pairs of non-negative integers $(a,b,c)$ such that $a+b+c=n$ is $\binom{n+2}{2}$ by stars and bars and so your sum can be written as

$$S(x)= \sum_{r=0}^{\infty} \binom{r+2}{2} \frac{x^r}{1+r} = \sum_{\color{blue}{r=1}}^{\infty} \binom{r+1}{2} \frac{x^{r-1}}{r} =\sum_{r=1}^{\infty} \binom{r+1}{r} \frac{x^r}{2x} = \frac1{2x}\left(\sum_{r=0}^{\infty} \binom{2+r-1}{r} x^r\right) - \frac1{2x}$$

This is the binomial series for $(1-x)^{-2}$ which converges for $|x|<1$. Thus $2xS(x)=\frac1{(1-x)^2}-1 = \frac{2x-x^2}{(x-1)^2}$. This gives $$S(x)=\frac{2-x}{2(1-x)^2}$$

And although this formula is only valid for $x\in (-1,1)$, putting $x=-1$ yields $S(-1)=\frac38$ which happens to be correct as per the calculus-based answer. I am not sure why this works but a little digging on the internet led me to Abel’s theorem. Perhaps some one with enough knowledge could use it to justify the steps. Or perhaps this idea is completely incorrect. I’d appreciate if someone could comment on the same.

user76284
  • 6,408
Sahaj
  • 5,355
  • 1
    This approach works nicely! For small $x$ the rearrangement of the terms is indeed correct and Abels Theorem then is applicable for $S$ and gives the desired result. – M.E.W. Sep 08 '24 at 17:05
1

Since the question of the value of the sum has already been addressed, let me tackle the question about the convergence.

Lemma. For any $a \in \mathbb{C} \setminus\{0,-1,-2,\ldots\}$,

$$\sum_{k=0}^{\infty} \frac{(-1)^k}{k+a} = \sum_{n=0}^{\infty} \frac{n!}{2^{n+1}a(a+1)\cdots(a+n)}.$$

Also, for any given positive integer $N$, there exist constants $c_0, \ldots, c_{N-1}$ such

$$\sum_{k=0}^{\infty} \frac{(-1)^k}{k+a} = \sum_{n=0}^{N-1} \frac{c_n}{a+n} + \mathcal{O}(a^{-N-1}) $$

as $\operatorname{Re}(a) \to \infty$.

I think this can be directly obtained by the Euler's series acceleration, although a direct proof is available (see below). Repeatedly applying this lemma then proves that OP's series converges.


Proof of Lemma. Assume first that $\operatorname{Re}(a) > 0$. For a given positive integer $K$,

$$ \begin{align*} \sum_{k=0}^{K-1} \frac{(-1)^k}{k+a} &= \sum_{k=0}^{K-1} (-1)^k \int_{0}^{1} x^{k+a-1} \, \mathrm{d}x \\ &= \int_{0}^{1} \sum_{k=0}^{K-1} (-1)^k x^{k+a-1} \, \mathrm{d}x \\ &= \int_{0}^{1} \frac{1 - (-x)^K}{1 + x} x^{a-1} \, \mathrm{d}x \\ &\to \int_{0}^{1} \frac{x^{a-1}}{1 + x} \, \mathrm{d}x \end{align*} $$

as $K \to \infty$ by the dominated convergence theorem. Now, by performing integration by parts, we get

$$ \begin{align*} \int_{0}^{1} \frac{x^{a-1}}{1 + x} \, \mathrm{d}x &= \left[ \frac{x^a}{a(x+1)} \right]_{0}^{1} + \frac{1}{a} \int_{0}^{1} \frac{x^{a}}{(x+1)^2} \, \mathrm{d}x \\ &= \frac{1}{2a} + \frac{1}{2^2 a(a+1)} + \frac{2!}{a(a+1)} \int_{0}^{1} \frac{x^{a+1}}{(x+1)^3} \, \mathrm{d}x \\ &\qquad \vdots \\ &= \sum_{n=0}^{N-1} \frac{n!}{2^{n+1}a(a+1)\cdots(a+n)} + \frac{N!}{a(a+1)\cdots(a+N-1)} \int_{0}^{1} \frac{x^{a+N-1}}{(x+1)^N} \, \mathrm{d}x \tag{*} \end{align*} $$

Since the factor $\frac{N!}{a(a+1)\cdots(a+N-1)}$ grow at most polynomially fast and $0 \leq \frac{x}{x+1} \leq \frac{1}{2}$ for $0 \leq x \leq 1$, the remainder term converges to $0$ as $N \to \infty$, proving the first assertion of the lemma when $\operatorname{Re}(a) > 0$. The general case then easily follows from the principle of analytic continuation.

The second assertion also easily follows from the formula $\text{(*)}$, first expanding the summation part using the partial fraction decomposition and then noting that the remainder term decays at least as fast as $a^{-N}$ as $\operatorname{Re}(a) \to \infty$ for each fixed $N$.

Sangchul Lee
  • 181,930
0

I think your idea is good. Assuming your sums don't include 0, you can actually use stars and bars to give an expression for $\Phi(n)$:

$\Phi(n) = \binom{n-1}{2}$. If you are summing over the 0s as well (i.e. if $(0,0,n)$ is a valid triple) then that will become $\Phi(n) = \binom{n+2}{2}$.

Now plug that into your expression and you get

$\sum_{i=0} \frac{(n+2)(n+1)(-1)^n}{2(n+1)} = \frac{1}{2} \sum_{i=0} (n+2)(-1)^n$.

Obviously this does not converge.

user76284
  • 6,408
blueslimr
  • 327
  • 8
  • 4
    This doesn't answer the question, unfortunately, because the divergence of the rearranged sum does not necessarily mean that the original sum diverges. – Greg Martin Sep 07 '24 at 17:29
  • Ah, yes, that's true. I was going off the suggestion, but I see that you are right. If we do the summation correctly we will actually start we will actually start summing over all triples of the form $(0,0,k)$ which will will be the alternating harmonic series (ignoring the +1) which does converge. As an aside: did you edit 'stars and bars' to 'sticks and stones'? Does 'stars and bars' have some negative connotation? I thought it simply referred to 'Old Glory'. – blueslimr Sep 07 '24 at 17:38
0

With no proof

If $k$ is the number of $\sum_{0}^\infty$, the result for $k>1$ is given by $$\large\frac {a_k}{2\,b_k}$$ where the $a_k$ and $b_k$ respectively correspond to sequences $A090633$ and $A090634$ in $OEIS$.

The first ones are $$\left\{\frac{1}{2},\frac{3}{8},\frac{7}{24},\frac{15}{64},\frac{31}{160},\frac{21}{128},\frac{127}{896}, \frac{255}{2048},\frac{511}{4608}\right\}$$