Proving several error terms for the divisor function $d(n)$

analytic-number-theoryasymptoticsnumber theoryprime numberssolution-verification

given the divisor function $d(n) = \#\{d|n\}$ I am trying to show the following:

  1. $d(n) = O(\sqrt{n})$
  2. $d(n) = O\Big(\exp\Big(\frac{c \log n}{\log \log n}\Big)\Big)$ for some constant $c > 0$
  3. $d(n) = O(n^{\epsilon})$ for any $\epsilon > 0$

I believe I have show one. I don't know where to start with two. Three I think I have made some progress.

My attempts are below:

  1. $d(n) = \sum_{d|n}1 = \sum_{ab = n}1 \leq 2\sum_{a \leq \sqrt{n}}1 + O(1) \leq 2 \sqrt{n} + O(1)$. Can I then just say this is $O(\sqrt{n})$?
  2. Let $\epsilon > 0$. In part three I show that each prime greater than $\exp(\frac{1}{\epsilon})$ contributes at most $1$ to the product. Therefore we consider small primes. By taylor expansion we can estimate $p_{j}^{\epsilon \alpha_{j}} \geq 1 + \epsilon \alpha_{j} \log p_{j}.$ Hence $\frac{\alpha_{j}+1}{p_{j}^{\epsilon \alpha_{j}}} \leq \frac{\alpha_{j}+1}{1 + \epsilon \alpha_{j} \log p_{j}}$. I am not sure what I can then bound this quantity by. If I find a suitable bound for this, I am thinking can I put this into the product defined in part three.
  3. If we consider $\frac{d(n)}{n^{\epsilon}}$ and write $n = p_{1}^{\alpha_{1}}p_{2}^{\alpha_{2}}\dots p_{k}^{\alpha_{k}}$ where $\alpha_{j} > 0$ and $p_{j}$ are distinct primes. Then we can rewrite the $n^{\epsilon}$ are a product of primes. By the fundamental theorem of arithmetic we can rewrite the divisor function as $\prod_{j=1}^{k}(\alpha_{j} + 1)$.

Therefore $\frac{d(n)}{n^{\epsilon}} = \prod_{j=1}^{k}\frac{\alpha_{j}+1}{p_{j}^{\epsilon \alpha_{j}}}$

We can now fix a prime $p_{j}$ and consider a single term $\frac{a_{j}+1}{p_{j}^{\epsilon \alpha_{j}}}$. For large $\alpha_{j}$ the denominator will dominate. For small $\alpha_{j}$ the epsilon in the denominator means that perhaps sometimes the numerator can dominate for sufficiently small $p_{j}$.

Splitting into cases I get:

  • Suppose $p_{j} \geq \exp(\frac{1}{\epsilon}).$ Then $p_{j}^{\epsilon \alpha_{j}} \geq \exp(\alpha_{j}) \geq 1 + \alpha_{j}$ by the taylor expansion of the exponential function. Therefore all large primes give a contribution of at most $1$ in the product.

For $p_{j} < \exp(\frac{1}{\epsilon})$, $ \frac{a_{j}+1}{p_{j}^{\epsilon \alpha_{j}}} \leq C_{p_{j}, \epsilon}$ (the constant doesn't depend on $a_{j}$ since $\frac{a + 1}{p_{j}^{\epsilon a}} \rightarrow 0$ as $a \to \infty$.) Therefore, each small prime gives a bounded contribution to the product.

Can I then say that the number of small primes is bounded (by finding a bound for $\exp(\frac{1}{\epsilon})$? Therefore the whole product is bounded.

Hence the result follows.

Thanks.

Best Answer

For the first question, the answer is yes. For the third question (i.e. the bold question), yes it works because you get $\frac{d(n)}{n^{\varepsilon}}\leqslant C_{\varepsilon}$ where $\displaystyle C_{\varepsilon}=\prod_{p<e^{1/\varepsilon}}C_{p,\varepsilon} $ does not depend of $n$. As for $2.$, I'll use $3.$. First notice that the bound $C_{\varepsilon}:=\sup\limits_{n\geqslant 1}\frac{d(n)}{n^{\varepsilon}}$ is reached for $\displaystyle n_{\varepsilon}:=\prod_p p^{\alpha_{p,\varepsilon}}$ where $\displaystyle\alpha_{p,\varepsilon}:=\left\lfloor \frac{1}{p^{\varepsilon}-1} \right\rfloor$. Indeed, if $f(\alpha):=\frac{\alpha+1}{p^{\alpha\varepsilon}}$, then $$ f(\alpha+1)\geqslant f(\alpha)\iff\frac{\alpha+2}{\alpha+1}\geqslant p^{\varepsilon}\iff\alpha\leqslant\alpha_{p,\varepsilon} $$ and $f$ reaches its maximum at $\alpha=\alpha_{p,\varepsilon}$. Now let $x_k:=\left(1+\frac{1}{k}\right)^{1/\varepsilon}$, then $$ \alpha_{p,\varepsilon}=k\iff \frac{1}{p^{\varepsilon}-1}-1<k\leqslant\frac{1}{p^{\varepsilon}-1}\iff x_{k+1}<p\leqslant x_k $$ for $k\geqslant 1$. Let $k_0:=\alpha_{2,\varepsilon}$, then there is no prime $p$ such that $p\leqslant x_{k_0+1}$ because $x_{k_0+1}<2$ so we have $$ n_{\varepsilon}=\prod_{k\leqslant k_0}\left(\prod_{x_{k+1}<p\leqslant x_k}p\right)^k $$ From this expression we can deduce the two following estimations : $$ \ln n_{\varepsilon}=\vartheta(x_1)+\mathcal{O}\left(x_1^{3/4}\right) \ \ \text{ and }\ \ \ln d(n_{\varepsilon})=(\ln 2)\pi(x_1)+\mathcal{O}\left(x_1^{3/4}\right) $$ Indeed, $\displaystyle\ln n_{\varepsilon}=\sum_{k\leqslant k_0}k(\vartheta(x_k)-\vartheta(x_{k+1}))=\sum_{k\leqslant k_0}\vartheta(x_k)$ and, using $x_2\leqslant x_1^{\frac{\ln 3}{\ln 2}-1}$, we have $$ \sum_{2\leqslant k\leqslant k_0}\vartheta(x_k)\ll k_0\vartheta(x_2)\ll k_0 x_2\ln x_2\ll x_2(\ln x_2)^2\ll x_1^{\frac{\ln 3}{\ln 2}-1}(\ln x_1)^2\ll x_1^{3/4} $$ because $\frac{\ln 3}{\ln 2}-1\approx 0.58\leqslant 0.75$. As for the other approximation, we have $$ \ln d(n_{\varepsilon})=\sum_{k\leqslant k_0}\ln(k+1)(\pi(x_k)-\pi(x_{k+1}))=\sum_{k\leqslant k_0}\ln\left(1+\frac{1}{k}\right)\pi(x_k) $$ and $$ \sum_{2\leqslant k\leqslant k_0}\ln\left(1+\frac{1}{k}\right)\pi(x_k)\ll\frac{k_0 x_2}{\ln x_2}\ll x_1^{3/4} $$ using the same arguments. Now let $R(x)$ be such that $\pi(x)-{\rm li}(x)\ll R(x)$ and $\vartheta(x)-x\ll R(x)$, then $$ \ln d(n)\leqslant \ln C_{\varepsilon}+\varepsilon\ln n\leqslant\ln d(n_{\varepsilon})-\varepsilon\ln n_{\varepsilon}+\varepsilon\ln n\leqslant(\ln 2)\pi(x_1)-\varepsilon\vartheta(x_1)+\varepsilon\ln n+\mathcal{O}\left(x_1^{3/4}\right)$$ We then use the bound $R(x)\gg x^{4/5}$ and we obtain $$ \ln d(n)\leqslant (\ln 2){\rm li}(x_1)-\varepsilon x_1+\varepsilon\ln n+\mathcal{O}(R(x)) $$ In order to cancel the first order terms, we chose $\varepsilon:=\frac{\ln 2}{\ln\ln n}$ so that $x_1=2^{1/\varepsilon}=\ln n$. We thus have $\ln d(n)\leqslant(\ln 2){\rm li}(\ln n)+\mathcal{O}(R(\ln n))$ and, using the bound $R(x)\ll\frac{x}{(\ln x)^2}$, we finally get $$ \ln d(n)\leqslant \frac{(\ln 2)\ln n}{\ln \ln n}+\mathcal{O}\left(\frac{\ln n}{(\ln \ln n)^2}\right)\underset{n\rightarrow +\infty}{\sim}\frac{(\ln 2)\ln n}{\ln \ln n} $$

Related Question