Brought to you by:
Paper

Approximating $\ln 2$ by numbers in the field $ {\mathbb{Q}}(\sqrt{2}\,)$

and

© 2018 RAS(DoM) and LMS
, , Citation M. Yu. Luchin and V. Kh. Salikhov 2018 Izv. Math. 82 549 DOI 10.1070/IM8518

1064-5632/82/3/549

Abstract

Using a new integral construction combining the idea of symmetry suggested by the second author in 2007 and the integral introduced by Marcovecchio in 2009, we obtain a new bound for the approximation of $\ln{2}$ by numbers in the field $\mathbb{Q}(\sqrt{2}\,)$.

Export citation and abstract BibTeX RIS

The research was partially supported by the Russian Foundation for Basic Research (project no. 12-01-00171).

§ 1. Introduction. Integral construction. Arithmetic part

The object of this paper is to prove the following result.

Theorem 1.  Let $\mu{\kern1pt}{>}{\kern1pt}7.98101$, $p_1,p_2,p_3,p_4 {\kern1pt}{\in}{\kern1pt}\mathbb{Z}$, $(p_3,p_4){\kern1pt}{\ne}{\kern1pt}(0,0)$, $P{\kern1pt}{=}\max_{1 {\kern.7pt}{\leq}{\kern.7pt} i {\kern.7pt}{\leq}{\kern.7pt} 4}|p_i|$, and $P>P_0(\mu)$. Then the following inequality holds:

Equation (1)

A similar bound for $\mu=12.4288$ was obtained in [1], and this result was improved in [2]; the corresponding value of $\mu$ was $9.307$.

The derivation of the new bound (1) is related to an application of the following integral construction. Let $h,j,k,l,m,q\in\mathbb{Z^+}$, $h+j+q=k+l+m$, $h+j-k\geq 0$, $k+l-j\geq0$ and $k+m-h\geq 0$. Let $x\in\mathbb{C}$, $\operatorname{Re}x> 0$ and $x\ne 1$. Consider the integral

Equation (2)

The result in Theorem 1 is obtained by taking

Equation (3)

Equation (4)

The integral (2) differs from the integral introduced by Marcovecchio in [3], p. 148, formula (5), only by the factor $\sqrt{{s}/{(s-1)}}$ in the denominator of the integrand. The integral (2) was considered for the first time in [4]. To make the picture complete, we describe a brief scheme of some transformations of this integral (see [4], Russian pp. 484, 485). We denote the integrand of (2) by $G(s,t)$. Then

Equation (5)

Evaluating the residue in (5) and making the change $s={z^2}/{(z^2-1)}$ in the integral with respect to the variable $s$, we obtain the relation

Equation (6)

where

Equation (7)

Equation (8)

By choosing for $s\in[0,-\infty)$ a value $z\in[0,1)$, we see from (5) and (6) that

Equation (9)

We write

Equation (10)

let $K$ be the ring of numbers of the form $a+b\sqrt{2}$, where $a, b\in\mathbb{Z}$, and for positive integers $M\in\mathbb{N}$ write $q_M=\operatorname{\text{lcm}}(1,2,\dots,M)$ with $q_0= 1$.

Lemma 1.  Let $M_0=\max(2k+2l-2j,\, h+j-k,\, k+m-h)$, $m\geq q$. Then the following representation holds for all $l_1\leq k+m-h$:

Equation (11)

where the $a(l_1), b(l_1)$ belong to $\mathbb{K}$.

Proof.  For $N\in\mathbb{Z^+}$ we write

Since the integrand (8) of the integral in (10) is even, we have the following expansion into a sum of simplest fractions:

Equation (12)

where $P(z)\in\mathbb{K}[z]$, $\deg{P(z)}=2k+2l-2j-2$, and, moreover,

By Leibniz' formula, we see from (12) that

where

Therefore,

where the $k_{\nu}(\overline{m})$ belong to $\mathbb{K}$.

However, $m_2+1.5m_3+m_4\leq 1.5(h+j-k+l_1+1-\nu)$. Hence,

that is,

Equation (13)

By (12), we have

Equation (14)

Obviously,

Further, it follows from the definition of $M_0$ that $q_{M_0}\int_0^1P(z)\, dz\equiv A_1\in\mathbb{K}$. It can also readily be seen that

for all $\nu=2,\dots,h+j-k+l_1+1$. Then it follows from (13) and (14) that

whence, since $m\geq q$, (11) follows. $\square$

Corollary 1.  The integral (2) admits the following representation for $m\geq q$:

Equation (15)

Proof.  We have

where $C(l_1)\in\mathbb{K}$. Applying Lemma 1, we see from (7) and (9) that

where $a,b\in\mathbb{K}$, and this proves Corollary 1. $\square$

Along with the family of parameters (4), we shall use a more general situation in which

Equation (16)

where $h', j', k', l', m', q'\in\mathbb{Z^+}$. It is convenient to denote the integral (2) for parameters of the form (16) and for $x$ of the form (3) as follows:

Equation (17)

For the family of parameters (16) we write

Equation (18)

Let $p$ be a prime, $p>\sqrt{Mn}$, and $\omega=\{{n}/{p}\}$ the fractional part of the number ${n}/{p}$. Consider the inequalities

Equation (19)

We denote by $\Delta_n$ the product of all primes $\smash{p>\sqrt{Mn}}$ such that $\omega=\{{n}/{p}\}$ satisfies at least one of the inequalities (19). The following lemma sharpens the result obtained in Corollary 1.

Lemma 2.  When $m\geq q$ the integral (17) admits the representation

Equation (20)

where $A_n, B_n\in\mathbb{K}$, $n\in\mathbb{N}$.

Proof.  The representation (20) follows from (15) by the standard procedure of refining the denominator (see, for example, Lemma 3 in [6]). The inequalities (19) were obtained for the integral (2) for the first time in [4], Russian p. 491, inequalities (11). The inequalities (19) differ somewhat from those considered, for the same purpose, by Marcovecchio in [3], inequalities (31). $\square$

The following lemma contains the final version of a linear form of the form (20) which we shall use to prove Theorem 1.

Lemma 3.  The following representation holds ( see (17)):

Equation (21)

where the $\Lambda_i(n)$ belong to $\mathbb{Z}$ and $\Delta_n$ is defined by the inequalities (19) for the family of parameters (4).

Proof.  It was proved in [4], equation (10), that the equation $J_n(h', j', k', l', m', q')=x^{l-q}J_n(k', m', h', q', j', l')$ holds (see (17)), that is,

For the family of parameters (4) we have from (18) that

However, then

By Lemma 2,

where $A_n, B_n\in\mathbb{K}$, which implies (21). $\square$

We conclude §1 by proving the following important lemma.

Lemma 4.  Let $n,d\in\mathbb{N}$, $\theta\in\mathbb{R}$, $\sqrt{d}\notin\mathbb{N}$, and

where the $\Lambda_i(n)$ belong to $\mathbb{Z}$, and let $\Lambda(n)=\max_{1 \leq i \leq 4}|\Lambda_i(n)|$. Let

and for some constant $\gamma_3>\gamma_2$ and every $\varepsilon_1, \varepsilon_2> 0$ let there be an $N=N(\varepsilon_1, \varepsilon_2)$ such that the following inequalities hold for every $n\geq N$ and at least one of the values $m\in\{n, n+1\}$:

Equation (22)

Further, let $\gamma_1+\gamma_2\,{>}\,0$, $\mu\,{>}\,2(\gamma_1+\gamma_3)/(\gamma_3-\gamma_2)$, $p_1,p_2,p_3,p_4\,{\in}\,\mathbb{Z}$, $(p_3, p_4)\,{\ne}\,(0, 0)$, $P=\max_{1 \leq i \leq4}|p_i|$ and $P>P_0(\mu)$. Then

Equation (23)

Proof.  We fix a $\tau\,{\in}\,(0,(\gamma_3-\gamma_2)/2)$ such that $2(\gamma_1+\gamma_3)/(\gamma_3-\gamma_2-2\tau)\,{<}\,\mu$ and write

Equation (24)

Let us choose $\varepsilon_1={\varepsilon}/{3}$, $\varepsilon_2={\tau}/{2}$, $N=N({\varepsilon}/{3},{\tau}/{2})$ in the inequalities (22). We introduce an $n_0\in\mathbb{N}$, where $n_0\geq N$ and the following family of conditions holds: for all $n\in\mathbb{N}$, $n>n_0$,

Equation (25)

Equation (26)

Equation (27)

Equation (28)

Equation (29)

Equation (30)

Equation (31)

Let $P_0(\mu)=\exp\bigl(({(\gamma_3-\gamma_2)}/{2}-\tau)n_0\bigr)$ and $P>P_0(\mu)$. We define $n\in\mathbb{N}$, $n> n_0$, by the inequalities

Equation (32)

Then the inequalities (22) hold for at least one of the values $m\in\{n, n+1\}$. We write (see (25))

Let us first consider the case in which the inequalities (22) hold for $m=n$. Two situations are possible: $\omega_1=\omega_2$, $\omega_1\ne\omega_2$.

1) Let $\omega_1=\omega_2$. We have

We have $|L_n|>e^{-(\gamma_3+{\varepsilon}/{3})n}$ by (22), and $|\Lambda_1(n)\sqrt{d}+\Lambda_2(n)|<e^{(\gamma_1+{\varepsilon}/{3})n}$ by (28). Hence, by (32), the condition $\gamma_1+\gamma_3>\gamma_1+\gamma_2> 0$ and the inequality (24) yield that

which coincides with the inequality (23).

2) Let $\omega_1\ne \omega_2$. We consider

where

Obviously,

Therefore,

Equation (33)

We claim that

By the inequality (33), it suffices to prove that

Equation (34)

Applying the inequalities (22), (26), (32), (27) and (29), we obtain that

This proves the inequality (34), and therefore $|\beta_2|<\Omega/2$. However, $\Omega=|\beta_1-\beta_2|\leq |\beta_1|+|\beta_2|$. Hence, $|\beta_1|\geq \Omega-|\beta_2|>\Omega/2$. Using the inequalities (33), (26), (27), (28) and (32) in succession and taking into account that $\gamma_1+\gamma_2> 0$, we have

which coincides with the inequality (23).

Let us now consider the case in which the inequalities (22) hold for $m=n+ 1$. Again, two situations are possible: $\omega_1=\omega_2$, $\omega_1\ne\omega_2$.

3) Let $\omega_1=\omega_2$. As in part 1), we have

Applying the inequalities (22), (28) and (30), we obtain that

4) Let $\omega_1\ne \omega_2$. The arguments here are similar to that in part 2), where $n\to n+1$, in particular, in the inequalities (33) and (34). Applying the inequalities (22), (26), (32) and (27), we obtain that

which proves the inequality (34) with $n\to n+1$, and simultaneously the inequality $|\beta_1|>\Omega/2$. Then, applying the inequalities (26), (27), (28), (31) and (32) in succession, we obtain

which coincides with the inequality (23).

This completes the proof of the lemma. $\square$

Remark 1.  Assertions similar to Lemma 4 were used in [1], [2] and [5].

§ 2. Asymptotics

To prove Theorem 1, we shall apply Lemma 4 to the linear form (21). We need to evaluate the constants $\gamma_1$, $\gamma_2$ and $\gamma_3$. In this section, we evaluate $\gamma_1$ and $\gamma_3$ and, in the next section, the constant $\gamma_2$. To evaluate both the constants $\gamma_1$ and $\gamma_3$, we apply the saddle-point method. We have (see (17) and (21))

where

Equation (35)

The saddle points are the solutions of the system $f_s'(s, t)= 0$, $f_t'(s, t)= 0$ that differ from the zeros of the function $f(s, t)$. In [4], Russian p. 492, equations (12), this system was solved in the general case for the integral (17). For the function $f(s, t)$ written above we have three saddle points:

Equation (36)

Equation (37)

$(s_3, t_3)=(\overline{s}_2, \overline{t}_2)$, the complex conjugate of $(s_2, t_2)$. We write $\xi=(s, t)\in\mathbb{C}^2$.

Lemma 5.  Let $\xi^0$ be a non-degenerate saddle point of the function $S(\xi)$, let $\gamma$ be a two-dimensional smooth complex manifold with boundary, let $\xi^0$ be an interior point of $\gamma$, let the functions $\varphi(\xi)$ and $S(\xi)$ be holomorphic at the point $\xi^0$, and let also $\max_{\xi\in\gamma}\operatorname{Re} S(\xi)$ be attained only at the point $\xi^0$, let

Then, as $\lambda\to+\infty$,

Equation (38)

Proof.  This assertion is proved in the Fedoryuk's monograph [7], p. 259, Proposition 1.1. $\square$

Lemma 6.  For the linear form (21) we have the equation

Equation (39)

Proof.  Consider the function

Let $L_1$ be the circle $|\tau-{1}/{x}|={1}/{t_1}-{1}/{x}$ and $L_2$ the circle $|\delta-{1}/{x}|={1}/{s_1}-{1}/{x}$. We note that, by (3) and (36), we have ${1}/{x}<{1}/{t_1}<{1}/{s_1}< 1$. Obviously, $\max_{(\delta, \tau)\in L_2\times L_1}\ln{|g(\delta, \tau)|}$ is attained only at the point $({1}/{s_1},{1}/{\tau_1})$. We denote by $L_1^*$ the image of the circle $L_1$ under the map $t={1}/{\tau}$ and by $L_2^*$ the image of the circle $L_2$ under the map $s={1}/{\delta}$. Then it follows from the definition of the function $g(\delta, \tau)$ that $\max_{(s, t)\in L_2^*\times L_1^*}\ln{|f(s, t)|}$ is attained only at the point $(s_1, t_1)$. For the integral $J_n$, by (35), we have

Equation (40)

where $A_n,B_n \in\mathbb{Q}\cdot \sqrt{2}\oplus\mathbb{Q}$ (see Corollary 1).

We claim that

Equation (41)

where the circles $L_1^*$ and $L_2^*$ are traversed in the positive direction. Equations of the form (41) are standard and occur in many papers, for example, in [3] and [8].

In the proof of Lemma 1, for the integral in (10) we have, by (12) and (14)

where $A(l_1),B(l_1)\in\mathbb{Q}\cdot \sqrt{2}\oplus\mathbb{Q}$, $A(l_1)=k_1/2=-(\operatorname{res}_{z=z_0}R_{l_1}(z))/2$ and $z_0=3+2\sqrt{2}$, and then it follows from (7), (9) and (40) that $A_n=(\operatorname{res}_{z=z_0}R(z))/2$. Applying (6), we obtain

where $l$ stands for the contour going around the point $z_0$ in the positive direction and mapping onto the circle $L_2^*$ under the map $s={z^2}/{(z^2-1)}$. Thus, the formula (41) is proved.

Let $\gamma_1$ be a small arc of the circle $L_1^*$ with centre at the point $t_1$, let $\gamma_2$ be a small arc of the circle $L_2^*$ with centre at the point $s_1$, let $\gamma=\gamma_2\times\gamma_1$, let $\Gamma=(L_2^*\times L_1^*)\setminus\gamma$; let $\xi^0=(s_1, t_1)$, and let $\max_{\xi\in\Gamma}|f(\xi)|=F<|f(\xi^0)|$. We can define some branch $\ln{f(\xi)}=\ln{|f(\xi)|}+ih(\xi)$ on $\gamma$ holomorphic at the point $\xi^0$.

It follows from (35) and (36) that $f(\xi^0)=-|f(\xi^0)|$, and therefore one can choose $h(\xi^0)=\pi$. Further, the function $\varphi(\xi)$ is holomorphic at the point $\xi^0$ since $\operatorname{Re}s> 1$, $s\in\gamma_2$. Obviously, $\varphi(\xi^0)\ne 0$. By (41) we have

Equation (42)

We apply Lemma 5 to the first integral in (42) with $\lambda=n$ and $S(\xi)=\ln{f(\xi)}$. The conditions of Lemma 5 are satisfied, since the remaining non-degeneracy condition for the saddle point $\xi^0$ can readily be verified: $\det{S_{\xi\xi}''(\xi^0)}\approx4.38471\times10^{11}\ne 0$. Then by (38) we obtain

We estimate the second integral in (42) trivially:

for some positive constant $C$. Then $\lim_{n\to\infty}({1}/{n})\ln{|A_n|}=\ln{|f(s_1, t_1)|}$.

To complete the proof of the lemma, it remains to evaluate the limit

We note that $\lim_{n\to\infty}({1}/{n})\ln{(q_{56n})}=56$ and evaluate $\Delta=\lim_{n\to\infty}({1}/{n})\ln{\Delta_n}$. Let us write out the inequalities (19) for the family of parameters (4):

Equation (43)

The set $E$ of numbers $\omega$ satisfying at least one of the inequalities (43) has the form

Let $\psi(x)={\Gamma'(x)}/{\Gamma(x)}$, where $\Gamma(x)$ stands for the gamma function. Then, in the standard way (see Lemma 6 in [9]), we obtain

Equation (44)

This completes the proof of the lemma. $\square$

Lemma 7.  For the linear form (21) let

Equation (45)

where the function $f(s,t)$ is defined in (35), the point $(s_2,t_2)$ is of the form (37), and $\Delta=\lim_{n\to\infty}({1}/{n})\ln{\Delta_n}$ is evaluated in (44).

Futher, let $\varepsilon_1,\varepsilon_2> 0$. Then there is an $N\in\mathbb{N}$, $N=N(\varepsilon_1, \varepsilon_2)$, such that the following inequalities hold for all $n\geq N$ and for at least one of the values $m\in\{n, n+1\}$:

Equation (46)

Proof.  Let $l_1$ be a ray, in the complex plane $s$, which issues from zero and passes through the point $s_2$ and let $l_2$ be an analogous ray, in the plane $t$, which passes through the point $t_2$. By (35), we have

Let $J^{(1)}=A+Bi$. Then $J^{(2)}=A-Bi$, that is, $J_n=({1}/{(2\pi i)})(J^{(1)}-J^{(2)})={B}/{\pi}=-({1}/{\pi})\operatorname{Im}J^{(2)}$. Let us move the integration with respect to $s$ in the integral $J^{(2)}$: $(0,-\infty)\to l_1$, and let us similarly move the integration with respect to the variable $t$: $(0, -i\infty)\to l_2$. Since there are no singular points of the integrand in the domains between the indicated rays, it follows that $J^{(2)}=\int_{l_1}ds\int_{l_2}G(s,t)\, dt$. Hence,

Equation (47)

As in Lemma 6, we write $\xi=(s,t)$ and $\xi^0=(s_2,t_2)$.

Simple computer calculations show that $\max_{\xi\in\,l_1\times l_2}|f(\xi)|$ is attained only at the point $\xi^0$. Let $\gamma_1$ be a small segment of $l_1$ containing $s_2$ as an interior point, let $\gamma_2$ be a similar segment of $l_2$ containing $t_2$ as an interior point, and let $\gamma=\gamma_1\times\gamma_2$, $\Gamma=(l_1\times l_2)\setminus\gamma$ and $\max_{\xi\in\Gamma}|f(\xi)|=F<|f(\xi^0)|$.

As in Lemma 6, one can apply Lemma 5 to the integral $\int_{\gamma}G(\xi)\,d\xi$, since

In our situation, the equation (38) becomes

We have $h(\xi^{(0)})\equiv\omega\approx 1.9062$. Let $\varphi(\xi^0)(\det{S_{\xi\xi}''(\xi^0)})^{-{1}/{2}}=re^{i\alpha}$, $r> 0$. Then (see (47))

Obviously,

Let us write $\omega_0={(\pi-\omega)}/{2}$. We claim that for every $n\in\mathbb{N}$

Indeed, let $n\omega+\alpha=2\pi n_1+\alpha_1$, $(n+1)\omega+\alpha=2\pi n_2+\alpha_2$, where $n_1,n_2\in \mathbb{Z}$, $\alpha_1,\alpha_2\in[0,2\pi)$. If $|\sin{(n\omega+\alpha)}|<\sin{(\omega_0)}$, then

In this case, $\alpha_2\notin A$, that is, $|\sin{((n+1)\omega+\alpha)}|\geq \sin{(\omega_0)}$. Let $m\in\{n, n+1\}$ be the value for which $|\sin{(m\omega+\alpha)}|\geq \sin{(\omega_0)}$. We choose $C_1=r\sin{(\omega_0)}$ and $C_2=4r$. Then for $n\geq N_1$, we obtain from (17) that the integral $J_m$ has the bounds

Since $\lim_{n\to\infty}({1}/{n})\ln({q_{56n}}/{\Delta_n})=56-\Delta$, it follows that for all $m\geq N_2$ we have

Let ${C_1}/{m}\,{>}\,e^{-({\varepsilon_1}/{2})m}$, ${C_2}/{m}\,{<}\,e^{({\varepsilon_2}/{2})m}$ and $N\,{=}\,\max{(N_1, N_2, N_3)}$ for all $n\,{\geq}\,N_3$. Then, by (21), for $n\geq N$ we have

and similarly $|L_m|>e^{-(\gamma_3+\varepsilon_1)m}$. This completes the proof of the lemma. $\square$

§ 3. Evaluation of the constant $\gamma_2$. Completion of the proof of Theorem 1

For the integral (2) we consider the function

Equation (48)

where $a(\alpha)={\alpha}/{(\alpha-1)}$, $\alpha\in(0,1)$, $l_t$ is the circle in the complex plane $t$ whose diameter is the segment $[{x}/{2},2x]$ on the real line, and the integration with respect to $l_t$ is carried out in the negative direction.

In this section, we write

In the following lemma we establish a relationship between the function $g(x, \alpha)$ and the integral (2).

Lemma 8.  The integral (2) and the function (48) satisfy the relation

Equation (49)

where $T$ is the operator $D_{k+m-h}x^{j}D_{h+j-k}$.

Proof.  By (48), we obtain in the standard way (see, for example, [3], pp. 167–169) that

Equation (50)

In (50) we make the change of variables $s_1=s$, $t_1={xs}/{t}$. For all $s< 0$ we obtain the integration with respect to the variable $t_1$ in the negative direction over the circle $L_{t_1}$ whose diameter is the segment $[2s,{s}/{2}]$. Further, $dt\,ds=-({xs_1}/{t_1^2})\, dt_1\, ds_1$, $s-t=({s_1}/{t_1})(t_1-x)$ and $t-x=({x}/{t_1})(s_1-t_1)$. Therefore,

The formula (50) becomes

Equation (51)

where the integration over $L_{t_1}$ is carried out in the positive direction.

Let us multiply both the sides of the identity (51) by $x^j$ and apply the operator $D_{k+m-h}$ to both sides. One can now replace the integration over $L_{t_1}$ by integration over the line $(-i\infty, i\infty)$ since both integrals are equal to $2\pi i\cdot \operatorname{res}_{t_1=s_1}G(s_1,t_1)$. Finally, passing to the limit as $\alpha\to 1-0$, we obtain the identity (49). $\square$

Our next task is to evaluate the function $g(x, \alpha)$.

Lemma 9.  The function (48) satisfies the relation

Equation (52)

Proof.  We evaluate the inner integral in (48) using Cauchy's residue theorem:

Let us make the change $u={s}/{(s-1)}$ in the integral thus obtained. Setting $s={u}/{(u-1)}$, $ds=-{du}/{(1-u)^2}$, $1-s={1}/{(1-u)}$ and $u\in[0, \alpha]$, we obtain

Finally, let us make the change $u=z^2$, $z\in[0, \sqrt{\alpha}\,]$. We have

This completes the proof of the lemma. $\square$

The following lemma gives a representation of the integral (2) in a form convenient for the evaluation of the constant $\gamma_2$. We restrict ourselves to the case $l\leq j$, which holds for the family (4) of parameters.

Lemma 10.  Let $l\leq j$. Then the integral (2) satisfies the relation

Equation (53)

Proof.  We write $A=z^2-1$, $B=1/(x-1)$, $l-j=-d$, $d\in\mathbb Z^+$. Then

Further,

Therefore,

Hence, it follows from (52) that

By (49), we have

We claim that $S_1=S_2= 0$. Consider the sum $S_1$. For all $\nu=0,1,\dots,d-1$, we obtain from Leibniz' formula that

Therefore, $D_{k+m-h}(P_{\nu}(x))= 0$. Hence, $S_1= 0$. Similarly, $S_2= 0$. Further, in the sum $S_3$ we set $\mu=d+\nu$, $\nu=0,1,\dots,k+l-j-1$. Then

Finally,

that is,

We have $J=S_3+S_4$, which coincides with the assertion of the lemma. $\square$

To evaluate the integral (2) by applying the formula (53), two more lemmas will be useful.

Lemma 11.  Let $M\in\mathbb{N}$, $a,b\in\mathbb{R}$. Then

Equation (54)

Proof.  We proceed by induction on $M$. When $M= 1$ we have

which coincides with (54) when $M= 1$.

Let us make the induction step $M\to M+1$. Using the induction assumption, we obtain

We write

We need to prove that $C_{\rho}=d_{\rho}$ for all $\rho=0,1,\dots,M+1$.

The following equation holds:

Further,

Finally, when $\rho\in[1, M]$ we need to show that

Consider two cases.

1. Let $b-a+M\ne 0$. Here

Therefore,

2. Let $b-a+M= 0$. Here

This completes the proof of the lemma. $\square$

Lemma 12.  The following equation holds for every $N\in\mathbb{N}$ and for arbitrary analytic functions $u=u(x)$ and $\vartheta=\vartheta(x)$:

Equation (55)

Proof.  By the definition of the operator $D_N$, it suffices to show that

Equation (56)

In turn, to prove the formula (56), it suffices to prove that the coefficients of $u^{(r)}\,\vartheta^{(N-r)}$, $r=0,1,\dots,N-1$, on both the sides of the equation (56) coincide, that is,

Equation (57)

We shall prove this by induction on $r$. When $r= 0$ we have

To carry out the induction step $(r-1)\to r$, note that

By the induction assumption,

and then the equation (57) holds. $\square$

In what follows, it is useful to note the asymptotics of generalized binomial coefficients:

For $x\in\mathbb{R}$ we introduce the function

Equation (58)

Obviously, the function $x^*$ is odd.

Lemma 13.  Let $n\in\mathbb{N}$, $n\to+\infty$, $b=b_0n+O(1)$, $r=r_0n+O(1)$, $b_0,r_0\in\mathbb{R}$, $r\in\mathbb{Z^+}$ and ${\binom{b}{r}}\ne 0$. Then the following equation holds:

Equation (59)

Proof.  It is clear that $r_0\geq 0$. If $r_0= 0$, then

and the formula (59) holds. Let $r_0> 0$ everywhere below. We consider several cases.

1) Let $b_0>r_0$. The following formula holds:

It follows from Stirling's formula that $\ln{\Gamma(x+1)}\,{=}\,x\ln{x}-x+O(\ln{x})$ as $x\,{\to}\,+\infty$. Taking into account that $b$, $b-r$, $r\to +\infty$ as $n\to+\infty$, we obtain

Similar formulae hold for $\Gamma(b-r+1)$ and $\Gamma(r+1)$. Therefore,

and the formula (59) is proved in the case under consideration.

2) Let $b_0=r_0$. Here

for some constant $C_1\ne 0$, and the formula (59) is verified trivially.

3) Let $0<b_0<r_0$. In this case,

where, as above, the constant $C_2$ is non-zero,

and we have used the fact that the function (58) is odd.

4) Let $b_0= 0$. As in cases 2) and 3), the following equation holds:

that is,

The last case remains.

5) Let $b_0< 0$. Here

This completes the proof of the lemma. $\square$

Let us apply the results thus obtained to the linear form (21). We first evaluate the integral $J_n$ in (35) using the formula (53):

Equation (60)

where $T$ stands for the operator $D_{56n}x^{38n}D_{20n}$.

We write $A=(3+2\sqrt{2})^{-140n}2^{-125n}$ and

Equation (61)

Equation (62)

Equation (63)

Then it follows from (21) and (60)–(63) that

Equation (64)

Let us combine the evaluation of the operators $D_M$ by the ordinary Leibniz formula and by the formula (54). In the latter case, we denote the operator $D_M$ by $D_{M}'$.

Let us begin with the evaluation of $\Sigma_{1,\nu}$. We have

By Lemma 11,

Non-zero summands occur in the last sum only when $r_2\in[6n, 26n+\nu-r_1]$. Further, it follows from (3) that

Therefore, it follows from (61) for $\nu=0,1,\dots,12n-1$ that

Equation (65)

Let us now evaluate $\Sigma_{2,\nu}$ for $\nu=0,1,\dots,7n-1$. We write

where $\Lambda= \Lambda_0n+O(1)$, $\Lambda\in\mathbb{Z^+}$, will be chosen below to optimize the bound for $\gamma_2$.

We have

As in the evaluation of $\Sigma_{1,\nu}$, we have

Therefore, it follows from (62) that

Equation (66)

Finally, let us calculate the value of $\Sigma_{3}$ using Lemma 12. We write

Then

Similarly,

Equation (67)

We note that $v=({1}/{2})\ln{\bigl(x+x-1+2\sqrt{x(x-1)}\bigr)=({1}/{2})\ln{\sqrt{2}}} =({1}/{4})\ln{2}$.

Hence, the coefficient at $\ln{2}$ in the linear form (21) is

Equation (68)

Remark 2.  The formula (68) enables one to find the constant $\gamma_1$ (see Lemma 6) using the terms of some positive series rather than by applying the saddle-point method. The former method is very simple, and therefore it is popular. It has been used many times in recent years; see, for example, [3], [6] and [9]. Therefore, we restrict ourselves to rather brief comments.

Since ${1}/{x}< 1$, it follows that

is a convergent series. Therefore, differentiating termwise, we obtain the positive series

The desired asymptotic behaviour is given by the maximal term of the series. Let us find this term by solving the equation

We have $s'\approx 3160.03$. Correspondingly, the index of the maximal term of the series is $s=s'n+O(1)$. Therefore, as in the papers indicated above,

by Lemma 13, and, using the equation (68) we obtain the result coinciding with that of Lemma 6. We note that the reasoning used in Remark 2 does not involve the evaluation of the constant $\gamma_2$.

Let us return to the proof of Theorem 1. The first term in the sum (67) is evaluated in the same way as $\Sigma_{2,\nu}$, where $\nu=7n-{1}/{2}$. The half-integer value of $\nu$ enables us to extend the bounds of variation of $r_1$, $r_2$, $\rho$. Instead of (66), we obtain

Equation (69)

Let us now evaluate the summands $S_3(\lambda_1)$ in the sum in (67). As above, let

We obtain in succession

Therefore,

Equation (70)

It remains to evaluate $S_3^*(\lambda)$. Let

We obtain in the standard way that

Thus, in the second case, we have

Equation (71)

All the summands in (64) evaluated using the formulae (65), (66), (70) and (71) are of the form $(3-2\sqrt{2})^{N}R$ or $(3-2\sqrt{2})^{N}\sqrt{2}\,R$, and the summands in the formula (69) are of the form $(3-2\sqrt{2})^{N}R\ln{2}$ or $(3-2\sqrt{2})^{N}R\sqrt{2}\,\ln{2}$, where $R\in\mathbb{Q}$ and $N\in\mathbb{Z^+}$. We note that $(3+2\sqrt{2})^{N}=A_N+B_N\sqrt{2}$ and $(3-2\sqrt{2})^{N}=A_N-B_N\sqrt{2}$ for $N\in\mathbb{N}$, where $A_N,B_N\in\mathbb{N}$. Correspondingly, $\sqrt{2}(3-2\sqrt{2})^N=-2B_N+A_N\sqrt{2}$. Obviously, $\lim_{N\to\infty}({1}/{N})\ln{A_N}=\lim_{N\to\infty}({1}/{N})\ln{B_N}=3+2\sqrt{2}$.

The total number of summands in $L_n$ is estimated as $O(n^5)$. Thus, for $\gamma_2$ we have the bound $\lim_{n\to\infty}\sup({1}/{n})\ln{\Lambda}\leq 56-\Delta+\lim_{n\to\infty}({1}/{n})\ln{|S|}$, where $S$ is the summand of maximal modulus among all above sums after replacing $3-2\sqrt{2}$ by $3+2\sqrt{2}$. The asymptotic behaviour of the binomial coefficients is calculated using Lemma 13.

Computer calculations show that the corresponding maximal summand is attained in the sum (69) for the following values of parameters: $\Lambda=\Lambda'n+O(1)$, $r_1=r_1'n+O(1)$, $r_2=r_2'n+O(1)$ and $\rho=\rho'n+O(1)$, where

Hence by Lemma 13, it follows from (69) that

Then, by Lemma 4, the inequality (1) holds for

This completes the proof of Theorem 1.

§ 4. Concluding remarks

4.1. The number $\ln{2}$ occupies a special position in the theory of Diophantine approximation. As Nesterenko said in the paper [9], ``$\ln{2}$ is a natural model for comparing the different methods developed for estimating the irrationality exponent for logarithms of rational numbers''. The best estimate for the irrationality measure of the number $\ln{2}$ is due to Marcovecchio [3]: $\mu\left(\ln{2}\right)\leq 3.57455390{\dots}\,$, and the estimate for the measure of quadratic irrationality is due to Polyansky [10]: $\mu_2(\ln{2})\leq 12.84161{\dots}\,$ . For obvious reasons, the estimate obtained in this paper is between these values.

4.2. In [1], to obtain an estimate of the form (1), the classical hypergeometric construction with integer parameters was used, while in [2] an analogous integral (but with half-integral parameters) was applied. In essence, the integral (2) is a linear combination of hypergeometric integrals with half-integral parameters (see the equations (7)–(9) in this paper). This linear combination is arranged in such a way that the coefficients of the corresponding linear form (see (35) and (21)) have a relatively small common denominator. We note that an integral of the form (2) was first applied in the paper [6] to obtain a new bound for the irrationality measure of the number ${\pi}/{\sqrt{3}}$.

4.3. The use of a combined differentiation with the help of Lemma 11, which enables us to reduce the value of the constant $\gamma_2$, also makes an important contribution.

4.4. The most difficult and laborious part of the work was to obtain a sufficiently small value of the constant $\gamma_2$. It is probable that the bound given in this paper is not definitive and that the methods developed in §3 can be improved.

4.5. The first author has obtained several new results using the integral (2). In particular, he managed to improve the bound for the approximation of the number $\pi$ by numbers in the field $\mathbb{Q}(\sqrt{3})$. These results are currently being prepared for publication.

The authors dedicate this paper to the centenary jubilee of Professor A. B. Shidlovskii, who was the teacher of the second author for many years.

Please wait… references are loading.
10.1070/IM8518