1 Introduction

Let \(n\ge 2\) be an integer, and consider the Hamiltonian system

$$\begin{aligned} {\left\{ \begin{array}{ll} \displaystyle z^{2}\frac{dq}{dz} =\ \nabla _{p}\mathcal {H}(z,q,p), &{}\\ \displaystyle z^{2}\frac{dp}{dz} =- \nabla _{q}\mathcal {H}(z,q,p), &{} \end{array}\right. } \end{aligned}$$
(1)

where \(q=(q_2,\ldots ,q_n)\), \(p=(p_2,\ldots ,p_n)\). Here

$$\begin{aligned} \nabla _{q}:=\left( \frac{\partial }{\partial q_2},\ldots , \frac{\partial }{\partial q_n} \right) , \quad \nabla _{p}:=\left( \frac{\partial }{\partial p_2},\ldots , \frac{\partial }{\partial p_n} \right) . \end{aligned}$$

The system (1) is equivalent to an autonomous one

$$\begin{aligned} {\left\{ \begin{array}{ll} \dot{q}_1=\ H_{p_1}, &{}\dot{q}=\ \nabla _{p}H, \\ \dot{p}_1= -H_{q_1}, &{}\dot{p}= -\nabla _{q}H, \end{array}\right. } \end{aligned}$$
(2)

where \(q_1=z\) and \(H(q_1,q,p_1,p) :=q_1^2 p_1+\mathcal {H}(q_1,q,p)\) or \(H(q_1,q,p_1,p) :=\) \( p_1+ q_1^{-2}\mathcal {H}(q_1,q,p)\). We say that the Hamiltonian system (2) is \(C^\omega \)-Liouville integrable if there exist first integrals \(\phi _j \in C^\omega \) \((j=1,\ldots ,n)\) which are functionally independent on an open dense set and Poisson commuting, i.e., \(\{ \phi _j, \phi _k \} =0\), \(\{ H, \phi _k \} =0\), where \(\{ \cdot , \cdot \}\) denotes the Poisson bracket. The Hamiltonian H is a first integral of this autonomous system. We abbreviate \(C^\omega \)-Liouville integrable to \(C^\omega \)-integrable or integrable if there is no fear of confusion.

In [2] Bolsinov and Taimanov showed a non \(C^\omega \)-integrability of some Hamiltonian system related with geodesic flow on a Riemannian manifold. Then Gorni and Zampieri showed similar results in the local setting, namely for a Hamiltonian system being singular at the origin they showed the non \(C^\omega \)-integrability (cf. [3, 5, 6]). In this paper we study the nonintegrability from a semi-global point of view. Namely we consider Hamiltonian system which is singular at the origin \(q_1=0\) as well as \(q_1=1\). We shall show that the system is integrable near the origin, while it is not integrable in the domain containing both \(q_1=0\) and \(q_1=1\). The Hamiltonian function is given by the arbitrary small non zero perturbation of an integrable Hamiltonian of the confluent generalized hypergeometric system (cf. Sect. 2).

More precisely, we consider

$$\begin{aligned} H =\sum _{j\in J'}\mu _{j}q_jp_j +\frac{q_1^2}{(q_1-1)^2} \sum _{j\in J }\mu _{j}q_jp_j +q_1^2p_1, \end{aligned}$$
(3)

where \(\mu _j\) are complex constants and J and \(J'\) are the sets of multi-indices such that

$$\begin{aligned} J \ne \emptyset , J'\ne \emptyset , J\cap J'=\emptyset , J\cup J'=\{2,\ldots ,n\}. \end{aligned}$$
(4)

The Hamiltonian is derived from the generalized hypergeometric system by confluence of singularities (cf. Sect. 2). The Hamiltonian system (2)–(3) determines the Hamiltonian vector field

$$\begin{aligned} \chi _{H}= & {} \ q_1^2\frac{\partial }{\partial q_1} - 2q_1p_1\frac{\partial }{\partial p_1} + \frac{2q_1}{(q_1-1)^3}\left( \sum _{j\in J }\mu _{j}q_jp_j \right) \frac{\partial }{\partial p_1} \nonumber \\&+ \sum _{j\in J'}\mu _{j}\left( q_j\frac{\partial }{\partial q_j} -p_j\frac{\partial }{\partial p_j} \right) + \frac{q_1^2}{(q_1-1)^2} \sum _{j\in J }\mu _{j}\left( q_j\frac{\partial }{\partial q_j} -p_j\frac{\partial }{\partial p_j} \right) . \end{aligned}$$
(5)

Let

$$\begin{aligned} H_1:=\sum _{j=2}^{n} p_j^2B_j(q_1,p). \end{aligned}$$
(6)

Note that \(H_1\) does not depend on q. Suppose that the nonresonance condition (NRC) holds:

$$\begin{aligned} \forall \gamma =(\gamma _2,\ldots ,\gamma _n) \in \mathbb {Z}^{n-1}{\setminus }\{0\},\ \sum _{j=2}^{n}\mu _{j}\gamma _j\ne 0, \end{aligned}$$
(7)

i.e. \(\mu _{j}\)’s are linearly independent over \(\mathbb {Z}^{n-1}\). Moreover, assume

(TC): For \(k\in J'\), the equation

$$\begin{aligned} q_1^2\frac{d}{dq_1}v - 2\mu _{k}v = B_k(q_1,0) \end{aligned}$$
(8)

has no solution v holomorphic at \(q_1=0\), and for \(k\in J \), the equation

$$\begin{aligned} q_1^2\frac{d}{dq_1}w - 2\mu _{k}\frac{q_1^2 w}{(q_1-1)^2} = B_k(q_1,0) +\mu _{k} \frac{q_1B_k(0,0)}{(q_1-1)^2} + B_k(0,0) \end{aligned}$$
(9)

has no solution w holomorphic at \(q_1=1\).

Let \(\varOmega _1\subset {\mathbb C}\) be a domain containing \(\{q_1 = 0, 1\}\), and \(\varOmega _2\subset {\mathbb C}^{2n-1}\) be a neighborhood of \((p_1,q,p)=(0,0,0)\) and define \(\varOmega :=\varOmega _1\times \varOmega _2\). Then we have

Theorem 1

Assume that (NRC) and (TC) are satisfied. Then, there exists \(\varOmega \) such that the Hamiltonian system (2) is not \(C^\omega \)-integrable in \(\varOmega \). More precisely, for every first integral \(\phi \) satisfying \(\chi _{H+H_1}\phi =0\) and holomorphic in \(\varOmega \), there exists a holomorphic function \(\psi \) defined in some neighborhood of the origin \(t=0\) such that \(\phi (q_1,q,p_1,p)=\psi (H+H_1)\) in some neighborhood of the origin.

In spite of the non integrability shown in Theorem 1 we have the integrability about a singular point of \(\chi _{H+H_1}\). We recall that the Hamiltonian system corresponding to \(H+H_1\) has irregular singularities at \(q_1=0\) and \(q_1=1\). We have

Proposition 1

Suppose that \(H_1(q_1,p)\) be independent of \(p_\nu \) for every \(\nu \in J'\). Then, \(\chi _{H+H_1}\) is analytically Liouville-integrable in some neighborhood of the origin.

Remark

(i) In Sect. 5 we show that (TC) holds on an open dense set in the set of analytic functions. (TC) also implies that \(H_1\) could be replaced by \(\varepsilon H_1\) with an arbitrary small \(\varepsilon \ne 0\). On the other hand, it is necessary in Theorem 1 that \(H_1\) does not vanish identically because H is integrable in view of Lemma 1 (cf. Sect. 3). Hence the non-integrability occurs by an arbitrary small non-zero generic perturbation.

By Proposition 1 we see that our class of Hamiltonians contains subclass for each of which the integrability at the origin holds. Hence the (non-) integrability in Theorem 1 is caused by the interference of singular points.

(ii) Of course, a globally integrable system is locally integrable. So, it is sufficient for the proof of Theorem 1 to prove the local non-integrability.

(iii) In these days, monodromy is usually treated from the point of view of the differential Galois theory (for example, see [7]) because of enrichment of the theory however, we treat it from another point of view.

2 Confluence of singularities

In this section we deduce (3) from the genelarized hypergeometric system

$$\begin{aligned} (z-C)\frac{dv}{dz}=Av, \end{aligned}$$
(10)

where \(C=\mathrm{diag}(\varLambda _1,{}^t\varLambda _1)\), \(\varLambda _1\) being \((n-1)\times (n-1)\) matrix with eigenvalues \(\lambda _2,\ldots ,\lambda _n\) such that \(\lambda _j\ne 0\) for all j (cf. [1, 4] ). For the sake of simplicity, we assume \(\varLambda _1=\mathrm{diag}(\lambda _2,\ldots ,\lambda _n)\). We assume \(A=\mathrm{diag}(A_1,A_1)\), where \(A_1\) is an \((n-1)\times (n-1)\) constant matrix satisfying \(\varLambda _1A_1=A_1\varLambda _1\). For simplicity, we further assume \(A_1=\mathrm{diag}(\tau _2,\ldots ,\tau _n)\).

Let \(v={}^t(q,p)\in \mathbb {C}^{2(n-1)}\). Define

$$\begin{aligned} H=\langle (z-\varLambda _1)^{-1}p,A_1q\rangle , \end{aligned}$$
(11)

where \(\langle (x_2,\ldots ,x_n),{}^t(y_2,\ldots ,y_n)\rangle :=\sum _{2\le k\le n}x_ky_k\). Then, (10) is written in the Hamiltonian system

$$\begin{aligned} \frac{dq}{dz}= H_p(z,q,p),\ \frac{dp}{dz}=-H_q(z,q,p). \end{aligned}$$
(12)

Now we operate the confluence of regular singularities. Let \(v_{\nu }\) and \((Av)_{\nu }\) denote the \(\nu \)th entry of v and Av, respectively. Then we can write (12) in the form

$$\begin{aligned} (z-\lambda _{\nu })\frac{dv_{\nu }}{dz}=(Av)_{\nu }. \end{aligned}$$

Substituting \(z=1/\zeta \), we have

$$\begin{aligned} -\zeta ^2\frac{dv_{\nu }}{d\zeta } =(\zeta ^{-1}-\lambda _{\nu })^{-1}(Av)_{\nu } . \end{aligned}$$
(13)

In the following, \(a\mapsto b\) denotes the replacement of a by b.

Let \(\zeta \mapsto \epsilon ^{-1}\eta \); and \(\lambda _{\nu }\mapsto \epsilon \lambda _{\nu }\) for \(\nu \in J\), \(\lambda _{\nu }\mapsto \lambda _{\nu }\) for \(\nu \in J'\). Multiply the \(\nu \)th row of A in (13) by \(\epsilon ^{-1}\) if \(\nu \in J'\) and take the limit \(\epsilon \rightarrow 0\). Then (12) is reduced to the Hamiltonian system

$$\begin{aligned} -\eta ^2\frac{dq}{d\eta }=\mathfrak {A}A_1q,\ -\eta ^2\frac{dp}{d\eta }=-{}^tA_1\mathfrak {A}p, \end{aligned}$$
(14)

where \(\mathfrak {A}=\mathrm{diag}(\mathfrak {A}_2,\ldots ,\mathfrak {A}_{n})\) and

$$\begin{aligned} \mathfrak {A}_{\nu }:= {\left\{ \begin{array}{ll} -\lambda _{\nu }^{-1} &{}(\nu \in J'), \\ (\eta ^{-1}-\lambda _{\nu })^{-1} &{}(\nu \in J). \end{array}\right. } \end{aligned}$$
(15)

Note that (14) is irregular singular at \(\eta =0\).

In order to introduce another singular point, choose any \(a\ne 0\) such that \(a\ne \lambda _j^{-1}\) for all j and put \(\zeta =\eta -a\). Let \(\zeta \mapsto \epsilon ^{-1}\zeta \) and \((A)_{\nu }\mapsto \epsilon ^{-1}(A)_{\nu }\). Make substitution \(a\mapsto \epsilon ^{-1}a\) for \(j\in J'\) and \(a\mapsto a\) for \(j\in J\) and take the limit \(\epsilon \rightarrow 0\). Then (12) is reduced to a Hamiltonian system with irregular points at 0 and \(-a\). Set \(a=-1\). Finally, by transforming to the autonomous system and putting \(\mu _j:=\mu _{j}\), we obtain (3).

3 Proof of Proposition 1

Let H and \(H_1\) be given by (3) and (6), respectively. First we show

Lemma 1

If \(k\in J\), then \(\chi _{H}\) has first integrals

$$\begin{aligned} q_k \exp \left( \frac{\mu _{k}}{q_1-1} \right) , \quad p_k \exp \left( -\frac{\mu _{k}}{q_1-1} \right) , \end{aligned}$$
(16)

while, for \(k\in J'\) it has

$$\begin{aligned} q_k \exp \left( \frac{\mu _{k}}{q_1} \right) , \quad p_k \exp \left( -\frac{\mu _{k}}{q_1} \right) . \end{aligned}$$
(17)

Note that \(\chi _{H}\) is analytically integrable at \(q_1=0\) or \(q_1=1\), because \(q_kp_k\) is an analytic first integral about the singular point \(q_1=0\) or \(q_1=1\).

Proof of Lemma 1

The assertion is easily verified in view of the definition of first integrals.

Remark

Lemma 1 says that in the \(C^\infty \) class the Hamiltonian is superintegrable. The perturbation in Proposition 1 breaks some first integrals, but not all of them. The remaining ones are not either sufficiently regular for integrability near both points.

Proof of Proposition 1

We have \(H_1\) not depending on \(p_k\), \(k\in J'\), \(q_1, q_k\), \(k=2,\ldots ,n\) by hypothesis and (6). So the dynamical equations give that \(q_k\), \(k \in J'\), \(q_1, p_k\), \(k=2,\ldots ,n\) are first integrals of \(H_1\). Thus in particular

$$\begin{aligned} p_k q_k, \ (k\in J'), \quad p_k \exp \left( -\frac{\mu _{k}}{q_1-1} \right) , \quad (k\in J) \end{aligned}$$
(18)

are first integrals of \(H_1\), and are analytic at 0. As these are also first integrals of H, they are in involution and first integrals of \(H+ H_1\). This ends the proof.

4 Proof of Theorem 1

Let \(\phi =:u\) be a holomorphic first integral in \(\varOmega \) and expand u at \(p=0\)

$$\begin{aligned} u=\sum _{\alpha }u_{\alpha }(q_1,q,p_1)p^{\alpha }. \end{aligned}$$
(19)

Substitute (19) into \(\chi _{H+H_1}u=0\) and compare the powers like \(p^0=1\) of both sides. Then we have the equation of \(u_0=u_0(q_1,q,p_1)\)

$$\begin{aligned} \left\{ q_1^2p_1,u_0\right\} + \sum _{j\in J'} \mu _{j}q_j \frac{\partial }{\partial q_j}u_0 + \frac{q_1^2}{(q_1-1)^2} \sum _{j\in J} \mu _{j}q_j \frac{\partial }{\partial q_j}u_0 = 0. \end{aligned}$$
(20)

Indeed, no constant term in p appears from \(\chi _{H_1}u\) in view of the definition of \(\chi _{H_1}\).

Substituting the expansion \(u_0=\sum _{\beta }u_{0,\beta }(q_1,p_1)q^{\beta }\) into (20), we see that \(U_0:=u_{0,0}\) satisfies \(\{q_1^2p_1, U_0\}=0\), namely

$$\begin{aligned} \left( q_1\frac{\partial }{\partial q_1} -2p_1\frac{\partial }{\partial p_1} \right) U_0=0. \end{aligned}$$
(21)

Substitute the expansion \(U_0=\sum _{\nu ,\mu } c_{\nu \mu }q_1^{\mu }p_1^{\nu }\) into (21). Then we have

\(\sum _{\nu ,\mu }c_{\nu ,\mu }(\mu -2\nu )q_1^{\mu }p_1^{\nu } =0\). It follows that \(c_{\nu ,\mu }=0\) for \(\mu \ne 2\nu \). Hence we obtain

$$\begin{aligned} U_0 = \sum _{\nu }c_{\nu ,2\nu }q_1^{2\nu }p_1^{\nu } = \sum _{\nu }c_{\nu ,2\nu }(q_1^{2}p_1)^{\nu }. \end{aligned}$$
(22)

It follows that there exists a function of one variable t, \(\phi _0(t)\) holomorphic in some neighborhood of \(t=0\) such that \(U_0=\phi _0(q_1^2p_1)\).

Next, we focus on the equation of \(u_{0,\beta }\) with \(\beta \ne 0\)

$$\begin{aligned} \{q_1^2p_1,u_{0,\beta }\} + \sum _{j\in J'}\mu _{j} \beta _ju_{0,\beta } +\frac{q_1^2}{(q_1-1)^2} \sum _{j\in J }\mu _{j} \beta _ju_{0,\beta } = 0. \end{aligned}$$

Expand

$$\begin{aligned} u_{0,\beta } = \sum _{\nu }\omega _{\beta ,\nu }(q_1)p_1^{\nu }, \end{aligned}$$
(23)

and consider the equation of \(\omega _{\beta ,\nu }\). If \(\nu =0\), then, by comparing the coefficients of \(p_1^0=1\), we have

$$\begin{aligned} q_1^2\frac{d}{dq_1}\omega _{\beta ,0} + \left( \sum _{j\in J'}\mu _{j}\beta _j +\frac{q_1^2}{(q_1-1)^2} \sum _{j\in J }\mu _{j}\beta _j \right) \omega _{\beta ,0} = 0. \end{aligned}$$
(24)

Since \(\beta \ne 0\), it follows from (NRC), (7), that either \(A':= \sum _{j\in J'}\mu _{j}\beta _j\ne 0\) or \( A := \sum _{j\in J }\mu _{j}\beta _j\ne 0\) is valid. If \(A'\ne 0\), then we have \(\omega _{\beta ,0}=0\) in some neighborhood of \(q_1=0\). Indeed, by subsituting the expansion \( \omega _{\beta ,0} =\sum _{l=0}^{\infty }C_lq_1^{l}\) into (24) and by using the relations

$$\begin{aligned} q_1^2\frac{d}{dq_1}\omega _{\beta ,0} =\sum _{l=0}^{\infty }C_l{l}q_1^{l+1} \end{aligned}$$

and

$$\begin{aligned} \frac{q_1^2}{(q_1-1)^2} \sum _{j\in J }\mu _{j} \beta _j\omega _{\beta ,0} =\sum _{l=0}^{\infty } C_l'\ {q_1}^{l+2} \end{aligned}$$

for some \(C_l'\), we obtain

$$\begin{aligned} C_0A'&=0\text {\quad i.e. }C_0=0, \\ C_1A'&+C_0\cdot 0=0\text {\quad i.e. }C_1=0, \\ C_2A'&+C_0' +C_1=0\text {\quad i.e. }C_2=0, \\&\cdots \end{aligned}$$

Note that \(C_0'=0\) since \(C_0=0\). Hence we have \(\omega _{\beta ,0}=0\).

In the case where \(A'=0\) and \(A\ne 0\), (24) is written in

$$\begin{aligned} (q_1-1)^2\frac{d}{dq_1}\omega _{\beta ,0} + A\omega _{\beta ,0} = 0. \end{aligned}$$
(25)

Similarly to the case \(A'\ne 0\), we obtain \(\omega _{\beta ,0}=0\) in some neighborhood of \(q_1=1\). Therefore, we have \(\omega _{\beta ,0}=0\) in \(\varOmega _1\).

Next, by comparing the coefficients of \(p_1^1=p_1\), we have the equation of \(\omega _{\beta ,1}(q_1)\)

$$\begin{aligned} \left( q_1^2\frac{d}{dq_1}-2q_1 \right) \omega _{\beta ,1} + \left( A' +\frac{q_1^2}{(q_1-1)^2} A \right) \omega _{\beta ,1} =0. \end{aligned}$$
(26)

Similarly to the above, \(A'\ne 0\) implies \(\omega _{\beta ,1}=0\) near \(q_1=0\), while \(A'=0\) and \(A\ne 0\) imply \(\omega _{\beta ,1}=0\) near \(q_1=1\). Hence we have \(\omega _{\beta ,1}=0\) in \(\varOmega _1\). By the same argument we obtain \(\omega _{\beta ,\nu }=0\) in \(\varOmega _1\) for all \(\nu \in \mathbb {N}\cup \{0\}\). It follows that \(u_{0,\beta }=0\) for all \(\beta \ne 0\).

Therefore, we have

$$\begin{aligned} u_0 = u_{0,0}(q_1^2p_1) + \sum _{\beta \ne 0} u_{0,\beta }(q_1^2p_1)q^{\beta } = \phi _0(q_1^2p_1) \end{aligned}$$
(27)

for some \(\phi _0(t)\) of one variable being analytic at \(t=0\). Note that

$$\begin{aligned} u|_{p=0}-\phi _0(H+H_1)|_{p=0}= & {} u_0(q_1,p_1)-\phi _0(H|_{p=0}) \\= & {} \phi _0(q_1^2p_1)-\phi _0(q_1^2p_1) \equiv 0. \end{aligned}$$

Hence, without loss of generality, we may assume \(u|_{p=0}=0\).

Next we consider \(u_{\alpha }=u_{\alpha }(q_1,p_1,q)\) for \(|\alpha |=1\). Write \(\alpha =e_k\) \((2\le k\le n)\) where \(e_k:=(0,\ldots ,0,1,0,\ldots ,0)\) is the kth unit vector. Then, \(u_{\alpha }\) satisfies

$$\begin{aligned}&\left\{ q_1^2p_1,u_{\alpha }\right\} + \sum _{j\in J'} \mu _{j} \left( q_j\frac{\partial }{\partial q_j}-\delta _{k,j} \right) u_{\alpha } \nonumber \\&\quad + \frac{q_1^2}{(q_1-1)^2} \sum _{j\in J } \mu _{j} \left( q_j\frac{\partial }{\partial q_j}-\delta _{k,j} \right) u_{\alpha } = 0, \end{aligned}$$
(28)

where \(\delta _{k,j}\) is the Kronecker’s delta, \(\delta _{k,j}=1\) if \(k=j\), and =0 if otherwise. Note that, because \(u_0=0\), \(\chi _{H_1}\) gives no term.

Substitute the expansion \(u_{\alpha } =\sum _{\beta }u_{\alpha ,\beta }(q_1,p_1)q^{\beta }\) into (28), and compare the powers like \(q^0=1\). Then we have the equation of \(u_{\alpha ,0}\)

$$\begin{aligned} \left\{ q_1^2p_1,u_{\alpha ,0}\right\} - \mu _{k}\left( \sum _{j\in J'}\delta _{k,j} \right) u_{\alpha ,0} - \frac{q_1^2}{(q_1-1)^2}\left( \sum _{j\in J } \mu _{j}\delta _{k,j} \right) u_{\alpha ,0} = 0. \end{aligned}$$
(29)

If \(k\in J'\), then

$$\begin{aligned} \left\{ q_1^2p_1,u_{\alpha ,0}\right\} - \mu _{k}u_{\alpha ,0} = 0. \end{aligned}$$

Because \(\mu _{k}\ne 0\) by (NRC) condition, we have \(u_{\alpha ,0}=0\).

On the other hand, if \(k\in J \), then

$$\begin{aligned} \left\{ q_1^2p_1,u_{\alpha ,0}\right\} - \frac{q_1^2}{(q_1-1)^2} \mu _{k}u_{\alpha ,0} = 0. \end{aligned}$$

By considering the equation around \(q_1=1\) together with (NRC) condition we obtain \(u_{\alpha ,0}=0\).

Next we consider \(u_{\alpha ,\beta }\)(\(\beta \ne 0\)) (\(\alpha =(\alpha _2,\ldots ,\alpha _n)\), \(\alpha _j=\delta _{j,k}\)).

$$\begin{aligned}&\left\{ q_1^2p_1,u_{\alpha ,\beta }\right\} + \sum _{j\in J'} \mu _{j} (\beta _j-\alpha _j)u_{\alpha ,\beta } \nonumber \\&\quad +\, \frac{q_1^2}{(q_1-1)^2} \sum _{j\in J } \mu _{j} (\beta _j-\alpha _j)u_{\alpha ,\beta } = 0. \end{aligned}$$
(30)

If \(\beta \ne \alpha \), then (NRC) condition yields \(u_{\alpha ,\beta }=0\), by the similar argument as in the above. If \(\beta =\alpha \), then we have \(\{q_1^2p_1,u_{\alpha ,\alpha }\}=0\). Hence, there exists \(\phi _\alpha (t)\) of one variable t such that \(u_{\alpha ,\alpha }= \phi _{\alpha }(q_1^2p_1)\). Therefore we obtain

$$\begin{aligned} u= \sum _{|\alpha |=1} \phi _{\alpha }(q_1^2p_1) q^{\alpha }p^{\alpha }+O(|p|^2). \end{aligned}$$
(31)

Now we consider the equation for \(u_{\alpha }\) when \(|\alpha |=2\). We substitute (19) and (31) into the equation \(\chi _{H+H_1}u=0\) and compare the powers like \(p^\alpha \) \((|\alpha |=2)\). In order to get the expressions of the powers like \(p^\alpha \), we note that the following terms appear from \(\chi _{H}u\):

$$\begin{aligned}&\left\{ q_1^2p_1,u_{\alpha }\right\} + \sum _{j\in J'} \mu _{j}\left( q_j\frac{\partial }{\partial q_j}-\alpha _j \right) u_{\alpha } + \frac{q_1^2}{(q_1-1)^2}\sum _{j\in J } \mu _{j}\left( q_j\frac{\partial }{\partial q_j}-\alpha _j \right) u_{\alpha } \nonumber \\&\quad +\, \frac{2q_1}{(q_1-1)^3}\sum _{j\in J } \mu _{j}q^{\alpha } \frac{\partial }{\partial p_1}\phi _{\alpha -e_j}. \end{aligned}$$
(32)

On the other hand, the following terms appear from \(\chi _{H_1}u\).

$$\begin{aligned}&\sum _{\nu }\frac{\partial }{\partial p_\nu } \left( \sum _{j}p_j^2B_j(q_1,p)\right) \frac{\partial }{\partial q_\nu } (\phi _{e_\nu } q_\nu p_\nu )\nonumber \\&\quad - \frac{\partial }{\partial q_1} \left( \sum _{j}p_j^2B_j(q_1,p)\right) \frac{\partial }{\partial p_1} (\sum _{|\alpha |=1} \phi _\alpha q^\alpha p^\alpha ). \end{aligned}$$
(33)

Note that the second term in (33) is \(O(|p|^3)\). Hence it does not appear in the recurrence formula because \(|\alpha |=2\). Moreover, since we consider terms of \(O(|p|^2)\), the first term yields

$$\begin{aligned} 2 \sum _\nu \phi _{e_\nu } B_\nu (q_1,0) \delta _{\alpha ,2e_\nu }. \end{aligned}$$
(34)

Therefore, by comparing the powers like \(p^\alpha \) in \(\chi _{H+H_1}u=0\) we have

$$\begin{aligned}&\left\{ q_1^2p_1,u_{\alpha }\right\} + \sum _{j\in J'} \mu _{j}\left( q_j\frac{\partial }{\partial q_j}-\alpha _j \right) u_{\alpha } \nonumber \\&\quad + \frac{q_1^2}{(q_1-1)^2}\sum _{j\in J } \mu _{j}\left( q_j\frac{\partial }{\partial q_j}-\alpha _j \right) u_{\alpha } \nonumber \\&\quad + \frac{2q_1}{(q_1-1)^3} q^{\alpha } \sum _{j\in J } \mu _{j} \frac{\partial }{\partial p_1}\phi _{\alpha -e_j} + 2 \sum _\nu \phi _{e_\nu } B_\nu (q_1,0) \delta _{\alpha ,2e_\nu } =0. \end{aligned}$$
(35)

Expand \(u_{\alpha }\) with respect to q, \(u_{\alpha } =\sum _{\beta } u_{\alpha ,\beta }(q_1,p_1)q^{\beta }\) and insert the expansion into (35). By comparing the power of \(q^{\beta }\) we obtain the recurrence relation for \(u_{\alpha ,\beta }(q_1,p_1)\). We consider 4 cases:

  1. (i)

    \(\alpha \ne 2e_\nu \) for every \(\nu \) and \(\beta \ne \alpha \).

  2. (ii)

    \(\alpha =2e_k\) for some k and \(\beta \ne \alpha ,0\).

  3. (iii)

    \(\alpha =2e_k\) for some k and \(\beta =0\).

  4. (iv)

    \(\beta =\alpha \).

Case (i): We note that the fourth and the fifth terms of the left-hand side of (35) yield no term in the recurrence relation for \(u_{\alpha ,\beta }\). Indeed, the fourth term is a monomial of \(q^\alpha \). Hence, \(u_{\alpha ,\beta }\) satisfies

$$\begin{aligned} \left\{ q_1^2p_1,u_{\alpha ,\beta }\right\} + \sum _{j\in J'} \mu _{j}(\beta _j-\alpha _j ) u_{\alpha ,\beta } +\frac{q_1^2}{(q_1-1)^2}\sum _{j\in J } \mu _{j}( \beta _j-\alpha _j ) u_{\alpha ,\beta } = 0. \end{aligned}$$
(36)

By virtue of (NRC) and \(\beta \ne \alpha \), either \(\sum _{j\in J'}\mu _j (\beta _j-\alpha _j )\ne 0\) or

\(\sum _{j\in J } \mu _j ( \beta _j-\alpha _j )\ne 0\) holds. One can easily show that \(u_{\alpha ,\beta } = 0\) by the holomorphy of \(u_{\alpha ,\beta }\).

Case (ii): Because the fourth and fifth terms of the left-hand side of (35) do not yield terms by the assumption \(\beta \ne \alpha ,0\), we see that \(u_{\alpha ,\beta }\) satisfies (36). Therefore, we have \(u_{\alpha ,\beta }=0\).

Case (iii): Let \(k\in J'\). Because the fourth term of the left-hand side of (35) is a monomial \(q^\alpha \), \(u_{\alpha ,0}\) satisfies

$$\begin{aligned} \left\{ q_1^2p_1,u_{\alpha ,0}\right\} -2\mu _{k}u_{\alpha ,0} + 2\phi _{e_k}(q_1^2p_1)B_k(q_1,0) = 0. \end{aligned}$$
(37)

Expand \(u_{\alpha ,0}(q_1,p_1) =\sum _{\nu }u_{\alpha ,0,\nu }(q_1)p_1^{\nu }\) and compare the constant terms in \(p_1\) of both sides of (37). Then we have

$$\begin{aligned} q_1^2\frac{d}{dq_1}u_{\alpha ,0,0} - 2\mu _{k}u_{\alpha ,0,0} + 2\phi _{e_k}(0)B_k(q_1,0) = 0. \end{aligned}$$
(38)

If \(\phi _{e_k}(0)\ne 0\), then \(v:= u_{\alpha ,0,0}/(-2\phi _{e_k}(0))\) satisfies

$$\begin{aligned} q_1^2\frac{d}{dq_1}v - 2\mu _{k}v = B_k(q_1,0), \end{aligned}$$

which contradicts (TC). Hence, \(\phi _{e_k}(0)=0\) and (38) reduces to

$$\begin{aligned} q_1^2\frac{d}{dq_1}u_{\alpha ,0,0} - 2\mu _{k}u_{\alpha ,0,0} = 0. \end{aligned}$$

(NRC) condition implies \(2\mu _{k}\ne 0\), and the holomorphcity of \(u_{\alpha ,0,0}\) at \(q_1=0\) tells us \(u_{\alpha ,0,0}=0\).

Next, \(u_{\alpha ,0,1}\) satisfies

$$\begin{aligned} \left( q_1^2\frac{d}{dq_1}-2q_1\right) u_{\alpha ,0,1} - 2\mu _{k}u_{\alpha ,0,1} + 2B_k(q_1,0)\phi _{e_k}'(0)q_1^2 =0. \end{aligned}$$
(39)

Since \(u_{\alpha ,0,1}(q_1)=O(q_1^2)\), we put \(u_{\alpha ,0,1}(q_1)= q_1^2\tilde{u}_{\alpha ,0,1}(q_1)\) with \(\tilde{u}:=\tilde{u}_{\alpha ,0,1}(q_1)\) satisfying

$$\begin{aligned} q_1^2\frac{d}{dq_1} \tilde{u} -2\mu _{k}\tilde{u} =-2B_k(q_1,0)\phi _{e_k}'(0). \end{aligned}$$

If \(\phi _{e_k}'(0)\ne 0\), then, by putting \(v=\tilde{u}/(-2\phi _{e_k}'(0))\), we have a contradiction to (TC). Therefore, \(\phi _{e_k}'(0)=0\) and \(\tilde{u}=0\).

Similarly we can show \(u_{\alpha ,0,\nu }=0\) and \(\phi _{e_k}^{(\nu )}(0)=0\) for \(\nu \in \mathbb {N}\cup \{0\}\), which implies \(u_{\alpha ,0}=0\) and \(\phi _{e_k}=0\) for every \(k\in J'\).

Let \(k\in J\). Then \(u_{\alpha ,0}\) satisfies

$$\begin{aligned} \{q_1^2p_1,u_{\alpha ,0}\} -2\mu _{k}\frac{q_1^2}{(q_1-1)^2}u_{\alpha ,0} +2\phi _{e_k}(q_1^2p_1)B_k(q_1,0) =0. \end{aligned}$$

Expand \(u_{\alpha ,0}(q_1,p_1)=\sum _{\nu }u_{\alpha ,0,\nu }(q_1)p_1^{\nu }\). Then \(u_{\alpha ,0,0}\) satisfies

$$\begin{aligned} q_1^2\frac{d}{dq_1}u_{\alpha ,0,0} - 2\mu _{k}\frac{q_1^2}{(q_1-1)^2}u_{\alpha ,0,0} + 2\phi _{e_k}(0)B_k(q_1,0) =0. \end{aligned}$$
(40)

If \(\phi _{e_k}(0)\ne 0\), then, by (40) we have \(B_k(0,0) =0\). On the other hand, \(v:= u_{\alpha ,0,0}/(-2\phi _{e_k}(0))\) satisfies

$$\begin{aligned} q_1^2\frac{d}{dq_1}v - 2\mu _{k}\frac{q_1^2}{(q_1-1)^2}v = B_k(q_1,0), \end{aligned}$$

which contradicts (TC). So, \(\phi _{e_k}(0)=0\) and (40) reduces to

$$\begin{aligned} (q_1-1)^2\frac{d}{dq_1}u_{\alpha ,0,0} - 2\mu _{k}u_{\alpha ,0,0} =0. \end{aligned}$$

Again we have \(u_{\alpha ,0,0}=0\).

Next, consider the equation of \(u_{\alpha ,0,1}\)

$$\begin{aligned} \left( q_1^2\frac{d}{dq_1}-2q_1\right) u_{\alpha ,0,1} - 2\mu _{k} \frac{q_1^2}{(q_1-1)^2}u_{\alpha ,0,1} = -2\phi _{e_k}'(0)q_1^2B_k(q_1,0). \end{aligned}$$
(41)

Observing \(u_{\alpha ,0,1}(0)=0\), we put \(u_{\alpha ,0,1}(q_1)=cq_1+q_1^2 v\). Substituting it into (41), we have \(c=-2\phi _{e_k}'(0)B_k(0,0)\) and v satisfies

$$\begin{aligned}&-2\phi _{e_k}'(0)\left\{ B_k(q_1,0) + B_k(0,0) + 2B_k(0,0)\mu _{k} \frac{q_1}{(q_1-1)^2} \right\} \\&\quad =\left( q_1^2\frac{d}{dq_1} -2\mu _{k} \frac{q_1}{(q_1-1)^2} \right) v. \end{aligned}$$

By use of (TC), we obtain \(\phi _{e_k}'(0)=0\) and \(u_{\alpha ,0,1}=0\).

In general, \(u_{\alpha ,0,\nu }\) (\(\nu \ge 2\)) satisfies

$$\begin{aligned} \left( q_1^2\frac{d}{dq_1}-2\nu q_1\right) u_{\alpha ,0,\nu } - 2\mu _{k} \frac{q_1^2}{(q_1-1)^2}u_{\alpha ,0,\nu } = -2\frac{\phi _{e_k}^{(\nu )}(0)}{\nu !}q_1^{2\nu } B_k(q_1,0). \end{aligned}$$
(42)

Since we easily see \(u_{\alpha ,0,\nu }=O(q^{2\nu -1})\), we put \(u_{\alpha ,0,\nu }=cq_1^{2\nu -1}+q_1^{2\nu }w\). Then we have \(c=-2\phi _{e_k}^{(\nu )}(0)B_k(0,0)/\nu !\) and w satisfies

$$\begin{aligned}&-\frac{2\phi _{e_k}'(0)}{\nu !}\left\{ B_k(q_1,0) + B_k(0,0) + 2B_k(0,0)\mu _{k} \frac{q_1}{(q_1-1)^2} \right\} \\&\quad = \left( q_1^2\frac{d}{dq_1} -2\mu _{k} \frac{q_1^2}{(q_1-1)^2} \right) w. \end{aligned}$$

By virtue of (TC), we obtain \(\phi _{e_k}^{(\nu )}(0)=0\) and \(w=0\). Therefore, \(u_{\alpha ,0,\nu }=0\) for all \(\nu \in \mathbb {N}\cup \{0\}\). Because of analyticity, we have \(u_{\alpha ,0}=0\) and \(\phi _{e_k}=0\) for every \(k\in J\). Consequently, \(\phi _{e_k}=0\) holds for all \(k\in J'\cup J\).

Case (iv): Because \(\phi _{e_k}=0\) for every k by what we have proved in the above, the fourth and fifth terms of the left-hand side of (35) do not yield terms in the recurrence relation. Hence, \(u_{\alpha ,\alpha }\) satisfies \(\{q_1^2p_1,u_{\alpha ,\alpha }\}=0\). It follows that there exists a function of one variable \(\phi _{\alpha }(t)\) such that \(u_{\alpha ,\alpha }=\phi _{\alpha }(q_1^2p_1)\).

Therefore we have proved

$$\begin{aligned} u= \sum _{|\alpha |=2} \phi _{\alpha }(q_1^2p_1)\ q^{\alpha }p^{\alpha } + O(|p|^3). \end{aligned}$$

Finally we shall prove

Lemma 2

Suppose

$$\begin{aligned} u= \sum _{|\alpha |=\nu } \phi _{\alpha }(q_1^2p_1) q^{\alpha }p^{\alpha } + O(|p|^{\nu +1}) \end{aligned}$$
(43)

for some \(\nu \ge 1\). Then we have

  1. (i)

    \(\phi _{\alpha }=0\) for all \(\alpha \) satisfying \(|\alpha |=\nu \).

  2. (ii)

    For every \(\alpha \) satisfying \(|\alpha |=\nu +1\), there exists a holomorphic function \(\phi _{\alpha }\) of one variable such that

    $$\begin{aligned} u= \sum _{|\alpha |=\nu +1} \phi _{\alpha }(q_1^2p_1) q^{\alpha }p^{\alpha } + O(|p|^{\nu +2}). \end{aligned}$$
    (44)

We have already proved (43) for \(\nu =1,2\). Note that the lemma ends the proof of Theorem 1 because we have \(u=0\) as an analytic function of q and p.

Proof of Lemma 2

By comparing the coefficients of \(p^\alpha \) in \(\chi _{H+H_1}u=0\) we have

$$\begin{aligned}&\left\{ q_1^2p_1,u_{\alpha }\right\} + \sum _{J'}\mu _{j} \left( q_j\frac{\partial }{\partial q_j} -\alpha _j \right) u_{\alpha } \nonumber \\&\quad + \frac{q_1^2}{(q_1-1)^2} \sum _{J }\mu _{j} \left( q_j\frac{\partial }{\partial q_j} -\alpha _j \right) u_{\alpha } \nonumber \\&\quad + \frac{2q_1}{(q_1-1)^3} \left( \sum _{J }\mu _{j}q_jp_j \right) \frac{\partial }{\partial p_1} u_{\gamma } + \sum _{j,\gamma } \frac{\partial H_1}{\partial p_j} \frac{\partial }{\partial q_j} u_{\gamma } = 0, \end{aligned}$$
(45)

where \(|\gamma |<|\alpha |\) and \(\alpha =\gamma +e_j\).

Let \(|\alpha |=\nu +1\). Substituting the expansion \(u_{\alpha }=\sum _{\beta } u_{\alpha ,\beta }(q_1,p_1)q^{\beta }\) into (45) and by using (43), we obtain the relation for \(u_{\alpha ,\beta }\)

$$\begin{aligned}&\left\{ q_1^2p_1,u_{\alpha ,\beta }\right\} + \sum _{J'}\mu _{j} ( \beta _j-\alpha _j )u_{\alpha ,\beta } + \frac{q_1^2}{(q_1-1)^2} \sum _{J }\mu _{j} ( \beta _j-\alpha _j )u_{\alpha ,\beta }\nonumber \\&\quad + 2\frac{q_1}{(q_1-1)^3} \sum _{J }\mu _{j} \frac{\partial }{\partial p_1} \phi _{\alpha -e_j}(q_1^2p_1)\delta _{\alpha ,\beta } \nonumber \\&\quad + 2\sum _{j\in J'\cup J} \delta _{\alpha -2e_j,\beta } B_j(q_1,0) \phi _{\alpha -e_j} (\alpha _j-1) = 0. \end{aligned}$$
(46)

Indeed, because it is easy to show the expressions up to the fourth term in the left-hand side of (46), we consider the fifth term, which corresponds to the fifth term in the left-hand side of (45). In view of (43) we may consider \(2\sum _{j} p_j B_j(q_1,0)\) in \(\frac{\partial H_1}{\partial p_j}\) because other terms have no effect to (45). Hence we may consider terms containing \(p^{\alpha -e_j}\) in \(\frac{\partial }{\partial q_j} u_{\gamma }\). By (43) the coefficient of the term containing \(p^{\alpha -e_j}\) is \((\alpha _j-1)q^{\alpha -2e_j}B_j(q_1,0)\phi _{\alpha -e_j}\). Hence we have the desired expression.

Set \(B':=\sum _{\in J'}\mu _{j} (\beta _j-\alpha _j)\) and \(B :=\sum _{j\in J }\mu _{j} (\beta _j-\alpha _j)\). We consider 4 cases.

Case (1) The case where \(\alpha -2e_j\ne \beta \) for \(j=2,\ldots ,n\) and \(B'\ne 0\). Clearly we have \(\beta \ne \alpha \). It follows that the fourth and the fifth terms in the left-hand side of (46) vanish. Hence we have \(u_{\alpha ,\beta }=0\) by considering (46) at \(q_1=0\).

Case (2) The case where \(\alpha -2e_j\ne \beta \) for \(j=2,\ldots ,n\), \(\beta \ne \alpha \) and \(B' =0\). By (NRC) we have \(B\ne 0\). Hence the fourth and the fifth terms in the left-hand side of (46) vanish. We have \(u_{\alpha ,\beta }=0\) by considering (46) at \(q_1=1\).

Case (3) The case where \(\alpha -2e_k=\beta \) for some k. Clearly, we have \(\beta \ne \alpha \). Assume \(k\in J\). Then, for every \(j\in J'\) we have \(j\ne k\), and hence \(\alpha _j=\beta _j\), which implies \(B'=0\). Equation (46) is reduced to

$$\begin{aligned} \left\{ q_1^2p_1, u_{\alpha ,\beta } \right\} - 2\mu _{k}\frac{q_1^2}{(q_1-1)^2} u_{\alpha ,\beta } + 2(\alpha _k-1)\phi _{\alpha -e_k}B_k(q_1,0) = 0. \end{aligned}$$

Expand \( u_{\alpha ,\beta } =\sum _{\nu =0}^{\infty } u_{\alpha ,\beta ,\nu }(q_1)p_1^{\nu }\). We will show that \(\phi _{\alpha -e_k}\) vanishes.

Indeed, \(v:= u_{\alpha ,\beta ,0}\) satisfies

$$\begin{aligned} q_1^2\frac{dv}{dq_1} - 2\mu _{k} \frac{q_1^2}{(q_1-1)^2}v = -2(\alpha _k-1)\phi _{\alpha -e_k}(0) B_k(q_1,0). \end{aligned}$$

Note that \(\alpha _k =2+\beta _k \ge 2\). If \(\phi _{\alpha -e_k}(0)\ne 0\), then \(w :=v/(-2(\alpha _k-1)\phi _{\alpha -e_k}(0))\) is a holomorphic solution at \(q_1=0\) of the equation

$$\begin{aligned} q_1^2\frac{dw}{dq_1} - 2\mu _{k} \frac{q_1^2}{(q_1-1)^2}w = B_k(q_1,0). \end{aligned}$$

Because one can verify \(B_k(0,0)=0\), we have a contradiction to (TC). Hence we have \(\phi _{\alpha -e_k}(0)=0\) and \(u_{\alpha ,\beta ,0}=0\).

Next, \(v=u_{\alpha ,\beta ,1}\) satisfies

$$\begin{aligned} q_1^2\frac{dv}{dq_1} - 2\mu _{k} \frac{q_1^2}{(q_1-1)^2}v - 2q_1 v = -2(\alpha _k-1) \phi _{\alpha -e_k}'(0) q_1^{2}B_k(q_1,0). \end{aligned}$$

By comparing the coefficients of \(q_1^2\) of both sides we see that \(v=O(q_1^2)\). Similarly to the above, \(w:= v q_1^{-2}\) leads to a contradiction to (TC). Hence, we have \(\phi _{\alpha -e_k}'(0)=0\) and \(u_{\alpha ,\beta ,1}=0\).

In general, \(v=u_{\alpha ,\beta ,\nu }\)(\(\nu \ge 2\)) satisfies

$$\begin{aligned} q_1^2\frac{dv}{dq_1} - 2\mu _{k} \frac{q_1^2}{(q_1-1)^2}v - 2q_1\nu v = -\frac{2(\alpha _k-1)}{\nu !} \phi _{\alpha -e_k}^{(\nu )}(0) q_1^{2\nu }B_k(q_1,0). \end{aligned}$$

Similarly to the above, we have \(\phi _{\alpha -e_k}^{(\nu )}(0)=0\) and \(u_{\alpha ,\beta ,\nu }=0\). Therefore, \(\phi _{\alpha -e_k}=0\) and \(u_{\alpha ,\beta }=0\) for \(k\in J \).

Let \(k\in J'\). Equation (46) is reduced to

$$\begin{aligned} \left\{ q_1^2p_1,u_{\alpha ,\beta }\right\} -2\mu _{k} u_{\alpha ,\beta } + 2(\alpha _k-1)\phi _{\alpha -e_k}B_k(q_1,0) =0. \end{aligned}$$

The holomorphicity of \(u_{\alpha ,\beta }\) at \(q_1=0\) and (TC) implies \(\phi _{\alpha -e_k}(0)=0\) and \(u_{\alpha ,\beta }=0\) for \(k\in J'\). Therefore, \(\phi _{\alpha }=0\) for \(k\in J'\). Because \(\phi _{\alpha }=0\) for \(k\in J \), we have \(\phi _{\alpha }=0\) for all \(\alpha \) with \(|\alpha |=\nu \).

Case (4) The case \(\beta =\alpha \). We have \(\{q_1^2p_1,u_{\alpha ,\alpha }\}=0\), since we have proved \(\phi _\gamma =0\) for \(|\gamma |=\nu \). Hence, there exists \(\phi _\alpha \) such that \(u_{\alpha ,\alpha }= \phi _{\alpha }(q_1^2p_1)\).

Consequently, we have proved the lemma. \(\square \)

5 Properties of (TC)

We will show that (TC) holds for almost all \(B_k(q_1,0)\). Set \(q_1=t\), \(B_k(t,0)=: a(t)\) and \(c:= \mu _{k}\), and write (8) in the form

$$\begin{aligned} t^2\frac{d}{d t}v - 2c v = a(t). \end{aligned}$$
(47)

Clearly, if a(t) is a constant function, then (TC) does not hold since (47) has a constant solution \(v=-a(0)/(2c)\). We first prove

Proposition 2

Suppose that a(t) is a polynomial of degree \(\ell \ge 1\). Then (47) has an analytic solution at \(t=0\) if and only if (47) has a polynomial solution v of degree \(\ell -1\). The set of a(t) for which (47) has a polynomial solution is contained in the set of codimension one of the set of polynomials of degree \(\ell \).

Remark

For a given polynomial v of degree \(\ell -1\), define a(t) by (47). Clearly the set of a’s such that (47) has a polynomial solution is an infinite set.

Proof of Proposition 2

Let \(a(t)=\sum _{j=0}^\ell a_j t^j\) \((a_\ell \ne 0)\) and let \(v(t)= \sum _{j=0}^\infty v_j t^j\) be the analytic solution of (47). By inserting the expansions into (47) and by comparing the powers of t we obtain

$$\begin{aligned} v_0= -a_0/(2c), \quad v_n=(n-1)v_{n-1}/(2c) - a_n/(2c), \ n=1,2,\ldots \end{aligned}$$
(48)

If \(n>\ell \), then we have \(v_n=(n-1)v_{n-1}/(2c)\). Therefore, if \(v_\ell =0\), then \(v_n=0\) for \(n>\ell \). Hence v is a polynomial. On the other hand, if \(v_\ell \ne 0\), then \(v_n = (2c)^{\ell -n} (n-1)(n-2)\cdots \ell v_\ell \). It follows that v(t) is not analytic in any neighborhood of the origin, which contradicts to the assumption. Hence v is a polynomial of degree \(\ell -1\). The converse statement is trivial.

We will show the latter half. By the recurrence formula (48), one easily sees that \(v_\ell \) is a nontrivial linear function of \(a_0, \ldots , a_\ell \). Hence the condition \(v_\ell =0\) is satisfied for a polynomial a(t) on the set of codimension 1. This completes the proof. \(\square \)

Example

We give an example of \(B_k(q_1,0)\)’s satisfying the condition (TC) in Theorem 1. We use the notation in Proposition 2. If \(k\in J'\), then we look for \(a(t) \equiv B_k(t,0)\) such that \(a(t) = \alpha t + \beta t^2\) for some complex constants \(\alpha \) and \(\beta \). In order to verify that (47) has no solution v being analytic at \(t=0\), we expand \(v(t)= \sum _{j=0}^\infty v_j t^j\) and consider the recurrence relation (48). We assume that \(c= \mu _{k}\ne 0\). Clearly, we have \(v_1= -\alpha /(2c)\) and \(v_1 - 2c v_2 =\beta \). It follows that \(v_2 = -(\alpha /(2c) +\beta )/(2c)\). For \(n \ge 3\), we have \(v_n=(n-1)v_{n-1}/(2c)\), which implies \(v_n = (n-1)! (2c)^{2-n} v_2\). Therefore, if \(v_2\ne 0\), then v does not converge. Hence (47) has no analytic solution. We observe that \(v_2\ne 0\) holds if \(\alpha /(2c) +\beta \ne 0\).

Next we assume \(k\in J\), and we consider (9) in (TC). (9) is rewritten in (53) which follows. We look for b(t) such that \(b(t)= \gamma t^2 + \delta t^3\) for some complex constants \(\gamma \) and \(\delta \). We set \(q_1=t+1\). Since \(b(0)=0\), we have \(a(0)=0\). Hence, by (53) we have the relation

$$\begin{aligned} a(t+1) =a(q_1)= (\gamma + \delta t )(t+1)^2= q_1^2 (\gamma -\delta + \delta q_1). \end{aligned}$$

In order to verify (TC) we argue as in the above. We expand w(t) in the series \(w(t) = w_2 t^2 + w_3 t^3 + \cdots \) and we subsitutute it into (53). By comparing the powers of \(t^2\) of both sides we have \(w_2= -\gamma /(2c)\). Similarly, we have \(w_3= -(\gamma /c + \delta )/(2c)\). If \(\gamma + c\delta \ne 0\), then we have \(w_3\ne 0\) and we see that the formal power series expansion of \(w(t) = w_2 t^2 + w_3 t^3 + \cdots \) diverges. Hence we have the desired property. Consequently, we choose \(B_k(q_1,0)= \alpha q_1 + \beta q_1^2\) with \(\alpha /(2c) +\beta \ne 0\) for \(k\in J'\), and \(B_k(q_1,0)= q_1^2(\gamma -\delta + \delta q_1)\) with \(\gamma + c\delta \ne 0\) for \(k\in J\). Then we see that (TC) is satisfied.

Next we study (TC) when a(t) is an analytic function. By replacing v(t) and a(t) with \(v(t)-v(0)\) and \(a(t)-a(0)\), \((2c v(0)= -a(0))\), respectively, we may assume that \(v(0)=0\) and \(a(0)=0\) in (47). Then we have

Proposition 3

The set of analytic functions a(t)’s at the origin such that (47) has an analytic solution v is contained in the set of codimension 1 of the set of germs of analytic functions at \(t=0\).

Proof

Let v be the analytic solution of (47) at \(t=0\). Set \(v(t)=t\tilde{v}(t)\) and \(a(t)=t\tilde{a}(t)\). Then

$$\begin{aligned} t^2\frac{d}{d t}\tilde{v} + t\tilde{v} -2c \tilde{v} = \tilde{a}(t). \end{aligned}$$
(49)

We make the (formal) Borel transform \({\mathcal B}(\tilde{v})\) to (49)

$$\begin{aligned} {\mathcal {B}}(\tilde{v})(z)\equiv \widehat{\tilde{v}}(z):= \sum _{n=1}^\infty v_n \frac{z^{n-1}}{(n-1)!}. \end{aligned}$$
(50)

Because \(\tilde{v}(t)\) and \(\tilde{a}(t)\) are analytic at \(t=0\), it follows that \({\mathcal B}(\tilde{v})(z)\) and \({\mathcal B}(\tilde{a})(z)\) are entire functions of exponential type of order 1. Recalling that \({\mathcal {B}} \left( (t^2\frac{d}{d t} + t)\tilde{v}\right) (z) =z{\mathcal {B}}(\tilde{v})(z)\) we have

$$\begin{aligned} (z -2c){\mathcal {B}}(\tilde{v}) = {\mathcal {B}}(\tilde{a})(z). \end{aligned}$$
(51)

It follows that

$$\begin{aligned} {\mathcal {B}}(\tilde{a})(2c)=0. \end{aligned}$$
(52)

This shows that the germ \(\{ a_n \}_{n=1}^\infty \) of a(t) at \(t=0\) is contained in the hyperplane. This ends the proof. \(\square \)

Next we consider (9) in (TC). We set \(t=q_1-1\), \(a(t+1):= B_k(t+1,0)\), \(c=\mu _{k}\) and \(a(0)=B_k(0,0)\). Then (9) can be written in

$$\begin{aligned} \left( t^2\frac{d}{dt} - 2c \right) w = \frac{t^2}{(t+1)^2} a(t+1) + \frac{a(0)}{(t+1)^2} (t^2+c(t+1))=:b(t). \end{aligned}$$
(53)

This equation has the same form as (47). We determine w(0) by \(-2c w(0)=b(0)\). If we make the appropriate change of unknown functions w and b as before, one may assume that \(w(0)=0\) and \(b(0)=0\). In view of the definition of b(t) we have \(c a(0)=0\). Hence we have \(a(0)=0\). It follows that \(b(t)=t^2 a(t+1)/(t+1)^2\). In the following we assume \(w(0)=0\) and \(a(0)=0\). Then we have

Proposition 4

Suppose that a(t) is holomorphic in a connected domain containing \(t=0\) and \(t=1\). Then the set of a(t) for which (53) has an analytic solution is contained in the set of codimension one of the set of germs of analytic functions at \(t=0\).

Proof

Let w(t) be an analytic solution of (53) at \(t=0\). We set \(\alpha :=a'(0)\) and \(a(z) = \alpha z +A(z)z^2\) for some analytic function A(z). Then, by the general formula w is given by

$$\begin{aligned} w= \exp \left( - \frac{2c}{t} \right) \left( K + \int _\tau ^t \exp \left( \frac{2c}{s}\right) \left( \frac{\alpha }{s+1} +A(s+1)\right) ds \right) , \end{aligned}$$
(54)

where K and \(\tau \ne 0\) are some constants. We take a smooth curve \(\gamma \) which connects \(\tau \) and the origin such that it stays in the half space, \(\mathfrak {R}\, (c/t)<0\) near the origin. Then the limit

$$\begin{aligned}&\int _\tau ^0 \exp \left( \frac{2c}{s}\right) \left( \frac{\alpha }{s+1} +A(s+1)\right) ds \qquad \qquad \nonumber \\&\quad := \lim _{t\in \gamma , t\rightarrow 0} \int _\tau ^t \exp \left( \frac{2c}{s}\right) \left( \frac{\alpha }{s+1} +A(s+1)\right) ds \end{aligned}$$
(55)

exists and it is a non-constant analytic function of \(\tau \). If the condition

$$\begin{aligned} K + \int _\tau ^0 \exp \left( \frac{2c}{s}\right) \left( \frac{\alpha }{s+1} +A(s+1)\right) ds \ne 0 \end{aligned}$$
(56)

holds, then, by taking the limit \(t\rightarrow 0\), \(\mathfrak {R}\, (c/t)<0\) in (54) we see that w(t) tends to infinity, which contradicts to the analyticity of w at the origin. Hence we have

$$\begin{aligned} K= \int ^\tau _0 \exp \left( \frac{2c}{s}\right) \left( \frac{\alpha }{s+1} +A(s+1)\right) ds. \end{aligned}$$
(57)

By substituting (57) to (54) we have

$$\begin{aligned} w(t) = \exp \left( - \frac{2c}{t} \right) \int _0^t \left( \frac{2c}{s}\right) \left( \frac{\alpha }{s+1} +A(s+1)\right) ds. \end{aligned}$$
(58)

We take t sufficiently close to the origin such that the Taylor expansion \(A(s+1)= \sum _{n=0}^\infty a_n s^n\) converges for \(|s|\le |t|\). Because \(w(te^{2\pi i})=w(t)\) holds by the analyticity of w, it follows that

$$\begin{aligned} \int _t^{te^{2\pi i}} \exp \left( \frac{2c}{s}\right) \left( \frac{\alpha }{s+1} +A(s+1)\right) ds =0. \end{aligned}$$
(59)

By calculating the residue we have \(\int _t^{te^{2\pi i}} \exp \left( \frac{2c}{s}\right) \frac{\alpha }{s+1}ds = 2\pi i \alpha (1-e^{-2c})\). The non-resonance condition implies \(c=\mu _{k}\ne 0\), and hence \(1-e^{-2c}\ne 0\). Hence, by (59) the germ of \(A(z)/\alpha \) at \(z=1\) (in case \(\alpha \ne 0\)) or that of A(z) at \(z=1\)( in case \(\alpha =0\)) is contained in some hyperplane of the set of germs of analytic functions.

We recall that A(z) is analytic in some domain containing \(z=0\) and \(z=1\). We will show that by the analytic continuation from \(z=1\) to \(z=0\) the germ of A(z) at \(z=1\) is transformed to that of A(z) at \(z=0\) by an infinite matrix. If we can prove this, then the germ of A(z) or \(A(z)/\alpha \) at \(z=0\) is contained in some hyperplane. In view of \(a(z) = \alpha z +A(z)z^2\), the germ of a(z) at \(z=0\) is contained in some hyperplane.

We take a rectifiable curve which connects \(z=1\) and \(z=0\). First we consider the analytic continuation from \(z=1\) to \(z=z_0\), where \(z_0\) is contained in the disk centered at \(z=1\) in which A(z) is analytic. Let \(A(z) = \sum _{n=0}^\infty a_n (z-1)^n\) be the expansion at \(z=1\). Then the Taylor expansion of A(z) at \(z=z_0\) is given by

$$\begin{aligned} \sum _{k=0}^\infty \frac{(z-z_0)^k}{k!}\sum _{n=k}^\infty a_n (z_0-1)^{n-k}\frac{n!}{(n-k)!}. \end{aligned}$$
(60)

It follows that the germ at \(z=z_0\) is given by

$$\begin{aligned} \left( \sum _{n=k}^\infty a_n {n\atopwithdelims ()k} (z_0-1)^{n-k} \right) _{k=0}^\infty . \end{aligned}$$
(61)

Hence the germ at \(z=1\) is transformed to the one in (60) by the infinite matrix

$$\begin{aligned} {\mathcal A}:= \left( (z_0-1)^{n-k} {n\atopwithdelims ()k} \right) _{k\downarrow 0,1,\ldots ; n\rightarrow 0,1,\ldots }, \end{aligned}$$
(62)

where we set the (kn)-component \((k>n)\) to be zero. Note that if \(|z_0-1|\) is sufficiently small, then \({\mathcal A}\) defines a continuous linear operator on the space of sequences with an appropriate norm. Therefore, if the germ of A(z) at \(z=1\) is contained in the hyperplane, then the germ of A(z) at \(z=z_0\) is contained in some hyperplane. By finite times of analytic continuation we see that the germ of A(z) at \(z=0\) is contained in some hyperplane. This completes the proof. \(\square \)