1 Introduction and Main Results

Let \(\Sigma \) be a smooth, compact surface with boundary \(\partial \Sigma \). For a Riemannian metric g on \(\Sigma \) let \(G^g:\Sigma \times \Sigma \rightarrow \mathbb {R}\) be the Green function for the Laplace-Beltrami operator \(-\Delta _g\) with Neumann boundary conditions and mean value 0; i.e. for each \(\xi \in \Sigma \) the function \(G^g(\cdot ,\xi )\) is the unique solution of

$$\begin{aligned}\left\{ \begin{aligned} -\Delta _g G^g(\cdot ,\xi )&= \delta _\xi -\frac{1}{|\Sigma |_g}{} & {} \qquad \text {in }{\mathring{\Sigma }}:= \Sigma \setminus \partial \Sigma ,\\ \partial _{\nu _g} G^g(\cdot ,\xi )&= 0{} & {} \qquad \text {on }\partial \Sigma ,\\ \int _\Sigma G^g(\cdot ,\xi )\,dv_g&= 0. \end{aligned} \right. \end{aligned}$$

Here \(\nu _g\) is the unit outward normal on \(\partial \Sigma \), \(dv_g\) is the area element of \(\Sigma \), \(|\Sigma |_g=\int _{\Sigma } dv_g\) and \(\delta _{\xi }\) is the Dirac distribution on \(\Sigma \) concentrated at \(\xi \in \Sigma \). Setting \(\kappa (\xi )=2\pi \) for \(\xi \in {\mathring{\Sigma }}\) and \(\kappa (\xi )=\pi \) for \(\xi \in \partial \Sigma \), the Robin function \(R^g:\Sigma \rightarrow \mathbb {R}\) is defined by

$$\begin{aligned} R^g(\xi ):= \lim _{x\rightarrow \xi }\left( G^g(x,\xi ) + \frac{1}{\kappa (\xi )}\log d_g(x,\xi )\right) , \end{aligned}$$

where \(d_g\) denotes the distance with respect to the metric g. Given integers \(m\ge l\ge 0\) with \(m\ge 1\) and real numbers \(\sigma _i\ne 0\), \(i=1,\dots ,m\), we set

$$\begin{aligned} X:= {\mathring{\Sigma }}^l\times (\partial \Sigma )^{m-l}\qquad \text {and}\qquad \Delta _X:= \{x\in X:\hbox {} x_i=x_j\; \hbox {for some}\; i\ne j\}, \end{aligned}$$

and define, for an arbitrary given function \(h\in {{\mathcal {C}}}^2(\Sigma ^m, \mathbb {R})\):

(1.1)

We will prove that \(f_g\) is a Morse function for a “generic” metric g on \(\Sigma \), in a sense that will be made precise below.

Functions of the form (1.1) play a crucial role in understanding the blow-up behavior of solutions of mean field equations like

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta _g u= \lambda \left( \frac{V e^u}{\int _{\Sigma } Ve^u\, dv_g} - \frac{1}{|\Sigma |_g}\right) &{} \text { in } {\mathring{\Sigma }}, \\ \partial _{\nu _g} u = 0 &{} \text { on } \partial \Sigma , \\ \int _{\Sigma } u\, dv_g = 0. &{} \end{array}\right. } \end{aligned}$$
(1.2)

For a compact surface without boundary it has been proven in [7] that a nondegenerate or more generally, an isolated stable critical point \(x=(x_1,\dots ,x_m)\) of \(f_g\), with \(l=m\) and \(h(x) = 2 \sum _{i=1}^m\sigma _i\log V(x_i)\), gives rise to solutions \((\lambda ,u)\) of (1.2) with \(\lambda \) close to \(8\pi m\) and such that u blows up as \(\lambda \rightarrow 8\pi m\) precisely at \(x_1,\dots ,x_m\in \Sigma \). Similar results have been obtained in the case of mean field equations on a bounded domain in \(\mathbb {R}^2\) with Dirichlet boundary conditions in [8, 13], and for solutions of Gelfand’s problem and steady states of the Keller–Segel system (see [2, 5, 6, 10] for instance).

Now we state our main results. We fix an integer \(k\ge 0\) and \(0<\alpha <1\), and let \(\textrm{Riem}^{k+2,\alpha }(\Sigma )\) be the space of Riemannian metrics of class \({{\mathcal {C}}}^{k+2,\alpha }\) on \(\Sigma \), i.e. the space of symmetric and positive definite sections \(\Sigma \rightarrow (T\Sigma \otimes T\Sigma )^*\) of class \({{\mathcal {C}}}^{k+2,\alpha }\).

Theorem 1.1

If \(\sigma _i\ne 0\) for \(i=1,\dots ,m\) and \(\sigma _1+\dots +\sigma _m\ne 0\) then for any \(g\in \textrm{Riem}^{k+2,\alpha }(\Sigma )\) the set

$$\begin{aligned} {{\mathcal {M}}}_g^{k+2,\alpha }(\Sigma ):= \{\psi \in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+):{f_{\psi g}\text { is a Morse function }} \} \end{aligned}$$

is a residual subset of \({{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\).

As a consequence of Theorem 1.1 for any Riemannian metric g on \(\Sigma \), there exists a metric \(\widetilde{g}\) which is arbitrarily close to g and conformal to g, and such that \(f_{\widetilde{g}}\) is a Morse function.

Theorem 1.2

If \(\sigma _i>0\) for all \(i=1,\dots ,m\) then the set

$$\begin{aligned} \textrm{Riem}_{Morse}^{k+2,\alpha }(\Sigma ):= \{g\in \textrm{Riem}^{k+2,\alpha }(\Sigma ):f_g \text { is a Morse function} \} \end{aligned}$$

is an open and dense subset of \(\textrm{Riem}^{k+2,\alpha }(\Sigma )\).

Remark 1.3

In [1] the authors considered functions \(f_g\) of the form (1.1) on a surface \(\Sigma \) without boundary and with \(h(x_1,\dots ,x_m)=\sum _{i=1}^m\log V(x_i)\) where \(V\in {{\mathcal {C}}}^2(\Sigma ,\mathbb {R}_+)\). This is motivated by the mean field equation (1.2). They proved that \(f_g\) is a Morse function for V in an open and dense subset of \({{\mathcal {C}}}^2(\Sigma ,\mathbb {R}_+)\). Also related is the paper [3], which deals with functions like \(f_g\) where G is the Green function of the Dirichlet Laplace operator on a bounded smooth domain \(\Omega \subset \mathbb {R}^n\). In [3], it is proved that \(f_g\) is a Morse function for a generic domain \(\Omega \). The present paper seems to be the first investigating the Morse property of \(f_g\) as in (1.1) as a function of the Riemannian metric.

2 Preliminaries

Riemann surfaces are locally conformally flat, so there exist isothermal coordinates where the metric is conformal to the Euclidean metric (see [4, 11, 15]). We modify the isothermal coordinates applied in [7, 9, 16] to a Riemann surface with boundary. For \(\xi \in \Sigma \) there exists a local chart \(y_{\xi }:\Sigma \supset U(\xi )\rightarrow B^\xi \subset \mathbb {R}^2\) transforming g to \(e^{\varphi _\xi \circ y_\xi }\langle \cdot ,\cdot \rangle _{\mathbb {R}^2}\). We may assume that \(y_\xi (\xi )=0\). For \(\xi \in {\mathring{\Sigma }}\) we may also assume that \(\overline{U(\xi )}\subset {\mathring{\Sigma }}\) and that the image of \(y_{\xi }\) is a disc \(B^\xi :=\{ y\in \mathbb {R}^2: |y|< 2r_{\xi }\}\) of radius \(2r_\xi >0\). For \(\xi \in \partial \Sigma \), by Lemma 4 in [16], there exists an isothermal coordinate system \(\left( U(\xi ), y_{\xi }\right) \) near \(\xi \) such that the image of \(y_{\xi }\) is a half disk \({B}_{2r_{\xi }}^{+}:=\{y\in \mathbb {R}^2: |y|<2r_{\xi }, y_2\ge 0\}\) of radius \(2r_{\xi }>0\) and \(y_{\xi }\left( U(\xi )\cap \partial \Sigma \right) = {{B}_{2r_{\xi }}^{+}} \cap \partial \mathbb {R}^{2}_+\). For any \(x \in \) \(y_{\xi }^{-1}\left( {{B}_{2r_{\xi }}^{+}} \cap \partial \mathbb {R}^{2}_+\right) \), we have

$$\begin{aligned} \left( y_{\xi }\right) _*(\nu _g(x))=\left. -\exp \left( -\frac{\varphi _{\xi }(y)}{2}\right) \frac{\partial }{ \partial y_2 }\right| _{ y=y_{\xi }(x)}. \end{aligned}$$
(2.1)

In this case, we take \(B^\xi = {B}_{2r_{\xi }}^{+}\). For \(\xi \in \Sigma \) and \(0<r\le 2r_\xi \) we set

$$\begin{aligned} B_r^\xi := B^\xi \cap \{ y\in \mathbb {R}^2: |y|< r\}\quad \text {and}\quad U_{r}(\xi ):=y_\xi ^{-1}(B_{r}^{\xi }). \end{aligned}$$

Recall that \(\varphi _\xi :B^\xi \rightarrow \mathbb {R}\) is related to K, the Gaussian curvature of \(\Sigma \), by the equation

$$\begin{aligned} -\Delta \varphi _\xi (y) = 2K\big (y^{-1}_\xi (y)\big ) e^{\varphi _\xi (y)} \quad \text {for all }y\in B^\xi . \end{aligned}$$

We can assume that \(y_\xi \) and \(\varphi _\xi \) depend smoothly on \(\xi \), and that \(\varphi _\xi (0)=0\) and \(\nabla \varphi _\xi (0)=0\).

Observe that we have for \(\zeta \in U(\xi )\),

$$\begin{aligned} \lim _{x\rightarrow \zeta }\frac{d_g(x,\zeta )}{\big |\varphi _\xi (x)-\varphi _\xi (\zeta )\big |} = e^{\frac{1}{2}\varphi _\xi (\zeta )}, \end{aligned}$$

which implies

$$\begin{aligned} R^g(\zeta ) = \lim _{x\rightarrow \zeta }\left( G^g(x,\zeta )+\frac{1}{\kappa (\zeta )}\log \big |y_\xi (x)-y_\xi (\zeta )\big |\right) + \frac{1}{2\kappa (\zeta )}\varphi _\xi \big (y_\xi (\zeta )\big ), \end{aligned}$$
(2.2)

and in particular, using \(\varphi _\xi \big (y_\xi (\xi )\big )=\varphi _\xi (0)=0\):

$$\begin{aligned} R^g(\xi ) = \lim _{x\rightarrow \xi }\left( G^g(x,\xi )+\frac{1}{\kappa (\xi )}\log \big |y_\xi (x)\big |\right) . \end{aligned}$$

Now we construct a regular part \(H^g(x,\xi )\) of the Green function \(G^g(x,\xi )\) such that \(H^g(\xi ,\xi )=R^g(\xi )\). Let \(\chi \in {{\mathcal {C}}}^\infty (\mathbb {R},[0,1])\) be such that

$$\begin{aligned} \chi (s) = {\left\{ \begin{array}{ll} 1 &{}\quad \text {if }|s|\le 1,\\ 0 &{}\quad \text {if }|s|\ge 2. \end{array}\right. } \end{aligned}$$

For \(\xi \in {\mathring{\Sigma }}\) we choose \(\delta _\xi = \min \{\frac{1}{2} r_\xi ,\frac{1}{2}{{\,\textrm{dist}\,}}(x,\partial \Sigma )\}\), and for \(\xi \in \partial \Sigma \) we set \(\delta _\xi :=\frac{1}{2} r_\xi \). Next we define the cut-off function \(\chi _\xi \in {{\mathcal {C}}}^\infty (\Sigma ,[0,1])\) by

$$\begin{aligned} \chi _\xi (x):= {\left\{ \begin{array}{ll} \chi \big (|y_\xi (x)|/\delta _\xi \big ) &{}\quad \text {if }x\in U(\xi )\\ 0&{}\quad \text {if }x\in \Sigma \setminus U(\xi ) \end{array}\right. }. \end{aligned}$$

Then, for \(\xi \in \Sigma \) the function \(H^g_{\xi }:=H^g(\cdot ,\xi ):\Sigma \rightarrow \mathbb {R}\) is defined to be the unique solution of the Neumann problem

$$\begin{aligned} \left\{ \begin{aligned}&\Delta _g H^g_{\xi } = \frac{1}{\kappa (\xi )} (\Delta _g\chi _\xi )\log \big |y_\xi \big |+\frac{2}{\kappa (\xi )} \left\langle \nabla ^g\chi _\xi ,\nabla ^g\log |y_\xi |\right\rangle _g+\frac{1}{|\Sigma |_g}&\quad \text {in }{\mathring{\Sigma }},\\&\partial _{ \nu _g} H^g_{\xi } = \frac{1}{\kappa (\xi )}\big (\partial _{\nu _g }\chi _\xi \big ) \log |y_\xi | + \frac{1}{\kappa (\xi )}\chi _\xi \partial _{\nu _g} \log |y_\xi |&\quad \text {on }\partial \Sigma ,\\&\int _{\Sigma } H^g_{\xi } dv_g = \frac{1}{\kappa (\xi )}\int _{\Sigma } \chi _{\xi }\log |y_\xi |\, dv_g. \end{aligned} \right. \end{aligned}$$
(2.3)

Lemma 2.1

For \(g\in \textrm{Riem}^{k+2,\alpha }(\Sigma )\) the function \(H^g\) is of class \({{\mathcal {C}}}^{k+3,\alpha }\) in any compact subsets of \(\Sigma \times {\mathring{\Sigma }}\) and in \(\Sigma \times \partial \Sigma \). Moreover, it satisfies

$$\begin{aligned} G^g(x,\xi ) = -\frac{1}{\kappa (\xi )}\chi _\xi (x)\cdot \log \big |y_\xi (x)\big | + H^g(x,\xi ) \end{aligned}$$

and \(H^g(\xi ,\xi )=R^g(\xi )\). Consequently \(R^g\) is of class \({{\mathcal {C}}}^{k+3,\alpha }\) in any compact subsets of \({\mathring{\Sigma }}\) and in \(\partial \Sigma \).

Proof

First we observe that \(e^{\varphi _{\xi }}\in {{\mathcal {C}}}^{k+2,\alpha }(B^{\xi }, \mathbb {R})\) because \(g\in \textrm{Riem}^{k+2,\alpha }(\Sigma ) \). For \(\xi \in {\mathring{\Sigma }}\), by the choice of \(\delta _{\xi }\) we have that \(-\Delta _g H^g_\xi \) is of class \({{\mathcal {C}}}^{k+2,\alpha }\) in \(\Sigma \) and \(\partial _{\nu _g } H^g(x,\xi ) \equiv 0\) on \(\partial \Sigma \). Now the Schauder estimate for the Neumann problem (see [14, 16], for instance) implies that the solution of (2.3) uniquely exists in \({{\mathcal {C}}}^{k+4,\alpha }(\Sigma ,\mathbb {R})\).

For \(x\in U(\xi )\cap \partial \Sigma \), setting \(y=y_{\xi }(x)\) we have:

$$\begin{aligned} \partial _{ \nu _g } \log |y_{\xi }(x)| {\mathop {=}\limits ^{(2.1)}} - e^{-\frac{1}{2} {\varphi }_{\xi }(y)}\frac{\partial }{\partial y_2} \log |y| = - e^{-\frac{1}{2} {\varphi }_{\xi }(y)}\frac{y_2}{|y|^2}\equiv 0. \end{aligned}$$

Clearly, \(\partial _{\nu _g} \chi (|y_{\xi }|(x))=0\) for \(x \in \partial \Sigma \cap U_{\delta _{\xi }}(\xi )\). It follows that \(\partial _{\nu _g} H^g_\xi \) is of class \({{\mathcal {C}}}^{k+2,\alpha }\) on \(\partial \Sigma \). Moreover \(\Delta _{g}H^g_\xi \) is of class \({{\mathcal {C}}}^{k+2,\alpha }\) in \(\Sigma \). Consequently (2.3) has a unique solution \(H^g_\xi \in {{\mathcal {C}}}^{k+3,\alpha }(\Sigma ,\mathbb {R})\) by the Schauder estimates.

Finally \(H^g_\xi \) is uniformly bounded in \({{\mathcal {C}}}^{k+3, \alpha }\) for \(\xi \) in any compact subsets of \({\mathring{\Sigma }}\) and in \(\partial \Sigma \). Therefore \( H^g(\xi ,\xi )\) is in \({{\mathcal {C}}}^{k+3, \alpha }\) in any compact subsets of \({\mathring{\Sigma }}\) and in \(\partial \Sigma \), and \(H^g(\xi ,\xi )=R^g(\xi )\) by (2.2). \(\square \)

For \(g\in \textrm{Riem}^{k+2,\alpha }(\Sigma )\) we now consider the map

$$\begin{aligned} {{\mathcal {H}}}^g: \Sigma \times \Sigma \times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+) \rightarrow \mathbb {R},\quad {{\mathcal {H}}}^g(x,\xi ,\psi ):= H^{\psi g}(x,\xi ). \end{aligned}$$

Proposition 1

The map \({{\mathcal {H}}}^g\) is \({{\mathcal {C}}}^1\) in \(\Sigma \times {\mathring{\Sigma }}\times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\) and in \(\Sigma \times \partial \Sigma \times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\). Moreover, we have

$$\begin{aligned} ~ D_{\psi } {{\mathcal {H}}}^g(x,\xi , 1)[\theta ] = - \frac{1}{|\Sigma |_g}\int _{\Sigma } (G^{g}(z,x) + G^g(z,\xi )) \theta (z) dv_g(z), \end{aligned}$$
(2.4)

for any \(\theta \in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R})\).

Proof

For \(g\in \textrm{Riem}^{k+2,\alpha }(\Sigma )\) and \(\psi \in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\) we clearly have \(\psi g\in \textrm{Riem}^{k+2,\alpha }(\Sigma )\). By a direct calculation we obtain the following equations for \(H^{\psi g}_\xi - H^g_\xi \):

$$\begin{aligned} \left\{ \begin{aligned} -\Delta _{g} (H^{\psi g}_\xi - H^g_\xi )&= \frac{1}{|\Sigma |_g} - \frac{\psi }{ |\Sigma |_{\psi g}}{} & {} \text {in }{\mathring{\Sigma }},\\ \partial _{ \nu _g} (H^{\psi g}_\xi - H^g_\xi )&= 0{} & {} \text {on }\partial \Sigma ,\\ \int _{\Sigma } (H^{\psi g}_\xi -H^{{g}}_\xi )\, dv_g&= -\int _{\Sigma } G^{\psi g}_\xi (\psi -1) dv_g. \end{aligned} \right. \end{aligned}$$

An expansion of \(|\Sigma |_{\psi g}= \int _{\Sigma } dv_{\psi g}\) yields:

$$\begin{aligned} -\Delta _g(H^{\psi g}_\xi - H^g_\xi ) = \frac{1-\psi }{|\Sigma |_g}+ \frac{\int _{\Sigma }(\psi -1) dv_g}{|\Sigma |^2_{g}} + {\mathcal {O}}(\Vert \psi -1\Vert _{{{\mathcal {C}}}^{k+2,\alpha }}^2) \quad \text {as } \Vert \psi -1\Vert _{{{\mathcal {C}}}^{k+2,\alpha }} \rightarrow 0. \end{aligned}$$

Recall that \(G^{\psi g}_\xi = - \frac{4}{\sigma (\xi )}\chi (|y_{\xi }|)\log {|y_{\xi }|}+H^{\psi g}_\xi \), so the representation formula gives:

$$\begin{aligned} H^{\psi g}_\xi (x) - H^g_\xi (x)= & {} \frac{1}{|\Sigma |_g} \int _{\Sigma } G^{\psi g}(z,\xi )(\psi (z)-1)\,dv_g(z) \nonumber \\ {}{} & {} -\frac{1}{|\Sigma |_{\psi g}} \int _{\Sigma } G^g(z,x)(\psi (z)-1)\,dv_g(z). \end{aligned}$$
(2.5)

By standard elliptic estimates (see [14, 16]) there exists a constant C such that

$$\begin{aligned} \big \Vert H^{\psi g}_\xi - H^g_\xi \big \Vert _{{{\mathcal {C}}}^{k+4,\alpha }} \le C\cdot \Vert \psi -1\Vert _{{{\mathcal {C}}}^{k+2,\alpha }}, \end{aligned}$$

thus

$$\begin{aligned} H^{\psi g}_\xi \rightarrow H^g_\xi \quad \text { in }{{\mathcal {C}}}^{k+4,\alpha }\text { as }\psi \rightarrow 1 \text { in }{{\mathcal {C}}}^{k+2,\alpha }. \end{aligned}$$
(2.6)

According to the construction of \(H^g(x,\xi )\), the convergence in (2.6) is uniform for \(\xi \) in any compact subset of \({\mathring{\Sigma }}\), and in \(\partial \Sigma \). It follows that \({{\mathcal {H}}}_g(x,\xi ,\cdot )\) is continuous at 1, uniformly for \(x\in \Sigma \) and \(\xi \) in any compact subsets of \({\mathring{\Sigma }}\) or \(\xi \in \partial \Sigma \). Using \( {{\mathcal {H}}}_{g}(x,\xi ,\psi )= {{\mathcal {H}}}_{\psi g} (x,\xi ,1)\) we see that \({{\mathcal {H}}}_g(x,\xi , \cdot )\) is continuous at every \(\psi \in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\).

Next we prove that \({{\mathcal {H}}}^g(x,\xi ,\psi )\) is \({{\mathcal {C}}}^1\) with respect to \(\psi \). We fix \(\theta \in {{\mathcal {C}}}^{2+k,\alpha }(\Sigma , \mathbb {R})\) and consider the metric \((1+t\theta )g\) with t sufficiently small so that \(1+t\theta > 0\). Then

$$\begin{aligned} w_\xi ^t(x):= & {} \frac{1}{t}\big (H^{(1+t\theta )g}(x,\xi )-H^g(x,\xi )\big )\nonumber \\ {}= & {} \frac{1}{t}\big ({{\mathcal {H}}}^g(x,\xi ,1+t\theta )-{{\mathcal {H}}}^g(x,\xi ,1)\big ) \end{aligned}$$
(2.7)

satisfies the following equations as \(t\rightarrow 0\):

$$\begin{aligned} ~ \left\{ \begin{aligned} -\Delta _{g}w_\xi ^t&= \frac{1}{|\Sigma |^2_g}\int _{\Sigma } \theta \, dv_{g} -\frac{\theta }{|\Sigma |_g} + {{\mathcal {O}}}(t){} & {} \quad \text {in }{\mathring{\Sigma }},\\ \partial _{\nu _g} w_t(x,\xi )&= 0{} & {} \quad x\in \partial \Sigma ,\\ \int _{\Sigma } w_\xi ^t\, dv_{g}&=- \int _{\Sigma } G^{(1+t\theta )g}_\xi \cdot \theta \, dv_g \end{aligned} \right. \end{aligned}$$
(2.8)

where \({{\mathcal {O}}}(t)\) is defined with respect to the \({{\mathcal {C}}}^{k+2,\alpha }\)-norm. Applying the standard elliptic estimates, \(w_\xi ^t\) converges as \(t\rightarrow 0\) to some function \(w_\xi ^0(x) = D_\psi {{\mathcal {H}}}^g(x,\xi ,1)[\theta ]\) in \({{\mathcal {C}}}^{k+4,\alpha }(\Sigma , \mathbb {R})\). Moreover, \(w_\xi ^0\) satisfies the equations

$$\begin{aligned} ~ \left\{ \begin{aligned} -\Delta _{g}w_\xi ^0&= \frac{1}{|\Sigma |^2_g}\int _{\Sigma } \theta \, dv_{g} -\frac{\theta }{|\Sigma |_g}{} & {} \quad \text { in }{\mathring{\Sigma }}, \\ \partial _{\nu _g} w_\xi ^0&= 0{} & {} \quad \text { on }\partial \Sigma ,\\ \int _{\Sigma } w_\xi ^0\, dv_{g}&= -\int _{\Sigma } G^{g}_\xi \cdot \theta \, dv_g. \end{aligned} \right. \end{aligned}$$
(2.9)

The representation formula now implies (2.4):

$$\begin{aligned} D_\psi {{\mathcal {H}}}^g(x,\xi ,1)[\theta ] = w_\xi ^0(x) = -\frac{1}{|\Sigma |_{g}}\int _{\Sigma } \big (G^{g}(z,x) +G^{g}(z,\xi )\big ) \cdot \theta (z)\, dv_g(z). \end{aligned}$$

Replacing g by \(\psi g\) and \(\theta \) by \(\frac{\theta }{\psi }\), we obtain the derivative of \({{\mathcal {H}}}_g(x,\xi ,\psi )\) for arbitrary \(\psi \in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\):

$$\begin{aligned} D_{\psi }{{\mathcal {H}}}^g(x,\xi ,\psi )[\theta ]= & {} \lim _{t\rightarrow 0}\frac{1}{t}\big (H^{(\psi +t\theta ) g}(x,\xi )-H^{\psi g}(x,\xi )\big ) \nonumber \\= & {} -\frac{1}{|\Sigma |_{\psi g}}\int _{\Sigma } \left( G^{\psi g} (z,x)+G^{\psi g}(z,\xi ) \right) \cdot \frac{\theta (z) }{\psi (z)}dv_{\psi g}(z) \nonumber \\= & {} -\frac{1}{|\Sigma |_{\psi g}}\int _{\Sigma } \left( G^{\psi g} (z,x)+G^{\psi g}(z,\xi ) \right) \cdot \theta (z) dv_{g}(z). \qquad \end{aligned}$$
(2.10)

In order to see that \({{\mathcal {H}}}_g(x,\xi ,\psi )\) is continuously Fréchet differentiable with respect to \(\psi \) it is sufficient to prove that \(D_{\psi }{{\mathcal {H}}}_g(x,\xi ,\psi )\) is continuous in \(\psi \) as a linear operator on \({{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\). For \(\psi _1,\psi _2\in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\) there holds

$$\begin{aligned} \begin{aligned}&\big | D_{\psi }{{\mathcal {H}}}_g(x,\xi ,\psi _1)[\theta ]-D_{\psi }{{\mathcal {H}}}_g(x,\xi ,\psi _2)[\theta ] \big |\\ {}&= \left| \frac{1}{|\Sigma |_{\psi _1 g}}\int _{\Sigma } \left( G^{\psi _1 g} (z,x) +G^{\psi _1 g}(z,\xi ) \right) \theta (z) dv_{g}(z)\right. \\&\quad \left. - \frac{1}{|\Sigma |_{\psi _2 g}}\int _{\Sigma } \left( G^{\psi _2 g} (z,x)+G^{\psi _2 g}(z,\xi ) \right) \theta (z) dv_{g}(z)\right| \\&\le C\cdot \Vert \theta \Vert _{{{\mathcal {C}}}^{k+2,\alpha }} \cdot \Vert \psi _1-\psi _2\Vert _{{{\mathcal {C}}}^{k+2,\alpha }} \end{aligned} \end{aligned}$$

where we applied (2.10); here \(C>0\) is a constant. Therefore \({{\mathcal {H}}}^g\) is \({{\mathcal {C}}}^1\) in \(\Sigma \times {\mathring{\Sigma }}\times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\) and in \(\Sigma \times \partial \Sigma \times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+).\) \(\square \)

3 Proof of Theorem 1.1

The proof is based on the following transversality theorem [12, Theorem 5.4].

Theorem 3.1

  Let \(M,\Psi ,N\) be Banach manifolds of class \({{\mathcal {C}}}^r\) for some \(r\in \mathbb {N}\), let \({{\mathcal {D}}}\subset M\times \Psi \) be open, let \({{\mathcal {F}}}: {{\mathcal {D}}}\rightarrow N\) be a \({{\mathcal {C}}}^r\) map, and fix a point \(z\in N\). Assume for each \((y,\psi )\in {{\mathcal {F}}}^{-1}(z)\) that:

  1. (1)

    \(D_y{{\mathcal {F}}}(y,\psi ): T_y M\rightarrow T_{z} N\) is semi-Fredholm with index \(<r\);

  2. (2)

    \(D{{\mathcal {F}}}(y,\psi ): T_y M \times T_\psi \Psi \rightarrow T_{z} N\) is surjective;

  3. (3)

    \({{\mathcal {F}}}^{-1}(z)\rightarrow \Psi \), \((y,\psi )\mapsto \psi \), is \(\sigma \)-proper.

Then

$$\begin{aligned} \Psi _{reg}:=\{ \psi \in \Psi : z \text { is a regular value of } {{\mathcal {F}}}(\cdot , \psi )\} \end{aligned}$$

is a residual subset of \(\Psi \).

Proof of Theorem 1.1

We set \(\Psi := {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\) and consider the functions \(f_{\psi g}:X \setminus \Delta _X \rightarrow \mathbb {R}\) for \(\psi \in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\) first in a local isothermal chart. Let \(y_{\xi _i}:\Sigma \supset U(\xi _i) \rightarrow B^{\xi _i} \subset \mathbb {R}^2\) be isothermal charts of \(\Sigma \) as in Sect. 2 with \(\xi _1,\dots ,\xi _l\in {\mathring{\Sigma }}\), \(\xi _{l+1},\dots ,\xi _m\in \partial \Sigma \). For simplicity of notation we set \(Y_i:=y_{\xi _i}:U_i:=U(\xi _i) \rightarrow B_i:=B^{\xi _{i}}\subset \mathbb {R}^2\) for \(i=1,\dots ,l\), and \(Y_i:=\pi _1\circ y_{\xi _i}:U_i:=U(\xi _i)\cap \partial \Sigma \rightarrow B_i \subset \mathbb {R}\) for \(i=l+1,\dots ,m\); here \(\pi _1:\mathbb {R}^2\rightarrow \mathbb {R}\) is the projection onto the first component. For \(i=l+1,\dots ,m\) we thus have \(y_{\xi _i}(x) = (Y_i(x),0)\). Then

$$\begin{aligned} Y= & {} Y_1\times \dots \times Y_m: X \supset U:=U_1\times \cdots \times U_m \rightarrow \mathbb {R}^{l+m},\ \ (x_1,\dots ,x_m) \\ {}\mapsto & {} \big (Y_1(x_1),\dots ,Y_m(x_m)\big ), \end{aligned}$$

is a chart of X. Set \(M=N:=\mathbb {R}^{l+m}\) and \(V:= Y(U\cap X{\setminus } \Delta _X) \subset \mathbb {R}^{l+m}=M\) so that \({{\mathcal {D}}}:= V \times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\) is an open subset of \(\mathbb {R}^{l+m} \times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\).

We will apply Theorem 3.1 to prove that \(0\in \mathbb {R}^{l+m}\) is a regular value of

$$\begin{aligned} \nabla \big (f_{\psi g}\circ Y^{-1}\big ): \mathbb {R}^{l+m}\supset V=Y(U\cap X\setminus \Delta _X) \rightarrow \mathbb {R}^{l+m} \end{aligned}$$

for \(\psi \) in a residual subset \(\Psi _U \subset \Psi \). This implies that the restriction of \(f_{\psi g}\) to \(U\cap X\setminus \Delta _X\) is a Morse function for \(\psi \in \Psi _U\). Then Theorem 1.1 follows because X is covered by finitely many neighborhoods U as above and because the intersection of finitely many residual sets is a residual set.

It remains to prove that the map

$$\begin{aligned} {{\mathcal {F}}}_g: \mathbb {R}^{l+m}\times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+) \supset {{\mathcal {D}}}\rightarrow \mathbb {R}^{l+m},\quad (y,\psi ) \mapsto \nabla \big (f_{\psi g}\circ Y^{-1}\big )(y), \end{aligned}$$

satisfies the assumptions of Theorem 3.1 with \(r=1\). Concerning the differentiability it is clear that \({{\mathcal {F}}}_g\) is \({{\mathcal {C}}}^1\) as a function of \(y\in V\). In order to see that \({{\mathcal {F}}}_g\) is also \({{\mathcal {C}}}^1\) in \(\psi \), by Lemma 2.1 it is sufficient to prove that \(\nabla _y \left( H^{\psi g}\left( Y_i^{-1}(\cdot ),Y_j^{-1}(\cdot )\right) \right) \) is \({{\mathcal {C}}}^1\) in \(\psi \). We recall that \(w^t_{\xi }\) is defined by (2.7). Applying the representation formula of \(w^t_{\xi }\) and Lebesgue’s dominated convergence theorem, we have for \((y,\psi )\in {{\mathcal {D}}}\) and \(\theta \in {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R})\):

$$\begin{aligned}{} & {} D_{\psi }\big |_{\psi =1}\nabla _y H^{\psi g}(Y_i^{-1}(y_i), Y_j^{-1}(y_j) )[\theta ]\\ {}{} & {} \quad =\lim _{t\rightarrow 0}\nabla _y \left( \left. w^t_{\xi }(x) \right| _{x= Y^{-1}_i(y_i),\xi =Y^{-1}_{j}(y_j)}\right) \\{} & {} {\mathop {=}\limits ^{(2.5)}} \lim _{t\rightarrow 0}\left( -\frac{1}{|\Sigma |_g}\int _{\Sigma } \nabla _y G_{Y_{j}^{-1}(y_j)}^{(1+t\theta )g}\theta dv_g\right. \\ {}{} & {} \left. \qquad +\int _{\Sigma } \frac{1}{t} \left( \frac{1}{|\Sigma |_g}-\frac{1+t\theta }{|\Sigma |_{(1+t\theta )g}}\right) \nabla _y G^g_{Y^{-1}_i(y_i)}dv_g \right) \\{} & {} \quad = - \frac{1}{|\Sigma |_g}\int _{\Sigma } \nabla _y \big (G^{g}(z,Y_i^{-1}(y_i) ) + G^g(z,Y_j^{-1}(y_j))\big )\cdot \theta (z) \,dv_g(z)\\{} & {} \quad =\nabla _y D_{\psi } {{\mathcal {H}}}^g(Y_i^{-1}(y_i), Y_j^{-1}(y_j), 1) [\theta ], \end{aligned}$$

where we used Proposition 1. Since \( D_{\psi }{{\mathcal {F}}}_g(y,\psi )[\theta ] = D_{\psi }{{\mathcal {F}}}_{\psi g}(y, 1)\left[ \frac{\theta }{\psi }\right] \) we obtain

$$\begin{aligned} D_{\psi }{{\mathcal {F}}}_g(y,\psi )[\theta ] = - \frac{1}{|\Sigma |_{\psi g}}\int _{\Sigma } \nabla _y \big (G^{\psi g }(z,Y_i^{-1}(y_i) ) + G^{\psi g}(z,Y_j^{-1}(y_j))\big )\cdot \theta (z) \,dv_{g}(z). \end{aligned}$$

As in the proof of Proposition 1 we deduce that \( {{\mathcal {F}}}_g(y,\psi )\) is \({{\mathcal {C}}}^1\) on U.

Now we need to check the assumptions (1)-(3) of Theorem 3.1. Obviously \(D_y{{\mathcal {F}}}_g(y,\psi ):\mathbb {R}^{l+m}\rightarrow \mathbb {R}^{l+m}\) is a Fredholm operator of index \(0<1\), hence (1) holds. Also, (3) is easy to prove: For \(j\in \mathbb {N}\) the set

$$\begin{aligned} M_j:= \big \{Y(x)\in \mathbb {R}^{l+m}: x\in U,\ d_g\big (x,\Delta _X \cup \partial U\big ) \ge 2^{-j}\big \} \subset Y(U) \subset \mathbb {R}^{l+m} \end{aligned}$$

is compact as a continuous image of a compact set. Therefore the map

$$\begin{aligned} M_j\times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+) \rightarrow {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+),\quad (y,\psi ) \mapsto \psi , \end{aligned}$$

is proper, hence its restriction to \({{\mathcal {F}}}_g^{-1}(0) \cap \big ( M_j\times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\big )\) is proper. Since \(V = \bigcup _{j=1}^\infty M_j\), so \({{\mathcal {D}}}=\bigcup _{j=1}^\infty \big (M_j \times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\big )\) it follows that the map

$$\begin{aligned} {{\mathcal {F}}}_g^{-1}(0) = \bigcup _{j=1}^\infty \Big ({{\mathcal {F}}}_g^{-1}(0) \cap \big ( M_j\times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+)\big )\Big ) \rightarrow {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}_+), \quad (y,\psi ) \mapsto \psi , \end{aligned}$$

is \(\sigma \)-proper.

Finally we prove the surjectivity of the derivative \(D{{\mathcal {F}}}_g(y,\psi ): \mathbb {R}^{l+m}\times {{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}) \rightarrow \mathbb {R}^{l+m}\) at a point \((y,\psi )\in {{\mathcal {F}}}_g^{-1}(0)\). In fact, we shall prove that \(D_\psi {{\mathcal {F}}}_g(y,\psi ):{{\mathcal {C}}}^{k+2,\alpha }(\Sigma ,\mathbb {R}) \rightarrow \mathbb {R}^{l+m}\) is onto. Since

$$\begin{aligned} D_{\psi } {{\mathcal {F}}}_g(y, \psi )[\theta ] = D_{\psi } {{\mathcal {F}}}_{\psi g}(y, 1)\left[ \theta / \psi \right] \qquad \text {for } \theta \in {{\mathcal {C}}}^{2+k,\alpha }(\Sigma , \mathbb {R}) \end{aligned}$$

it is sufficient to consider the case \(\psi \equiv 1\). We observe for \(\theta \in {{\mathcal {C}}}^{2+k,\alpha }(\Sigma , \mathbb {R})\):

$$\begin{aligned} D_{\psi }\big |_{\psi =1}\nabla _y G^{\psi g}(Y_i^{-1}(y_i), Y_j^{-1}(y_j) )[\theta ] = D_{\psi }\big |_{\psi =1}\nabla _y H^{\psi g}(Y_i^{-1}(y_i), Y_j^{-1}(y_j) )[\theta ]. \end{aligned}$$

Now Proposition 1 yields for \(\theta \in {{\mathcal {C}}}^{2+k,\alpha }(\Sigma , \mathbb {R})\) with \(\textrm{supp}(\theta )\subset \Sigma \setminus \big \{Y_1^{-1}(y_1),\dots ,Y_m^{-1}(y_m)\big \}\):

Consider an element \(u=(u_1,\cdots ,u_m)\in \mathbb {R}^{l+m}\) with \(u_1,\dots ,u_l\in \mathbb {R}^2\), \(u_{l+1},\dots ,u_m\in \mathbb {R}\), that is orthogonal to the range of \(D_\psi {{\mathcal {F}}}_g(y,1)\), i.e. it satisfies for every \(\theta \in {{\mathcal {C}}}^{2+k,\alpha }(\Sigma , \mathbb {R})\) with \(\textrm{supp}(\theta ) \subset \Sigma {\setminus } \big \{Y_1^{-1}(y_1),\dots ,Y_m^{-1}(y_m)\big \}\):

This implies, using \(\sum _{j=1}^m\sigma _j\ne 0\):

$$\begin{aligned} \sum _{i=1}^m \sigma _i\big \langle u_i,\nabla _{y_i} G^g(z, Y_i^{-1}(y_i))\big \rangle = 0 \qquad \text {for } z\in \Sigma \setminus \big \{Y_1^{-1}(y_1),\dots ,Y_m^{-1}(y_m)\big \}.\nonumber \\ \end{aligned}$$
(3.1)

Setting \(\kappa _i=2\pi \) for \(i=1,\dots ,l\) and \(\kappa _i=\pi \) for \(i=l+1,\dots ,m\) we have

$$\begin{aligned} G^g(z,Y_i^{-1}(y_i)) = H^g\big (z,Y_i^{-1}(y_i)\big ) - \frac{1}{\kappa _i} \chi \big (4|Y_i(z)-y_i|/\delta _{\xi _i}\big )\log |Y_i(z)-y_i|. \end{aligned}$$

Now we define \(z_i(t):= Y_i^{-1}(y_i+tu_i)\) for \(i\in \{1,\dots ,m\}\) and observe that

$$\begin{aligned} \nabla _{y_j}G^g\big (z_i(t),Y_j^{-1}(y_j)\big ) = {{\mathcal {O}}}(1)\qquad \text {as } t\rightarrow 0 \text { for }j\ne i \end{aligned}$$

whereas

$$\begin{aligned} \big \langle u_i,\nabla _{y_i} G^g(z_i(t), Y_i^{-1}(y_i))\big \rangle \rightarrow \infty \qquad \text {as } t\rightarrow 0. \end{aligned}$$

Equation (3.1) implies \(u_i=0\) because \(\sigma _i \ne 0\). This holds for all i, hence \(u=0\) and \(D_\psi {{\mathcal {F}}}_g(y,\psi )\) must be onto.

4 Proof of Theorem 1.2

Theorem 1.2 follows easily from the following lemma.

Lemma 4.1

  If \(\sigma _i>0\) for \(i=1,\dots ,m\) then \(\displaystyle \lim _{x\rightarrow \partial (X\setminus \Delta _X)}|\nabla ^g f_g(x)|_g=\infty \).

Proof of Theorem 1.2

Lemma 4.1 implies that the set of critical points of \(f_g\) is compact. It follows that if \(f_g\) is a Morse function, then it has only finitely many critical points, and any \({{\mathcal {C}}}^2\)-perturbation of \(f_g\) is also a Morse function. Therefore \(\textrm{Riem}_{Morse}^{k+2,\alpha }(\Sigma )\) is an open subset of \(\textrm{Riem}^{k+2,\alpha }(\Sigma )\). By Theorem 1.1 it is also a dense subset.

Proof of Lemma 4.1

Consider a sequence \(x^n \in X\setminus \Delta _X\) such that

$$\begin{aligned} x^n \rightarrow x^\infty \in \partial (X\setminus \Delta _X) = \Delta _X\cup \big ((\partial \Sigma )^l\times (\partial \Sigma )^{m-l}\big ) \subset \Sigma ^l\times (\partial \Sigma )^{m-l}. \end{aligned}$$

Case 1. \(x^\infty \in \Delta _X\subset {\mathring{\Sigma }}^l\times (\partial \Sigma )^{m-l}\).

Then, there exists \(\xi \in \Sigma \) such that the set \(I:=\big \{i\in \{1,\dots ,m\}: x^\infty _i=\xi \big \}\) contains at least two elements. Moreover, if \(\xi \in {\mathring{\Sigma }}\), then \(I\subset \{1,\cdots ,l\}\); if \(\xi \in \partial \Sigma \), then \(I\subset \{l+1,\cdots , m\}\). In either case, we have that

$$\begin{aligned} \big |\nabla ^g_{x_i}R^g(x^n_i)\big | = {{\mathcal {O}}}(1)\text { and } \big |\nabla ^g_{x_i}G^g(x^n_i,x^n_j)\big | = {{\mathcal {O}}}(1)\quad \text { as } n\rightarrow +\infty , \end{aligned}$$

for \(i\in I\) and \(j\notin I\), and \(\big |\nabla ^g_{x_i}H^g(x^n_i,x^n_j)\big | = {{\mathcal {O}}}(1)\) for \(i,j \in I\). Let \(y_\xi :U(\xi )\rightarrow B^\xi \subset \mathbb {R}^2\) be the isothermal chart with \(y_{\xi }(\xi )=0\) introduced in Sect. 2. Then for \(i\in I\), setting \(y^n_j:= y_\xi (x^n_j) \in \mathbb {R}^2\) for \(j\in I\) and assuming \(y^n_i \ne 0\), we obtain as \(n\rightarrow \infty \):

$$\begin{aligned} \begin{aligned} \big |\nabla ^g f_g(x^n)\big |_g&\ge \big |\nabla ^g_{x_i} f_g(x^n)\big |_g = 2\left| \nabla ^g_{x_i}\sum _{j\in I\setminus \{i\}}\sigma _i \sigma _jG^g(x^n_i,x^n_j)\right| _g + {{\mathcal {O}}}(1)\\&= \frac{2}{\kappa (\xi )}\left| \nabla ^g_{x_i}\sum _{j\in I\setminus \{i\}}\sigma _i\sigma _j\log \big |y_\xi (x^n_i)-y_\xi (x^n_j)\big |\right| _g + {{\mathcal {O}}}(1)\\&\ge \frac{2c}{\kappa (\xi )}\left| \nabla _{y_i}\sum _{j\in I\setminus \{i\}}\sigma _i\sigma _j\log \big |y^n_i-y^n_j\big |\right| + {{\mathcal {O}}}(1)\\&= \frac{2c}{\kappa (\xi )}\left| \sum _{j\in I\setminus \{i\}}\sigma _i\sigma _j\frac{y^n_i-y^n_j}{|y^n_i-y^n_j|^2}\right| + {{\mathcal {O}}}(1)\\&\ge \frac{2c}{\kappa (\xi )}\left| \sum _{j\in I\setminus \{i\}}\sigma _i\sigma _j\left\langle \frac{y^n_i-y^n_j}{|y^n_i-y^n_j|^2},\frac{y^n_i}{|y^n_i|}\right\rangle \right| + {{\mathcal {O}}}(1)\ \end{aligned} \end{aligned}$$

The second inequality is a consequence of the fact that there exists \(c>0\) such that for any function \(\textrm{g}: U(\xi )\rightarrow \mathbb {R}\):

$$\begin{aligned} \big |\nabla ^g_x \textrm{g}(x)\big |_g \ge c\big |\nabla \big (\textrm{g}\circ y^{-1}_\xi \big )\big (y_{\xi }(x)\big )\big | \quad \text {for } x\in U(\xi ). \end{aligned}$$

Now for \(n\in \mathbb {N}\) we choose \(i(n)\in I\) with \(|y^n_{i(n)}| \ge |y^n_j|\) for all \(j\in I\). This implies \(y^n_{i(n)} \ne 0\) and

$$\begin{aligned} \big \langle y^n_{i(n)}-y^n_j,y^n_{i(n)}\big \rangle \ge \frac{1}{2}\big |y^n_{i(n)}-y^n_j\big |^2 > 0\quad \text {for } j\in I\setminus \{i(n)\}, \end{aligned}$$

hence

$$\begin{aligned} \left\langle \frac{y^n_{i(n)}-y^n_j}{|y^n_{i(n)}-y^n_j|^2},\frac{y^n_{i(n)}}{|y^n_{i(n)}|}\right\rangle \ge \frac{1}{2|y^n_{i(n)}|} \quad \text {for }j\in I\setminus \{i(n)\}. \end{aligned}$$

As a consequence we obtain, using \(\sigma _i\sigma _j>0\) for all \(i,j\in I\):

$$\begin{aligned} \big |\nabla ^g f_g(x^n)\big |_g\ge & {} \big |\nabla ^g_{x_{i(n)}} f_g(x^n)\big |_g \ge \frac{c}{\kappa (\xi )}\pi \sum _{j\in I\setminus \{i(n)\}}\sigma _{i(n)}\sigma _j\frac{1}{|y^n_{i(n)}|}+ {{\mathcal {O}}}(1) \\\rightarrow & {} \infty \quad \text {as }n\rightarrow \infty . \end{aligned}$$

Case 2. \(x^{\infty }\notin \Delta _X\). Then there exists \(i\in \{1,...,l\}\) such that \(x^n_i\rightarrow \xi \) for some \(\xi \in \partial \Sigma \).

We fix an isothermal chart \((y_{\xi }, U({\xi }))\) around \(\xi \) as introduced in Sect. 2. For any \(\zeta \in U_{r_{\xi }}({\xi }) \), we decompose \(G^g(\cdot ,\zeta )\) as follows:

$$\begin{aligned} G^g(\cdot ,\zeta )= \tilde{H}^g(\cdot ,\zeta )-\frac{1}{\kappa (\zeta )}\chi (4|y_{\xi }(\cdot ))-y_{\xi }(\zeta )|/r_{\xi })\log {|y_{\xi }(\cdot )-y_{\xi }(\zeta ) |}. \end{aligned}$$
(4.1)

Equation (2.2) implies \(R^g(\zeta )=\tilde{H}^g(\zeta ,\zeta )+ \frac{1}{2\kappa (\zeta )}\varphi _\xi \big (y_\xi (\zeta )\big )\). Let \(\partial _1,\partial _2\) be the standard basis of \(\mathbb {R}^2\) and define \(\partial _{\zeta _1},\partial _{\zeta _2} \in T_{\zeta }\Sigma \) be the corresponding basis of the tangent space of \(\Sigma \) at \(\zeta \in U(\xi )\).

Fix \(\zeta \in U_{r_{\xi }}(\xi )\cap {\mathring{\Sigma }}\) and set \(\delta :=y_{\xi }(\zeta )_2>0\). The representation formula for \(\tilde{H}^g\) yields:

$$\begin{aligned} \partial _{\zeta _1} \tilde{H}^g(\eta ,\zeta ) |_{\eta =\zeta }= & {} \frac{1}{|\Sigma |_g} \int _{\Sigma } \partial _{\zeta _1} \tilde{H}^g(\cdot ,\zeta )dv_g - \int _{\Sigma }G^g(\cdot ,\zeta ) \Delta _g \partial _{\zeta _1}\tilde{H}^g(\cdot ,\zeta ) dv_g\nonumber \\{} & {} + \int _{\partial \Sigma } G^g(\cdot ,\zeta )\partial _{\nu _g} \partial _{\zeta _1} \tilde{H}^g(\cdot ,\zeta ) ds_g\nonumber \\= & {} \frac{1}{2\pi } \int _{\partial \Sigma } G^g(\zeta ,x) \partial _{\nu _g}\partial _{\zeta _1}\nonumber \\ {}{} & {} \times \left( \chi \left( {4|y_{\xi }(x) -y_{\xi }(\zeta )|} /r_{\xi }\right) \log {|y_{\xi }(x)-y_{\xi }(\zeta ) |}\right) ds_g(x) + {{\mathcal {O}}}(1) \nonumber \\= & {} - \frac{1}{2\pi }\int _{B^{\xi }\cap \partial \mathbb {R}^2_{+}} G^g\left( \zeta , y_{\xi }^{-1}(y)\right) \chi (4|y-y_{\xi }(\zeta )|/ r_{\xi }) \nonumber \\ {}{} & {} \times \frac{ 2(y_1-y_{\xi }(\zeta )_1)(y_2-y_{\xi }(\zeta )_2)}{ |y-y_{\xi }(\zeta )|^4} dy + {{\mathcal {O}}}(1)\nonumber \\= & {} -\frac{1}{2\pi \delta }\int ^{\varepsilon /\delta }_{-\varepsilon /\delta } G^g\left( \zeta , y_{\xi }^{-1}\big ((\delta s,0)+ (y_{\xi }(\zeta )_1, 0)\big )\right) \nonumber \\{} & {} \times \frac{2s}{(s^2+1)^2} ds+{{\mathcal {O}}}\left( 1\right) . \end{aligned}$$
(4.2)

as \(\delta \rightarrow 0\). Here \(\varepsilon \in (0,\frac{r_{\xi }}{16})\) is chosen sufficiently small. Decomposing \(G^g\) as in (4.1), we deduce for \(\zeta \in U(\xi )\) with \(y_{\xi }(\zeta )_2< \frac{r_{\xi }}{16}\):

$$\begin{aligned} G^g\left( \zeta , y_{\xi }^{-1}\big ((\delta s,0)+ (y_{\xi }(\zeta )_1, 0)\big )\right)= & {} \tilde{H}^g\left( \zeta , y_{\xi }^{-1}\big ((\delta s,0)+ (y_{\xi }(\zeta )_1, 0)\big )\right) \\ {}{} & {} -\frac{1}{\pi }\log |(\delta s, -\delta )|, \end{aligned}$$

Now we apply the mean value theorem for \(\tilde{H}^g\) and obtain as \(\delta s\rightarrow 0\):

$$\begin{aligned} \tilde{H}^g\big (\zeta , y_{\xi }^{-1}\big ((\delta s,0)+ (y_{\xi }(\zeta )_1, 0)\big )\big )= & {} \tilde{H}^g\left( \zeta ,y_{\xi }^{-1}\big (y_{\xi }(\zeta )_1,0\big )\right) \nonumber \\ {}{} & {} + {{\mathcal {O}}}\big (|\delta s|\sup _{x\in \partial \Sigma } \Vert \nabla _x\tilde{H}^g(\cdot ,x)\Vert _{{{\mathcal {C}}}(\Sigma )}\big ) \end{aligned}$$
(4.3)

This implies as \(\delta \rightarrow 0\):

$$\begin{aligned}{} & {} \left| -\frac{1}{2\pi \delta }\int ^{\varepsilon /\delta }_{-\varepsilon /\delta } G^g(\zeta , y_{\xi }^{-1} \big ((\delta s,0)+ (y_{\xi }(\zeta )_1, 0)\big ) \frac{2s}{ (s^2+1)^2} ds\right| \\{} & {} \quad \le \left| -\frac{1}{2\pi \delta }\tilde{H}^g\big (\zeta ,y_{\xi }^{-1}\big (y_{\xi }(\zeta )_1,0\big )\big ) \int _{|s|\le \varepsilon /\delta }\frac{2s}{(s^2+1)^2} ds\right| \\{} & {} \qquad + \left| {{\mathcal {O}}}\left( \int _{|s|\le \varepsilon /\delta } \frac{2s^2}{(s^2+1)^2} ds\right) + \frac{1}{2\pi ^2\delta }\int _{|s|\le \varepsilon /\delta } \log (\delta \sqrt{s^2+1}) \frac{2s}{(s^2+1)^2} ds\right| \\{} & {} \quad \le {{\mathcal {O}}}(1). \end{aligned}$$

Now (4.2) yields \(\partial _{\zeta _1}\tilde{H}^g(\zeta ,\zeta )={{\mathcal {O}}}(1)\) for \(\zeta \in U_{r_{\xi }}(\xi )\cap \partial \Sigma \). Consequently, for \(\zeta \in U_{r_{\xi }}(\xi )\) with \(y_{\xi }(\zeta )_2< \frac{r_{\xi }}{16}\), we have proven that:

$$\begin{aligned} \partial _{\zeta _1}R^g(\zeta )= {{\mathcal {O}}}(1)\qquad \text {as } |y_{\xi }(\zeta )|\rightarrow 0. \end{aligned}$$
(4.4)

The representation formula of \(\tilde{H}^g\) yields for \(\zeta \in U_{r_{\xi }}(\xi )\cap {\mathring{\Sigma }}\) as \(\delta \rightarrow 0\):

$$\begin{aligned} \partial _{\zeta _2} \tilde{H}^g(\eta ,\zeta ) |_{\eta =\zeta }{} & {} =\frac{1}{|\Sigma |_g} \int _{\Sigma } \partial _{\zeta _2} \tilde{H}^g(\cdot ,\zeta )dv_g - \int _{\Sigma }G^g(\cdot ,\zeta ) \Delta _g \partial _{\zeta _2}\tilde{H}^g(\cdot ,\zeta ) dv_g\\{} & {} \quad + \int _{\partial \Sigma } G^g(\cdot ,\zeta )\partial _{\nu _g} \partial _{\zeta _2} \tilde{H}^g(\cdot ,\zeta ) ds_g\\{} & {} = \frac{1}{2\pi } \int _{\partial \Sigma } G^g(\zeta ,x) \partial _{\nu _g} \partial _{\zeta _2}\\{} & {} \quad \times \left( \chi \left( {4|y_{\xi }(x)-y_{\xi }(\zeta )|} /r_{\xi }\right) \log {|y_{\xi }(x)-y_{\xi }(\zeta ) |}\right) ds_g(x)+ {{\mathcal {O}}}(1) \\{} & {} \le \frac{1}{2\pi }\int _{B^{\xi }\cap \partial \mathbb {R}^2_{+}} G^g(\zeta , y_{\xi }^{-1}(y)) \chi (4|y-y_{\xi }(\zeta )|/r_{\xi })\\{} & {} \quad \times \frac{(y_1- y_{\xi }(\zeta )_1)^2- (y_2- y_{\xi }(\zeta )_2)^2}{ |y-y_{\xi }(\zeta )|^4} dy+{{\mathcal {O}}}(1) \\{} & {} \le \frac{1}{2\pi \delta }\int ^{\varepsilon /\delta }_{-\varepsilon /\delta }G^g(\zeta , y_{\xi }^{-1}((\delta s,0)+(y_{\xi }(\zeta )_1, 0)) \frac{s^2-1}{ (s^2+1)^2}ds +{{\mathcal {O}}}(1) \\{} & {} {\mathop {\le }\limits ^{(4.3)}} \frac{1}{2\pi \delta } \tilde{H}^g(\zeta ,y_{\xi }^{-1}(y_{\xi }(\zeta )_1,0))\int _{|s|\le \varepsilon /\delta } \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad + {{\mathcal {O}}}\left( 1+\int _{|s|\le \varepsilon /\delta } \frac{|s(s^2-1)|}{(s^2+1)^2} ds\right) \\{} & {} \quad + \frac{\log (\delta ^{-1})}{2\pi ^2\delta }\int _{|s|\le \varepsilon /\delta } \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad + \frac{1}{2\pi ^2\delta }\int _{|s|\le \varepsilon /\delta }\log \left( \frac{1}{\sqrt{s^2+1}}\right) \frac{s^2-1}{(s^2+1)^2} ds +{{\mathcal {O}}}(1)\\{} & {} \le \frac{1}{2\pi ^2\delta }\int _{|s|\ge 1} \log \left( \frac{\sqrt{s^{-2}+1}}{\sqrt{s^2+1}}\right) \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad - \frac{1}{2\pi ^2\delta }\int _{|s|\ge \varepsilon /\delta }\log \left( \frac{1}{\sqrt{s^2+1}}\right) \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad - \frac{\log (\delta ^{-1})\varepsilon }{\pi ^2(\delta ^2+\varepsilon ^2)}+{{\mathcal {O}}}(\delta ^{-1/2}) +{{\mathcal {O}}}(\log (\delta ^{-1})+1) \\{} & {} \le - \frac{1}{2\pi ^2\delta }\int _{|s|\ge 1} \log \left( s^{2}\right) \frac{s^2-1}{(s^2+1)^2} ds + {{\mathcal {O}}}(\delta ^{-\frac{1}{2}})\le -\frac{1}{4\pi \delta }. \end{aligned}$$

The last inequality used the identity

$$\begin{aligned} \int _{|s|\ge 1} \log (s^2) \frac{s^2-1}{(s^2+1)^2} ds =\pi . \end{aligned}$$

From the above estimate we deduce for \(\zeta \in {\mathring{\Sigma }}\cap U_{r_{\xi }}(\xi )\):

$$\begin{aligned} \partial _{\zeta _2} R^g(\zeta )\le -\frac{1}{2\pi |y_{\xi }(\zeta )_2|}\qquad \text {as }|y_{\xi }(\zeta )_2|\rightarrow 0. \end{aligned}$$
(4.5)

Now if \(I:=\big \{i\in \{1,\dots ,m\}: x^n_i\rightarrow \xi \big \}\) contains only a single element i then (4.5) yields

$$\begin{aligned} \begin{aligned} \big |\nabla ^g f_g(x^n)\big |_g&\ge \big |\nabla ^g_{x_{i}} f_g(x^n)\big |_g \ge \big |\sigma _i^2\nabla ^g_{x_{i}} R^g(x_i^n) + 2\sigma _i\sum _{j\ne i} \nabla ^g_{x_{i}} G^g(x^n_i,x^n_j)+ \nabla ^g_{x_{i}}h(x^n)\big |_g\\&\ge \sigma _i^2 |\nabla ^g_{x_i}R^g(x^n_i)|+{{\mathcal {O}}}(1) \ge \frac{ c\sigma _i^2 }{2\pi |y_{\xi }(x^n_i)_2|}+{{\mathcal {O}}}(1)\\&\rightarrow \infty \qquad \text {as }n\rightarrow \infty . \end{aligned} \end{aligned}$$

Here \(c>0\) is a constant such that for any function \(F:U(\xi )\rightarrow \mathbb {R}\):

$$\begin{aligned} \big |\nabla ^g_x F(x)\big |_g \ge c\big |\nabla \big (F\circ y^{-1}_\xi \big )\big (y_{\xi }(x)\big )\big | \quad \text {for } x\in U(\xi ). \end{aligned}$$
(4.6)

Next we consider the case that I contains at least two elements. Then \(\xi ^n_i\in U(\xi )\) for n large and \(i\in I\). By a direct calculation we obtain for \(j\in I\setminus \{i\}\), \(\iota =1,2\) and \(n\rightarrow \infty \):

$$\begin{aligned} \partial _{\zeta _\iota } G^g(x^n_j,\zeta ) |_{\zeta =x^n_{i}}= & {} \partial _{\zeta _\iota } \tilde{H}^g(x^n_j,\zeta ) |_{\zeta =x^n_{i}}\nonumber \\{} & {} + \frac{1}{\kappa (x^n_i)} \log \frac{1}{|y_{\xi }(x^n_j)- y_{\xi }(x_{i}^n)|} \partial _{\zeta _\iota }\chi (4|y_{\xi }(x^n_j)- y_{\xi }(\zeta )|/r_{\xi })|_{\zeta =x_{i}^n}\nonumber \\{} & {} + \left. \frac{1}{\kappa (x^n_i)} \chi (4|y_{\xi }(x^n_j)- y_{\xi }(x_{i}^n)|/r_{\xi })\partial _{\zeta _\iota } \log \frac{1}{|y_{\xi }(x^n_j)- y_{\xi }(\zeta )|}\right| _{\zeta =x_{i}^n}\nonumber \\= & {} \partial _{\zeta _\iota } \tilde{H}^g(x^n_j,\zeta ) |_{\zeta =x^n_{i}} -\frac{1}{\kappa (x^n_i)}\frac{(y_{\xi }(x_{i}^n)- y_{\xi }(x^n_j) )_{\iota }}{|y_{\xi }(x^n_j) - y_{\xi }(x_{i}^n)|^2} +{{\mathcal {O}}}(1). \end{aligned}$$
(4.7)

For \(n\in \mathbb {N}\) we set

$$\begin{aligned} \varrho _n:=\max \Big (\big \{ y_{\xi }(x^{n}_i)_2: i\in I\big \}\cup \big \{ |y_{\xi }(x_{i}^n)_1-y_{\xi }(x^{n}_j)_1|: i,j\in I\text { with } i\ne j \big \}\Big ). \end{aligned}$$

If there exists \(i(n)\in I\) such that \(\varrho _n=y_{\xi }\big (x^n_{i(n)}\big )_2\), then \(i(n)\in \{1,\dots ,l\}\) satisfies

$$\begin{aligned} y_{\xi }(x^n_{i(n)})_2-y_{\xi }(x^n_{j})_2\ge 0 \text { and } y_{\xi }(x^n_{i(n)})_2\ge |(y_{\xi }(x_{i}^n)_1-y_{\xi }(x^{n}_j)_1| \quad \text {for every }i,j\in I. \end{aligned}$$

Given \(\zeta =x^n_{i(n)}\) and \(\eta = x^n_j\) with \(j\in I\cap \{1,\ldots ,l\}{\setminus }\{i(n)\})\), we will calculate the upper bound of \(\partial _{\zeta _2}\tilde{H}^g(\eta ,\zeta )\) as \(n\rightarrow 0\). Setting \(a=y_{\xi }(\eta )_2>0\) and \(b=y_{\xi }(\zeta )_1-y_{\xi }(\eta )_1\) we have for \(|b|\le \varrho _n\) as \(n\rightarrow \infty \):

$$\begin{aligned} \partial _{\zeta _2} \tilde{H}^g(\eta ,\zeta ){} & {} =\frac{1}{|\Sigma |_g} \int _{\Sigma } \partial _{\zeta _2} \tilde{H}^g(\cdot ,\zeta )dv_g -\int _{\Sigma }G^g(\cdot ,\eta ) \Delta _g \partial _{\zeta _2}\tilde{H}^g(\cdot ,\zeta ) dv_g\\{} & {} \quad + \int _{\partial \Sigma } G^g(\cdot ,\eta )\partial _{\nu _g} \partial _{\zeta _2} \tilde{H}^g(\cdot ,\zeta ) ds_g\\{} & {} = {{\mathcal {O}}}(1) +\frac{1}{2\pi } \int _{\partial \Sigma } G^g(\eta ,x) \partial _{\nu _g}\partial _{\zeta _2} \big (\chi \big ({4|y_{\xi }(x)-y_{\xi }(\zeta )|} /r_{\xi }\big )\\{} & {} \quad \times \log {|y_{\xi }(x)-y_{\xi }(\zeta ) |}\big ) ds_g(x)\\{} & {} \le {{\mathcal {O}}}(1) + \frac{1}{2\pi }\int _{B^{\xi }\cap \partial \mathbb {R}^2_{+}} G^g(\eta , y_{\xi }^{-1}(y)) \chi (4|y-y_{\xi }(\zeta )|/r_{\xi })\\{} & {} \quad \times \frac{(y_1- y_{\xi }(\zeta )_1)^2-(y_2- y_{\xi }(\zeta )_2)^2}{ |y-y_{\xi }(\zeta )|^4} dy\\{} & {} \le {{\mathcal {O}}}(1) +\frac{1}{2\pi \varrho _n }\int ^{\varepsilon /\varrho _n}_{-\varepsilon /\varrho _n} G^g(\eta , y_{\xi }^{-1}((\varrho _n s,0)+ (y_{\xi }(\zeta )_1, 0)) \frac{ s^2-1}{ (s^2+1)^2} ds \\{} & {} {\mathop {\le }\limits ^{(4.3)}}\frac{1}{2\pi \varrho _n} \tilde{H}^g(\eta ,y_{\xi }^{-1}(y_{\xi }(\zeta )_1,0)) \int _{|s|\le \varepsilon /\varrho _n} \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad + {{\mathcal {O}}}\left( 1+\int _{|s|\le \varepsilon /\varrho _n} \frac{|s(s^2-1)|}{(s^2+1)^2} ds\right) \\{} & {} \quad -\frac{\log (a)}{2\pi ^2\varrho _n}\int _{|s|\le \varepsilon /\varrho _n} \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad + \frac{1}{2\pi ^2\varrho _n} \int _{|s|\le \varepsilon /\varrho _n}\log \left( \frac{1}{\sqrt{1+(\varrho _n s+b )^2/a^2}}\right) \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \le -\frac{1}{2\pi ^2\varrho _n}\int _{|s|\ge \varepsilon /\varrho _n} \left( \log (a^{-1})+ \log \left( \frac{1}{\sqrt{1+(\varrho _n s+b )^2/a^2}}\right) \right) \\{} & {} \quad \times \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad +\frac{1}{2\pi ^2\varrho _n}\int _{|s|\ge 1} \log \left( \frac{\sqrt{1+(\varrho _n s^{-1}+b )^2/a^2}}{\sqrt{1+(\varrho _n s+b )^2/a^2}}\right) \frac{s^2-1}{(s^2+1)^2} ds\\{} & {} \quad +{{\mathcal {O}}}\big (1+\log (\varrho _n^{-1})\big )\\{} & {} \le {{\mathcal {O}}}\big (\log (\varrho _n^{-1})\big ) \end{aligned}$$

Here we used the inequalities:

$$\begin{aligned} \log \left( \frac{\sqrt{1+(\varrho _n s^{-1}+b )^2/a^2}}{\sqrt{1+(\varrho _n s+b )^2/a^2}}\right) \le 0, \text { for }|b|\le \varrho _n, |s|\ge 1 \end{aligned}$$

and

$$\begin{aligned} 0\le \log \sqrt{a^2+(\varrho _n s+b)^2}\frac{s^2-1}{(s^2+1)^2} \le C \frac{\varrho _n ^{\frac{1}{2} }}{s^{\frac{3}{2}}} \end{aligned}$$

for some constant \(C>0\), any \(s\in \{ s\in \mathbb {R}:|s|\ge \varepsilon /\varrho _n,\ a^2+(\varrho _ns+b)^2\ge 1\}\). We notice that we have for \(j\in I\cap \{l+1,\ldots ,m\}\):

$$\begin{aligned} |\partial _{\zeta _2} \tilde{H}^g(x^n_j,\zeta ) |_{\zeta =x^n_{i(n)}}|= |\partial _{\zeta _2} \tilde{H}^g(\zeta ,x^n_j) |_{\zeta =x^n_{i(n)}}|\le {{\mathcal {O}}}(1) \qquad \text {as }n\rightarrow \infty ; \end{aligned}$$
(4.8)

and for \(j\in I\cap \{1,\dots ,l \}\setminus \{ i(n)\}\), \(\varrho _n=y_{\xi }(x^n_{i(n)})\):

$$\begin{aligned} \big |\partial _{\zeta _2} \tilde{H}^g(x^n_j,\zeta ) |_{\zeta =x^n_{i(n)}}\big |\le {{\mathcal {O}}}\big (\log (\varrho _n^{-1})\big ) \qquad \text {as }n\rightarrow \infty . \end{aligned}$$
(4.9)

Now (4.5)-(4.9) imply for \(n\rightarrow \infty \):

$$\begin{aligned} \begin{aligned} |\nabla ^g f_g(x^n)|_g&\ge \left| \nabla ^g_{x_{i(n)}}\left( \sigma _i^2 R^g(x^n_{i(n)})+2 \sigma _{i(n)} \sum _{j\ne i(n)} \sigma _jG^g(x^n_{i(n)}, x^n_j)+ h(x^n)\right) \right| _g\\&\ge c \left| \partial _{(x_{i(n)})_2} \left( \sigma ^2_{i(n)}R^g(\xi ^n_{i(n)})+ 2\sigma _{i(n)}\sum _{j\in I\setminus \{i(n)\} }\sigma _j G^g(\xi _{i(n)}, \xi _j)) \right) \right| +{{\mathcal {O}}}(1)\\&\ge {{\mathcal {O}}}\big (\log (\varrho _n^{-1})\big ) + \sigma ^2_{i(n)} \frac{ c }{2\pi \varrho _n}\\&\rightarrow \infty . \end{aligned} \end{aligned}$$

Here \(c>0\) is as in (4.6). If \(\varrho _n>\max \{y_{\xi }(x^n_i): i\in I\}\), then we take \(i(n)\in I\) such that \(y_{\xi }(x^n_{i(n)})_1=\max \{ y_{\xi }(x^n_i): i\in I\}\). The proof of Lemma 6 in [16] implies for \(x\in U(\xi )\) and \(y^*_{\xi }(x):= (y_{\xi }(x)_1, -y_{\xi }(x)_2)\):

$$\begin{aligned} \left\| G^g(\cdot ,x)+\frac{1}{2\pi }\log \big |y_{\xi }(\cdot )-y_{\xi }(x)\big | +\frac{1}{2\pi }\log \big |y_{\xi }(\cdot )-y^{*}_{\xi }(x)\big |\right\| _{{{\mathcal {C}}}^1(U_{r_{\xi }}(\xi ))}\le C \end{aligned}$$

where \(C>0\) depends only on \((\Sigma ,g)\) and on \(\xi \). Therefore we have for \(j\in I\setminus \{i(n)\}\) as \(n\rightarrow \infty \):

$$\begin{aligned} \partial _{\zeta _1} G(x^n_j, \zeta )\big |_{\zeta =x^n_{i(n)}}\le & {} -\frac{1}{2\pi } \frac{y_{\xi }(x^n_{i(n)})_1- y_{\xi }(x^n_j)_1}{|y_{\xi }(x^n_{i(n)})- y_{\xi }(x^n_j)|^2}\nonumber \\ {}{} & {} -\frac{1}{2\pi } \frac{y_{\xi }(x^n_{i(n)})_1- y_{\xi }(x^n_j)_1}{|y_{\xi }(x^n_{i(n)})- y^*_{\xi }(x^n_j)|^2}+ {{\mathcal {O}}}(1). \end{aligned}$$
(4.10)

We can take \(i'(n)\in I{\setminus }\{ i(n)\}\) such that \(\varrho _n= y_{\xi }(x^n_{i(n)})_1-y_{\xi }(x^n_{i'(n)})_1\) by the assumption. The inequality (4.10) yields

$$\begin{aligned} \partial _{\zeta _1} G(x^n_{i'(n)}, \zeta )\big |_{\zeta =x^n_{i(n)}}\le -\frac{1}{4\pi \varrho _n}\qquad \text {as }n\rightarrow \infty . \end{aligned}$$
(4.11)

From (4.4) together with (4.10) and (4.11) we derive the following estimate for the gradient of \(f_g\) as \(n\rightarrow \infty \):

$$\begin{aligned} \begin{aligned} |\nabla ^g f_g(x^n)|_g&\ge \left| \nabla ^g_{x_{i(n)}}\left( \sigma _i^2 R^g(x^n_{i(n)})+2 \sigma _{i(n)} \sum _{j\ne i(n)} \sigma _jG^g(x^n_{i(n)}, x^n_j)+ h(x^n)\right) \right| _g\\&\ge c \left| \partial _{(x_{i(n)})_1} \left( \sigma ^2_{i(n)}R^g(\xi ^n_{i(n)})+ 2\sigma _{i(n)}\sum _{j\in I\setminus \{i(n)\} }\sigma _j G^g(\xi _{i(n)}, \xi _j)) \right) \right| +{{\mathcal {O}}}(1)\\&\ge {{\mathcal {O}}}(1)+ \sigma _{i(n)}\sigma \big (i'(n)\big ) \frac{ c }{2 \pi \varrho _n}\\&\rightarrow \infty \end{aligned} \end{aligned}$$

Again \(c>0\) is as in (4.6). This completes the proof of Lemma 4.1.