1 Introduction

In their ground breaking work [2], Colding and Minicozzi proved that complete minimal surfaces embedded in \( \mathbb {R}^3\) with finite topology are proper. Based on the techniques in [2], Meeks and Rosenberg [5] then proved that complete minimal surfaces with positive injectivity embedded in \({\mathbb {R}}^3\) are proper. More recently, Meeks and Tinaglia [7] proved that complete constant mean curvature surfaces embedded in \(\mathbb {R}^3\) are proper if they have finite topology or have positive injectivity radius.

In contrast to the above results, in this paper we prove the following existence theorem for non-proper, complete, simply-connected surfaces embedded in \({\mathbb H}^2\times {\mathbb R}\) with constant mean curvature \(H\in (0,1/2)\). The convention used here is that the mean curvature function of an oriented surface M in an oriented Riemannian three-manifold N is the pointwise average of its principal curvatures.

The catenoids in \({\mathbb H}^2\times {\mathbb R}\) mentioned in the next theorem are defined at the beginning of Sect. 2.1.

Theorem 1.1

For any \(H\in (0,1/2)\) there exists a complete, stable, simply-connected surface \(\Sigma _H\) embedded in \({\mathbb H}^2\times {\mathbb R}\) with constant mean curvature H satisfying the following properties:

  1. (1)

    The closure of \(\Sigma _H\) is a lamination with three leaves, \(\Sigma _H\), \(C_1\) and \(C_2\), where \(C_1\) and \(C_2\) are stable catenoids of constant mean curvature H in \({\mathbb {H}}^3\) with the same axis of revolution L. In particular, \(\Sigma _{H}\) is not properly embedded in \({\mathbb H}^2\times {\mathbb R}\).

  2. (2)

    Let \(K_L\) denote the Killing field generated by rotations around L. Every integral curve of \(K_L\) that lies in the region between \(C_1\) and \(C_2\) intersects \(\Sigma _H\) transversely in a single point. In particular, the closed region between \(C_1\) and \(C_2\) is foliated by surfaces of constant mean curvature H, where the leaves are \(C_1\) and \(C_2\) and the rotated images \(\Sigma _H ({\theta })\) of \(\Sigma \) around L by angle \(\theta \in [0,2\pi )\).

When \(H=0\), Rodríguez and Tinaglia [10] constructed non-proper, complete minimal planes embedded in \({\mathbb H}^2\times {\mathbb R}\). However, their construction does not generalize to produce complete, non-proper planes embedded in \({\mathbb H}^2\times {\mathbb R}\) with non-zero constant mean curvature. Instead, the construction presented in this paper is related to the techniques developed by the authors in [3] to obtain examples of non-proper, stable, complete planes embedded in \({\mathbb {H}}^3\) with constant mean curvature H, for any \(H\in [0,1)\).

There is a general conjecture related to Theorem 1.1 and the previously stated positive properness results. Given X a Riemannian three-manifold, let \({\mathrm{Ch}}(X):= \inf _{S\in \mathcal {S}} \frac{\text {Area}(\partial S)}{\text {Volume}(S)},\) where \(\mathcal {S}\) is the set of all smooth compact domains in X. Note that when the volume of X is infinite, \(\hbox {Ch}(X)\) is the Cheeger constant.

Conjecture 1.2

Let X be a simply-connected, homogeneous three-manifold. Then for any \(H\ge \frac{1}{2}\hbox {Ch}(X)\), every complete, connected H-surface embedded in X with positive injectivity radius or finite topology is proper. On the other hand, if \(\hbox {Ch}(X)>0\), then there exist non-proper complete H-planes in X for every \(H\in [0,\frac{1}{2} \hbox {Ch}(X))\).

By the work in [2], Conjecture 1.2 holds for \(X=\mathbb {R}^3\) and it holds in \({\mathbb {H}}^3\) by work in progress in [6]. Since the Cheeger constant of \({\mathbb H}^2\times {\mathbb R}\) is 1, Conjecture 1.2 would imply that Theorem 1.1 (together with the existence of complete non-proper minimal planes embedded in \({\mathbb H}^2\times {\mathbb R}\) found in [10]) is a sharp result.

2 Preliminaries

In this section, we will review the basic properties of H-surfaces, a concept that we next define. We will call a smooth oriented surface \(\Sigma _H\) in \({\mathbb H}^2\times {\mathbb R}\) an H-surface if it is embedded and its mean curvature is constant equal to H; we will assume that \(\Sigma _H\) is appropriately oriented so that H is non-negative. We will use the cylinder model of \({\mathbb H}^2\times {\mathbb R}\) with coordinates \((\rho , \theta , t)\); here \(\rho \) is the hyperbolic distance from the origin (a chosen base point) in \(\mathbb H^2_0\), where \(\mathbb H^2_t\) denotes \(\mathbb H^2\times \{t\}\). We next describe the H-catenoids mentioned in the Introduction.

The following H-catenoids family will play a particularly important role in our construction.

2.1 Rotationally invariant vertical H-catenoids \(\mathcal {C}^H_d\)

We begin this section by recalling several results in [8, 9]. Given \(H\in (0,\frac{1}{2})\) and \(d\in [-2H,\infty )\), let

$$\begin{aligned} \eta _d=\cosh ^{-1}\left( \frac{2dH+\sqrt{1-4H^2+d^2}}{1-4H^2}\right) \end{aligned}$$

and let \(\lambda _d:[\eta _d,\infty )\rightarrow [0,\infty )\) be the function defined as follows.

$$\begin{aligned} \lambda _d(\rho )= \int ^{\rho }_{\eta _d} \frac{d+2H\cosh r}{\sqrt{\sinh ^2 r - ( d+2H\cosh r)^2}}dr. \end{aligned}$$
(1)

Note that \(\lambda _d(\rho )\) is a strictly increasing function with \(\lim _{\rho \rightarrow \infty }\lambda _d(\rho )= \infty \) and derivative \(\lambda '_d(\eta _d)=\infty \) when \(d\in (-2H,\infty )\).

In [8] Nelli, Sa Earp, Santos and Toubiana proved that there exists a 1-parameter family of embedded H-catenoids \(\{\mathcal {C}^H_d \ | \ d\in (-2H,\infty )\}\) obtained by rotating a generating curve \(\lambda _d(\rho )\) about the t-axis. The generating curve \(\widehat{\lambda }_d \) is obtained by doubling the curve \((\rho , 0, \lambda _d(\rho ))\), \(\rho \in [\eta _d,\infty )\), with its reflection \((\rho , 0, -\lambda _d(\rho ))\), \(\rho \in [\eta _d,\infty )\). Note that \(\widehat{\lambda }_d\) is a smooth curve and that the necksize, \(\eta _d\), is a strictly increasing function in d satisfying the properties that \(\eta _{-2H}=0\) and \(\lim _{d\rightarrow \infty }\eta _d=\infty \).

If \(d=-2H\), then by rotating the curve \((\rho , 0, \lambda _d(\rho ))\) around the t-axis one obtains a simply-connected H-surface \(E_H\) that is an entire graph over \(\mathbb H^2_0\). We denote by \(-E_H\) the reflection of \(E_H\) across \(\mathbb H^2_0\).

We next recall the definition of the mean curvature vector.

Definition 2.1

Let M be an oriented surface in an oriented Riemannian three-manifold and suppose that M has non-zero mean curvature H(p) at p. The mean curvature vector at p is \( \mathbf {H}(p):=H(p)N(p)\), where N(p) is its unit normal vector at p. The mean curvature vector \(\mathbf {H}(p)\) is independent of the orientation on M.

Note that the mean curvature vector \(\mathbf {H}\) of \(\mathcal {C}^H_d\) points into the connected component of \({\mathbb H}^2\times {\mathbb R}-\mathcal {C}^H_d\) that contains the t-axis. The mean curvature vector of \(E_H\) points upward while the mean curvature vector of \(-E_H\) points downward.

In order to construct the examples described in Theorem 1.1, we first obtain certain geometric properties satisfied by H-catenoids. For example, in the following lemma, we show that for certain values of \(d_1\) and \(d_2\), the catenoids \(\mathcal {C}^H_{d_1}\) and \(\mathcal {C}^H_{d_2}\) are disjoint.

Given \(d\in (-2H,\infty )\), let \(b_{d}(t):=\lambda _d^{-1}(t)\) for \(t\ge 0\); note that \(b_d(0)=\eta _{d}\). Abusing the notation let \(b_d(t):=b_d(-t)\) for \(t\le 0\).

Lemma 2.1

(Disjoint H-catenoids) Given \(d_1>2\), there exist \(d_0>d_1\) and \(\delta _0>0\) such that for any \(d_2\in [d_0,\infty )\), then

$$\begin{aligned} \inf _{t\in \mathbb R}( b_{d_2}(t)-b_{d_1}(t))\ge \delta _0. \end{aligned}$$

In particular, the corresponding H-catenoids are disjoint, i.e. \(\mathcal {C}^H_{d_1}\cap \mathcal {C}^H_{d_2}=\emptyset \).

Moreover, \(b_{d_2}(t)-b_{d_1}(t)\) is decreasing for \(t>0\) and increasing for \(t<0\). In particular,

$$\begin{aligned} \sup _{t\in \mathbb R}( b_{d_2}(t)-b_{d_1}(t))=b_{d_2}(0)-b_{d_1}(0)= \eta _{d_2}-\eta _{d_1}. \end{aligned}$$

The proof of the above lemma requires a rather lengthy computation that is given in the Appendix.

We next recall the well-known mean curvature comparison principle.

Proposition 2.2

(Mean curvature comparison principle) Let \(M_1\) and \(M_2\) be two complete, connected embedded surfaces in a three-dimensional Riemannian manifold. Suppose that \(p\in M_1\cap M_2\) satisfies that a neighborhood of p in \(M_1\) locally lies on the side of a neighborhood of p in \(M_2\) into which \(\mathbf {H}_2(p)\) is pointing. Then \(|H_1|(p)\ge |H_2|(p)\). Furthermore, if \(M_1\) and \(M_2\) are constant mean curvature surfaces with \(|H_1|=|H_2|\), then \(M_1 =M_2\).

3 The examples

For a fixed \(H\in (0,1/2)\), the outline of construction is as follows. First, we will take two disjoint H-catenoids \(\mathcal {C}_1\) and \(\mathcal {C}_2\) whose existence is given in Lemma 2.1. These catenoids \(\mathcal {C}_1\), \( \mathcal {C}_2\) bound a region \(\Omega \) in \({\mathbb H}^2\times {\mathbb R}\) with fundamental group \(\mathbb {Z}\). In the universal cover \(\widetilde{\Omega }\) of \(\Omega \), we define a piecewise smooth compact exhaustion \(\Delta _1\subset \Delta _2\subset \cdots \subset \Delta _n\subset \cdots \) of \(\widetilde{\Omega }\). Then, by solving the H-Plateau problem for special curves \(\Gamma _n\subset \partial \Delta _n\), we obtain minimizing H-surfaces \(\Sigma _n\) in \(\Delta _n\) with \(\partial \Sigma _n=\Gamma _n\). In the limit set of these surfaces, we find an H-plane \(\mathcal {P}\) whose projection to \(\Omega \) is the desired non-proper H-plane \(\Sigma _H\subset {\mathbb {H}}^2\times \mathbb {R}\).

3.1 Construction of \(\widetilde{\Omega }\)

Fix \(H\in (0,\frac{1}{2})\) and \(d_1,d_2\in (2,\infty )\), \(d_1<d_2\), such that by Lemma 2.1, the related H-catenoids \(\mathcal {C}^H_{d_1}\) and \(\mathcal {C}^H_{d_2}\) are disjoint; note that in this case, \(\mathcal {C}_{d_1}^H\) lies in the interior of the simply-connected component of \({\mathbb H}^2\times {\mathbb R}-\mathcal {C}^H_{d_2}\). We will use the notation \(\mathcal {C}_i:= \mathcal {C}^H_{d_i}\). Recall that both catenoids have the same rotational axis, namely the t-axis, and recall that the mean curvature vector \(\mathbf {H}_i\) of \(\mathcal {C}_i\) points into the connected component of \({\mathbb H}^2\times {\mathbb R}-\mathcal {C}_i\) that contains the t-axis. We emphasize here that H is fixed and so we will omit describing it in future notations.

Let \(\Omega \) be the closed region in \({\mathbb H}^2\times {\mathbb R}\) between \(\mathcal {C}_1\) and \(\mathcal {C}_2\), i.e., \(\partial \Omega = \mathcal {C}_1\cup \mathcal {C}_2\) (Fig. 1-left). Notice that the set of boundary points at infinity \(\partial _{\infty }\Omega \) is equal to \(S^1_{\infty }\times \{-\infty \} \cup S^1_{\infty }\times \{\infty \}\), i.e., the corner circles in \(\partial _\infty {\mathbb H}^2\times {\mathbb R}\) in the product compactification, where we view \({\mathbb {H}}^2\) to be the open unit disk \(\{(x,y)\in \mathbb {R}^2 \mid x^2+y^2<1\}\) with base point the origin \(\vec {0}\).

Fig. 1
figure 1

The induced coordinates \((\rho , \widetilde{\theta }, t)\) in \(\widetilde{\Omega }\)

By construction, \(\Omega \) is topologically a solid torus. Let \(\widetilde{\Omega }\) be the universal cover of \(\Omega \). Then, \(\partial \widetilde{\Omega } = \widetilde{\mathcal {C}}_1\cup \widetilde{\mathcal {C}}_2\) (Fig. 1-right), where \(\widetilde{\mathcal {C}}_1,\widetilde{\mathcal {C}}_2\) are the respective lifts to \(\widetilde{\Omega }\) of \({\mathcal {C}}_1, {\mathcal {C}}_2\). Notice that \(\widetilde{\mathcal {C}}_1\) and \(\widetilde{\mathcal {C}}_2\) are both H-planes, and the mean curvature vector \(\mathbf {H}\) points outside of \(\widetilde{\Omega }\) along \(\widetilde{\mathcal {C}}_1\) while \(\mathbf {H}\) points inside of \(\widetilde{\Omega }\) along \(\widetilde{\mathcal {C}}_2\). We will use the induced coordinates \((\rho , \widetilde{\theta }, t)\) on \(\widetilde{\Omega }\) where \(\widetilde{\theta }\in (-\infty , \infty )\). In particular, if

$$\begin{aligned} \Pi :\widetilde{\Omega }\rightarrow \Omega \end{aligned}$$
(2)

is the covering map, then \(\Pi (\rho _o, \widetilde{\theta }_o, t_o)= (\rho _o, \theta _o, t_o)\) where \(\theta _o\equiv \widetilde{\theta _o} \mod 2\pi \).

Recalling the definition of \(b_i(t)\), \(i=1,2\), note that a point \((\rho , \theta , t)\) belongs to \(\Omega \) if and only if \(\rho \in [b_1(t), b_2(t)]\) and we can write

$$\begin{aligned} \widetilde{\Omega }= \{ (\rho , \widetilde{\theta }, t) \ | \ \rho \in [b_1(t), b_2(t)], \ \widetilde{\theta }\in \mathbb R,\ t\in \mathbb R\}. \end{aligned}$$

3.2 Infinite bumps in \(\widetilde{\Omega }\)

Let \(\gamma \) be the geodesic through the origin in \(\mathbb H^2_0\) obtained by intersecting \(\mathbb H^2_0\) with the vertical plane \(\{\theta =0\} \cup \{\theta =\pi \}\). For \(s\in [0,\infty )\), let \(\varphi _s\) be the orientation preserving hyperbolic isometry of \(\mathbb H^2_0\) that is the hyperbolic translation along the geodesic \(\gamma \) with \(\varphi _s(0,0)=(s,0)\). Let

$$\begin{aligned} \widehat{\varphi }_s:{\mathbb H}^2\times {\mathbb R}\rightarrow {\mathbb H}^2\times {\mathbb R}, \quad \widehat{\varphi }_s(\rho ,\theta ,t)=(\varphi _s(\rho ,\theta ),t) \end{aligned}$$
(3)

be the related extended isometry of \({\mathbb H}^2\times {\mathbb R}\).

Let \(\mathcal {C}_d\) be an embedded H-catenoid as defined in Sect. 2.1. Notice that the rotation axis of the H-catenoid \(\widehat{\varphi }_{s_0}(\mathcal {C}_d)\) is the vertical line \(\{ (s_0, 0, t) \mid t\in \mathbb R\}\).

Let \(\delta :=\inf _{t\in \mathbb R}(b_2(t)-b_1(t))\), which gives an upper bound estimate for the asymptotic distance between the catenoids; recall that by our choices of \(\mathcal {C}_1,\mathcal {C}_2\) given in Lemma 2.1, we have \(\delta >0\). Let \(\delta _1=\frac{1}{2}\min \{\delta , \eta _1\}\) and let \(\delta _2=\delta -\frac{\delta _1}{2}\). Let \(\widehat{\mathcal {C}}_{1}:=\widehat{\varphi }_{\delta _1}(\mathcal {C}_1)\) and \(\widehat{\mathcal {C}}_{2}:=\widehat{\varphi }_{-\delta _2}(\mathcal {C}_2)\). Note that \(\delta _1+\delta _2>\delta \).

Claim 3.1

The intersection \(\Omega \cap \widehat{\mathcal {C}}_{i}\), \(i=1,2\), is an infinite strip.

Proof

Given \(t\in \mathbb R\), let \(\mathbb H^2_t\) denote \(\mathbb H^2\times \{t\}\). Let \(\tau ^i_t:=\mathcal {C}_i\cap \mathbb H^2_t\) and \(\widehat{\tau }^i_t:=\widehat{\mathcal {C}}_i\cap \mathbb H^2_t\). Note that for \(i=1,2\), \(\tau ^i_t\) is a circle in \(\mathbb H^2_t\) of radius \(b_i(t)\) centered at (0, 0, t) while \(\widehat{\tau }^1_t\) is a circle in \(\mathbb H^2_t\) of radius \(b_1(t)\) centered at \(p_{1,t}:=(\delta _1,0,t)\) and \(\widehat{\tau }^2_t\) is a circle in \(\mathbb H^2_t\) of radius \(b_2(t)\) centered at \(p_{2,t}:=(-\delta _2,0,t)\). We claim that for any \(t\in \mathbb R\), the intersection \(\widehat{\tau }^i_t\cap \Omega \) is an arc with end points in \(\tau ^i_t\), \(i=1,2\). This result would give that \(\Omega \cap \widehat{\mathcal {C}}_{i}\) is an infinite strip. We next prove this claim.

Consider the case \(i=1\) first. Since \(\delta _1<\eta _1\le b_1(t)\), the center \(p_{1,t}\) is inside the disk in \(\mathbb H^2_t\) bounded by \(\tau ^1_t\). Since the radii of \(\tau ^1_t\) and \(\widehat{\tau }^1_t\) are both equal to \(b_1(t)\), then the intersection \(\tau ^1_t\cap \widehat{\tau }^1_t\) is nonempty. It remains to show that \(\widehat{\tau }^1_t\cap \tau ^2_t=\emptyset \), namely that \(b_1(t)+\delta _1<b_2(t)\). This follows because

$$\begin{aligned} \delta _1<\delta =\inf _{t\in \mathbb R}(b_2(t)-b_1(t)). \end{aligned}$$

This argument shows that \(\Omega \cap \widehat{\mathcal {C}}_{1}\) is an infinite strip.

Consider now the case \(i=2\). Since \(\delta _2<\delta < b_2(t)\), the center \(p_{2,t}\) is inside the disk in \(\mathbb H^2_t\) bounded by \(\tau ^2_t\). Since the radii of \(\tau ^2_t\) and \(\widehat{\tau }^2_t\) are both equal to \(b_2(t)\), then the intersection \(\tau ^2_t\cap \widehat{\tau }^2_t\) is nonempty. It remains to show that \(\tau ^1_t\cap \widehat{\tau }^2_t=\emptyset \), namely that \(b_2(t)-\delta _2>b_1(t)\). This follows because

$$\begin{aligned} b_2(t)-b_1(t)\ge \inf _{t\in \mathbb R}(b_2(t)-b_1(t))=\delta >\delta _2 \end{aligned}$$

This completes the proof that \(\Omega \cap \widehat{\mathcal {C}}_{2}\) is an infinite strip and finishes the proof of the claim. \(\square \)

Fig. 2
figure 2

The position of the bumps \(\mathcal {B}^\pm \) in \(\widetilde{\Omega }\) is shown in the picture. The small arrows show the mean curvature vector direction. The H-surfaces \(\Sigma _n\) are disjoint from the infinite strips \(\mathcal {B}^\pm \) by construction

Now, let \(Y^+:=\Omega \cap \widehat{\mathcal {C}}_2\) and let \(Y^-:=\Omega \cap \widehat{\mathcal {C}}_1\). In light of Claim 3.1 and its proof, we know that \(Y^+\cap \mathcal {C}_1=\emptyset \) and \(Y^-\cap \mathcal {C}_2=\emptyset \).

Remark 3.2

Note that by construction, any rotational surface contained in \(\Omega \) must intersect \(\widehat{\mathcal {C}}_{1}\cup \widehat{\mathcal {C}}_{2}\). In particular, \(Y^+\cup Y^-\) intersects all H-catenoids \(\mathcal {C}_d\) for \(d\in (d_1,d_2)\) as the circles \(\mathcal {C}_d\cap \mathbb H^2_t\) intersect either the circle \(\widehat{\tau }^2_t\) or the circle \(\widehat{\tau }^1_t\) for some \(t>0\) since \(\delta _1+\delta _2>\delta \).

In \(\widetilde{\Omega }\), let \(\mathcal {B}^+\) be the lift of \(Y^+\) in \(\widetilde{\Omega }\) which intersects the slice \(\{\widetilde{\theta }=-10\pi \}\). Similarly, let \(\mathcal {B}^-\) be the lift of \(Y^-\) in \(\widetilde{\Omega }\) which intersects the slice \(\{\widetilde{\theta }=10\pi \}\). Note that each lift of \(Y^+\) or \(Y^-\)is contained in a region where the \(\widetilde{\theta }\) values of their points lie in ranges of the form \((\theta _0-\pi ,\theta _0+ \pi )\) and so \(\mathcal {B}^+\cap \mathcal {B}^-=\emptyset \). See Fig. 2.

The H-surfaces \(\mathcal {B}^\pm \) near the top and bottom of \(\widetilde{\Omega }\) will act as barriers (infinite bumps) in the next section, ensuring that the limit H-plane of a certain sequence of compact H-surfaces does not collapse to an H-lamination of \(\widetilde{\Omega }\) all of whose leaves are invariant under translations in the \(\widetilde{\theta }\)-direction.

Next we modify \(\widetilde{\Omega }\) as follows. Consider the component of \(\widetilde{\Omega }-(\mathcal {B}^+\cup \mathcal {B}^-)\) containing the slice \(\{\widetilde{\theta }=0\}\). From now on we will call the closure of this region \(\widetilde{\Omega }^*\).

3.3 The compact exhaustion of \(\widetilde{\Omega }^*\)

Consider the rotationally invariant H-planes \(E_H,-E_H\) described in Sect. 2. Recall that \(E_H\) is a graph over the horizontal slice \(\mathbb H^2_0\) and it is also tangent to \(\mathbb H^2_0\) at the origin. Given \(t\in \mathbb R\), let \(E_H^t=-E_H+(0,0,t)\) and \(-E_H^t=E_H-(0,0,t)\). Both families \(\{E_H^t\}_{t\in \mathbb R}\) and \(\{-E_H^t\}_{t\in \mathbb R}\) foliate \({\mathbb H}^2\times {\mathbb R}\). Moreover, there exists \(n_0\in \mathbb N\) such that for any \(n>n_0\), \(n\in \mathbb N\), the following holds. The highest (lowest) component of the intersection \(S_n^+:=E_H^n\cap \Omega \) (\(S_n^-:=-E_H^n\cap \Omega \)) is a rotationally invariant annulus with boundary components contained in \(\mathcal {C}_1\) and \(\mathcal {C}_2\). The annulus \(S_n^+\) lies “above” \(S_n^-\) and their intersection is empty. The region \(\mathcal {U}_n\) in \(\Omega \) between \(S_n^+\) and \(S_n^-\) is a solid torus, see Fig. 3-left, and the mean curvature vectors of \(S_n^+\) and \(S_n^-\) point into \(\mathcal {U}_n\).

Fig. 3
figure 3

\(\mathcal {U}_n = \Omega \cap \widehat{\mathcal {U}}_n\) and \(\widetilde{\mathcal {U}}_n\) denotes its universal cover. Note that \(\partial \widetilde{\mathcal {U}}_n\subset \widetilde{\mathcal {C}}_1\cup \widetilde{\mathcal {C}}_2\cup \widetilde{S}^+_n\cup \widetilde{S}^-_n\)

Let \(\widetilde{\mathcal {U}}_n\subset \widetilde{\Omega }\) be the universal cover of \(\mathcal {U}_n\), see Fig. 3-right. Then, \(\partial \widetilde{\mathcal {U}}_n- \partial \widetilde{\Omega }= \widetilde{S}_n^+\cup \widetilde{S}_n^-\) where can view \(\widetilde{S}_n^\pm \) as a lift to \(\widetilde{\mathcal {U}}_n\) of the universal cover of the annulus \(S_n^\pm \). Hence, \(\widetilde{S}_n^\pm \) is an infinite H-strip in \(\widetilde{\Omega }\), and the mean curvature vectors of the surfaces \(\widetilde{S}_n^+, \widetilde{S}_n^-\) point into \(\widetilde{\mathcal {U}}_n\) along \(\widetilde{S}_n^\pm \). Note that each \(\widetilde{\mathcal {U}}_n\) has bounded t-coordinate. Furthermore, we can view \(\widetilde{\mathcal {U}}_n\) as \((\mathcal {U}_n\cap \mathcal {P}_0)\times \mathbb R\), where \(\mathcal {P}_0\) is the half-plane \(\{\theta =0\}\) and the second coordinate is \(\widetilde{\theta }\). Abusing the notation, we redefine \(\widetilde{\mathcal {U}}_n\) to be \(\widetilde{\mathcal {U}}_n\cap \widetilde{\Omega }^*\), that is we have removed the infinite bumps \(\mathcal {B}^\pm \) from \(\widetilde{\mathcal {U}}_n\).

Now, we will perform a sequence of modifications of \(\widetilde{\mathcal {U}}_n\) so that for each of these modifications, the \(\widetilde{\theta }\)-coordinate in \(\widetilde{\mathcal {U}}_n\) is bounded and so that we obtain a compact exhaustion of \(\widetilde{\Omega }^*\). In order to do this, we will use arguments that are similar to those in Claim 3.1. Recall that the necksize of \(\mathcal {C}_2\) is \(\eta _2=b_2(0)\). Let \(\widehat{\mathcal {C}}_3= \widehat{\varphi }_{\eta _2}(\mathcal {C}_2)\), see equation (3) for the definition of \(\widehat{\varphi }_{\eta _2}\). Then, \(\widehat{\mathcal {C}}_3\) is a rotationally invariant catenoid whose rotational axis is the line \((\eta _2, 0)\times \mathbb R\) (Fig. 4-left).

Lemma 3.3

The intersection \(\widehat{\mathcal {C}}_3\cap \Omega \) is a pair of infinite strips.

Proof

It suffices to show that \(\widehat{\mathcal {C}}_3\cap \mathcal {C}_1\) and \(\widehat{\mathcal {C}}_3\cap \mathcal {C}_2\) each consists of a pair of infinite lines. Now, consider the horizontal circles \(\tau ^1_t, \tau ^2_t\), and \(\widehat{\tau }^3_t\) in the intersection of \(\mathbb H^2_t\) and \(\mathcal {C}_1,\mathcal {C}_2\), and \(\widehat{\mathcal {C}}_3\) respectively, where \(\mathbb H^2_t= \mathbb H^2\times \{t\}\). For any \(t\in \mathbb R\), \(\tau ^i_t\) is a circle of radius \(b_i(t)\) in \(\mathbb H^2_t\) with center (0, 0, t). Similarly, \(\widehat{\tau }^3_t\) is a circle of radius \(b_2(t)\) in \(\mathbb H^2_t\) with center \((\eta _2,0,t)\), see Fig. 4-right. Hence, it suffices to show that for any \(t\in \mathbb R\) each of the intersection \(\tau ^1_t\cap \widehat{\tau }^3_t\) and \(\tau ^2_t\cap \widehat{\tau }^3_t\) consists of two points.

Fig. 4
figure 4

\(\tau ^i_t=\mathcal {C}_i\cap \mathbb H^2_t\) is a round circle of radius \(b_i(t)\) with center O. \(\widehat{\tau }^3_t=\widehat{\mathcal {C}}_3\cap \mathbb H^2_t\) is a round circle of radius \(b_2(t)\) with center \(C=(\eta _2,0,t)\)

By construction, it is easy to see \(\tau ^2_t\cap \widehat{\tau }^3_t\) consists of two points. This is because \(\tau ^2_t\) and \(\widehat{\tau }^3_t\) have the same radius, \(b_2(t)\) and \(\eta _2+b_2(t)>b_2(t)\) and \(\eta _2-b_2(t)>-b_2(t)\). Therefore, it remains to show that \(\tau ^1_t\cap \widehat{\tau }^3_t\) consists of two points. By construction, this would be the case if \(\eta _2-b_2(t)<b_1(t)\) and \(\eta _2-b_2(t)>-b_1(t)\). The first inequality follows because \(\eta _2=\inf _{t\in \mathbb R}b_2(t)\). The second inequality follows from Lemma 2.1 because

$$\begin{aligned} \eta _2>\eta _2-\eta _1=\sup _{t\in \mathbb R}(b_2(t)-b_1(t)). \end{aligned}$$

\(\square \)

Now, let \(\widehat{\mathcal {C}}_3\cap \Omega = T^+\cup T^-\), where \(T^+\) is the infinite strip with \(\theta \in (0,\pi )\), and \(T^-\) is the infinite strip with \(\theta \in (-\pi ,0)\). Note that \(T^\pm \) is a \(\theta \)-graph over the infinite strip \(\widehat{\mathcal {P}}_0=\Omega \cap \mathcal {P}_0\) where \(\mathcal {P}_0\) is the half plane \(\{\theta =0\}\). Let \(\mathcal {V}\) be the component of \(\Omega -\widehat{C}_3\) containing \(\widehat{\mathcal {P}}_0\). Notice that the mean curvature vector \(\mathbf {H}\) of \(\partial \mathcal {V}\) points into \(\mathcal {V}\) on both \(T^+\) and \(T^-\).

Consider the lifts of \(T^+\) and \(T^-\) in \(\widetilde{\Omega }\). For \(n\in \mathbb {Z}\), let \(\widetilde{T}^+_n\) be the lift of \(T^+\) which belongs to the region \(\widetilde{\theta }\in (2n\pi , (2n+1)\pi )\). Similarly, let \(\widetilde{T}^-_n\) be the lift of \(T^-\) which belongs to the region \(\widetilde{\theta }\in ((2n-1)\pi , 2n\pi )\). Let \(\mathcal {V}_n\) be the closed region in \(\widetilde{\Omega }\) between the infinite strips \(\widetilde{T}^-_{-n}\) and \(\widetilde{T}^+_n\). Notice that for n sufficiently large, \(\mathcal {B}^\pm \subset \mathcal {V}_n\).

Next we define the compact exhaustion \(\Delta _n\) of \(\widetilde{\Omega }^*\) as follows: \(\Delta _n:=\widetilde{\mathcal {U}}_n\cap \mathcal {V}_n\). Furthermore, the absolute value of the mean curvature of \(\partial \Delta _n\) is equal to H and the mean curvature vector \(\mathbf {H}\) of \(\partial \Delta _n\) points into \(\Delta _n\) on \(\partial \Delta _n-[(\partial \Delta _n\cap \widetilde{\mathcal {C}}_1)\cup \mathcal {B}^-]\).

3.4 The sequence of H-surfaces

We next define a sequence of compact H-surfaces \(\{\Sigma _n\}_{n\in \mathbb {N}}\) where \(\Sigma _n\subset \Delta _n\). For each n sufficiently large, we define a simple closed curve \(\Gamma _n\) in \(\partial \Delta _n\), and then we solve the H-Plateau problem for \(\Gamma _n\) in \(\Delta _n\). This will provide an embedded H-surface \(\Sigma _n\) in \(\Delta _n\) with \(\partial \Sigma _n=\Gamma _n\) for each n.

The Construction of \(\Gamma _n\) in \(\partial \Delta _n:\)

First, consider the annulus \(\mathcal {A}_n=\partial \Delta _n-(\widetilde{\mathcal {C}}_1\cup \widetilde{\mathcal {C}}_2\cup \mathcal {B}^+\cup \mathcal {B}^-)\) in \(\partial \Delta _n\). Let \(\widehat{l}_n^+= \widetilde{\mathcal {C}}_1\cap \widetilde{T}^+_n\), and \(\widehat{l}_n^-= \widetilde{\mathcal {C}}_2\cap \widetilde{T}^-_{-n}\) be the pair of infinite lines in \(\widetilde{\Omega }\). Let \(l^\pm _n=\widehat{l}^\pm _n\cap \mathcal {A}_n\). Let \(\mu _n^+\) be an arc in \(\widetilde{S}^+_n\cap \mathcal {A}_n\), whose \(\widetilde{\theta }\) and \(\rho \) coordinates are strictly increasing as a function of the parameter and whose endpoints are \(l^+_n\cap \widetilde{S}^+_n\) and \(l^-_n\cap \widetilde{S}^+_n\) (Fig. 5-left). Similarly, define \(\mu _n^-\) to be a monotone arc in \(\widetilde{S}^-_n\cap \mathcal {A}_n\) whose endpoints are \(l^+_n\cap \widetilde{S}^-_n\) and \(l^-_n\cap \widetilde{S}^-_n\). Note that these arcs \(\mu ^+_n\) and \(\mu ^-_n\) are by construction disjoint from the infinite bumps \(\mathcal {B}^\pm \). Then, \(\Gamma _n=\mu ^+_n\cup l^+_n \cup \mu ^-_n\cup l^-_n\) is a simple closed curve in \(\mathcal {A}_n\subset \partial \Delta _n\) (Fig. 5-right).

Fig. 5
figure 5

In the left, \(\mu ^n_+\) is pictured in \(\widetilde{S}^+_n\). On the right, the curve \(\Gamma _n\) is described in \(\partial \Delta _n\)

Next, consider the following variational problem (H-Plateau problem): Given the simple closed curve \(\Gamma _n\) in \(\mathcal {A}_n\), let M be a smooth compact embedded surface in \(\Delta _n\) with \(\partial M=\Gamma _n\). Since \(\Delta _n\) is simply-connected, M separates \(\Delta _n\) into two regions. Let Q be the region in \(\Delta _n-\Sigma \) with \(Q\cap \widetilde{\mathcal {C}}_2\ne \emptyset \), the “upper” region. Then define the functional \(\mathcal {I}_H=\hbox {Area}(M)+2H\,\hbox {Volume}(Q)\).

By working with integral currents, it is known that there exists a smooth (except at the 4 corners of \(\Gamma _n\)), compact, embedded H-surface \(\Sigma _n\subset \Delta _n\) with \(\text {Int}(\Sigma _n)\subset \text {Int}(\Delta _n)\) and \(\partial \Sigma _n=\Gamma _n\). Note that in our setting, \(\Delta _n\) is not H-mean convex along \(\Delta _n\cap \widetilde{\mathcal {C}}_1\). However, the mean curvature vector along \(\Sigma _n\) points outside Q because of the construction of the variational problem. Therefore \(\Delta _n\cap \widetilde{\mathcal {C}}_1\) is still a good barrier for solving the H-Plateau problem. In fact, \(\Sigma _n\) can be chosen to be, and we will assume it is, a minimizer for this variational problem, i.e., \(I(\Sigma _n)\le I(M)\) for any \(M\subset \Delta _n\) with \(\partial M =\Gamma _n\); see for instance [12, Theorem 2.1] and [1, Theorem 1]. In particular, the fact that \(\text {Int}(\Sigma _n)\subset \text {Int}(\Delta _n)\) is proven in Lemma 3 of [4]. Moreover, \(\Sigma _n\) separates \(\Delta _n\) into two regions.

Similarly to Lemma 4.1 in [3], in the following lemma we show that for any such \(\Gamma _n\), the minimizer surface \(\Sigma _n\) is a \(\widetilde{\theta }\)-graph.

Lemma 3.4

Let \(E_n:=\mathcal {A}_n\cap \widetilde{T}^+_n\). The minimizer surface \(\Sigma _n\) is a \(\widetilde{\theta }\)-graph over the compact disk \(E_n\). In particular, the related Jacobi function \(J_n\) on \(\Sigma _n\) induced by the inner product of the unit normal field to \(\Sigma _n\) with the Killing field \( \partial _{\widetilde{\theta }}\) is positive in the interior of \(\Sigma _n\).

Proof

The proof is almost identical to the proof of Lemma 4.1 in [3], and for the sake of completeness, we give it here. Let \(T_\alpha \) be the isometry of \(\widetilde{\Omega }\) which is a translation by \(\alpha \) in the \(\widetilde{\theta }\) direction, i.e.,

$$\begin{aligned} T_{\alpha }(\rho , \widetilde{\theta }, t)=(\rho , \widetilde{\theta }+\alpha , t). \end{aligned}$$
(4)

Let \(T_\alpha (\Sigma _n)=\Sigma ^\alpha _n\) and \(T_\alpha (\Gamma _n)=\Gamma ^\alpha _n\). We claim that \(\Sigma ^\alpha _n\cap \Sigma _n=\emptyset \) for any \(\alpha \in {\mathbb {R}}\setminus \{0\}\) which implies that \(\Sigma _n\) is a \(\widetilde{\theta }\)-graph; we will use that \(\Gamma ^\alpha _n \) is disjoint from \(\Sigma _n\) for any \(\alpha \in {\mathbb {R}}\setminus \{0\}\).

Arguing by contradiction, suppose that \(\Sigma ^\alpha _n\cap \Sigma _n\ne \emptyset \) for a certain \(\alpha \ne 0\). By compactness of \(\Sigma _n\), there exists a largest positive number \(\alpha '\) such that \(\Sigma ^{\alpha '}_n\cap \Sigma _n\ne \emptyset \). Let \(p\in \Sigma ^{\alpha '}_n\cap \Sigma _n\). Since \(\partial \Sigma ^{\alpha '}_n \cap \partial \Sigma _n =\emptyset \) and the interior of \(\Sigma _n\), respectively \(\Sigma ^{\alpha '}_n\), lie in the interior of \(\Delta _n\), respectively \(T_{\alpha '}(\Delta _n)\), then \(p\in \mathrm{Int}(\Sigma ^{\alpha '}_n) \cap \mathrm{Int}(\Sigma _n)\). Since the surfaces \(\mathrm{Int}(\Sigma ^{\alpha '}_n)\), \(\mathrm{Int}(\Sigma _n)\) lie on one side of each other and intersect tangentially at the point p with the same mean curvature vector, then we obtain a contradiction to the mean curvature comparison principle for constant mean curvature surfaces, see Proposition 2.2. This proves that \(\Sigma _n\) is graphical over its \(\widetilde{\theta }\)-projection to \(E_n\).

Since by construction every integral curve, \((\overline{\rho },s,\overline{t})\) with \(\overline{\rho }, \overline{t}\) fixed and \((\overline{\rho },s_0, \overline{t})\in E_n\) for a certain \(s_0\), of the Killing field \(\partial _{\widetilde{\theta }}\) has non-zero intersection number with any compact surface bounded by \(\Gamma _n\), we conclude that every such integral curve intersects both the disk \(E_n\) and \(\Sigma _n\) in single points. This means that \(\Sigma _n\) is a \(\widetilde{\theta }\)-graph over \(E_n\) and thus the related Jacobi function \(J_n\) on \(\Sigma _n\) induced by the inner product of the unit normal field to \(\Sigma _n\) with the Killing field \(\partial _{\widetilde{\theta }}\) is non-negative in the interior of \(\Sigma _n\). Since \(J_n\) is a non-negative Jacobi function, then either \(J_n\equiv 0\) or \(J_n>0\). Since by construction \(J_n\) is positive somewhere in the interior, then \(J_n\) is positive everywhere in the interior. This finishes the proof of the lemma. \(\square \)

4 The proof of Theorem 1.1

With \(\Gamma _n\) as previously described, we have so far constructed a sequence of compact stable H-disks \(\Sigma _n\) with \(\partial \Sigma _n = \Gamma _n \subset \partial \Delta _n\). Let \(J_n\) be the related non-negative Jacobi function described in Lemma 3.4.

By the curvature estimates for stable H-surfaces given in [11], the norms of the second fundamental forms of the \(\Sigma _n\) are uniformly bounded from above at points which are at intrinsic distance at least one from their boundaries. Since the boundaries of the \(\Sigma _{n}\) leave every compact subset of \(\widetilde{\Omega }^*\), for each compact set of \(\widetilde{\Omega }^*\), the norms of the second fundamental forms of the \(\Sigma _n\) are uniformly bounded for values n sufficiently large and such a bound does not depend on the chosen compact set. Standard compactness arguments give that, after passing to a subsequence, \(\Sigma _n\) converges to a (weak) H-lamination \(\widetilde{\mathcal {L}}\) of \(\widetilde{\Omega }^*\) and the leaves of \(\widetilde{\mathcal {L}}\) are complete and have uniformly bounded norm of their second fundamental forms, see for instance [5].

Let \(\beta \) be a compact embedded arc contained in \(\widetilde{\Omega }^*\) such that its end points \(p_+\) and \(p_-\) are contained respectively in \(\mathcal {B}^+\) and \(\mathcal {B}^-\), and such that these are the only points in the intersection \([\mathcal {B}^+\cup \mathcal {B}^-]\cap \beta \). Then, for n-sufficiently large, the linking number between \(\Gamma _n\) and \(\beta \) is one, which gives that, for n sufficiently large, \(\Sigma _n\) intersects \(\beta \) in an odd number of points. In particular \(\Sigma _n\cap \beta \ne \emptyset \) which implies that the lamination \(\widetilde{\mathcal {L}}\) is not empty.

Remark 4.1

By Remark 3.2, a leaf of \(\widetilde{\mathcal {L}}\) that is invariant with respect to \(\widetilde{\theta }\)-translations cannot be contained in \(\widetilde{\Omega }^*\). Therefore none of the leaves of \(\widetilde{\mathcal {L}}\) are invariant with respect to \(\widetilde{\theta }\)-translations.

Let \(\widetilde{L}\) be a leaf of \(\widetilde{\mathcal {L}}\) and let \(J_{\widetilde{L}}\) be the Jacobi function induced by taking the inner product of \(\partial _{\widetilde{\theta }}\) with the unit normal of \(\widetilde{L}\). Then, by the nature of the convergence, \(J_{\widetilde{L}}\ge 0\) and therefore since it is a Jacobi field, it is either positive or identically zero. In the latter case, \(\widetilde{\mathcal {L}}\) would be invariant with respect to \(\widetilde{\theta }\)-translations, contradicting Remark 4.1. Thus, by Remark 4.1, we have that \(J_{\widetilde{L}}\) is positive and therefore \(\widetilde{L}\) is a Killing graph with respect to \(\partial _{\widetilde{\theta }}\).

Claim 4.2

Each leaf \(\widetilde{L}\) of \(\widetilde{\mathcal {L}}\) is properly embedded in \(\widetilde{\Omega }^*\).

Proof

Arguing by contradiction, suppose there exists a leaf \(\widetilde{L}\) of \(\widetilde{\mathcal {L}}\) that is NOT proper in \(\widetilde{\Omega }^*\). Then, since the leaf \(\widetilde{L}\) has uniformly bounded norm of its second fundamental form, the closure of \(\widetilde{L}\) in \(\widetilde{\Omega }^*\) is a lamination of \(\widetilde{\Omega }^*\) with a limit leaf \(\Lambda \), namely \(\Lambda \subset \overline{\widetilde{L}}-\widetilde{L}\). Let \(J_{\Lambda }\) be the Jacobi function induced by taking the inner product of \(\partial _{\widetilde{\theta }}\) with the unit normal of \(\Lambda \).

Just like in the previous discussion, by the nature of the convergence, \(J_{\Lambda }\ge 0\) and therefore, since it is a Jacobi field, it is either positive or identically zero. In the latter case, \(\Lambda \) would be invariant with respect to \(\widetilde{\theta }\)-translations and thus, by Remark 4.1, \(\Lambda \) cannot be contained in \(\widetilde{\Omega }^*\). However, since \(\Lambda \) is contained in the closure of \(\widetilde{ L}\), this would imply that \(\widetilde{L}\) is not contained in \(\widetilde{\Omega }^*\), giving a contradiction. Thus, \(J_{\Lambda }\) must be positive and therefore, \(\Lambda \) is a Killing graph with respect to \(\partial _{\widetilde{\theta }}\). However, this implies that \(\widetilde{L}\) cannot be a Killing graph with respect to \(\partial _{\widetilde{\theta }}\). This follows because if we fix a point p in \(\Lambda \) and let \(U_p\subset \Lambda \) be neighborhood of such point, then by the nature of the convergence, \(U_p\) is the limit of a sequence of disjoint domains \(U_{p_n}\) in \(\widetilde{L}\) where \(p_n\in \widetilde{L}\) is a sequence of points converging to p and \(U_{p_n}\subset \widetilde{L}\) is a neighborhood of \(p_n\). While each domain \(U_{p_n}\) is a Killing graph with respect to \(\partial _{\widetilde{\theta }}\), the convergence to \(U_p\) implies that their union is not. This gives a contradiction and proves that \(\Lambda \) cannot be a Killing graph with respect to \(\partial _{\widetilde{\theta }}\). Since we have already shown that \(\Lambda \) must be a Killing graph with respect to \(\partial _{\widetilde{\theta }}\), this gives a contradiction. Thus \(\Lambda \) cannot exist and each leaf \(\widetilde{L}\) of \(\widetilde{\mathcal {L}}\) is properly embedded in \(\widetilde{\Omega }^*\). \(\square \)

Arguing similarly to the proof of the previous claim, it follows that a small perturbation of \(\beta \), which we still denote by \(\beta \) intersects \(\Sigma _n\) and \(\widetilde{\mathcal {L}}\) transversally in a finite number of points. Note that \(\widetilde{\mathcal {L}}\) is obtained as the limit of \(\Sigma _n\). Indeed, since \(\Sigma _n\) separates \(\mathcal {B}^+\) and \(\mathcal {B}^-\) in \(\widetilde{\Omega }^*\), the algebraic intersection number of \(\beta \) and \(\Sigma _n\) must be one, which implies that \(\beta \) intersects \(\Sigma _n\) in an odd number of points. Then \(\beta \) intersects \(\widetilde{\mathcal {L}}\) in an odd number of points and the claim below follows.

Claim 4.3

The curve \(\beta \) intersects \(\widetilde{\mathcal {L}}\) in an odd number of points.

In particular \(\beta \) intersects only a finite collection of leaves in \(\widetilde{\mathcal {L}}\) and we let \(\mathcal {F}\) denote the non-empty finite collection of leaves that intersect \(\beta \).

Definition 4.1

Let \((\rho _1, \widetilde{\theta }_0, t_0)\) be a fixed point in \(\widetilde{\mathcal {C}}_1\) and let \(\rho _2(\widetilde{\theta }_0, t_0)>\rho _1\) such that \((\rho _2(\widetilde{\theta }_0, t_0), \widetilde{\theta }_0, t_0)\) is in \(\widetilde{\mathcal {C}}_2\). Then we call the arc in \(\widetilde{\Omega }\) given by

$$\begin{aligned} (\rho _1+s(\rho _2-\rho _1), \widetilde{\theta }_0, t_0), \quad s\in [0,1]. \end{aligned}$$
(5)

the vertical line segment based at \((\rho _1, \widetilde{\theta }_0, t_0)\).

Claim 4.4

There exists at least one leaf \(\widetilde{L}_{\beta }\) in \(\mathcal {F}\) that intersects \(\beta \) in an odd number of points and the leaf \(\widetilde{L}_{\beta }\) must intersect each vertical line segment at least once.

Proof

The existence of \(\widetilde{L}_{\beta }\) follows because otherwise, if all the leaves in \(\mathcal {F}\) intersected \(\beta \) in an even number of points, then the number of points in the intersection \(\beta \cap \mathcal {F}\) would be even. Given \(\widetilde{L}_{\beta }\) a leaf in \(\mathcal {F}\) that intersects \(\beta \) in an odd number of points, suppose there exists a vertical line segment which does not intersect \(\widetilde{L}_{\beta }\). Then since by Claim 4.2 \(\widetilde{L}_{\beta }\) is properly embedded, using elementary separation arguments would give that the number of points of intersection in \(\beta \cap \widetilde{L}_{\beta }\) must be zero mod 2, that is even, contradicting the previous statement. \(\square \)

Let \(\Pi \) be the covering map defined in equation (2) and let \(\mathcal {P}_H:=\Pi (\widetilde{L}_{\beta })\). The previous discussion and the fact that \(\Pi \) is a local diffeomorphism, implies that \(\mathcal {P}_H\) is a stable complete H-surface embedded in \(\Omega \). Indeed, \(\mathcal {P}_H\) is a graph over its \(\theta \)-projection to \(\mathrm{Int}(\Omega )\cap \{(\rho ,0,t)\mid \rho >0, \, t\in \mathbb {R}\}\), which we denote by \(\theta (\mathcal {P}_H)\). Abusing the notation, let \(J_{\mathcal {P}_H}\) be the Jacobi function induced by taking the inner product of \(\partial _{\theta }\) with the unit normal of \(\mathcal {P}_H\), then \(J_{\mathcal {P}_H}\) is positive. Finally, since the norm of the second fundamental form of \(\mathcal {P}_H\) is uniformly bounded, standard compactness arguments imply that its closure \(\overline{\mathcal {P}}_H\) is an H-lamination \(\mathcal {L}\) of \(\Omega \), see for instance [5].

Claim 4.5

The closure of \(\mathcal {P}_H\) is an H-lamination of \(\Omega \) consisting of itself and two H-catenoids \(L_1, L_2\subset \Omega \) that form the limit set of \(\mathcal {P}_H\).

Remark 4.6

Note that these two H-catenoids are not necessarily the ones which determine \(\partial \Omega \).

Proof

Given \((\rho _1, \widetilde{\theta }_0, t_0)\in {\widetilde{\mathcal {C}}}_1\), let \(\widetilde{\gamma }\) be the fixed vertical line segment in \(\widetilde{\Omega }\) based at \((\rho _1, \widetilde{\theta }_0, t_0)\), let \(\widetilde{p}_0\) be a point in the intersection \(\widetilde{L}_\beta \cap \widetilde{\gamma }\) (recall that by Claim 4.4 such intersection is not empty) and let \(p_0=\Pi (\widetilde{p}_0)\in \Pi (\widetilde{\gamma })\cap \mathcal {P}_H\). Then, by Claim 4.4, for any \(i\in \mathbb N\), the vertical line segment \(T_{2\pi i}(\widetilde{\gamma })\) intersects \(\widetilde{L}_\beta \) in at least a point \(\widetilde{p}_i\), and \(\widetilde{p}_{i+1}\) is above \(\widetilde{p}_i\), where T is the translation defined in equation (4). Namely, \(\widetilde{p}_0=(r_0, \widetilde{\theta }_0, t_0)\), \(\widetilde{p}_i=(r_i, \widetilde{\theta }_0+2\pi i, t_0)\) and \(r_i<r_{i+1}<\rho _2(\widetilde{\theta }_0, t_0)\). The point \(\widetilde{p_i}\in \widetilde{L}_\beta \) corresponds to the point \(p_i= \Pi (\widetilde{p}_i)=(r_i, \widetilde{\theta }_0\, \text {mod}\, 2\pi , t_0)\in \mathcal {P}_H\). Let \(r(2):=\lim _{i\rightarrow \infty }r_i\) then \(r(2)\le \rho _2(\widetilde{\theta }_0,t_0)\) and note that since \(\lim _{i\rightarrow \infty }(r_{i+1}-r_i)=0\), then the value of the Jacobi function \(J_{\mathcal {P}_H}\) at \( p_i\) must be going to zero as i goes to infinity. Clearly, the point \(Q:=(r(2), \widetilde{\theta }_0 \, \text {mod}\, 2\pi , t_0)\in \Omega \) is in the closure of \(\mathcal {P}_H\), that is \(\mathcal {L}\). Let \(L_2\) be the leaf of \(\mathcal {L}\) containing Q. By the previous discussion \(J_{L_2}(Q)=0\). Since by the nature of the convergence, either \(J_{L_2}\) is positive or \(L_2\) is rotational, then \(L_2\) is rotational, namely an H-catenoid.

Arguing similarly but considering the intersection of \(\widetilde{L}_\beta \) with the vertical line segments \(T_{-2\pi i}(\widetilde{\gamma })\), \(i\in \mathbb N\), one obtains another H-catenoid \(L_1\), different from \(L_2\), in the lamination \(\mathcal {L}\). This shows that the closure of \(\mathcal {P}_H\) contains the two H-catenoids \(L_1\) and \(L_2\).

Let \(\Omega _g\) be the rotationally invariant, connected region of \(\Omega -[L_1\cup L_2]\) whose boundary contains \(L_1\cup L_2\). Note that since \(\mathcal {P}_H\) is connected and \(L_1\cup L_2\) is contained in its closure, then \(\mathcal {P}_H\subset \Omega _g\). It remains to show that \(\mathcal {L}=\mathcal {P}_H\cup L_1\cup L_2\), i.e. \(\overline{\mathcal {P}}_H-\mathcal {P}_H=L_1\cup L_2\). If \(\overline{\mathcal {P}}_H-\mathcal {P}_H\ne L_1\cup L_2\) then there would be another leaf \(L_3\in \mathcal {L}\cap \Omega _g\) and by previous argument, \(L_3\) would be an H-catenoid. Thus \(L_3\) would separate \(\Omega _g\) into two regions, contradicting that fact that \(\mathcal {P}_H\) is connected and \(L_1\cup L_2\) are contained in its closure. This finishes the proof of the claim. \(\square \)

Note that by the previous claim, \(\mathcal {P}_H\) is properly embedded in \(\Omega _g\).

Claim 4.7

The H-surface \(\mathcal {P}_H\) is simply-connected and every integral curve of \(\partial _\theta \) that lies in \(\Omega _g\) intersects \(\mathcal {P}_H\) in exactly one point.

Proof

Let \(D_g:=\mathrm{Int}(\Omega _g)\cap \{(\rho ,0,t)\mid \rho >0, \, t\in \mathbb {R}\}\), then \(\mathcal {P}_H\) is a graph over its \(\theta \)-projection to \(D_g\), that is \(\theta (\mathcal {P}_H)\). Since \(\theta :\Omega _g\rightarrow D_g\) is a proper submersion and \(\mathcal {P}_H\) is properly embedded in \(\Omega _g\), then \(\theta (\mathcal {P}_H)=D_g\), which implies that every integral curve of \(\partial _\theta \) that lies in \(\Omega _g\) intersects \(\mathcal {P}_H\) in exactly one point. Moreover, since \(D_g\) is simply-connected, this gives that \(\mathcal {P}_H\) is also simply-connected. This finishes the proof of the claim. \(\square \)

From this claim, it clearly follows that \(\Omega _g\) is foliated by H-surfaces, where the leaves of this foliation are \(L_1\), \(L_2\) and the rotated images \(\mathcal {P}_H ({\theta })\) of \(\mathcal {P}_H\) around the t-axis by angles \(\theta \in [0,2\pi )\). The existence of the examples \(\Sigma _H\) in the statement of Theorem 1.1 can easily be proven by using \(\mathcal {P}_H\). We set \(\Sigma _H=\mathcal {P}_H\), and \(C_i=L_i\) for \(i=1,2\). This finishes the proof of Theorem 1.1.

5 Appendix: Disjoint H-catenoids

In this section, we will show the existence of disjoint H-catenoids in \({\mathbb H}^2\times {\mathbb R}\). In particular, we will prove Lemma 2.1. Given \(H\in (0,\frac{1}{2})\) and \(d\in [-2H,\infty )\), recall that \(\eta _d=\cosh ^{-1}(\frac{2dH+\sqrt{1-4H^2+d^2}}{1-4H^2})\) and that \(\lambda _d:[\eta _d,\infty )\rightarrow [0,\infty )\) is the function defined as follows.

$$\begin{aligned} \lambda _d(\rho )= \int ^{\rho }_{\eta _d} \frac{d+2H\cosh r}{\sqrt{\sinh ^2 r - ( d+2H\cosh r)^2}}dr. \end{aligned}$$
(6)

Recall that \(\lambda _d(\rho )\) is a monotone increasing function with \(\lim _{\rho \rightarrow \infty }\lambda _d(\rho )= \infty \) and that \(\lambda '_d(\eta _d)=\infty \) when \(d\in (-2H,\infty )\). The H-catenoid \(\mathcal {C}^H_d\), \( d\in (-2H,\infty )\), is obtained by rotating a generating curve \(\widehat{\lambda }_d(\rho )\) about the t-axis. The generating curve \(\widehat{\lambda }_d \) is obtained by doubling the curve \((\rho , 0, \lambda _d(\rho ))\), \(\rho \in [\eta _d,\infty )\), with its reflection \((\rho , 0, -\lambda _d(\rho ))\), \(\rho \in [\eta _d,\infty )\).

Finally, recall that \(b_{d}(t):=\lambda _d^{-1}(t)\) for \(t\ge 0\), hence \(b_d(0)=\eta _{d}\), and that abusing the notation \(b_d(t):=b_d(-t)\) for \(t\le 0\).

Lemma 2.1 (Disjoint H-catenoids) Given \(d_1>2\) there exist \(d_0>d_1\) and \(\delta _0>0\) such that for any \(d_2\in [d_0,\infty )\) and \(t>0\) then

$$\begin{aligned} \inf _{t\in {\mathbb {R}}}( b_{d_2}(t)-b_{d_1}(t))\ge \delta _0. \end{aligned}$$

In particular, the corresponding H-catenoids are disjoint, i.e., \(\mathcal {C}^H_{d_1}\cap \mathcal {C}^H_{d_2}=\emptyset \).

Moreover, \(b_{d_2}(t)-b_{d_1}(t)\) is decreasing for \(t>0\) and increasing for \(t<0\). In particular,

$$\begin{aligned} \sup _{t\in {\mathbb {R}}}( b_{d_2}(t)-b_{d_1}(t))=b_{d_2}(0)-b_{d_1}(0)= \eta _{d_2}-\eta _{d_1}. \end{aligned}$$

Proof

We begin by introducing the following notations that will be used for the computations in the proof of this lemma,

$$\begin{aligned} c:=\cosh r=\frac{e^r+e^{-r}}{2},\, s:=\sinh r=\frac{e^r-e^{-r}}{2}. \end{aligned}$$

Recall that \(c^2-s^2=1\) and \(c-s=e^{-r}\). Using these notations,

$$\begin{aligned} \lambda _d(\rho )= \int ^{\rho }_{\eta _d} \frac{d+2H\cosh r}{\sqrt{\sinh ^2 r - ( d+2H\cosh r)^2}}\ dr \end{aligned}$$
(7)

can be rewritten as

$$\begin{aligned} \lambda _d(\rho )= \int ^{\rho }_{\eta _d} \frac{d+2H(s+e^{-r})}{\sqrt{s^2 - ( d+2Hc)^2}}\ dr=f_d(\rho )+J_d(\rho ), \end{aligned}$$
(8)

where

$$\begin{aligned} f_d(\rho )=\int ^{\rho }_{\eta _d} \frac{2Hs}{\sqrt{s^2 - ( d+2Hc)^2}} \ dr \ \ \text{ and } \ \ J_d(\rho )= \int ^{\rho }_{\eta _d} \frac{d+2He^{-r}}{\sqrt{s^2 - ( d+2Hc)^2}} \ dr \end{aligned}$$

First, by using a series of substitutions, we will get an explicit description of \(f_d(\rho )\). Then, we will show that for \(d>2\), \(J_d(\rho )\) is bounded independently of \(\rho \) and d.

Claim 4.8

$$\begin{aligned} f_d(\rho )=\frac{2H }{\sqrt{1-4H^2}}\cosh ^{-1} \left( \frac{(1-4H^2)\cosh \rho - 2dH }{\sqrt{d^2+1-4H^2}} \right) . \end{aligned}$$
(9)

Remark 4.9

After finding \(f_d(\rho )\), we used Wolfram Alpha to compute the derivative of \(f_d(\rho )\) and verify our claim. For the sake of completeness, we give a proof.

Proof of Claim 4.8

The proof is a computation with requires several integrations by substitution. Consider

$$\begin{aligned} \int \frac{2Hs}{\sqrt{s^2 - ( d+2Hc)^2}} \ dr \end{aligned}$$

By using the fact that \(s^2=c^2-1\) and applying the substitution \(\{u=c,du=\frac{dc}{dr}dr=sdr\}\) we obtain that

$$\begin{aligned} \int \frac{2Hs}{\sqrt{s^2 - ( d+2Hc)^2}} \ dr=\int \frac{2H}{\sqrt{u^2-1 - ( d+2Hu)^2}}\ du. \end{aligned}$$

Note that

$$\begin{aligned} \begin{aligned}&u^2-1 - ( d+2Hu)^2=u^2-1-(d^2+4dHu+4H^2u^2)\\&=(1-4H^2)u^2-4dHu-d^2-1\\&=(1-4H^2)\left( u^2-\frac{4dH}{1-4H^2}u+\frac{4d^2H^2}{(1-4H^2)^2}\right) -\frac{4d^2H^2}{1-4H^2}-d^2-1\\&=(1-4H^2)\left[ \left( u-\frac{2dH}{(1-4H^2)}\right) ^2-\left( \frac{4d^2H^2}{(1-4H^2)^2}+\frac{d^2+1}{1-4H^2}\right) \right] \\&=(1-4H^2)\left[ \left( u-\frac{2dH}{(1-4H^2)}\right) ^2-\left( \frac{4d^2H^2+(1-4H^2)(d^2+1)}{(1-4H^2)^2}\right) \right] \\&=(1-4H^2)\left[ \left( u-\frac{2dH}{(1-4H^2)}\right) ^2-\left( \frac{d^2+1-4H^2}{(1-4H^2)^2}\right) \right] . \end{aligned} \end{aligned}$$

Therefore, by applying a second substitution, \(\{w=u-\frac{2dH}{(1-4H^2)}, dw=du\}\), and letting \(a^2=(\frac{d^2+1-4H^2}{(1-4H^2)^2})\) we get that

$$\begin{aligned} \int \frac{2H}{\sqrt{u^2-1 - ( d+2Hu)^2}}\ du=\int \frac{2H}{\sqrt{1-4H^2}\sqrt{w^2-a^2}}\ dw \end{aligned}$$

By using the fact that \(\sec ^2x-1=\tan ^2x\) and applying a third substitution, \(\{w=a\sec t, dw=a\sec t\tan t dt\}\), we obtain that

$$\begin{aligned} \begin{aligned} \int \frac{2Ha\sec t\tan t}{\sqrt{1-4H^2}\sqrt{a^2\sec ^2 t-a^2}} dt&=\int \frac{2H\sec t}{\sqrt{1-4H^2}}\ dt\\&=\frac{2H }{\sqrt{1-4H^2}}\ln |\sec t+\tan t| \end{aligned} \end{aligned}$$

Therefore

$$\begin{aligned} \begin{aligned} \int \frac{2H}{\sqrt{1-4H^2}\sqrt{w^2-a^2}}\ dw&=\frac{2H }{\sqrt{1-4H^2}}\ln |\frac{w}{a}+\sqrt{\frac{w^2}{a^2}-1}|\\&=\frac{2H }{\sqrt{1-4H^2}}\cosh ^{-1}\left( \frac{w}{a}\right) \end{aligned} \end{aligned}$$

Since \(w=u-\frac{2dH}{(1-4H^2)}\) then

$$\begin{aligned} \begin{aligned} \int \frac{2H}{\sqrt{u^2-1 - ( d+2Hu)^2}}\ du&=\frac{2H }{\sqrt{1-4H^2}}\cosh ^{-1}\left( \frac{u-\frac{2dH}{(1-4H^2)}}{a}\right) \\&=\frac{2H }{\sqrt{1-4H^2}}\cosh ^{-1}\left( \frac{u-\frac{2dH}{(1-4H^2)}}{\frac{\sqrt{d^2+1-4H^2}}{(1-4H^2)}}\right) \\&=\frac{2H }{\sqrt{1-4H^2}}\cosh ^{-1}\left( \frac{(1-4H^2)u- 2dH }{\sqrt{d^2+1-4H^2}} \right) . \end{aligned} \end{aligned}$$

Finally, since \(u=\cosh r\)

$$\begin{aligned} \begin{aligned} \int ^{\rho }_{\eta _d} \frac{2Hs}{\sqrt{s^2 - ( d+2Hc)^2}}&=\frac{2H }{\sqrt{1-4H^2}}\cosh ^{-1}\left( \frac{(1-4H^2)\cosh r- 2dH }{\sqrt{d^2+1-4H^2}} \right) \bigg |_{\eta _d}^\rho \\&= \frac{2H }{\sqrt{1-4H^2}}\left( \cosh ^{-1}\left( \frac{(1-4H^2)\cosh \rho - 2dH }{\sqrt{d^2+1-4H^2}} \right) \right. \\&\left. -\,\cosh ^{-1}\left( \frac{(1-4H^2)\cosh \eta _d- 2dH }{\sqrt{d^2+1-4H^2}} \right) \right) \end{aligned} \end{aligned}$$

Recall that \(\eta _d=\cosh ^{-1} (\frac{2dH+\sqrt{1-4H^2+d^2}}{1-4H^2})\) and thus

$$\begin{aligned} \frac{(1-4H^2)\cosh \eta _d- 2dH }{\sqrt{d^2+1-4H^2}}= \frac{(1-4H^2)(\frac{2dH+\sqrt{1-4H^2+d^2}}{1-4H^2})- 2dH }{\sqrt{d^2+1-4H^2}}=1. \end{aligned}$$

This implies that

$$\begin{aligned} f_d(\rho )=\frac{2H }{\sqrt{1-4H^2}}\cosh ^{-1} \left( \frac{(1-4H^2)\cosh \rho - 2dH }{\sqrt{d^2+1-4H^2}} \right) . \end{aligned}$$

\(\square \)

By Claim 4.8 we have that

$$\begin{aligned} \begin{aligned} f_d(\rho )&=\frac{2H }{\sqrt{1-4H^2}}\left( \cosh ^{-1} \frac{(1-4H^2)\cosh \rho - 2dH }{\sqrt{d^2+1-4H^2}} \right) \\&= \frac{2H }{\sqrt{1-4H^2}}\left( \rho +\ln \frac{1-4H^2 }{\sqrt{d^2+1-4H^2}}\right) +g_d(\rho ), \end{aligned} \end{aligned}$$

where \(\lim _{\rho \rightarrow \infty }g_d(\rho )=0\).

Recall that \(\lambda _d(\rho )= f_d(\rho )+J_d(\rho )\) where

$$\begin{aligned} J_d(\rho )= \int ^{\rho }_{\eta _d} \frac{d+2He^{-r}}{\sqrt{s^2 - ( d+2Hc)^2}}\ dr =\int ^{\rho }_{\eta _d} \frac{d+2He^{-r}}{\sqrt{c^2-1 - ( d+2Hc)^2}}\ dr. \end{aligned}$$

Claim 4.10

$$\begin{aligned} \sup _{d\in (2,\infty ),\rho \in (\eta _d,\infty )}J_d(\rho )\le \pi \sqrt{1-2H}. \end{aligned}$$

Proof of Claim 4.10

Let

$$\begin{aligned} \alpha =\dfrac{2dH+\sqrt{1-4H^2+d^2}}{1-4H^2}\,\text { and }\,\beta =\dfrac{2dH-\sqrt{1-4H^2+d^2}}{1-4H^2} \end{aligned}$$

be the roots of \(c^2-1 - ( d+2Hc)^2\), i.e.

$$\begin{aligned} \begin{aligned} c^2-1 - ( d+2Hc)^2&= (1-4H^2)\left( c^2-\frac{4dH}{1-4H^2}c-\frac{1+d^2}{1-4H^2}\right) \\&=(1-4H^2)(c-\alpha )(c-\beta ). \end{aligned} \end{aligned}$$

Note that \(\alpha =\cosh {\eta _d}\) and that as \(H\in (0,\frac{1}{2})\), \(\beta<0<\alpha \). Furthermore, \(2He^{-r}<2H<1<d\). Thus we have,

$$\begin{aligned} \begin{aligned} J_d(\rho )&=\int ^{\rho }_{\eta _d} \frac{d+2He^{-r}}{\sqrt{1-4H^2}\sqrt{(c-\alpha )(c-\beta )}}dr \\&< \frac{2d}{\sqrt{1-4H^2}} \int _{\eta _d}^\infty \frac{dr}{\sqrt{(c-\alpha )(c-\beta )}}\\&< \frac{2d}{\sqrt{1-4H^2}\sqrt{\alpha -\beta }} \int _{\eta _d}^\infty \frac{dr}{\sqrt{c-\alpha }}, \end{aligned} \end{aligned}$$

where the last inequality holds because for \(r>\eta _d\), \(\cosh {r}>\alpha \) and thus \(\sqrt{\alpha -\beta }< \sqrt{c-\alpha }\). Notice that \(\alpha -\beta = \frac{2\sqrt{1-4H^2+d^2}}{1-4H^2}>\frac{2d}{1-4H^2}\). Therefore

$$\begin{aligned} \frac{2d}{\sqrt{1-4H^2}\sqrt{\alpha -\beta }}<\frac{2d}{\sqrt{1-4H^2}}\frac{\sqrt{1-4H^2}}{\sqrt{2d}}=\sqrt{2d} \end{aligned}$$

and

$$\begin{aligned} J_d(\rho )< \sqrt{2d} \int _{\eta _d}^\infty \frac{dr}{\sqrt{c-\alpha }}. \end{aligned}$$

Applying the substitution \(\{ u=c-\alpha , du=sdr=\sqrt{(u+\alpha )^2-1}dr\}\), we obtain that

$$\begin{aligned} \int _{\eta _d}^\infty \frac{dr}{\sqrt{c-\alpha }}= \int _0^\infty \frac{du}{\sqrt{u}\sqrt{(u+\alpha )^2-1}} \end{aligned}$$
(10)

Let \(\omega =\alpha -1\). Note that since \(d\ge 1\) then \(\alpha >1\) and we have that \((u+\alpha )^2-1> (u+\omega )^2\) as \(u>0\). This gives that

$$\begin{aligned} \int _0^\infty \frac{du}{\sqrt{u}\sqrt{(u+\alpha )^2-1}}< \int _0^\infty \frac{du}{\sqrt{u}(u+\omega )} \end{aligned}$$

Applying the substitution \(\{ v=\sqrt{u}, dv=\dfrac{du}{2\sqrt{u}}\}\) we get

$$\begin{aligned} \int _0^\infty \frac{du}{\sqrt{u}(u+\omega )}= \int _0^\infty \frac{2dv}{v^2+\omega }= \ \frac{2}{\sqrt{\omega }}\arctan {\frac{w}{\sqrt{\omega }}}\bigg |_0^\infty <\dfrac{\pi }{\sqrt{\omega }} \end{aligned}$$

and thus

$$\begin{aligned} J_d(\rho )< \sqrt{\frac{2d}{\omega }} \pi . \end{aligned}$$

Note that

$$\begin{aligned} \begin{aligned} \omega =\alpha -1&= \frac{2dH+\sqrt{1-4H^2+d^2}}{1-4H^2}-1\\&>\frac{(1+2H)d}{1-4H^2}-1=\frac{d}{1-2H}-1. \end{aligned} \end{aligned}$$

Since \(d>2\), we have \(2\omega >\dfrac{d}{1-2H}\) and \(\dfrac{d}{\omega }<2(1-2H)\). Then \( \sqrt{\dfrac{2d}{\omega }}<2\sqrt{1-2H}. \)

Finally, this gives that

$$\begin{aligned} J_d(\rho )<2\pi \sqrt{1-2H} \end{aligned}$$

independently on \(d>2\) and \(\rho >\eta _d\). This finishes the proof of the claim. \(\square \)

Using Claims 4.8 and 4.10, we can now prove the next claim.

Claim 4.11

Given \(d_2>d_1>2\) there exists \(T\in \mathbb R\) such for any \(t>T\), we have that

$$\begin{aligned} \begin{aligned} \frac{2H }{\sqrt{1-4H^2}}&(\lambda _{d_2}^{-1}(t)-\lambda _{d_1}^{-1}(t)) \\ >&\frac{1}{2} \ln \sqrt{\frac{d_2^2+1-4H^2 }{d_1^2+1-4H^2}}-2\pi \sqrt{1-2H}. \end{aligned} \end{aligned}$$

Proof of Claim 4.11

Recall that \(\lambda _d(\rho )=f_d(\rho )+J_d(\rho )\) and that by Claims 4.8 and 4.10 we have that

$$\begin{aligned} f_d(\rho )= \frac{2H }{\sqrt{1-4H^2}}\left( \rho +\ln \frac{1-4H^2 }{\sqrt{d^2+1-4H^2}}\right) +g_d(\rho ), \end{aligned}$$
(11)

where \(\lim _{\rho \rightarrow \infty }g_d(\rho )=0\), and that

$$\begin{aligned} \sup _{d\in (2,\infty ),\rho \in (\eta _d,\infty )}J_d(\rho )\le 2\pi \sqrt{1-2H} . \end{aligned}$$
(12)

Let \(\rho _i(t):=\lambda _{d_i}^{-1}(t)\), \(i=1,2\). Using this notation, since \(t=\lambda _1(\rho _1(t))=\lambda _2(\rho _2(t))\) we obtain that

$$\begin{aligned} \begin{aligned} 0&=\lambda _2(\rho _2(t))-\lambda _1(\rho _1(t))\\&=f_{d_2}(\rho _2(t))+J_{d_2}(\rho _2(t)) -f_{d_1}(\rho _1(t))-J_{d_1}(\rho _1(t))\\&=\frac{2H }{\sqrt{1-4H^2}}\left( \rho _2(t) +\ln \frac{1-4H^2 }{\sqrt{d_2^2+1-4H^2}}\right) +g_{d_2}(\rho _2(t))+J_{d_2}(\rho _2(t))\\&\quad -\frac{2H }{\sqrt{1-4H^2}}\left( \rho _1(t) -\ln \frac{1-4H^2 }{\sqrt{d_1^2+1-4H^2}}\right) -g_{d_1}(\rho _1(t))-J_{d_1}(\rho _1(t)) \end{aligned} \end{aligned}$$

Recall that \(\lim _{t\rightarrow \infty }\rho _i(t)=\infty \), \(i=1,2\), therefore given \(\varepsilon >0\) there exists \(T_\varepsilon \in \mathbb R\) such that for any \(t>T_\varepsilon \), \(|g_{d_i}(\rho _i(t))|\le \varepsilon \). Taking

$$\begin{aligned} 4\varepsilon <\ln \sqrt{\frac{d_2^2+1-4H^2 }{d_1^2+1-4H^2}} \end{aligned}$$

for \(t>T_\varepsilon \) we get that

$$\begin{aligned} \begin{aligned} \frac{2H }{\sqrt{1-4H^2}}&(\rho _2(t)-\rho _1(t)) \\&>\ln \sqrt{\frac{d_2^2+1-4H^2 }{d_1^2+1-4H^2}}+J_{d_1}(\rho _1(t))-J_{d_2}(\rho _2(t))-2\varepsilon \\&>\frac{1}{2} \ln \sqrt{\frac{d_2^2+1-4H^2 }{d_1^2+1-4H^2}}+J_{d_1}(\rho _1(t))-J_{d_2}(\rho _2(t)). \end{aligned} \end{aligned}$$

Notice that \(J_{d_1}(\rho _1(t))>0\) and that Claim 4.10 gives that

$$\begin{aligned} \sup _{\rho \in (\eta _{d_2},\infty )}J_{d_2}(\rho )\le 2\pi \sqrt{1-2H}. \end{aligned}$$

Therefore

$$\begin{aligned} \begin{aligned} \frac{2H }{\sqrt{1-4H^2}}&(\rho _2(t)-\rho _1(t)) \\&> \frac{1}{2} \ln \sqrt{\frac{d_2^2+1-4H^2 }{d_1^2+1-4H^2}}-2\pi \sqrt{1-2H}. \end{aligned} \end{aligned}$$

This finishes the proof of the claim. \(\square \)

We can now use Claim 4.11 to finish the proof of the lemma. Given \(d_1>2\) fix \(d_0>d_1\) such that

$$\begin{aligned} \begin{aligned} \frac{\sqrt{1-4H^2}}{4H} \left( \ln \sqrt{\frac{d_0^2+1-4H^2 }{d_1^2+1-4H^2}}-4\pi \sqrt{1-2H}\right) =1. \end{aligned} \end{aligned}$$

Then, by Claim 4.11, given \(d_2\ge d_0\) there exists \(T>0\) such that \(\lambda _{d_2}^{-1}(t)-\lambda _{d_1}^{-1}(t)>1\) for any \(t>T\). Notice that since for any \(\rho \in (\eta _2,\infty )\), \(\lambda '_{d_2}(\rho )>\lambda '_{d_1}(\rho )\), then there exists at most one \(t_0>0\) such that \(\lambda _{d_2}^{-1}(t_0)-\lambda _{d_1}^{-1}(t_0)=0\). Therefore, since there exists \(T>0\) such that \(\lambda _{d_2}^{-1}(t)-\lambda _{d_1}^{-1}(t)>1\) for any \(t>T\) and \(\lambda _{d_2}^{-1}(0)-\lambda _{d_1}^{-1}(0)=\eta _{d_2}-\eta _{d_1}>0\), this implies that there exists a constant \(\delta (d_2)>0\) such that for any \(t>0\),

$$\begin{aligned} \lambda _{d_2}^{-1}(t)-\lambda _{d_1}^{-1}(t)>\delta (d_2). \end{aligned}$$

A priori it could happen that \(\lim _{d_2\rightarrow \infty }\delta (d_2)=0\). The fact that \(\lim _{d_2\rightarrow \infty }\delta (d_2)>0\) follows easy by noticing that by applying Claim 4.11 and using the same arguments as in the previous paragraph there exists \(d_3> d_0\) such that for any \(d\ge d_3\) and \(t>0\),

$$\begin{aligned} \lambda _{d}^{-1}(t)-\lambda _{d_0}^{-1}(t)>0. \end{aligned}$$

Therefore, for any \(d\ge d_3\) and \(t>0\),

$$\begin{aligned} \lambda _{d}^{-1}(t)-\lambda _{d_1}^{-1}(t)> \lambda _{d_0}^{-1}(t)-\lambda _{d_1}^{-1}(t)>\delta (d_0) \end{aligned}$$

which implies that

$$\begin{aligned} \lim _{d_2\rightarrow \infty }\delta (d_2)\ge \delta (d_0)>0. \end{aligned}$$

Setting \(\delta _0=\inf _{d\in [d_0,\infty )}\delta (d_2)>0\) gives that

$$\begin{aligned} \inf _{t\in \mathbb R_{\ge 0}}( \lambda _{d_2}^{-1}(t)-\lambda _{d_1}^{-1}(t))\ge \delta _0. \end{aligned}$$

By definition of \(b_d(t)\) then

$$\begin{aligned} \inf _{t\in \mathbb R}( b_{d_2}(t)-b_{d_1}(t))= \inf _{t\in \mathbb R_{\ge 0}}( \lambda _{d_2}^{-1}(t)-\lambda _{d_1}^{-1}(t))\ge \delta _0. \end{aligned}$$

It remains to prove that \(b_2(t)-b_1(t)\) is decreasing for \(t>0\) and increasing for \(t<0\). By definition of \(b_d(t)\), it suffices to show that \(b_2(t)-b_1(t)\) is decreasing for \(t>0\). We are going to show \(\frac{d}{dt}(b_2(t)-b_1(t))<0\) when \(t>0\).

By definition of \(b_i\), for \(t>0\) we have that \(\lambda _i(b_i(t))=t\) and thus \(b_i'(t)=\frac{1}{\lambda _i'(b_i(t))}\). By definition of \(\lambda _d(t)\) for \(t>0\) the following holds,

$$\begin{aligned} b_1'(t)=\frac{1}{\lambda _1'(b_1(t))}>\frac{1}{\lambda _1'(b_2(t))}>\frac{1}{\lambda _2'(b_2(t))}=b_2'(t). \end{aligned}$$

The first inequality is due to the convexity of the function \(\lambda _1(t)\) and the second inequality is due to the fact that \(\lambda _1'(\rho )<\lambda _2'(\rho )\) for any \(\rho >\eta _2\). This proves that \(\frac{d}{dt}(b_2(t)-b_1(t))=b_2'(t)-b_1'(t)<0\) for \(t>0\) and finishes the proof of the claim.

\(\square \)

Note that if d is sufficiently close to \(-2H\) then \(\mathcal {C}^H_d\) must be unstable. This follows because as d approaches \(-2H\), the norm of the second fundamental form of \(\mathcal {C}^H_d\) becomes arbitrarily large at points that approach the “origin” of \({\mathbb H}^2\times {\mathbb R}\) and a simple rescaling argument gives that a sequence of subdomains of \(\mathcal {C}^H_d\) converge to a catenoid, which is an unstable minimal surface. This observation, together with our previous lemma suggests the following conjecture.

Conjecture: Given \(H\in (0,\frac{1}{2})\) there exists \(d_H>-2H\) such that the following holds. For any \(d>d'>d_H\), \(\mathcal {C}^H_{d}\cap \mathcal {C}^H_{d'}=\emptyset \), and the family \(\{\mathcal {C}^H_d \mid d\in [d_H,\infty )\}\) gives a foliation of the closure of the non-simply-connected component of \({\mathbb H}^2\times {\mathbb R}-\mathcal {C}^H_{d_H}\). The H-catenoid \(\mathcal {C}^H_d\) is unstable if \(d\in (-2H, d_H)\) and stable if \(d\in (d_H,\infty )\). The H-catenoid \(\mathcal {C}^H_{d_H}\) is a stable-unstable catenoid.