Skip to main content
Log in

On Rigidity and Convergence of Circle Patterns

  • Published:
Discrete & Computational Geometry Aims and scope Submit manuscript

Abstract

Two planar embedded circle patterns with the same combinatorics and the same intersection angles can be considered to define a discrete conformal map. We show that two locally finite circle patterns covering the unit disc are related by a hyperbolic isometry. Furthermore, we prove an analogous rigidity statement for the complex plane if all exterior intersection angles of neighboring circles are uniformly bounded away from 0. Finally, we study a sequence of two circle patterns with the same combinatorics each of which approximates a given simply connected domain. Assume that all kites are convex and all angles in the kites are uniformly bounded and the radii of one circle pattern converge to 0. Then a subsequence of the corresponding discrete conformal maps converges to a Riemann map between the given domains.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  1. Ahlfors, L.V.: Lectures on Quasiconformal Mappings. D. Van Nostrand Co., Toronto (1966)

    MATH  Google Scholar 

  2. Bobenko, A.I., Springborn, B.A.: Variational principles for circle patterns and Koebe’s theorem. Trans. Am. Math. Soc. 356(2), 659–689 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  3. Brightwell, G.R., Scheinerman, E.R.: Representations of planar graphs. SIAM J. Discrete Math. 6(2), 214–229 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bücking, U.: Approximation of conformal mappings by circle patterns. Geom. Dedic. 137, 163–197 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  5. Cannon, J.W.: The combinatorial Riemann mapping theorem. Acta Math. 173(2), 155–234 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  6. Gurel-Gurevich, O., Nachmias, A.: Recurrence of planar graph limits. Ann. Math. 177(2), 761–781 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  7. He, Z.-X.: Rigidity of infinite disk patterns. Ann. Math. 149(1), 1–33 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  8. He, Z.-X., Schramm, O.: Hyperbolic and parabolic packings. Discrete Comput. Geom. 14(2), 123–149 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  9. He, Z.-X., Schramm, O.: On the convergence of circle packings to the Riemann map. Invent. Math. 125(2), 285–305 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  10. He, Z.-X., Schramm, O.: The \(C^\infty \)-convergence of hexagonal disk packings to the Riemann map. Acta Math. 180(2), 219–245 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  11. Lan, S.-Y., Dai, D.-Q.: The \(C^\infty \)-convergence of SG circle patterns to the Riemann mapping. J. Math. Anal. Appl. 332(2), 1351–1364 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  12. Lehto, O., Virtanen, K.I.: Quasiconformal Mappings in the Plane. Springer, New York (1973)

    Book  MATH  Google Scholar 

  13. Lewin, L.: Polylogarithms and Associated Functions. North Holland Publishing, New York (1981)

    MATH  Google Scholar 

  14. Lyons, R., Peres, Y.: Probability on Trees and Networks. Cambridge Series in Statistical and Probabilistic Mathematics, vol. 42. Cambridge University Press, Cambridge (2016)

    Book  MATH  Google Scholar 

  15. Matthes, D.: Convergence in discrete Cauchy problems and applications to circle patterns. Conform. Geom. Dyn. 9, 1–23 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  16. Rivin, I.: Euclidean structures on simplicial surfaces and hyperbolic volume. Ann. Math. 139(3), 553–580 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  17. Rodin, B., Sullivan, D.: The convergence of circle packings to the Riemann mapping. J. Differ. Geom. 26(2), 349–360 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  18. Schramm, O.: How to cage an egg. Invent. Math. 107(3), 543–560 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  19. Schramm, O.: Square tilings with prescribed combinatorics. Isr. J. Math. 84(1–2), 97–118 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  20. Schramm, O.: Circle patterns with the combinatorics of the square grid. Duke Math. J. 86(2), 347–389 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  21. Springborn, B.A.: Variational Principles for Circle Patterns, Ph.D. thesis, Technische Universität Berlin, 2003, published online at http://opus.kobv.de/tuberlin/volltexte/2003/668/

  22. Stephenson, K.: Introduction to Circle Packing: The Theory of Discrete Analytic Functions. Cambridge University Press, Cambridge (2005)

    MATH  Google Scholar 

  23. Thurston, W.: The finite Riemann mapping theorem. In: Invited address at the International Symposioum in Celebration of the proof of the Bieberbach Conjecture, Purdue University (1985)

  24. Werness, B.: Discrete analytic functions on non-uniform lattices without global geometric control. http://arxiv.org/abs/1511.01209 (2015)

  25. Woess, W.: Random Walks on Infinite Graphs and Groups. Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge (2000)

    Book  MATH  Google Scholar 

Download references

Acknowledgements

The author is grateful to Boris Springborn for useful discussions and advice. The author would also like to thank the anonymous referees for their valuable comments which helped to improve the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ulrike Bücking.

Additional information

Editor in Charge: János Pach

This research was supported by the DFG Collaborative Research Center TRR 109 “Discretization in Geometry and Dynamics”.

Appendix A: Proof of Lemma 4.1

Appendix A: Proof of Lemma 4.1

In this appendix we prove Lemma 4.1 by suitably adapting the proof of Lemma 6.1 in [7] using estimates of Sect. 5.1.

Let \({\mathscr {C}}\) be an infinite circle pattern. Denote

$$\begin{aligned} \tau (v)=\frac{r(v)}{d(0,B(v))}\in (0,\infty ]\quad \text {for } v\in V\qquad \mathrm{and}\qquad \tau ({{\mathscr {C}}})= \limsup _{v_k\rightarrow \infty } \tau (v_k)\in [0,\infty ], \end{aligned}$$

where we have arranged the vertices of V into a sequence \((v_k)\). Note that if \({\mathscr {C}}\) is locally finite and \(f:{\mathbb {C}}\rightarrow {\mathbb {C}}\) is a similarity then \(\tau (f({{\mathscr {C}}}))= \tau ({{\mathscr {C}}})\).

Let \(v_0\in V\). Without loss of generality we may assume that the centers \(c(v_0)=0\) and \({\widetilde{c}}(v_0)=0\) are placed at the origin. Furthermore we may assume that \(r(v_0)={\widetilde{r}}(v_0)\) by suitable scaling. Let \(v_1\in V\) be another vertex.

Case (i): Assume that \(\tau ({{\mathscr {C}}})<\infty \). Then there exists a constant \(\delta >3\) and a finite subset \(V_0\subset V\), containing \(v_0\) and \(v_1\) such that for all \(v\in V{\setminus } V_0\) there holds

$$\begin{aligned} \frac{r(v)}{d(0,B(v))} =\tau (v)\le \frac{\delta }{3}. \end{aligned}$$

Let \(R_0>0\) be large enough so that \(B(v)\subset \mathbb {B}(R_0)\) for all \(v\in V_0\). Then for all \(R>R_0\) we have \(V_{\mathbb {c}(R)}\cap V_{\mathbb {c}(\delta R)} =\emptyset \) as \(\tau (v)\le \delta /3\).

Set \(R_j=\delta ^jR_0\) and \(V_j=V_{\mathbb {c}(R_j)}\) for \(j\ge 1\). Then for all \(0\le i_1<i_2<i_3\) the set \(V_{i_2}\) separates \( V_{i_1}\) and \(V_{i_3}\) and \(\textsc {Vel}(V_2,V_{2k-1})\ge \sum _{j=1}^{k-1} \textsc {Vel}(V_{2j},V_{2j+1})\) by Lemma 5.6. Furthermore, Lemma 5.7 implies that \(\textsc {Vel}(V_{2j},V_{2j+1})\ge C_2\) and thus \(\textsc {Vel}(V_2,V_{2k-1})\ge (k-1) C_2\). Choose \(k=2+\lceil C_6/C_2 \rceil \), where \(\lceil x \rceil \) is the smallest integer \(\ge x\). Then \(\textsc {Vel}(V_2,V_{2k-1})> C_6\). Now we apply Lemma 5.11 to the circle pattern \(\widetilde{\mathscr {C}}\). So there is \({\widetilde{R}}>0\) such that for all \({\widetilde{\rho }}\in [{\widetilde{R}},2{\widetilde{R}}]\) the set \({\widetilde{V}}_{\mathbb {c}({\tilde{\rho }})}= \{v\in V: {\widetilde{S}}(v)\cap \mathbb {c}({\widetilde{\rho }}) \not =\emptyset \}\) separates \(V_2\) and \(V_{2k-1}\). Then \({\widetilde{V}}_{\mathbb {c}({\tilde{\rho }})}\) also separates \(V_1\), and \(V_{2k}\) and \({\widetilde{V}}_{\mathbb {c}({\tilde{\rho }})}\cap V_1 \subset V_2\cap V_1 =\emptyset \) as well as \({\widetilde{V}}_{\mathbb {c}({\tilde{\rho }})}\cap V_{2k} \subset V_{2k-1}\cap V_{2k} =\emptyset \). This implies that \({\widetilde{B}}(v)\subset \mathbb {B}_{\mathrm{int}}({\widetilde{\rho }})\) for all \(v\in V_1\) and \({\widetilde{B}}(v)\subset {\mathbb {C}}{\setminus }\mathbb {B}_{\mathrm{int}}(2{\widetilde{\rho }})\) for all \(v\in V_{2k}\).

Consider the subgraph \(G_1\) of G consisting of all vertices v such that \(B(v)\cap \mathbb {B}(R_1)\not =\emptyset \). Then \({\widetilde{B}}(v)\subset \mathbb {B}({\widetilde{R}})\subset \mathbb {B}(2{\widetilde{R}})\). Denote by \(r_{\mathrm{hyp}}(B)\subset {\mathbb {D}}\) the hyperbolic radius of the disc B. We deduce from Lemma 3.9 that

$$\begin{aligned} r_{\mathrm{hyp}}\left( \frac{1}{2{\widetilde{R}}}{\widetilde{B}}(v)\right) \le r_{\mathrm{hyp}}\left( \frac{1}{{R_1}}{B}(v)\right) \end{aligned}$$

holds for all vertices v of \(G_1\).

Similarly, consider the subgraph \(G_2\) of G consisting of all vertices v such that \({\widetilde{B}}(v)\cap \mathbb {B}(2{\widetilde{R}})\not =\emptyset \). Then B(v) is contained in the interior of \(\mathbb {B}({R}_{2k})\) and for all vertices v of \(G_2\) we deduce from Lemma 3.9 that

$$\begin{aligned} r_{\mathrm{hyp}}\left( \frac{1}{2{\widetilde{R}}}{\widetilde{B}}(v)\right) \ge r_{\mathrm{hyp}}\left( \frac{1}{{R_{2k}}}{B}(v)\right) . \end{aligned}$$

As the discs \(\frac{1}{2{\widetilde{R}}}{\widetilde{B}}(v),\frac{1}{{R_1}}{B}(v), \frac{1}{2{\widetilde{R}}}{\widetilde{B}}(v),\frac{1}{{R_{2k}}}{B}(v)\) are all contained in \(\frac{1}{2}{\mathbb {D}}\) the hyperbolic and Euclidean radii are comparable. In particular, there exists an absolute constant \({\widehat{C}}_{0}>0\) such that

$$\begin{aligned} \frac{1}{2\tilde{R}}\tilde{r}(v_j)\le \hat{C}_{0} \frac{1}{{R_1}}{r}(v_j)\quad \text {and}\quad \hat{C}_{0} \frac{1}{2\tilde{R}}\tilde{r}(v_j) \ge \frac{1}{{R_{2k}}}{r}(v_j)= \frac{\delta ^{-2k+1}}{R_1} r(v_j) \end{aligned}$$

hold for \(j=0,1\). As \(r(v_0)={\widetilde{r}}(v_0)\) we see that \(\frac{1}{2{\widetilde{R}}}\le {\widehat{C}}_{0} \frac{1}{{R_1}}\) and \({\widehat{C}}_{0} \frac{1}{2{\widetilde{R}}}\ge \frac{1}{{R_{2k}}}\), therefore

$$\begin{aligned} C={\widehat{C}}_{0}^2\delta ^{2k-1} \ge {\widehat{C}}_{0}\frac{2{\widetilde{R}}}{{R_1}} \ge \frac{{\widetilde{r}}(v_1)}{r(v_1)} \ge \delta ^{-2k+1} \frac{2{\widetilde{R}}}{{\widehat{C}}_{0}{R_1}} \ge \frac{\delta ^{-2k+1}}{{\widehat{C}}_{0}^2}=\frac{1}{C}. \end{aligned}$$

Case (ii): Now assume that \(\tau ({{\mathscr {C}}})=\infty \). We consider the infinite set \(W=\{v\in V: \tau (v)\ge 1\}\). Define \({\widetilde{\tau }}\) analogously as \(\tau \). We start with the following claim:

$$\begin{aligned} \limsup _{v\in W, v\rightarrow \infty } {\widetilde{\tau }}(v)>0. \end{aligned}$$
(29)

Proof of (29) Suppose the contrary, that is \(\limsup _{v\in W, v\rightarrow \infty } {\widetilde{\tau }}(v)=0\). Then for all \(0<\varepsilon <1/2\) there is a finite subset \(V_0\ni v_0\), containing also all neighbors of \(v_0\), such that for all vertices \(v\in W{\setminus } V_0=:W^{\prime }\) there holds

$$\begin{aligned} \frac{{\widetilde{r}}(v)}{d(0,{\widetilde{B}}(v))} ={\widetilde{\tau }}(v)<\varepsilon . \end{aligned}$$
(30)

Denote by \({\widehat{G}}\) the subgraph of G which contains no vertices of \(W^{\prime }\) (or edges ending in \(W^{\prime }\)). Let \(R_0>0\) be large enough so that \(B(v)\subset \mathbb {B}(R_0)\) for all \(v\in V_0\). Set \(R_j=4^j R_0\) and \(V_j=V_{\mathbb {c}(R_j)}\) for \(j\ge 1\). Then \(V_i\cap V_j\subset W^{\prime }\) for all \(i\not =j\) and for all \(0\le i_1<i_2<i_3\) the set \(V_{i_2}\) separates \( V_{i_1}\) and \(V_{i_3}\).

Denote by \(\textsc {Vel}_{{\widehat{G}}}(V_i,V_j)\) the vertex extremal length between \(V_i\cap (V{\setminus } W^{\prime })\) and \(V_j\cap (V{\setminus } W^{\prime })\) in \({\widehat{G}}\). Set \(k=2+\lceil 2C_6/C_2\rceil \). Then we obtain, as in the case (i),

$$\begin{aligned} \textsc {Vel}_{{\widehat{G}}}(V_2,V_{2k-1}) \ge \sum _{i=1}^{k-1} \textsc {Vel}_{{\widehat{G}}}(V_{2i},V_{2i+1}) \ge (k-1) C_2 > 2C_6. \end{aligned}$$

Assume that there exists \({\widetilde{R}}>0\) such that for all \(\rho \in [{\widetilde{R}},2{\widetilde{R}}]\) the set \({\widetilde{V}}_{\mathbb {c}(\rho )}\) separates \(V_2\) from \(V_{2k-1}\) in G. Then \(V_2\cap V_{2k-1} \subset {\widetilde{V}}_{\mathbb {c}({\tilde{R}})}\cap {\widetilde{V}}_{\mathbb {c}(2{\tilde{R}})}\cap W^{\prime }=\emptyset \) due to (30). As \(R_0\) is arbitrary this implies that \(\tau ({{\mathscr {C}}})\le \frac{4^{2k-3}-1}{2}<\infty \) contradicting our assumption. It now remains to prove the existence of \({\widetilde{R}}>0\). \(\square \)

Proof

(Existence of\({\widetilde{R}}>0\)) Let \(U_{2,2k-1}\) be the subset of vertices which is separated from \(V_0\) by \(V_2\) and from \(\infty \) by \(V_{2k-1}\), that is \(U_{2,2k-1}=\{v\in V: S(v)\cap \{z\in {\mathbb {C}}: R_2<|z|<R_{2k-1}\}\not = \emptyset \}\). Denote \(W^{\prime \prime }=U_{2,2k-1}\cap W\). As k is fixed, \(\tau (v)\ge 1\) for \(v\in W\) and by condition (2) the number of vertices in \(W^{\prime \prime }\) is bounded by a universal constant, say \(|W^{\prime }|\le M\).

Define \({\widetilde{R}}=\min {\{ R: {\widetilde{B}}(v)\cap \mathbb {B}(R)\not =\emptyset \text { for all } v\in V_2\}} >0\). Without loss of generality we may assume that \({\widetilde{R}}=1\). Let \(\gamma ^*\in \varGamma _G^*(V_2,V_{2k-1})\). We deduce as in the proof of Lemma 5.11 that the diameter of \({\widetilde{{\mathbb {S}}}}(\gamma ^*)\) is \(\ge {\widetilde{R}}=1\).

Assume that there exists \(\rho _1\in [1,2]\) such that \({\widetilde{V}}_{\mathbb {c}(\rho _1)}\) does not separate \(V_2\) and \(V_{2k-1}\). Thus there exists a path \(\gamma _0\) from \(V_2\) to \(V_{2k-1}\) with \(\gamma _0\cap {\widetilde{V}}_{\mathbb {c}(\rho _1)}=\emptyset \). Again as in the proof of Lemma 5.11 this implies that \({\widetilde{S}}(v_0)\) is contained in the interior of \(\mathbb {B}(\rho _1)\), so \({\widetilde{{\mathbb {S}}}}(\gamma ^*)\cap \mathbb {B}(\rho _1)\supset {\widetilde{{\mathbb {S}}}}(\gamma ^*\cap \gamma _0)\not = \emptyset \).

Define \(\eta (v)= \min {\{2{\widetilde{r}}(v),6\}}\) if the area of the lunar region \(\textsc {Area}({\widetilde{B}}(v)\cap \mathbb {B}(3))\ge \frac{1}{5}\pi {\widetilde{r}}(v)^2\) analogously as in the proof of Lemma 5.11. If \(v\in V_{\mathbb {c}(3)}\) and if \(\textsc {Area}({\widetilde{B}}(v)\cap \mathbb {B}(3))<\frac{1}{5}\pi {\widetilde{r}}(v)^2\) let \(\eta (v)\) be the maximum of the length of the arc \(\mathbb {c}(3)\cap {\widetilde{S}}(v)\) and of the maximum of the distance of a point in \({\widetilde{S}}(v)\cap \mathbb {B}(3)\) to the arc \(\mathbb {c}(3)\cap {\widetilde{S}}(v)\) as in condition (3). Else set \(\eta (v)=0\). Then for any connected curve \(\gamma ^*\in \varGamma ^*_G(V_2,V_{2k-1})\) we have \(\sum _{v\in \gamma ^*}\eta (v) \ge 1\) as in the proof of Lemma 5.11. By our assumptions on \(V_2,V_{2k-1}\) and G every curve \(\gamma ^*\in \varGamma ^*_G(V_2,V_{2k-1})\) contains a connected subcurve, so the estimate holds for all such curves \(\gamma ^*\).

By our assumption we know that \(\frac{{\widetilde{r}}(v)}{d(0,{\widetilde{B}}(v))}<\varepsilon \) for \(v\in W^{\prime }\supset W^{\prime \prime }\), so \(\eta (v)< 6\varepsilon \). This implies \(\sum _{v\in W^{\prime \prime }} \eta (v)\le 6\varepsilon M\) as \(|W^{\prime \prime }|=M\) is constant.

Let \(\beta ^*\in \varGamma ^*_{{\tilde{G}}}(V_2\cap V{\setminus } W^{\prime }, V_{2k-1}\cap V{\setminus } W^{\prime })\). Then \(\gamma ^*=\beta ^*\cup W^{\prime \prime }\in \varGamma ^*_G(V_2,V_{2k-1})\) and \(\sum _{v\in \beta ^*}\eta (v) \ge 1-6\varepsilon M\). Choose \(\varepsilon >0\) small enough so that \(1-6\varepsilon M\ge 1/\sqrt{2}\). Then \((\sqrt{2}\eta )\) is \(\varGamma ^*_{{\tilde{G}}}(V_2\cap V{\setminus } W^{\prime }, V_{2k-1}\cap V{\setminus } W^{\prime })\)-admissible and we obtain

$$\begin{aligned} \textsc {Vel}_{{\tilde{G}}}(V_2,V_{2k-1})= & {} \textsc {Mod}(\varGamma ^*_{{\tilde{G}}}(V_2\cap V{\setminus } W^{\prime }, V_{2k-1}\cap V{\setminus } W^{\prime }))\\\le & {} \mathrm{area}(\sqrt{2}\eta )=2\,\mathrm{area}(\eta ). \end{aligned}$$

Now \(2\,\mathrm{area}(\eta )\le 2\cdot \frac{9}{4C_2}=2C_6\) can be bounded from above similarly as in the proof of Lemma 5.11. This contradicts our choice of k and finishes the proof on existence of \({\widetilde{R}}\). \(\square \)

Thus we have shown (29), that is there exists \(\delta \in (0,\frac{1}{3})\) such that

$$\begin{aligned} \limsup _{v\in W, v\rightarrow \infty } {\widetilde{\tau }}(v)>3\delta >0. \end{aligned}$$

Let \(u_k\) be a sequence of pairwise disjoint vertices satisfying \(\tau (u_k)\ge 1\) and \({\widetilde{\tau }}(u_k)\ge 3\delta \). As \({\mathscr {C}}\) and \(\widetilde{{\mathscr {C}}}\) are both locally finite in \({\mathbb {C}}\) we deduce that \(\text {dist}(0,B(u_k))\rightarrow \infty \) and \(\text {dist}(0,{\widetilde{B}}(u_k))\rightarrow \infty \).

Let \(\rho _0>0\) be large enough so that the discs \(B(v_0), B(v_1), {\widetilde{B}}(v_0), {\widetilde{B}}(v_1)\) are all contained in \(\mathbb {B}(\rho _0)\). Choose \(u_k=:u\) so that

$$\begin{aligned} d=d(0,B(u))\ge 100\rho _0/\delta ^2\quad \mathrm{and}\quad {\widetilde{d}}=d(0,{\widetilde{B}}(u))\ge 100\rho _0/\delta ^2. \end{aligned}$$

See Fig. 10 for an illustration. Let F be a Möbius transformation satisfying \(F(0)=0\) and \(F({\widetilde{{\mathbb {C}}}}{\setminus } B(u))={\mathbb {D}}\) and let \({\widetilde{F}}\) be a Möbius transformation satisfying \({\widetilde{F}}(0)=0\) and \({\widetilde{F}}({\widehat{{\mathbb {C}}}}{\setminus } {\widetilde{B}}(u))={\mathbb {D}}\). Then the hyperbolic distance between 0 and \(F(\infty )\) is bigger than the hyperbolic distance between 0 and \(r(u)/(d+\rho _0+r(u))\). Therefore

$$\begin{aligned} |F(\infty )|&= \frac{r(u)}{d+r(u)} =\frac{1}{\frac{1}{\tau (u)} +1} \ge \frac{1}{2}> \delta \qquad \mathrm{and} \\ |{\widetilde{F}}(\infty )|&= \frac{1}{\frac{1}{{\widetilde{\tau }}(u)} +1} \ge \frac{1}{\frac{1}{3\delta }+1} > \delta . \end{aligned}$$

by analogous reasoning. As \(\rho _0<\delta ^2 d/100\) and \({\widetilde{\rho }}_0<\delta ^2 {\widetilde{d}}/100\) the two discs \(F(\mathbb {B}(\rho ))\) and \({\widetilde{F}}({\widetilde{\mathbb {B}}}(\rho ))\) lie in \(\frac{\delta }{2}{\mathbb {D}}\). This is even more true for the discs \(F(B(v_0))\), \(F(B(v_1))\), \({\widetilde{F}}({\widetilde{B}}(v_0))\), \({\widetilde{F}}({\widetilde{B}}(v_1))\). From our assumption, \(d=d(0,B(u))\ge 100\rho _0/\delta ^2 \ge 100\rho _0\), we deduce that \(|F^{\prime }(z_1)/F^{\prime }(z_2)|\le 2\) holds for all \(z_1,z_2\in \mathbb {B}(\rho _0)\). An analogous statement holds for \({\widetilde{F}}\). Therefore, the ratios of radii are bounded:

$$\begin{aligned} \frac{\mathrm{radius}(F(B(v_1)))}{\mathrm{radius}(F(B(v_0)))} \bigg / \frac{r(v_1)}{r(v_0)} \in \Big [\frac{1}{4},4\Big ] \quad \mathrm{and}\quad \frac{\mathrm{radius}({\widetilde{F}}({\widetilde{B}}(v_1)))}{\mathrm{radius}({\widetilde{F}} ({\widetilde{B}}(v_0)))} \bigg / \frac{{\widetilde{r}}(v_1)}{{\widetilde{r}}(v_0)} \in \Big [\frac{1}{4},4\Big ]. \end{aligned}$$

As \(r(v_0)={\widetilde{r}}(v_0)\) it only remains to show that \(\frac{\mathrm{radius}(F(B(v_1)))}{\mathrm{radius}(F(B(v_0)))}\) and \(\frac{\mathrm{radius}({\widetilde{F}}({\widetilde{B}}(v_1)))}{\mathrm{radius}({\widetilde{F}} ({\widetilde{B}}(v_0)))}\) are comparable.

Fig. 10
figure 10

Illustration of the configuration of \(\mathbb {B}(\rho _0)\) and B(u)

First compare \(\frac{1}{\delta } F({\mathscr {C}})\) with \({\widetilde{F}}(\widetilde{{\mathscr {C}}})\). As \(F(\infty )>\delta \) and \(F({\mathscr {C}})\) is locally finite in \({\mathbb {C}}{\setminus }\{F(\infty )\}\), there is only a finite number of circles in \(\frac{1}{\delta }F({\mathscr {C}})\) which intersect \({\mathbb {D}}\). Thus Lemma 3.9 implies \(r_{\mathrm{hyp}}(\frac{1}{\delta } F(B(v_j))) \ge r_{\mathrm{hyp}}({\widetilde{F}}({\widetilde{B}}(v_j)))\) for \(j=0,1\). Analogously, we see that \(r_{\mathrm{hyp}}(F(B(v_j))) \le r_\mathrm{hyp}(\frac{1}{\delta } {\widetilde{F}}({\widetilde{B}}(v_j)))\) for \(j=0,1\). As in the proof of the first case (i) the claim now follows. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bücking, U. On Rigidity and Convergence of Circle Patterns. Discrete Comput Geom 61, 380–420 (2019). https://doi.org/10.1007/s00454-018-0022-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00454-018-0022-0

Keywords

Mathematics Subject Classification

Navigation