Skip to main content

Lagrangian Duality in Complex Pose Graph Optimization

  • Chapter
  • First Online:
Optimization and Its Applications in Control and Data Sciences

Part of the book series: Springer Optimization and Its Applications ((SOIA,volume 115))

  • 1403 Accesses

Abstract

Pose Graph Optimization (PGO) is the problem of estimating a set of poses from pairwise relative measurements. PGO is a nonconvex problem, and currently no known technique can guarantee the efficient computation of a global optimal solution. In this paper, we show that Lagrangian duality allows computing a globally optimal solution, under certain conditions that are satisfied in many practical cases. Our first contribution is to frame the PGO problem in the complex domain. This makes analysis easier and allows drawing connections with the recent literature on unit gain graphs. Exploiting this connection we prove nontrival results about the spectrum of the matrix underlying the problem. The second contribution is to formulate and analyze the properties of the Lagrangian dual problem in the complex domain. The dual problem is a semidefinite program (SDP). Our analysis shows that the duality gap is connected to the number of eigenvalues of the penalized pose graph matrix, which arises from the solution of the SDP. We prove that if this matrix has a single eigenvalue in zero, then (1) the duality gap is zero, (2) the primal PGO problem has a unique solution, and (3) the primal solution can be computed by scaling an eigenvector of the penalized pose graph matrix. The third contribution is algorithmic: we exploit the dual problem and propose an algorithm that computes a guaranteed optimal solution for PGO when the penalized pose graph matrix satisfies the Single Zero Eigenvalue Property (SZEP). We also propose a variant that deals with the case in which the SZEP is not satisfied. This variant, while possibly suboptimal, provides a very good estimate for PGO in practice. The fourth contribution is a numerical analysis. Empirical evidence shows that in the vast majority of cases (100 % of the tests under noise regimes of practical robotics applications) the penalized pose graph matrix does satisfy the SZEP, hence our approach allows computing the global optimal solution. Finally, we report simple counterexamples in which the duality gap is nonzero, and discuss open problems.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    We use the term “rotation subproblem” to denote the problem of associating a rotation to each node in the graph, using relative rotation measurements. This corresponds to disregarding the translation measurements in PGO.

  2. 2.

    We use the somehow standard term “composition” to denote the group operation for SE(2). For two poses T 1 (p 1, R 1) and T 2 (p 2, R 2), the composition is T 1 ⋅ T 2 = (p 1 + R 1 p 2, R 1 R 2) [16]. Similarly, the identity element is (02, I 2).

  3. 3.

    When composing measurements along the loop, edge direction is important: for two consecutive edges (i, k) and (k, j) along the loop, the composition is T ij  = T ik ⋅ T kj , while if the second edge is in the form (j, k), the composition becomes T ij  = T ik ⋅ T jk −1.

  4. 4.

    \(\tilde{V }\) can be computed from singular value decomposition of \(\tilde{W}(\lambda ^{\star })\).

  5. 5.

    This was not included in the first row of Fig. 7 as it does not provide a guess for the positions of the nodes.

References

  1. Aragues, R., Carlone, L., Calafiore, G., Sagues, C.: Multi-agent localization from noisy relative pose measurements. In: IEEE International Conference on Robotics and Automation (ICRA), Shanghai, pp. 364–369 (2011)

    Google Scholar 

  2. Arie-Nachimson, M., Kovalsky, S.Z., Kemelmacher-Shlizerman, I., Singer, A., Basri, R.: Global motion estimation from point matches. In: 3DIMPVT, 2012

    Google Scholar 

  3. Bandeira, A.S., Singer, A., Spielman, D.A.: A Cheeger inequality for the graph connection Laplacian. SIAM. J. Matrix Anal. Appl. 34 (4), 1611–1630 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  4. Barooah, P., Hespanha, J.P.: Estimation on graphs from relative measurements. Control Syst. Mag. 27 (4), 57–74 (2007)

    Article  Google Scholar 

  5. Biswas, P., Lian, T., Wang, T., Ye, Y.: Semidefinite programming based algorithms for sensor network localization. ACM Trans. Sensor Netw. 2 (2), 188–220 (2006)

    Article  Google Scholar 

  6. Borra, D., Carli, R., Lovisari, E., Fagnani, F., Zampieri, S.: Autonomous calibration algorithms for planar networks of cameras. In Proceedings of the 2012 American Control Conference, pp. 5126–5131 (2012)

    Google Scholar 

  7. Boyd, S., Vandenberghe, L.: Convex Optimization. Cambridge University Press, Cambridge (2004)

    Book  MATH  Google Scholar 

  8. Calafiore, G.C., Carlone, L., Wei, M.: Distributed optimization techniques for range localization in networked systems. In: IEEE Conference on Decision and Control, pp. 2221–2226 (2010)

    Google Scholar 

  9. Carlevaris-Bianco, N., Eustice, R.M.: Generic factor-based node marginalization and edge sparsification for pose-graph SLAM. In: IEEE International Conference on Robotics and Automation (ICRA), pp. 5728–5735 (2013)

    Google Scholar 

  10. Carlone, L.: Convergence analysis of pose graph optimization via Gauss-Newton methods. In: IEEE International Conference on Robotics and Automation (ICRA), pp. 965–972 (2013)

    Google Scholar 

  11. Carlone, L., Censi, A.: From angular manifolds to the integer lattice: guaranteed orientation estimation with application to pose graph optimization. IEEE Trans. Robot. 30 (2), 475–492 (2014)

    Article  Google Scholar 

  12. Carlone, L., Dellaert, F.: Duality-based verification techniques for 2D SLAM. In: International Conference on Robotics and Automation (ICRA), pp. 4589–4596 (2014)

    Google Scholar 

  13. Carlone, L., Aragues, R., Castellanos, J.A., Bona, B.: A linear approximation for graph-based simultaneous localization and mapping. In: Robotics: Science and Systems (RSS) (2011)

    Google Scholar 

  14. Carlone, L., Aragues, R., Castellanos, J.A., Bona, B.: A fast and accurate approximation for planar pose graph optimization. Intl. J. Robot. Res. 33 (7), 965–987 (2014)

    Article  Google Scholar 

  15. Carlone, L., Tron, R., Daniilidis, K., Dellaert, F.: Initialization techniques for 3D SLAM: a survey on rotation estimation and its use in pose graph optimization. In: IEEE International Conference on Robotics and Automation (ICRA) (2015)

    Google Scholar 

  16. Chirikjian, G.S.: Stochastic Models, Information Theory, and Lie Groups, Volume 2: Analytic Methods and Modern Applications (Applied and Numerical Harmonic Analysis). Birkhauser, Basel (2012)

    Book  MATH  Google Scholar 

  17. Chiuso, A., Picci, G., Soatto, S.: Wide-sense estimation on the special orthogonal group. Commun. Inf. Syst. 8, 185–200 (2008)

    MathSciNet  MATH  Google Scholar 

  18. Chung, F.R.K.: Spectral Graph Theory. CBMS Regional Conference Series in Mathematics, vol. 92 American Mathematical Society, Providence, RI (1996)

    Google Scholar 

  19. Costa, J., Patwari, N., Hero, A.: Distributed weighted-multidimensional scaling for node localization in sensor networks. ACM Trans. Sensor Netw. 2 (1), 39–64 (2006)

    Article  Google Scholar 

  20. Cucuringu, M., Lipman, Y., Singer, A.: Sensor network localization by eigenvector synchronization over the Euclidean group. ACM Trans. Sensor Netw. 8 (3), 19:1–19:42 (2012)

    Google Scholar 

  21. Cucuringu, M., Singer, A., Cowburn, D.: Eigenvector synchronization, graph rigidity and the molecule problem. Inf. Infer.: J. IMA 1 (1), 21–67 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  22. Dellaert, F., Carlson, J., Ila, V., Ni, K., Thorpe, C.E.: Subgraph-preconditioned conjugate gradient for large scale slam. In: IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS) (2010)

    Google Scholar 

  23. Doherty, L., Pister, K., El Ghaoui, L.: Convex position estimation in wireless sensor networks. In: IEEE INFOCOM, vol. 3, pp. 1655–1663 (2001)

    Google Scholar 

  24. Dubbelman, G., Browning, B.: Closed-form online pose-chain slam. In: IEEE International Conference on Robotics and Automation (ICRA) (2013)

    Google Scholar 

  25. Dubbelman, G., Esteban, I., Schutte, K.: Efficient trajectory bending with applications to loop closure. In: IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), pp. 1–7 (2010)

    Google Scholar 

  26. Dubbelman, G., Hansen, P., Browning, B., Dias, M.B.: Orientation only loop-closing with closed-form trajectory bending. In: IEEE International Conference on Robotics and Automation (ICRA) (2012)

    Google Scholar 

  27. Duckett, T., Marsland, S., Shapiro, J.: Fast, on-line learning of globally consistent maps. Auton. Robot. 12 (3), 287–300 (2002)

    Article  MATH  Google Scholar 

  28. Eckart, C., Young, G.: The approximation of one matrix by another low rank. Psychometrika 1, 211–218 (1936)

    Article  MATH  Google Scholar 

  29. Eren, T., Goldenberg, O.K., Whiteley, W., Yang, Y.R.: Rigidity, computation, and randomization in network localization. In: INFOCOM 2004. Twenty-third Annual Joint Conference of the IEEE Computer and Communications Societies, vol. 4, pp. 2673–2684. IEEE, Hong Kong (2004)

    Google Scholar 

  30. Eren, T., Whiteley, W., Belhumeur, P.N.: Using angle of arrival (bearing) information in network localization. In: IEEE Conference on Decision and Control, pp. 4676–4681 (2006)

    Google Scholar 

  31. Eustice, R.M., Singh, H., Leonard, J.J., Walter, M.R.: Visually mapping the RMS Titanic: conservative covariance estimates for SLAM information filters. Int. J. Robot. Res. 25 (12), 1223–1242 (2006)

    Article  Google Scholar 

  32. Fredriksson, J., Olsson, C.: Simultaneous multiple rotation averaging using Lagrangian duality. In: Asian Conference on Computer Vision (ACCV) (2012)

    Google Scholar 

  33. Frese, U., Larsson, P., Duckett, T.: A multilevel relaxation algorithm for simultaneous localisation and mapping. IEEE Trans. Robot. 21 (2), 196–207 (2005)

    Article  Google Scholar 

  34. Govindu, V.M.: Combining two-view constraints for motion estimation. In: IEEE Conference on Computer Vision and Pattern Recognition (CVPR), pp. 218–225 (2001)

    Google Scholar 

  35. Govindu, V.M.: Lie-algebraic averaging for globally consistent motion estimation. In: IEEE Conference on Computer Vision and Pattern Recognition (CVPR) (2004)

    Google Scholar 

  36. Grisetti, G., Stachniss, C., Burgard, W.: Non-linear constraint network optimization for efficient map learning. Trans. Intell. Transport. Syst. 10 (3), 428–439 (2009)

    Article  Google Scholar 

  37. Hartley, R., Trumpf, J., Dai, Y., Li, H.: Rotation averaging. IJCV 103 (3), 267–305 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  38. Hatanaka, T., Fujita, M., Bullo, F.: Vision-based cooperative estimation via multi-agent optimization. In: IEEE Conference on Decision and Control (2010)

    Book  Google Scholar 

  39. Hazewinkel, M. (ed.) Complex number. In: Encyclopedia of Mathematics. Springer, New York (2001)

    Google Scholar 

  40. Huang, S., Lai, Y., Frese, U., Dissanayake, G.: How far is SLAM from a linear least squares problem? In: IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), pp. 3011–3016 (2010)

    Google Scholar 

  41. Huang, S., Wang, H., Frese, U., Dissanayake, G.: On the number of local minima to the point feature based SLAM problem. In: IEEE International Conference on Robotics and Automation (ICRA), pp. 2074–2079 (2012)

    Google Scholar 

  42. Indelman, V., Nelson, E., Michael, N., Dellaert, F.: Multi-robot pose graph localization and data association from unknown initial relative poses via expectation maximization. In: IEEE International Conference on Robotics and Automation (ICRA) (2014)

    Google Scholar 

  43. Johannsson, H., Kaess, M., Fallon, M., Leonard, J.J.: Temporally scalable visual SLAM using a reduced pose graph. In: IEEE International Conference on Robotics and Automation (ICRA), pp. 54–61 (2013)

    Google Scholar 

  44. Kaess, M., Ranganathan, A., Dellaert, F.: iSAM: incremental smoothing and mapping. IEEE Trans. Robot. 24 (6), 1365–1378 (2008)

    Article  Google Scholar 

  45. Kaess, M., Johannsson, H., Roberts, R., Ila, V., Leonard, J., Dellaert, F.: iSAM2: incremental smoothing and mapping using the Bayes tree. Int. J. Robot. Res. 31, 217–236 (2012)

    Article  Google Scholar 

  46. Kim, B., Kaess, M., Fletcher, L., Leonard, J., Bachrach, A., Roy, N., Teller, S.: Multiple relative pose graphs for robust cooperative mapping. In: IEEE International Conference on Robotics and Automation (ICRA), Anchorage, Alaska, May 2010, pp. 3185–3192

    Google Scholar 

  47. Knuth, J., Barooah, P.: Collaborative 3D localization of robots from relative pose measurements using gradient descent on manifolds. In: IEEE International Conference on Robotics and Automation (ICRA), pp. 1101–1106 (2012)

    Google Scholar 

  48. Knuth, J., Barooah, P.: Collaborative localization with heterogeneous inter-robot measurements by Riemannian optimization. In: IEEE International Conference on Robotics and Automation (ICRA) (2013)

    Google Scholar 

  49. Knuth, J., Barooah, P.: Error growth in position estimation from noisy relative pose measurements. Robot. Auton. Syst. 61 (3), 229–224 (2013)

    Article  Google Scholar 

  50. Konolige, K.: Large-scale map-making. In: Proceedings of the 21th AAAI National Conference on AI, San Jose, CA (2004)

    Google Scholar 

  51. Kümmerle, R., Grisetti, G., Strasdat, H., Konolige, K., Burgard, W.: g2o: A general framework for graph optimization. In: Proceedings of the IEEE International Conference on Robotics and Automation (ICRA), Shanghai, May 2011

    Google Scholar 

  52. Levenberg, K.: A method for the solution of certain nonlinear problems in least squares. Quart. Appl. Math 2 (2), 164–168 (1944)

    MathSciNet  MATH  Google Scholar 

  53. Liu, M., Huang, S., Dissanayake, G., Wang, H.: A convex optimization based approach for pose SLAM problems. In: IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), pp. 1898–1903 (2012)

    Google Scholar 

  54. Lu, F., Milios, E.: Globally consistent range scan alignment for environment mapping. Auton. Robots 4, 333–349 (1997)

    Article  Google Scholar 

  55. Mao, G., Fidan, B., Anderson, B.: Wireless sensor network localization techniques. Comput. Networks 51 (10), 2529–2553 (2007)

    Article  MATH  Google Scholar 

  56. Martinec, D., Pajdla, T.: Robust rotation and translation estimation in multiview reconstruction. In: IEEE Conference on Computer Vision and Pattern Recognition (CVPR), pp. 1–8 (2007)

    Google Scholar 

  57. Olfati-Saber, R.: Swarms on sphere: a programmable swarm with synchronous behaviors like oscillator networks. In: IEEE Conference on Decision and Control, pp. 5060–5066 (2006)

    Google Scholar 

  58. Olson, E., Leonard, J., Teller, S.: Fast iterative alignment of pose graphs with poor initial estimates. In: IEEE International Conference on Robotics and Automation (ICRA), May 2006, pp. 2262–2269

    Google Scholar 

  59. Peters, J.R., Borra, D., Paden, B., Bullo, F.: Sensor network localization on the group of 3D displacements. SIAM J. Control Optim. (2014, submitted)

    Google Scholar 

  60. Piovan, G., Shames, I., Fidan, B., Bullo, F., Anderson, B.: On frame and orientation localization for relative sensing networks. Automatica 49 (1), 206–213 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  61. Reff, N.: Spectral properties of complex unit graphs (2011). arXiv 1110.4554

    Google Scholar 

  62. Rosen, D.M., Kaess, M., Leonard, J.J.: An incremental trust-region method for robust online sparse least-squares estimation. In: IEEE International Conference on Robotics and Automation (ICRA), St. Paul, MN, May 2012, pp. 1262–1269

    Google Scholar 

  63. Rosen, D.M., Kaess, M., Leonard, J.J.: RISE: an incremental trust-region method for robust online sparse least-squares estimation. IEEE Trans. Robot. 30 (5), 1091–1108 (2014)

    Article  Google Scholar 

  64. Russell, W.J., Klein, D.J., Hespanha, J.P.: Optimal estimation on the graph cycle space. IEEE Trans. Signal Process. 59 (6), 2834–2846 (2011)

    Article  MathSciNet  Google Scholar 

  65. Sarlette, A., Sepulchre, R.: Consensus optimization on manifolds. SIAM J. Control Optim. 48 (1), 56–76 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  66. Saunderson, J., Parrilo, P.A., Willsky, A.: Semidefinite descriptions of the convex hull of rotation matrices (2014). arXiv preprint: http://arxiv.org/abs/1403.4914

  67. Saunderson, J., Parrilo, P.A., Willsky, A.: Semidefinite relaxations for optimization problems over rotation matrices. In: IEEE Conference on Decision and Control, May 2014

    Google Scholar 

  68. Schrijver, A.: Theory of Linear and Integer Programming. Wiley, New York (1998)

    MATH  Google Scholar 

  69. Sharp, G.C., Lee, S.W., Wehe, D.K.: Multiview registration of 3D scenes by minimizing error between coordinate frames. IEEE Trans. Pattern Anal. Mach. Intell. 26 (8), 1037–1050 (2004)

    Article  MATH  Google Scholar 

  70. Singer, A.: A remark on global positioning from local distances. Proc. Natl. Acad. Sci. 105 (28), 9507–9511 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  71. Singer, A.: Angular synchronization by eigenvectors and semidefinite programming. Appl. Comput. Harmon. Anal. 30, 20–36 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  72. Singer, A., Shkolnisky, Y.: Three-dimensional structure determination from common lines in Cryo-EM by eigenvectors and semidefinite programming. SIAM J. Imag. Sci. 4 (2), 543–572 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  73. Stanfield, R.: Statistical theory of DF finding. J. IEE 94 (5), 762–770 (1947)

    Google Scholar 

  74. Thunberg, J., Montijano, E., Hu, X.: Distributed attitude synchronization control. In: IEEE Conference on Decision and Control (2011)

    Book  MATH  Google Scholar 

  75. Tron, R., Afsari, B., Vidal, R.: Intrinsic consensus on SO(3) with almost global convergence. In: IEEE Conference on Decision and Control (2012)

    Book  Google Scholar 

  76. Tron, R., Afsari, B., Vidal, R.: Riemannian consensus for manifolds with bounded curvature. IEEE Trans. Autom. Control 58 (4), 921–934 (2012)

    Article  MathSciNet  Google Scholar 

  77. Tron, R., Carlone, L., Dellaert, F., Daniilidis, K.: Rigid components identification and rigidity enforcement in bearing-only localization using the graph cycle basis. In: American Control Conference (2015)

    Google Scholar 

  78. Wang, H., Hu, G., Huang, S., Dissanayake, G.: On the structure of nonlinearities in pose graph SLAM. In: Robotics: Science and Systems (RSS) (2012)

    Google Scholar 

  79. Wang, L., Singer, A.: Exact and stable recovery of rotations for robust synchronization. Inf. Infer.: J. IMA pp. 1–53 (2013) doi: 10.1093/imaiai/drn000

  80. Zhang, S.: Quadratic maximization and semidefinite relaxation. Math. Programm. Ser. A 87, 453–465 (2000)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Luca Carlone .

Editor information

Editors and Affiliations

Appendix

Appendix

Proof of Proposition 1: Zero Cost in Trees

We prove Proposition 1 by inspection, providing a procedure to build an estimate that annihilates every summand in (8). The procedure is as follows:

  1. 1.

    Select a root node, say the first node (p i , r i ), with i = 1, and set it to the origin, i.e., p i  = 02, r i  = [1 0] (compare with (7) for θ i  = 0);

  2. 2.

    For each neighbor j of the root i, if j is an outgoing neighbor, set r j  = R ij r i , and p j  = p i + D ij r i , otherwise set r j  = R ji r i , and p j  = p i + D ji r j ;

  3. 3.

    Repeat point 2 for the unknown neighbors of every node that has been computed so far, and continue until all poses have been computed.

Let us now show that this procedure produces a set of poses that annihilates the objective in (8). According to the procedure, we set the first node to the origin: p 1 = 02, r 1 = [1 0]; then, before moving to the second step of the procedure, we rearrange the terms in (8): we separate the edges into two sets \(\mathcal{E} = \mathcal{E}_{1} \cup \bar{\mathcal{E}}_{1}\), where \(\mathcal{E}_{1}\) is the set of edges incident on node 1 (the root), and \(\bar{\mathcal{E}}_{1}\) are the remaining edges. Then the cost can be written as:

$$\displaystyle\begin{array}{rcl} f(p,r)& =& \sum _{(i,j)\in \mathcal{E}_{1}}\|p_{j} - p_{i} - D_{ij}r_{i}\|_{2}^{2} +\| r_{ j} - R_{ij}r_{i}\|_{2}^{2} + \\ & & +\sum _{(i,j)\in \bar{\mathcal{E}}_{1}}\|p_{j} - p_{i} - D_{ij}r_{i}\|_{2}^{2} +\| r_{ j} - R_{ij}r_{i}\|_{2}^{2}{}\end{array}$$
(48)

We can further split the set \(\mathcal{E}_{1}\) into edges that have node 1 as a tail (i.e., edges in the form (1, j)) and edges that have node 1 as head (i.e., (j, 1)):

$$\displaystyle\begin{array}{rcl} f(p,r)& =& \sum _{(1,j),j\in \mathcal{N}_{ 1}^{\text{out}}}\|p_{j} - p_{1} - D_{1j}r_{1}\|_{2}^{2} +\| r_{ j} - R_{1j}r_{1}\|_{2}^{2} + \\ & & +\sum _{ (j,1),j\in \mathcal{N}_{1}^{\text{in}}}\|p_{1} - p_{j} - D_{j1}r_{j}\|_{2}^{2} +\| r_{ 1} - R_{j1}r_{j}\|_{2}^{2} + \\ & & \qquad +\sum _{(i,j)\in \bar{\mathcal{E}}_{1}}\|p_{j} - p_{i} - D_{ij}r_{i}\|_{2}^{2} +\| r_{ j} - R_{ij}r_{i}\|_{2}^{2} {}\end{array}$$
(49)

Now, we set each node j in the first two summands as prescribed in step 2 of the procedure. By inspection one can verify that this choice annihilates the first two summands and the cost becomes:

$$\displaystyle\begin{array}{rcl} f(p,r) =\sum _{(i,j)\in \bar{\mathcal{E}}_{1}}\|p_{j} - p_{i} - D_{ij}r_{i}\|_{2}^{2} +\| r_{ j} - R_{ij}r_{i}\|_{2}^{2}& &{}\end{array}$$
(50)

Now we select a node k that has been computed at the previous step, but has some neighbor that is still unknown. As done previously, we split the set \(\bar{\mathcal{E}}_{1}\) into two disjoint subsets: \(\bar{\mathcal{E}}_{1} = \mathcal{E}_{k} \cup \bar{\mathcal{E}}_{k}\), where the set \(\mathcal{E}_{k}\) contains the edges in \(\bar{\mathcal{E}}_{1}\) that are incident on k, and \(\bar{\mathcal{E}}_{k}\) contains the remaining edges:

$$\displaystyle\begin{array}{rcl} f(p,r)& =& \sum _{\{(k,j),j\in \mathcal{N}_{ k}^{\text{out}}\}\cap \bar{\mathcal{E}}_{1}}\|p_{j} - p_{k} - D_{kj}r_{k}\|_{2}^{2} +\| r_{ j} - R_{kj}r_{k}\|_{2}^{2} + \\ & & +\sum _{\{ (j,k),j\in \mathcal{N}_{k}^{\text{in}}\}\cap \bar{\mathcal{E}}_{1}}\|p_{k} - p_{j} - D_{jk}r_{j}\|_{2}^{2} +\| r_{ k} - R_{jk}r_{j}\|_{2}^{2} + \\ & & \qquad +\sum _{(i,j)\in \bar{\mathcal{E}}_{k}}\|p_{j} - p_{i} - D_{ij}r_{i}\|_{2}^{2} +\| r_{ j} - R_{ij}r_{i}\|_{2}^{2} {}\end{array}$$
(51)

Again, setting neighbors j as prescribed in step 2 of the procedure, annihilates the first two summands in (51). Repeating the same reasoning for all nodes that have been computed, but still have unknown neighbors, we can easily show that all terms in (51) become zero (the assumption of graph connectivity ensures that we can reach all nodes), proving the claim.

Proof of Proposition 2: Zero Cost in Balanced Graphs

Similarly to section “Proof of Proposition 1: Zero Cost in Trees” in this Appendix, we prove Proposition 2 by showing that in balanced graphs one can always build a solution that attains zero cost.

For the assumption of connectivity, we can find a spanning tree \(\mathcal{T}\) of the graph, and split the terms in the cost function accordingly:

$$\displaystyle\begin{array}{rcl} f(p,r)& =& \sum _{(i,j)\in \mathcal{T}}\|p_{j} - p_{i} - D_{ij}r_{i}\|_{2}^{2} +\| r_{ j} - R_{ij}r_{i}\|_{2}^{2} + \\ & & +\sum _{(i,j)\in \bar{\mathcal{T}}}\|p_{j} - p_{i} - D_{ij}r_{i}\|_{2}^{2} +\| r_{ j} - R_{ij}r_{i}\|_{2}^{2}{}\end{array}$$
(52)

where \(\bar{\mathcal{T}}\doteq\mathcal{E}\setminus \mathcal{T}\) are the chords of the graph w.r.t. \(\mathcal{T}\).

Then, using the procedure in section “Proof of Proposition 1: Zero Cost in Trees” in this Appendix we construct a solution {r i , p i } that attains zero cost for the measurements in the spanning tree \(\mathcal{T}\). Therefore, our claim only requires to demonstrate that the solution built from the spanning tree also annihilates the terms in \(\bar{\mathcal{T}}\):

$$\displaystyle\begin{array}{rcl} f(p^{\star },r^{\star }) =\sum _{ (i,j)\in \bar{\mathcal{T}}}\|p_{j}^{\star } - p_{ i}^{\star } - D_{ ij}r_{i}^{\star }\|_{ 2}^{2} +\| r_{ j}^{\star } - R_{ ij}r_{i}^{\star }\|_{ 2}^{2}& &{}\end{array}$$
(53)

To prove the claim, we consider one of the chords in \(\bar{\mathcal{T}}\) and we show that the cost at {r i , p i } is zero. The cost associated to a chord \((i,j) \in \bar{\mathcal{T}}\) is:

$$\displaystyle{ \|p_{j}^{\star } - p_{ i}^{\star } - D_{ ij}r_{i}^{\star }\|_{ 2}^{2} +\| r_{ j}^{\star } - R_{ ij}r_{i}^{\star }\|_{ 2}^{2} }$$
(54)

Now consider the unique path \(\mathcal{P}_{ij}\) in the spanning tree \(\mathcal{T}\) that connects i to j, and number the nodes along this path as i, i + 1, , j − 1, j.

Let us start by analyzing the second summand in (54), which corresponds to the rotation measurements. According to the procedure in section “Proof of Proposition 1: Zero Cost in Trees” in this Appendix to build the solution for \(\mathcal{T}\), we propagate the estimate from the root of the tree. Then it is easy to see that:

$$\displaystyle{ r_{j}^{\star } = R_{ j-1j}\cdots R_{i+1i+2}R_{ii+1}r_{i}^{\star } }$$
(55)

where R ii+1 is the rotation associated to the edge (i, i + 1), or its transpose if the edge is in the form (i + 1, i) (i.e., it is traversed backwards along \(\mathcal{P}_{ij}\)). Now we notice that the assumption of balanced graph implies that the measurements compose to the identity along every cycle in the graph. Since the chord (i, j) and the path \(\mathcal{P}_{ij}\) form a cycle in the graph, it holds:

$$\displaystyle{ R_{j-1j}\cdots R_{i+1i+2}R_{ii+1} = R_{ij} }$$
(56)

Substituting (56) back into (55) we get:

$$\displaystyle{ r_{j}^{\star } = R_{ ij}r_{i}^{\star } }$$
(57)

which can be easily seen to annihilate the second summand in (54).

Now we only need to demonstrate that also the first summand in (54) is zero. The procedure in section “Proof of Proposition 1: Zero Cost in Trees” in this Appendix leads to the following estimate for the position of node j:

$$\displaystyle\begin{array}{rcl} p_{j}^{\star }& =& p_{ i}^{\star } + D_{ ii+1}r_{i}^{\star } + D_{ i+1i+2}r_{i+1}^{\star } + \cdots + D_{ j-1j}r_{j-1}^{\star } \\ & =& p_{i}^{\star } + D_{ ii+1}r_{i}^{\star } + D_{ i+1i+2}R_{ii+1}r_{i}^{\star } + \cdots + D_{ j-1j}R_{j-2j-1}\cdots R_{i+1i+2}R_{ii+1}r_{i}^{\star } \\ & =& p_{i}^{\star } + \left (D_{ ii+1} + D_{i+1i+2}R_{ii+1} + \cdots + D_{j-1j}R_{j-2j-1}\cdots R_{i+1i+2}R_{ii+1}\right )r_{i}^{\star } {}\end{array}$$
(58)

The assumption of balanced graph implies that position measurements compose to zero along every cycle, hence:

$$\displaystyle\begin{array}{rcl} \varDelta _{ij}& =& \varDelta _{ii+1} + R_{ii+1}\varDelta _{i+1i+2} + R_{i+1i+2}R_{ii+1}\varDelta _{i+2i+3} + \cdots \\ & & +R_{j-2j-1}\cdots R_{i+1i+2}R_{ii+1}\varDelta _{j-1j} {}\end{array}$$
(59)

or equivalently:

$$\displaystyle\begin{array}{rcl} D_{ij}& =& D_{ii+1} + D_{i+1i+2}R_{ii+1} + \cdots \\ & & +D_{j-1j}R_{j-2j-1}\cdots R_{i+1i+2}R_{ii+1}{}\end{array}$$
(60)

Substituting (60) back into (58) we obtain:

$$\displaystyle{ p_{j}^{\star } = p_{ i}^{\star } + D_{ ij}r_{i}^{\star } }$$

which annihilates the first summand in (54), concluding the proof.

Proof of Proposition 4: Properties of \(\mathcal{W}\)

Let us prove that \(\mathcal{W}\) has (at least) two eigenvalues in zero. We already observed that the top-left block of \(\mathcal{W}\) is \(\bar{\mathcal{L}} = \mathcal{L}\otimes I_{2}\), where \(\mathcal{L}\) is the Laplacian matrix of the graph underlying the PGO problem. The Laplacian \(\mathcal{L}\) of a connected graph has a single eigenvalue in zero, and the corresponding eigenvector is 1 n (see, e.g., [18, Sects. 1.2–1.3]), i.e., \(\mathcal{L}\cdot 1_{n} = 0\). Using this property, it is easy to show that the matrix N ≐[0 n  1 n ]I 2 is in the nullspace of \(\mathcal{W}\), i.e., \(\mathcal{W}N = 0\). Since N has rank 2, this implies that the nullspace of \(\mathcal{W}\) has at least dimension 2, which proves the first claim.

Let us now prove that the matrix \(\mathcal{W}\) is composed by 2 × 2 blocks \([\mathcal{W}]_{ij}\), with \([\mathcal{W}]_{ij} \in \alpha SO(2)\), \(\forall i,j = 1,\ldots,2n\), and \([\mathcal{W}]_{ii} =\alpha _{ii}I_{2}\) with α ii  ≥ 0. We prove this by direct inspection of the blocks of \(\mathcal{W}\). Given the structure of \(\mathcal{W}\) in (14), the claim reduces to proving that the matrices \(\bar{\mathcal{L}}\), \(\bar{Q}\), and \(\bar{A}^{\top }\bar{D}\) are composed by 2 × 2 blocks in α S O(2), and the diagonal blocks of \(\bar{\mathcal{L}}\) and \(\bar{Q}\) are multiples of the identity matrix. To this end, we start by observing that \(\bar{\mathcal{L}} = \mathcal{L}\otimes I_{2}\), hence all blocks in \(\bar{\mathcal{L}}\) are multiples of the 2 × 2 identity matrix, which also implies that they belong to α S O(2). Consider next the matrix \(\bar{Q}\doteq\bar{D}^{\top }\bar{D} +\bar{ U}^{\top }\bar{U}\). From the definition of \(\bar{D}\) it follows that \(\bar{D}^{\top }\bar{D}\) is zero everywhere, except the 2 × 2 diagonal blocks:

$$\displaystyle{ [\bar{D}^{\top }\bar{D}]_{ ii} =\sum _{j\in \mathcal{N}_{ i}^{\text{out}}}\|\varDelta _{ij}\|_{2}^{2}I_{ 2},\qquad i = 1,\ldots,n. }$$
(61)

Similarly, from simple matrix manipulation we obtain the following block structure of \(\bar{U}^{\top }\bar{U}\):

$$\displaystyle\begin{array}{rcl} [\bar{U}^{\top }\bar{U}]_{ ii}& =& d_{i}I_{2},\qquad i = 1,\ldots,n; \\{} [\bar{U}^{\top }\bar{U}]_{ ij}& =& -R_{ij},\qquad (i,j) \in \mathcal{E}; \\{} [\bar{U}^{\top }\bar{U}]_{ ij}& =& -R_{ji}^{\top },\qquad (j,i) \in \mathcal{E}; \\ {}[\bar{U}^{\top }\bar{U}]_{ ij}& =& 0_{2\times 2},\qquad \text{otherwise}.{}\end{array}$$
(62)

where d i is the degree (number of neighbours) of node i. Combining (61) and (62) we get the following structure for \(\bar{Q}\):

$$\displaystyle\begin{array}{rcl} [\bar{Q}]_{ii}& =& \beta _{i}I_{2},\qquad i = 1,\ldots,n; \\ {}[\bar{Q}]_{ij}& =& -R_{ij},\qquad (i,j) \in \mathcal{E}; \\ {}[\bar{Q}]_{ij}& =& -R_{ji}^{\top },\qquad (j,i) \in \mathcal{E}; \\ {}[\bar{Q}]_{ij}& =& 0_{2\times 2},\qquad \text{otherwise}.{}\end{array}$$
(63)

where we defined \(\beta _{i}\doteq d_{i} +\sum _{j\in \mathcal{N}_{ i}^{\text{out}}}\|\varDelta _{ij}\|_{2}^{2}\). Clearly, \(\bar{Q}\) has blocks in α S O(2) and the diagonal blocks are nonnegative multiples of I 2.

Now, it only remains to inspect the structure of \(\bar{A}^{\top }\bar{D}\). The matrix \(\bar{A}^{\top }\bar{D}\) has the following structure:

$$\displaystyle\begin{array}{rcl} [\bar{A}^{\top }\bar{D}]_{ ii}& =& \sum _{j\in \mathcal{N}_{ i}^{\text{out}}}D_{ij},\qquad i = 1,\ldots,n; \\ {}[\bar{A}^{\top }\bar{D}]_{ ij}& =& -D_{ji},\qquad (j,i) \in \mathcal{E}; \\ {}[\bar{A}^{\top }\bar{D}]_{ ij}& =& 0_{2\times 2},\qquad \text{otherwise}. {}\end{array}$$
(64)

Note that \(\sum _{j\in \mathcal{N}_{ i}^{\text{out}}}D_{ij}\) is the sum of matrices in α S O(2), hence it also belongs to α S O(2). Therefore, also all blocks of \(\bar{A}^{\top }\bar{D}\) are in α S O(2), thus concluding the proof.

Proof of Proposition 5: Cost in the Complex Domain

Let us prove the equivalence between the complex cost and its real counterpart, as stated in Proposition 5.

We first observe that the dot product between two 2-vectors \(x_{1},x_{2} \in \mathbb{R}^{2}\), can be written in terms of their complex representation \(\tilde{x}_{1}\doteq x_{1}^{\vee }\), and \(\tilde{x}_{2}\doteq x_{2}^{\vee }\), as follows:

$$\displaystyle{ x_{1}^{\top }x_{ 2} = \frac{\tilde{x}_{1}^{{\ast}}\tilde{x}_{2} +\tilde{ x}_{1}\tilde{x}_{2}^{{\ast}}} {2} }$$
(65)

Moreover, we know that the action of a matrix Z ∈ α S O(2) can be written as the product of complex numbers, see (18).

Combining (65) and (18) we get:

$$\displaystyle\begin{array}{rcl} x_{1}^{\top }Zx_{ 2}& \sim & \frac{\tilde{x}_{1}^{{\ast}}\;\tilde{z}\;\tilde{x}_{2} +\tilde{ x}_{1}\;\tilde{z}^{{\ast}}\;\tilde{x}_{2}^{{\ast}}} {2} {}\end{array}$$
(66)

where \(\tilde{z} = Z^{\vee }\). Furthermore, when Z is multiple of the identity matrix, it easy to see that z = Z is actually a real number, and Eq. (66) becomes:

$$\displaystyle\begin{array}{rcl} x_{1}^{\top }Zx_{ 1}& \sim & \tilde{x}_{1}^{{\ast}}\;z\;\tilde{x}_{ 1}{}\end{array}$$
(67)

With the machinery introduced so far, we are ready to rewrite the cost x Wx in complex form. Since W is symmetric, the product becomes:

$$\displaystyle\begin{array}{rcl} x^{\top }Wx =\sum _{ i=1}^{2n-1}\left [x_{ i}^{\top }[W]_{ ii}x_{i} +\sum _{ j=i+1}^{2n-1}2\;x_{ i}^{\top }[W]_{ ij}x_{j}\right ]& &{}\end{array}$$
(68)

Using the fact that [W] ii is a multiple of the identity matrix, \(\tilde{W}_{ii}\doteq[W]_{ii}^{\vee }\in \mathbb{R}^{}\), and using (67) we conclude \(x_{i}^{\top }[W]_{ii}x_{i} =\tilde{ x}_{i}^{{\ast}}\tilde{W}_{ii}\tilde{x}_{i}\). Moreover, defining \(\tilde{W}_{ij}\doteq[W]_{ij}^{\vee }\) (these will be complex numbers, in general), and using (66), Eq. (68) becomes:

$$\displaystyle\begin{array}{rcl} x^{\top }Wx& =& \sum _{ i=1}^{2n-1}\left [\tilde{x}_{ i}^{{\ast}}\tilde{W}_{ ii}\tilde{x}_{i} +\sum _{ j=i+1}^{2n-1}(\tilde{x}_{ i}^{{\ast}}\tilde{W}_{ ij}\tilde{x}_{j} +\tilde{ x}_{i}\tilde{W}_{ij}^{{\ast}}\tilde{x}_{ j}^{{\ast}})\right ] \\ & =& \sum _{i=1}^{2n-1}\left [\tilde{x}_{ i}^{{\ast}}\tilde{W}_{ ii}\tilde{x}_{i} +\sum _{j\neq i}\tilde{x}_{i}^{{\ast}}\tilde{W}_{ ij}\tilde{x}_{j}\right ] =\tilde{ x}^{{\ast}}\tilde{W}\tilde{x} {}\end{array}$$
(69)

where we completed the lower triangular part of \(\tilde{W}\) as \(\tilde{W}_{ji} =\tilde{ W}_{ij}^{{\ast}}\).

Proof of Proposition 6: Zero Eigenvalues in \(\tilde{W}\)

Let us denote with N 0 the number of zero eigenvalues of the pose graph matrix \(\tilde{W}\). N 0 can be written in terms of the dimension of the matrix (\(\tilde{W} \in \mathbb{C}^{(2n-1)\times (2n-1)}\)) and the rank of the matrix:

$$\displaystyle{ N_{0} = (2n - 1) -\mbox{ rank}(\tilde{W}) }$$
(70)

Now, recalling the factorization of \(\tilde{W}\) given in (25), we note that:

$$\displaystyle{ \mbox{ rank}(\tilde{W}) = \mbox{ rank}\left (\left [\begin{array}{cc} A&\tilde{D}\\ 0 & \tilde{U} \end{array} \right ]\right ) = \mbox{ rank}(A)+\mbox{ rank}(\tilde{U}) }$$
(71)

where the second relation follows from the upper triangular structure of the matrix. Now, we know from [68, Sect. 19.3] that the anchored incidence matrix A, obtained by removing a row from the the incidence matrix of a connected graph, is full rank:

$$\displaystyle{ \mbox{ rank}(A) = n - 1 }$$
(72)

Therefore:

$$\displaystyle{ N_{0} = n -\mbox{ rank}(\tilde{U}) }$$
(73)

Now, since we recognized that \(\tilde{U}\) is the complex incidence matrix of a unit gain graph (Lemma 1), we can use the result of Lemma 2.3 in [61], which says that:

$$\displaystyle{ \mbox{ rank}(\tilde{U}) = n - b, }$$
(74)

where b is the number of connected components in the graph that are balanced. Since we are working on a connected graph (Assumption 1), b can be either one (balanced graph or tree), or zero otherwise. Using (73) and (74), we obtain N 0 = b, which implies that N 0 = 1 for balanced graphs or trees, or N 0 = 0, otherwise.

Proof of Proposition 7: Spectrum of Complex and Real Pose Graph Matrices

Recall that any Hermitian matrix has real eigenvalues, and possibly complex eigenvectors. Let \(\mu \in \mathbb{R}^{}\) be an eigenvalue of \(\tilde{W}\), associated with an eigenvector \(\tilde{v} \in \mathbb{C}^{2n-1}\), i.e.,

$$\displaystyle\begin{array}{rcl} \tilde{W}\tilde{v}& =& \mu \tilde{v}{}\end{array}$$
(75)

From Eq. (75) we have, for i = 1, , 2n − 1,

$$\displaystyle\begin{array}{rcl} \sum _{j=1}^{2n-1}\tilde{W}_{ ij}\tilde{v}_{j} =\mu \tilde{ v}_{i} \Leftrightarrow \sum _{j=1}^{2n-1}[W]_{ ij}v_{j} =\mu v_{i}& &{}\end{array}$$
(76)

where v i is such that \(v_{i}^{\vee } =\tilde{ v}_{i}\). Since Eq. (76) holds for all i = 1, , 2n − 1, it can be written in compact form as:

$$\displaystyle{ Wv =\mu v }$$
(77)

hence v is an eigenvector of the real anchored pose graph matrix W, associated with the eigenvalue μ. This proves that any eigenvalue of \(\tilde{W}\) is also an eigenvalue of W.

To prove that the eigenvalue μ is actually repeated twice in W, consider now Eq. (75) and multiply both members by the complex number \(\mathrm{e}^{\text{j} \frac{\pi }{ 2} }\):

$$\displaystyle\begin{array}{rcl} \tilde{W}\tilde{v}\mathrm{e}^{\text{j} \frac{\pi } { 2} }& =& \mu \tilde{v}\mathrm{e}^{\text{j} \frac{\pi } {2} }{}\end{array}$$
(78)

For i = 1, , 2n − 1, we have:

$$\displaystyle\begin{array}{rcl} \sum _{j=1}^{2n-1}\tilde{W}_{ ij}^{{\ast}}\tilde{v}_{ j}\mathrm{e}^{\text{j} \frac{\pi } {2} } =\mu \tilde{ v}_{i}\mathrm{e}^{\text{j} \frac{\pi } { 2} } \Leftrightarrow \sum _{j=1}^{2n-1}[W]_{ij}w_{j} =\mu w_{i}& &{}\end{array}$$
(79)

where w i is such that \(w_{i}^{\vee } =\tilde{ v}_{j}\mathrm{e}^{\text{j} \frac{\pi } { 2} }\). Since Eq. (79) holds for all i = 1, , 2n − 1, it can be written in compact form as:

$$\displaystyle{ Ww =\mu w }$$
(80)

hence also w is an eigenvector of W associated with the eigenvalue μ.

Now it only remains to demonstrate that v and w are linearly independent. One can readily check that, if \(\tilde{v}_{i}\) is in the form \(\tilde{v}_{i} =\eta _{i}\mathrm{e}^{\text{j}\theta _{i}}\), then

$$\displaystyle{ v_{i} =\eta _{i}\left [\begin{array}{c} \cos (\theta _{i}) \\ \sin (\theta _{i}) \end{array} \right ]. }$$
(81)

Moreover, observing that \(\tilde{v}_{j}\mathrm{e}^{\text{j} \frac{\pi } { 2} } =\eta _{i}\mathrm{e}^{\text{j}(\theta _{i}+\pi /2)}\), then

$$\displaystyle{ w_{i} =\eta _{i}\left [\begin{array}{c} \cos (\theta _{i} +\pi /2) \\ \sin (\theta _{i} +\pi /2) \end{array} \right ] =\eta _{i}\left [\begin{array}{c} -\sin (\theta _{i}) \\ \cos (\theta _{i}) \end{array} \right ] }$$
(82)

From (81) and (82) is it easy to see that v w = 0, thus v, w are orthogonal, hence independent. To each eigenvalue μ of \(\tilde{W}\) there thus correspond an identical eigenvalue of W, of geometric multiplicity at least two. Since \(\tilde{W}\) has 2n − 1 eigenvalues and W has 2(2n − 1) eigenvalues, we conclude that to each eigenvalue μ of \(\tilde{W}\) there correspond exactly two eigenvalues of W in μ. The previous proof also shows how the set of orthogonal eigenvectors of W is related to the set of eigenvectors of \(\tilde{W}\).

Proof of Theorem 1: Primal-dual Optimal Pairs

We prove that, given \(\lambda \in \mathbb{R}^{n}\), if an \(\tilde{x}_{\lambda } \in \mathcal{X}(\lambda )\) is primal feasible, then \(\tilde{x}_{\lambda }\) is primal optimal; moreover, λ is dual optimal, and the duality gap is zero.

By weak duality we know that for any λ:

$$\displaystyle{ \mathcal{L}(x_{\lambda },\lambda ) \leq f^{\star } }$$
(83)

However, if x λ is primal feasible, by optimality of f , it must also hold

$$\displaystyle{ f^{\star } \leq f(x_{\lambda }) }$$
(84)

Now we observe that for a feasible x λ , the terms in the Lagrangian associated to the constraints disappear and \(\mathcal{L}(x_{\lambda },\lambda ) = f(x_{\lambda })\). Using the latter equality and the inequalities (83) and (84) we get:

$$\displaystyle{ f^{\star } \leq f(x_{\lambda }) = \mathcal{L}(x_{\lambda },\lambda ) \leq f^{\star } }$$
(85)

which implies f(x λ ) = f , i.e., x λ is primal optimal.

Further, we have that

$$\displaystyle{d^{\star } \geq \mathop{\mathrm{min}}\limits _{ x}\mathcal{L}(x,\lambda ) = \mathcal{L}(x_{\lambda },\lambda ) = f(x_{\lambda }) = f^{\star },}$$

which, combined with weak duality (d  ≤ f ), implies that d  = f and that λ attains the dual optimal value.

Numerical Data for the Toy Examples in Sect. 6

Ground truth nodes poses, written as x i  = [p i , θ i ]:

$$\displaystyle{ \begin{array}{ccccccc} x_{1} & =&[& 0.0000 & - 5.0000& 0.2451 &] \\ x_{2} & =&[& 4.7553 & - 1.5451& - 0.4496&] \\ x_{3} & =&[& 2.9389 & 4.0451 & 0.7361 &] \\ x_{4} & =&[& - 2.9389& 4.0451 & 0.3699 &] \\ x_{5} & =&[& - 4.7553& - 1.5451& - 1.7225&]\end{array} }$$
(86)

Relative measurements, for each edge (i, j), written as (i, j): [Δ ij , θ ij ]:

$$\displaystyle{ \begin{array}{ccccccc} (1,2)&:&[& 4.6606 & 1.2177 & 2.8186 &] \\ (2,3)&:&[& - 4.4199& 4.8043 & 0.1519 &] \\ (3,4)&:&[& - 4.1169& 4.9322 & 0.5638 &] \\ (4,5)&:&[& - 3.6351& - 5.0908& - 0.5855&] \\ (5,1)&:&[& 3.4744 & 5.9425 & 2.5775 &]\end{array} }$$
(87)

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Calafiore, G.C., Carlone, L., Dellaert, F. (2016). Lagrangian Duality in Complex Pose Graph Optimization. In: Goldengorin, B. (eds) Optimization and Its Applications in Control and Data Sciences. Springer Optimization and Its Applications, vol 115. Springer, Cham. https://doi.org/10.1007/978-3-319-42056-1_5

Download citation

Publish with us

Policies and ethics