Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-24T09:08:16.664Z Has data issue: false hasContentIssue false

FULL MEREOGEOMETRIES

Published online by Cambridge University Press:  18 May 2010

STEFANO BORGO*
Affiliation:
National Research Council, Institute of Cognitive Sciences and Technologies – ISTC–CNR, Laboratory for Applied Ontology and KRDB, Free University of Bolzano
CLAUDIO MASOLO*
Affiliation:
National Research Council, Institute of Cognitive Sciences and Technologies – ISTC–CNR, Laboratory for Applied Ontology
*
*NATIONAL RESEARCH COUNCIL, INSTITUTE OF COGNITIVE SCIENCES AND TECHNOLOGIES – ISTC–CNR, LABORATORY FOR APPLIED ONTOLOGY, POVO (TRENTO), ITALY E-mail:borgo@loa-cnr.it KRDB, FREE UNIVERSITY OF BOLZANO, BOZEN-BOLZANO, ITALY
*NATIONAL RESEARCH COUNCIL, INSTITUTE OF COGNITIVE SCIENCES AND TECHNOLOGIES – ISTC–CNR, LABORATORY FOR APPLIED ONTOLOGY, POVO (TRENTO), ITALY E-mail:borgo@loa-cnr.it KRDB, FREE UNIVERSITY OF BOLZANO, BOZEN-BOLZANO, ITALY

Abstract

We analyze and compare geometrical theories based on mereology (mereogeometries). Most theories in this area lack in formalization, and this prevents any systematic logical analysis. To overcome this problem, we concentrate on specific interpretations for the primitives and use them to isolate comparable models for each theory. Relying on the chosen interpretations, we introduce the notion of environment structure, that is, a minimal structure that contains a (sub)structure for each theory. In particular, in the case of mereogeometries, the domain of an environment structure is composed of particular subsets of Rn. The comparison of mereogeometrical theories within these environment structures shows dependencies among primitives and provides (relative) definitional equivalences. With one exception, we show that all the theories considered are equivalent in these environment structures.

Type
Research Article
Copyright
Copyright © Association for Symbolic Logic 2010

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

BIBLIOGRAPHY

Asher, N., & Vieu, L. (1995). Toward a geometry of common sense: A semantics and a complete axiomatization of mereotopology. In Mellish, C., editor, Proceedings of the International Joint Conference on Artificial Intelligence (IJCAI 1995). San Francisco: Morgan Kaufmann, pp. 846852.Google Scholar
Aurnague, M., Vieu, L., & Borillo, A. (1997). La Représentation Formelle des Concepts Spatiaux dans la Langue. In Denis, M., editor, Langage et Cognition Spatiale. Paris: Masson, pp. 69102.Google Scholar
Bennett, B. (2001). A categorical axiomatisation of region-based geometry. Fundamenta Informaticae, 46, 145158.Google Scholar
Bennett, B., Cohn, A. G., Torrini, P., & Hazarika, S. M. (2000a). Describing rigid body motions in a qualitative theory of spatial regions. In Kautz, H. A., & Porter, B., editors, Proceedings of the National Conference on Artificial Intelligence (AAAI 2000). Menlo Park: AAAI Press/MIT Press, pp. 503509.Google Scholar
Bennett, B., Cohn, A. G., Torrini, P., & Hazarka, S. M. (2000b). A foundation for region-based qualitative geometry. In Horn, W., editor, Proceedings of the European Conference on Artificial Intelligence (ECAI 2000). Amsterdam: IOS Press, pp. 204208.Google Scholar
Biacino, L., & Gerla, G. (1991). Connection structures. Notre Dame Journal of Formal Logic, 32, 242247.CrossRefGoogle Scholar
Biacino, L., & Gerla, G. (1996). Connection structures: Grzegorczyk’s and Whitehead’s definitions of point. Notre Dame Journal of Formal Logic, 37, 431439.CrossRefGoogle Scholar
Borgo, S., Guarino, N., & Masolo, C. (1996). A pointless theory of space based on strong connection and congruence. In Carlucci Aiello, L., Doyle, J., & Shapiro, S. C., editors, Proceedings of the International Conference on Principles of Knowledge Representation and Reasoning (KR 1996). San Francisco: Morgan Kaufmann, pp. 220229.Google Scholar
Casati, R., & Varzi, A. (1999). Parts and Places. The Structure of Spatial Representation. Cambridge: MIT Press.Google Scholar
Clarke, B. L. (1981). A calculus of individuals based on connection. Notre Dame Journal of Formal Logic, 22, 204218.CrossRefGoogle Scholar
Clarke, B. L. (1985). Individuals and points. Notre Dame Journal of Formal Logic, 26, 6175.Google Scholar
Cohn, A. G. (1995). Qualitative shape representation using connection and convex hulls. In Amsili, P., Borillo, M., & Vieu, L., editors, Proceedings of Time, Space and Movement: Meaning and Knowledge in the Sensible World. Toulouse: IRIT, pp. 316 (part C).Google Scholar
Cohn, A. G. (2001). Formalising bio-spatial knowledge. In Welty, C., & Smith, B., editors, Proceedings of the International Conference on Formal Ontology in Information Systems (FOIS 2001). New York: ACM Press, pp. 198209.Google Scholar
Cohn, A. G., Bennett, B., Gooday, J., & Gotts, N. (1997a). Representing and reasoning with qualitative spatial relations. In Stock, O., editor, Spatial and Temporal Reasoning. Dordrecht: Kluwer, pp. 97134.CrossRefGoogle Scholar
Cohn, A. G., Bennett, B., Gooday, J. M., & Gotts, N. (1997b). RCC: A calculus for region based qualitative spatial reasoning. GeoInformatica, 1, 275316.CrossRefGoogle Scholar
Cohn, A. G., & Hazarika, S. M. (2001). Qualitative spatial representation and reasoning: An overview. Fundamenta Informaticae, 46, 129.Google Scholar
Cohn, A. G., & Varzi, A. (2003). Mereotopological connection. Journal of Philosophical Logic, 32, 357390.CrossRefGoogle Scholar
Cristani, M., Cohn, A. G., & Bennet, B. (2000). Spatial locations via morpho-mereology. In Cohn, A. G., Giunchiglia, F., & Selman, B., editors, Proceedings of the International Conference on Principles of Knowledge Representation and Reasoning (KR 2000). San Francisco: Morgan Kaufmann, pp. 1525.Google Scholar
Davis, E. (2006). The expressivity of quantifying over regions. Journal of Logic and Computation, 16, 891916.CrossRefGoogle Scholar
Davis, E., Gotts, N., & Cohn, A. G. (1999). Constraint networks of topological relations and convexity. Constraints, 4, 241280.CrossRefGoogle Scholar
De Laguna, T. (1922). Point, line, and surface. A sets of solids. The Journal of Philosophy, 19, 449461.CrossRefGoogle Scholar
Donnelly, M. (2001). An Axiomatic Theory of Common-Sense Geometry, PhD Thesis, University of Texas.Google Scholar
Donnelly, M. (2004). On parts and holes: The spatial structure of the human Body. In Fieschi, M., Coiera, E., & Li, Y.-C. J., editors, Proceedings of MedInfo 2004. Amsterdam: IOS Press.Google Scholar
Dugat, V., Gambarotto, P., & Larvor, Y. (1999). Qualitative theory of shape and orientation. In Rodriguez, R. V., editor, Proceedings of the Workshop on Hot Topics in Spatial and Temporal Reasoning, International Joint Conference on Artificial Intelligence (IJCAI 1999). pp. 4553.Google Scholar
Düntsch, I., Wang, H., & McCloskey, S. (2001) A relation algebraic approach to the region connection calculus. Theoretical Computer Science, 255, 6383.CrossRefGoogle Scholar
Galton, A. (1996). Taking dimension seriously in qualitative spatial reasoning. In Wahlster, W., editor, Proceedings of the European Conference on Artificial Intelligence (ECAI 1996). Chichester: John Wiley & Sons, pp. 501505.Google Scholar
Galton, A. (2000). Qualitative Spatial Change. New York: Oxford University Press.CrossRefGoogle Scholar
Galton, A. (2004). Multidimensional mereotopology. In Dubois, D., Welty, C., & Williams, M.-A., editors, Proceedings of the International Conference on the Principles of Knowledge Representation and Reasoning (KR 2004). Menlo Park: AAAI Press, pp. 4554.Google Scholar
Gerla, G. (1994). Pointless geometries. In Buekenhout, F., editor, Handbook of Incidence Geometry. Amsterdam: Elsevier, pp. 10151031.Google Scholar
Gotts, N. (1996). Formalizing commonsense topology: The INCH calculus. In Kautz, H., & Selman, B., editors, Proceedings of the International Symposium on Artificial Intelligence and Mathematics (AI/MATH 1996). pp. 7275.Google Scholar
Grzegorczyk, A. (1960). Axiomatizability of geometry without points. Synthese, 12, 228235.CrossRefGoogle Scholar
Hodges, W. (1997). A Shorter Model Theory. Cambridge: Cambridge University Press.Google Scholar
Knauff, M., Rauh, R., & Renz, J. (1997). A cognitive assessment of topological spatial relations: Results from an empirical investigation. In Hirtle, S. C., & Frank, A. U., editors, Proceedings of the International Conference on Spatial Information Theory (COSIT 1997). Berlin: Springer, pp. 193206.Google Scholar
Li, S., & Ying, M. (2003). Region connection calculus: Its models and composition table. Artificial Intelligence, 145, 121146.CrossRefGoogle Scholar
Lobachevskii, N. I. (1835). New principles of geometry with complete theory of parallels (in Russian). Polnoe Sobranie Socinenij, 2.Google Scholar
Masolo, C., & Vieu, L. (1999). Atomicity vs. infinite divisibility of space. In Freksa, C., & Mark, D. M., editors, Proceedings of the International Conference on Spatial Information Theory (COSIT 1999). Berlin: Springer, pp. 233250.Google Scholar
Muller, P. (1998a). Éléments d’une théorie du mouvement pour la modélisation du raisonnement spatio-temporel de sens commun., PhD Thesis, Université Paul Sabatier.Google Scholar
Muller, P. (1998b). A qualitative theory of motion based on spatio-temporal primitives. In Cohn, A. G., Schubert, L., & Shapiro, S. C., editors, Proceedings of the International Conference on Principles of Knowledge Representation and Reasoning (KR 1998). San Francisco: Morgan Kaufmann, pp. 131141.Google Scholar
Munkres, J. (2000). Topology. Upper Saddle Rivere: Prentice Hall.Google Scholar
Nicod, J. (1924). La Géométrie dans le monde sensible. Paris: Presse Universitaire de France.Google Scholar
Pratt-Hartmann, I. (1999). First-order qualitative spatial representation languages with convexity. Journal of Spatial Cognition and Computation, 1, 181204.CrossRefGoogle Scholar
Pratt-Hartmann, I., & Lemon, O. (1997). Ontologies for plane, polygonal mereotopology. Notre Dame Journal of Formal Logic, 38, 225245.Google Scholar
Pratt-Hartmann, I., & Schoop, D. (1998). A complete axiom system for polygonal mereo-topology of the real plane. Journal of Philosophical Logic, 27, 621658.CrossRefGoogle Scholar
Pratt-Hartmann, I., & Schoop, D. (2000). Expressivity in polygonal, plane mereotopology. Journal of Symbolic Logic, 65, 822838.CrossRefGoogle Scholar
Pratt-Hartmann, I., & Schoop, D. (2002). Elementary polyhedral mereotopology. Journal of Philosophical Logic, 31, 469498.CrossRefGoogle Scholar
Randell, D. A., & Cohn, A. G. (1989). Modelling topological and metrical properties in physical processes. In Brachman, R. J., Levesque, H. J., & Reiter, R., editors, Proceedings of the International Conference on Principles of Knowledge Representation and Reasoning (KR 1989). San Francisco: Morgan Kaufmann, pp. 357368.Google Scholar
Randell, D. A., & Cohn, A. G. (1992). A spatial logic based on regions and connections. In Nebel, B., Rich, C., & Swartout, W., editors, Proceedings of the International Conference on Principles of Knowledge Representation and Reasoning (KR 1992). San Francisco: Morgan Kaufmann, pp. 165176.Google Scholar
Renz, J., Rauh, R., & Knauff, M. (2000). Towards cognitive adequacy of topological spatial relations. In Freksa, C., Brauer, W., Habel, C., & Wender, K. F., editors, Spatial Cognition II, Integrating Abstract Theories, Empirical Studies, Formal Methods, and Practical Applications. Berlin: Springer, pp. 184197.CrossRefGoogle Scholar
Roeper, P. (1997). Region-based topology. Journal of Philosophical Logic, 26, 251309.CrossRefGoogle Scholar
Schmidt, H.-J. (1979). Axiomatic Characterization of Physical Geometry. Berlin–Heidelberg: Springer-Verlag.Google Scholar
Schulz, S., & Hahn, U. (2001). Mereotopological reasoning about parts and (w)holes in bio-ontologies. In Welty, C., & Smith, B., editors, Proceedings of the International Conference on Formal Ontology in Information Systems (FOIS 2001). New York: ACM Press, pp. 210221.Google Scholar
Simons, P. (1987). Parts: A Study in Ontology. Oxford: Clarendon Press.Google Scholar
Smith, B. (1998). Basic concepts of formal ontology. In Guarino, N., editor, Proceedings of the International Conference on Formal Ontology in Information Systems (FOIS 1998). Amsterdam: IOS Press, pp. 1928.Google Scholar
Smith, B., & Varzi, A. (1999). The niche. Nous, 33, 214238.CrossRefGoogle Scholar
Stell, J. G. (2000). Boolean connection algebras; a new approach to the region-connection calculus. Artificial Intelligence, 122, 111136.CrossRefGoogle Scholar
Stock, O., editor. (1997). Spatial and Temporal Reasoning. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Tarski, A. (1956a). Foundations of the geometry of solids. In Corcoran, J., editor, Logic, Semantics, Metamathematics. Oxford: Oxford University Press, pp. 2430.Google Scholar
Tarski, A. (1956b). A general theorem concerning primitive notions of Euclidean geometry. Indagationes Mathematicae, 18, 468474.CrossRefGoogle Scholar
Van Benthem, J. (1983). The Logic of Time. Dordrecht: Kluwer.CrossRefGoogle Scholar
Vieu, L. (1997). Spatial representation and reasoning in artificial intelligence. In Stock, O., editor, Spatial and Temporal Reasoning. Dordrecht: Kluwer, pp. 541.CrossRefGoogle Scholar
Whitehead, A. N. (1929). Process and Reality. An Essay in Cosmology. New York: Macmillan.Google Scholar