Skip to main content
Log in

Realism, rhetoric, and reliability

  • S.I.: The Philosophy of Clark Glymour
  • Published:
Synthese Aims and scope Submit manuscript

Abstract

Ockham’s razor is the characteristic scientific penchant for simpler, more testable, and more unified theories. Glymour’s early work on confirmation theory (1980) eloquently stressed the rhetorical plausibility of Ockham’s razor in scientific arguments. His subsequent, seminal research on causal discovery (Spirtes et al. 2000) still concerns methods with a strong bias toward simpler causal models, and it also comes with a story about reliability—the methods are guaranteed to converge to true causal structure in the limit. However, there is a familiar gap between convergent reliability and scientific rhetoric: convergence in the long run is compatible with any conclusion in the short run. For that reason, Carnap (1945) suggested that the proper sense of reliability for scientific inference should lie somewhere between short-run reliability and mere convergence in the limit. One natural such concept is straightest possible convergence to the truth, where straightness is explicated in terms of minimizing reversals of opinion (drawing a conclusion and then replacing it with a logically incompatible one) and cycles of opinion (returning to an opinion previously rejected) prior to convergence. We close the gap between scientific rhetoric and scientific reliability by showing (1) that Ockham’s razor is necessary for cycle-optimal convergence to the truth, and (2) that patiently waiting for information to resolve conflicts among simplest hypotheses is necessary for reversal-optimal convergence to the truth.

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.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. More precisely, that is the second of three accounts presented in (Popper 1959).

  2. Cf. Niiniluoto (1987) for objections to Popper’s attempt to link corroboration to verisimilude.

  3. All of the results that follow still hold if (i) is weakened to: for each \(E, F \in \mathcal{I}\) and \(w \in E \cap F\), there exists \(G \in \mathcal{I}(w)\) such that \(G \subseteq E \cap F\).

  4. Schulte (2000) examines a more realistic framework with a similar structure, in which conservation laws are inferred from real physical data concerning particle reactions.

  5. This setup is essentially the same as the language learning paradigm in formal learning theory (Osherson et al. 1986).

  6. The “locking” terminology is from (Osherson et al. 1986).

  7. In this paper, the focus is on non-stochastic methods and theories. In the stochastic case, deductive methods guarantee an arbitrarily small chance of error in all worlds, whereas inductive methods allow for an arbitrarily high chance of error in some worlds.

  8. Local closure is the underlying, topological structure behind the characterization theorem in Angluin (1980), and the tip-offs in Martin and Osherson (1998).

  9. In fact, standard prior probabilities, conditioned on simple data, will result in posterior probabilities concentrated over such downward-closed disjunctions, due to the Bayes factor.

  10. In algebraic geometry, a natural question on a topological space is called a stratification of the space (Stacks Project 0000; Stratification 2012).

  11. In our general topological setting, “least recently” is a bit more subtle than it sounds. Suppose that the simplest articulation of paradigm P in information state E is \(P_{n}\) and the simplest articulation of paradigm T in E is \(T_{m}\). Say that P was re-articulated less recently than T iff there exists information state \(F \supseteq E\) such that the simplest articulations of T in F are all incompatible with \(T_{m}\), but \(P_{n}\) is also the simplest articulation of P compatible with F.

  12. Baltag et al. (2015) provide a stronger and more general result: every solvable problem has a locally closed refinement with a solution that satisfies postulates K1, K2, K3, K4, K6, K7, K8 of Gärdenfors (1988). It follows that the solution is cycle-free. Our result was arrived at independently, in the more restricted setting of natural problems.

References

  • Angluin, D. (1980). Inductive inference of formal languages from positive data. Information and Control, 45(2), 117–135.

    Article  Google Scholar 

  • Baltag, A., Gierasimczuk, N., & Smets, S. (2014). Epistemic topology (to appear). ILLC Technical Report.

  • Baltag, A., Gierasimczuk, N., & Smets, S. (2015). On the solvability of inductive problems: a study in epistemic topology (forthcoming). Proceedings of the Fifteenth Conference on Theoretical Aspects of Rationality and Knowledge.

  • Carlucci, L., & Case, J. (2013). On the necessity of U-Shaped learning. Topics in Cognitive Science, 5(1), 56–88.

    Article  Google Scholar 

  • Carlucci, L., Case, J., Jain, S., & Stephan, F. (2005). Non U-shaped vacillatory and team learning. In: Algorithmic Learning Theory, (pp. 241–255). Berlin: Springer.

  • Carnap, R. (1945). On inductive logic. Philosophy of Science, 12(2), 72.

    Article  Google Scholar 

  • Case, J., & Smith, C. (1983). Comparison of identification criteria for machine inductive inference. Theoretical Computer Science, 25(2), 193–220.

    Article  Google Scholar 

  • de Brecht, M., & Yamamoto, A. (2009). Interpreting learners as realizers for \({\varSigma }_{2}^{0}\)-measurable functions. 74, 39–44.

  • Fitzpatrick, S. (2013). Kelly on Ockham’s razor and truth-finding efficiency. Philosophy of Science, 80(2), 298–309.

    Article  Google Scholar 

  • Gärdenfors, P. (1988). Knowledge in Flux. Cambridge: MIT Press.

    Google Scholar 

  • Gierasimczuk, N. (2010). Knowing One’s Limits: Logical Analysis of Inductive Inference. : Institute for Logic, Language and Computation.

  • Glymour, C. (1980). Theory and Evidence. Princeton: Princeton University Press.

    Google Scholar 

  • Hamblin, C. L. (1973). Questions in montague english. Foundations of Language, 10, 41–53.

    Google Scholar 

  • Hintikka, J. (2007). Socratic Epistemology: Explorations of Knowledge-Seeking by Questioning. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Jain, S., Osherson, D. N., Royer, J. S., & Sharma, A. (1999). Systems that Learn: an Introduction to Learning Theory. Cambridge: MIT press.

    Google Scholar 

  • Kelly, K. T. (1996). The Logic of Reliable Inquiry. New York: Oxford University Press.

    Google Scholar 

  • Kelly, K. T. (2002). Efficient convergence implies Ockham’s razor. Proceedings of the 2002 International Workshop on Computational Models of Scientific Reasoning and Applications.

  • Kelly, K. T. (2007a). Ockham’s razor, empirical complexity, and truth-finding efficiency. Theoretical Computer Science, 383(2), 270–289.

    Article  Google Scholar 

  • Kelly, K. T. (2007b). Simplicity, truth, and the unending game of science.

  • Kelly, K. T. (2011). Simplicity, truth and probability. In: Bandyopadhyay, P. S., & Forster, M. (eds.), Handbook of the Philosophy of Science Volume 7: Philosophy of Statistics. North Holland: Elsevier.

  • Kelly, K. T., & Glymour, C. (2004). Why probability does not capture the logic of scientific justification. In: Hitchcock, C. (eds.), Debates in the Philosophy of Science, (pp. 94–114). London: Blackwell.

  • Kelly, K. T., & Mayo-Wilson, C. (2010). Causal conclusions that flip repeatedly and their justification. Proceedings of the Twenty-Sixth Conference on Uncertainty in Artificial Intelligence, (pp. 277–286).

  • Kuhn, T. S. (1962). The Structure of Scientific Revolutions. Chicago: University of Chicago Press.

    Google Scholar 

  • Lakatos, I. (1970). Falsification and the methodology of scientific research programmes. Criticism and the Growth of Knowledge, 4, 91–196.

    Article  Google Scholar 

  • Luo, W., & Schulte, O. (2006). Mind change efficient learning. Information and Computation, 204(6), 989–1011.

    Article  Google Scholar 

  • Martin, E., & Osherson, D. N. (1998). Elements of Scientific Inquiry. Cambridge: MIT Press.

    Google Scholar 

  • Martin, E., Sharma, A., & Stephan, F. (2006). Unifying logic, topology, and learning in parametric logic. Theoretical Computer Science, 350(1), 103–124.

    Article  Google Scholar 

  • Müller, F. M. (1884). The Katha Upanishad. In The Upanishads. New York: Oxford University Press.

    Google Scholar 

  • Niiniluoto, I. (1987). Truthlikeness (Vol. 185). Dordrecht: Springer Science & Business Media.

    Book  Google Scholar 

  • Osherson, D. N., Stob, M., & Weinstein, S. (1986). Systems that Learn: An Introduction to Learning Theory for Cognitive and Computer Scientists. Cambridge: The MIT Press.

    Google Scholar 

  • Popper, K. R. (1959). The Logic of Scientific Discovery. London: Hutchinson.

    Google Scholar 

  • Putnam, H. (1965). Trial and error predicates and the solution to a problem of mostowski. Journal of Symbolic logic, 30, 49–57.

    Article  Google Scholar 

  • Roberts, C. (2004). Context in dynamic interpretation. In: Horn, L., & Ward, G. (eds.), Handbook of Contemporary Pragmatic Theory, (pp. 197–220). London: Blackwell.

  • Robins, J. M., Scheines, R., Spirtes, P., & Wasserman, L. (2003). Uniform consistency in causal inference. Biometrika, 90(3), 491–515.

    Article  Google Scholar 

  • Schulte, O. (1999). Means-ends epistemology. The British Journal for the Philosophy of Science, 50(1), 1–31.

    Article  Google Scholar 

  • Schulte, O. (2000). Inferring conservation laws in particle physics: A case study in the problem of induction. The British Journal for the Philosophy of Science, 51(4), 771–806.

    Article  Google Scholar 

  • Schulte, O., Luo, W., & Greiner, R. (2007). Mind change optimal learning of Bayes net structure. In: 20th Annual Conference on Learning Theory (COLT), San Diego, CA, June 12–15.

  • Schulte, O., Frigo, G., Greiner, R., & Khosravi, H. (2010). The IMAP hybrid method for learning Gaussian Bayes nets. In: Advances in Artificial Intelligence, (pp. 123–134). Berlin:Springer.

  • Sharma, A., Stephan, F., & Ventsov, Y. (1997). Generalized notions of mind change complexity. Proceedings of the tenth annual conference on Computational learning theory, (pp. 96–108). ACM.

  • Spirtes, P., Glymour, C., & Scheines, R. (2000). Causation, Prediction, and Search (Vol. 81). Cambridge: MIT press.

    Google Scholar 

  • Stacks Project. Partitions and stratifications, http://stacks.math.columbia.edu/tag/09XY.

  • Stratification. (2012). The Encyclopedia of Mathematics. http://www.encyclopediaofmath.org/index.php?title=Stratification.

  • Vickers, S. (1996). Topology Via Logic. Cambridge: Cambridge University Press.

    Google Scholar 

  • Yamamoto, A., & de Brecht, M. (2010). Topological properties of concept spaces (full version). Information and Computation, 208(4), 327–340.

    Article  Google Scholar 

Download references

Acknowledgments

We are indebted to Thomas Icard for suggesting connections between our topological conception of simplicity and related work in the semantics of provability logic, which proved to be very fruitful. We are also indebted to Alexandru Baltag, Nina Gierasimczuk, and Sonja Smets, for comments and discussions, for informing us of related work by de Brecht and Yamamoto (2009), and for sharing their related results with us, particularly their sharpening of our proposition 12. We thank the anonymous referees. Referee number 3 was particularly thorough and perceptive. We are indebted for discussions following invited presentations of earlier versions of the preceding results at a number of venues: The University of Amsterdam (the 2011 Beth/Vienna Circle Lecture), the ILLC in Amsterdam (2012), the Sorbonne (2012, 2013), Ludwig-Maximilians Univerity in Munich (2012), the University of Copenhagen (2012), The Perimeter Institute for Theoretical Physics (2010), the CSLI at Stanford (2012), and the Argentinian Society for Philosophical Analysis (2012). This work was supported by John Templeton Foundation grant number RQ-8236.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hanti Lin.

Proofs and Lemmas

Proofs and Lemmas

Proof (proposition 1)

\(\Rightarrow :\) Suppose that \(A=O\cap C\) for \(O,C^c\) open. Then

which is an intersection of two closed sets and therefore closed. The final equality follows from the fact that \(\overline{O\cap C} \cap C^c \subseteq \overline{O} \cap \overline{C} \cap C^c = \overline{O} \cap C \cap C^c = \varnothing \).

\(\Leftarrow : \) Suppose that is closed. Since holds, for every A, we have that A is a difference of closed sets, and is, therefore, locally closed. \(\square \)

Proof (proposition 2)

\(\Rightarrow \) Suppose that \(\lambda \) solves \(\mathfrak {P}\). For each w, choose \(E_w \in \mathsf {Lock}(\lambda ,w)\). Then \(\{E_w : w\in W \}\) is a (countable) cover of W by locking information. Let \(F_w= \bigcup \{ E \in \mathcal{I}: E\subset E_w \text { and } \lambda (E) \ne \mathcal{Q}(w)\}\). Then \(E_w\setminus F_w\) is locally closed. We claim that \(A = \bigcup _{w\in A} E_w\setminus F_w\) for each \(A\in \mathcal{Q}\). Let \(w\in A\). Then \(w\in E_w\), and \(w\notin F_w\), since \(E_w\) is locking for w. So \(w\in E_w\setminus F_w\subseteq \bigcup _{v\in A} E_v\setminus F_v\). Suppose that \(v \in \bigcup _{w\in A} E_w \setminus F_w\). Then, for some \(w\in A\), \(v\in E_w\setminus F_w\). Suppose that \(\mathcal{Q}(v) = B\ne A\). Then there is \(E_v \in \mathsf {Lock}(\lambda ,v)\), and \(\lambda (E_v\cap E_w) = B \ne A\). But then \(v\in F_w\). Contradiction. We have shown that each \(A\in \mathcal{Q}\) is a countable union of locally closed propositions.

\(\Leftarrow :\) Suppose that for each answer A to \(\mathcal{Q}\) there is a countable collection \(\mathcal{C}_{A}\) of locally closed sets such that \(A = \bigcup \mathcal{C}_{A}\). Let \(\{C_{i}: i \in \mathbb {N}\}\) enumerate \(\bigcup _{A \in \mathcal{Q}}\mathcal{C}_{A}\). For each i, let \(E_{i}\), \(F_{i}\) be open such that \(C_{i} = E_{i} \setminus F_{i}\). Define:

$$\begin{aligned} \sigma (E) = {\left\{ \begin{array}{ll} \underset{i \in \mathbb {N}}{\min }\; E \subseteq E_i {\text { and }} E \nsubseteq F_i &{}\,\,\, {\text { if defined}}; \\ \omega &{}\,\,\, {\text { otherwise}}. \\ \end{array}\right. } \end{aligned}$$

Furthermore, define:

$$\begin{aligned} \lambda (E) = {\left\{ \begin{array}{ll} \mathcal{Q}(E_{\sigma (E)}\setminus F_{\sigma (E)})&{}\,\,\, {\text { if }} \sigma (E) < \omega ; \\ \mathcal{Q}(E) &{}\,\,\, {\text { otherwise}}. \\ \end{array}\right. } \end{aligned}$$

Let \(w\in W\). Let i be least in \(\mathbb {N}\) that \(w\in E_i\setminus F_i\). Then, for each \(j<i\), either \(w \notin E_j\), or \(w\in F_j\). Let:

$$\begin{aligned} E ~=~ \left( \underset{j<i \text { and } w\in F_j}{\bigcap } F_j\right) \cap E_i, \end{aligned}$$

under the usual convention that \(\bigcap _{j < 0} F_{j} = W\). Let \(F \in \mathcal{I}(w)\) such that \(F \subseteq E\). Then \(F \subseteq E\subseteq E_i\) and, since \(F\in \mathcal{I}(w)\), we have that \(F\nsubseteq F_i\). Furthermore, since \(F\in \mathcal{I}(w)\), it follows that \(F\nsubseteq E_j\), or \(F\subseteq F_j\), for all \(j<i\). So \(\lambda (F) = \mathcal{Q}(E_i\setminus F_i) = \mathcal{Q}(w)\). So \(E\in \mathsf {Lock}(\lambda ,w)\), as required. \(\square \)

Proof (proposition 3)

Recall the solution given in the proof of proposition 2. Let \(E \in \mathcal{I}\) be non-empty. Suppose that \(\lambda (E) = \mathcal{Q}(E)\). Then \(E \cap \lambda (E) = E \cap \mathcal{Q}(E) = E \ne \varnothing \). Suppose that \(\lambda (E) = \mathcal{Q}(E_{\sigma (E)}\setminus F_{\sigma (E)})\). Then \(E\subseteq E_{\sigma (E)}\) and \(E \not \subseteq F_{\sigma (E)}\), so \(E \cap \lambda (E) = E \cap \mathcal{Q}(E_{\sigma (E)}\setminus F_{\sigma (E)}) \ne \varnothing \). \(\square \)

Proof (proposition 4)

The result follows immediately from proposition 2 and the fact that a set is locally closed in the subspace topology \(\mathcal{I}|_{C}\) iff it is the intersection of C with a locally closed set. \(\square \)

Proof (proposition 5)

Let H be in \(\mathcal{Q}^{*}{\Vert }_{E}\). \(\Rightarrow \) Suppose that H is not downward closed in \(\mathcal{Q}^{*}{\Vert }_{E}\). So there exist AB in \(\mathcal{Q}^{*}{\Vert }_{E}\) such that \(A \subseteq H\) and \(B \prec A\), but \(B \not \subseteq H\). Since B and H are disjunctions of answers, there exists answer \(C \subseteq B\) such that C is disjoint from H. Then \(C \prec H\), so H is not minimal in \(\mathcal{Q}^{*}{\Vert }_{E}\).

\(\Leftarrow \) Suppose that H is not minimal in \(\mathcal{Q}^{*}{\Vert }_{E}\). Then there exists \(A \in \mathcal{Q}^{*}{\Vert }_{E}\) such that \(A \prec H\). Hence, A is disjoint from H, so H is not downward closed in \(\mathcal{Q}^{*}{\Vert }_{E}\). \(\square \)

Proof (proposition 6)

For the \(\Leftarrow \) direction of the first equivalence, suppose that , for all \(A\in \mathcal{Q}^*\), and for \(B\in \mathcal{Q}\), . Then . But since \(\mathcal{Q}\) is a partition, \(B\subseteq A_\prec \). For the \(\Rightarrow \) direction, suppose that \(\mathcal{Q}\) is homogeneous, \(A\in \mathcal{Q}^*\), and . Then by homogeneity, \(w\in \mathcal{Q}(w)\prec A\). So \(w\in A_\prec \).

The second equivalence is an immediate consequence of proposition 1. \(\square \)

Proof (proposition 7)

Let \(A,B,C\in \mathcal{Q}\), \(A\prec B\), and \(B\prec C\). Thus, \(A\subseteq \overline{B} \subseteq \overline{C}\). So it remains only to show that \(A\ne C\). Each proposition is disjoint from its frontier, so . Furthermore, if B locally closed is open, by proposition 1. So, since , is an open set containing B and disjoint from A. Therefore \(B\nprec A\), and \(A\ne C\) as required. \(\square \)

Proof (proposition 8)

Let A be in \(\mathcal{Q}{\Vert }_{E}\), and let B be in \(\mathcal{Q}^{*}{\Vert }_{E}\), for E in \(\mathcal{I}\). Then E is not disjoint from A, B. Suppose that \(A \prec _{W} B\). Then, in general, \(A \prec _{E} B\). Next, suppose that \(A \not \prec _{W} B\) but \(A \prec _{E} B\). Then \(A \cap E \prec _{W} B\). So \(\mathcal{Q}\) is not homogeneous for \(\mathfrak {I}\). \(\square \)

Proof (proposition 9)

Define the method:

$$\begin{aligned} \lambda (E) = {\left\{ \begin{array}{ll} \bigcup {Simp}(\mathcal{Q}{\Vert }_{E}) &{}\,\,\, {\text { if }}{Simp}(\mathcal{Q}{\Vert }_{E})\hbox { is co-initial in }\mathcal{Q}{\Vert }_{E}, \\ \bigcup \mathcal{Q}{\Vert }_{E} &{}\,\,\, {\text { otherwise}}. \\ \end{array}\right. } \end{aligned}$$

Clearly, \(\lambda \) is patient. Let \(w\in W\). Since \(\mathcal{Q}\) is natural, \(\mathcal{Q}(w)_\prec \) is closed, by proposition 6. so \(\mathcal{Q}(w)_\prec ^c\) is an open neighborhood of w. By normality, \(Q(w)_\succeq \) is an open neighborhood of w as well. Let \(E = Q(w)_\succeq \cap \mathcal{Q}(w)_\prec ^c\). Then \(\mathcal{Q}(w) = \bigcup Simp(\mathcal{Q}{\Vert }_{E}) = \lambda (E)\). So \(E\in \mathsf {Lock}(\lambda ,w)\), by proposition 8. \(\square \)

Proof (proposition 10)

Omitted. \(\square \)

Proof (proposition 11)

Suppose that \(\lambda \) is a non-Ockham solution to problem \(\mathfrak {P}\). Then there exists information \(E_0\in \mathcal{I}\) and answer A in \(\mathcal{Q}{\Vert }_{E_0}\) such that \(A \prec \lambda (E_{0})\). Let \(H = \lambda (E_{0})\). Let world w be in \(A\cap E_0\). Since \(\lambda \) is a solution, there exists \(E_1\) in \(\mathsf {Lock}(\lambda ,w)\). There exists world v in \(E_0 \cap E_1 \cap H\), since \(A \prec H\). There exists \(E_2 \in \mathsf {Lock}(\lambda , v)\), because \(\lambda \) is a solution, so \(\lambda (E_{2}) = \mathcal{Q}(v)\). Let \(B = \mathcal{Q}(v)\). Let \(e = (E_0, ~E_0 \cap E_1,~ E_0\cap E_1 \cap E_2)\). Then \(\lambda (e) = (H, A, B)\), since \(E_1\), \(E_2\) are locking. Note that \(B \subseteq H\), because v is in H. Furthermore, A is disjoint from H, and therefore from B, because \(A \prec H\). So (HAB) is a cycle sequence. \(\square \)

Proof (lemma 1)

Enumerate \(\mathcal{Q}\) as \(H_1, H_2, \ldots \). Define:

$$\begin{aligned} \preceq _0 = & {} \preceq \\ \preceq _{n+1} = & {} \{ (H_i, H_j) : H_i \preceq _n H_{n+1} \text { and } H_j |_n H_{n+1} \} \cup \preceq _n; \end{aligned}$$

where \(H_i |_n H_j\) holds iff \(H_i\) and \(H_j\) are not ordered by \(\preceq _n\).

We show, by induction on n, that \(\preceq _{n}\) is a partial order on \(\mathcal{Q}\) such that \(H_{\prec _n}\) is closed for all \(H\in \mathcal{Q}\). The base case follows from propositions 7 and 6. For induction, suppose that \(\preceq _{n}\) is a partial order.

Reflexivity Relation \(\preceq _{n}\) is reflexive, and \(\preceq _{n+1}\) extends \(\preceq _{n}\).

Antisymmetry Suppose that \(H_i \preceq _{n+1} H_j\) and \(H_j \preceq _{n+1} H_i\). If we have also that \(H_i \preceq _n H_j\) and \(H_j \preceq _n H_i\), then we are done. So suppose that \(H_i \npreceq _n H_j\) and \(H_j \preceq _n H_i\). Then \(H_i \preceq _n H_{n+1}\), and . But \(H_j\preceq _n H_{n+1}\), by the transitivity of \(\preceq _n\). Contradiction. Finally, suppose that \(H_i \npreceq _n H_j\) and \(H_j \npreceq _n H_i\). Then \(H_i \preceq _n H_{n+1}\); ; \(H_j \preceq _n H_{n+1}\); and . Contradiction.

Transitivity Suppose that \(H_i \preceq _{n+1} H_j\) and \(H_j \preceq _{n+1} H_k\). If \(H_i \preceq _{n} H_j\) and \(H_j \preceq _{n} H_k\), then we are done. So suppose that \(H_i\npreceq _n H_j,H_j\preceq _n H_k\). Then, \(H_i\preceq _n H_{n+1}\) and . If \(H_{n+1} \preceq _n H_k\) then, by transitivity of \(\preceq _n\), we are done. So suppose that . But then, \(H_i\preceq _{n+1} H_k\), as required. Now, suppose that \(H_i\preceq _n H_j,H_j\npreceq _n H_k\). Then, \(H_j\preceq _n H_{n+1}\) and . But then, \(H_i,H_j\preceq _{n+1} H_k\). Finally, suppose that \(H_i\npreceq _n H_j,H_j\npreceq _n H_k\). Then, \(H_i,H_j\preceq _n H_{n+1}\) and . So \(H_i\preceq _{n+1} H_k\), as required.

Next, we show that, if \(H_{\prec _n}\) is closed, for all \(H\in \mathcal{Q}\), then \(H_{\prec _{n+1}}\) is also closed, for all \(H\in \mathcal{Q}\). Consider arbitrary \(H \in \mathcal{Q}\). If H is ordered with \(H_{n+1}\) by \(\preceq _n\), then \(H_{\prec _{n+1}} = H_{\prec _n}\), and we are done. If , then \(H_{\prec _{n+1}} = H_{\prec _n} \cup H_{{n+1}_{\preceq _n}} = H_{\prec _n} \cup H_{{n+1}_{\prec _n}} \cup \overline{H}_{n+1}.\) The final equality follows from the fact that \(\overline{H}_{n+1} = {H_{n+1}}_\preceq \subseteq {H_{n+1}}_{\preceq _n}\). Then \(H_{\prec _{n+1}}\) is closed, since it is a finite union of closed sets.

Take the directed union of the orders:

$$\begin{aligned} \preceq _\omega= & {} \bigcup _{i} \preceq _i. \end{aligned}$$

We show that \(\preceq _\omega \) is a total order.

Reflexivity follows from the fact that \(\preceq \) is extended by \(\preceq _\omega \).

Anti-symmetry If \(H_i \preceq _\omega H_j\) and \(H_j \preceq _\omega H_i\), then \(H_i \preceq _n H_j\) and \(H_j\preceq _m H_i\), for some nm. Suppose that \(n\le m\). Then \(H_i \preceq _m H_j\), since \(\preceq _n \subseteq \preceq _m\). So \(H_i = H_j\), since \(\preceq _m\) is a partial order.

Transitivity Similarly, if \(H_i\preceq _\omega H_j\) and \(H_j \preceq _\omega H_k\), then for some m, \(H_i \preceq _m H_j\) and \(H_j \preceq _m H_k\). So by the transitivity of \(\preceq _m\), \(H_i\preceq _m H_k\). Since \(\preceq _m\) is extended by \(\preceq _\omega \), we have that \(H_i \preceq _\omega H_k\). Totality. Note that \(H_1, H_2, \ldots , H_m\) are totally ordered by \(\preceq _m\). So \(\preceq _\omega \) is total.

It remains to show that for \(H_i \in \mathcal{Q}\), \({H_i}_{\prec _\omega }\) is closed. That follows immediately from the fact that \({H_i}_{\prec _\omega } = {H_i}_{\prec _i}\), which we already showed to be closed. \(\square \)

Proof (proposition 12)

Apply lemma 1 to obtain a total ordering \(\preceq ^*\) of the elements of \(\mathcal{Q}\), such that \(H_{\prec ^*}\) is closed, for every \(H\in \mathcal{Q}\). Define:

$$\begin{aligned} \lambda (E) = \left\{ \begin{array}{l@{\quad }l} \min _{\preceq ^*} (\mathcal{Q}{\Vert }_{E}) &{}\text { if } \min _{\preceq ^*} (\mathcal{Q}{\Vert }_{E}) \ne \varnothing ; \\ \bigcup \mathcal{Q}{\Vert }_{E} &{}\text { otherwise. } \end{array} \right. \end{aligned}$$
(1)

Since \(H_{\prec ^*}\) is refutable, if H is true, then H eventually becomes minimal in \(\preceq ^*\). Once minimal, H will remain minimal until refuted—so \(\lambda \) will never perform a cycle. To see this suppose that \(c=(\lambda (E_0), \lambda (E_1), \lambda (E_2))\) is an output sequence. If \(\lambda (E_0) = \bigcup \mathcal{Q}{\Vert }_{E_0}\), then \(\lambda (E_1) \subseteq \lambda (E_0)\), and c is not a reversal sequence. So suppose that \(\lambda (E_0) = \min _{\preceq ^*} \mathcal{Q}{\Vert }_{E_0}\). But then if \(\lambda (E_0)\) is disjoint from \(\lambda (E_1)\), we have that \(\lambda (E_0) \notin \mathcal{Q}{\Vert }_{E_1}\). Since \(\lambda \) is clearly consistent, c cannot be a cycle sequence. We have shown that \(\lambda \) performs no cycles of length three. Since every cycle sequence contains a cycle sub-sequence of length three, \(\lambda \) performs no cycles. \(\square \)

Lemma 2

For arbitrary \(B\subseteq W,\)

$$\begin{aligned}&A_n|_B \cap \overline{A_{n-1}|_B \cap \overline{\ldots \cap \overline{A_1|_B \cap \overline{A_0|_B } } } }\\&\quad \subseteq ~ A_n|_B \cap \overline{A_{n-1} \cap \overline{\ldots \cap \overline{A_1 \cap \overline{A_0 } } } }. \end{aligned}$$

Proof (lemma 2)

\(A_1|_B \subseteq A_1|_B\). By the induction hypothesis, we have that:

$$\begin{aligned}&A_{n+1}|_B \cap \overline{A_n|_B \cap \overline{A_{n-1}|_B \cap \overline{\ldots \cap \overline{A_1|_B \cap \overline{A_0|_B } } } }}\\&\quad \subseteq ~ A_{n+1}|_B \cap \overline{A_n|_B \cap \overline{A_{n-1} \cap \overline{\ldots \cap \overline{A_1 \cap \overline{A_0 } } } }}\\&\quad \subseteq ~ A_{n+1}|_B \cap \overline{ A_n \cap \overline{A_{n-1} \cap \overline{\ldots \cap \overline{A_1 \cap \overline{A_0 } } } }}, \end{aligned}$$

where, at each step, we use the fact that if \(B\subseteq C\), then \(A\cap \overline{B} \subseteq A\cap \overline{C}\). \(\square \)

Say that reversal sequence \(a = (A_0, \ldots , A_n)\), with \(A_i\) in \(\mathcal{Q}^*\), is forcible in \(\mathfrak {P}\) iff, for every solution \(\lambda \) to \(\mathfrak {P}\), there is an input sequence e, such that \(\lambda (e)\leqslant a\).

Lemma 3

If \(\mathfrak {P}\) is solvable, then reversal sequence \(a = (A_0, \ldots , A_n)\) is forcible in \(\mathfrak {P}\) iff:

$$\begin{aligned} A_0 \cap \overline{ A_1 \cap \overline{\ldots \cap \overline{A_{n-1}\cap \overline{A_n}}}}\ne \varnothing . \end{aligned}$$

Proof (lemma 3)

\(\Leftarrow :\) Suppose that \(\lambda \) is a solution to \(\mathfrak {P}\), and \(A_0 \cap \overline{ A_1 \cap \overline{\ldots \cap \overline{A_{n-1}\cap \overline{A_n}}}}\ne \varnothing \). Let:

$$\begin{aligned} w_0\in & {} A_0 \cap \overline{ A_1 \cap \overline{\ldots \cap \overline{A_{n-1}\cap \overline{A_n}}}};\\ w_{i+1}\in & {} \bigcap _{j\le i} E_{j} \cap A_{i+1} \cap \overline{ A_{i+2} \cap \overline{\ldots \cap \overline{A_{n-1}\cap \overline{A_n}}}}; \end{aligned}$$

where \(E_j \in \mathsf {Lock}(\lambda , w_j)\). Letting \(e=(\bigcap _{j\le i} E_{j})_{i=0}^n\), we have that \(\lambda (e)\leqslant a\).

\(\Rightarrow :\) Suppose that \(\mathfrak {P}\) is solvable. We proceed by induction on n. In the base case, suppose that \(A_0=\varnothing \). By proposition 3, there is a solution \(\lambda \) such that \(\lambda (E)\ne \varnothing \) for all nonempty \(E\in \mathcal{I}\). So \(a=(A_0)\) is not forcible. For the inductive case, suppose that \(A_0 \cap \overline{ A_1 \cap \overline{\ldots \cap \overline{A_{n}\cap \overline{A_{n+1}}}}} = \varnothing \). Let:

$$\begin{aligned} C= & {} A_1 \cap \overline{\ldots \cap \overline{A_{n}\cap \overline{A_{n+1}}}} \end{aligned}$$

By Lemma 2:

$$\begin{aligned} A_1|_{\overline{C}^c} \cap \overline{\ldots \cap \overline{A_{n}|_{\overline{C}^c} \cap \overline{A_{n+1}|_{\overline{C}^c}}}}\subseteq & {} A_1|_{\overline{C}^c} \cap \overline{\ldots \cap \overline{A_{n} \cap \overline{A_{n+1}}}}\\= & {} \overline{C}^c \cap A_1 \cap \overline{\ldots \cap \overline{A_{n} \cap \overline{A_{n+1}}}}\\= & {} \overline{C}^c \cap C = \varnothing . \end{aligned}$$

So, by the induction hypothesis, \((A_1, ..., A_{n+1})\) is not forcible in the \(\mathfrak {P}|_{\overline{C}^c}\) subproblem. Let \(\lambda _2\) be a solution that never performs that reversal sequence in \(\mathfrak {P}|_{\overline{C}^c}\). Furthermore, let \(\lambda _{1}\) be a solution to the \(\mathfrak {P}|_{\overline{C}}\) subproblem. Both solutions exist by proposition 4, since \(\mathfrak {P}\) is solvable. Furthermore, we can assume that solution \(\lambda _{1}\) is consistent, by proposition 3. Define the method:

$$\begin{aligned} \lambda ^*(E) = {\left\{ \begin{array}{ll} \mathcal{Q}(\lambda _1(E \cap \overline{C})) &{}\,\,\, {\text { if }}\overline{C} \cap E \ne \varnothing ; \\ \mathcal{Q}(\lambda _2(E)) &{}\,\,\, {\text { otherwise}}. \\ \end{array}\right. } \end{aligned}$$

Since \(\overline{C}\) is refutable, it is possible to focus on the \(\mathfrak {P}|_{\overline{C}}\) subproblem, until \(\overline{C}\) is refuted. Since, by assumption, \(A_0\cap \overline{C} = \varnothing \), consistent method \(\lambda _1\) never outputs \(A_0\). Therefore, \(\lambda _1\) cannot perform the reversal sequence \((A_0, A_1, \ldots , A_{n+1})\). Furthermore, since \(\lambda _2\) never performs the reversal sequence \((A_1, ..., A_{n+1})\), it follows that \(\lambda ^*\) solves \(\mathfrak {P}\), and never performs the reversal sequence \((A_0, A_1, \ldots , A_{n+1})\). \(\square \)

Proof (proposition 14)

\(\Rightarrow :\) Suppose that \(\lambda \) is a solution to \(\mathfrak {P}\) that is not patient. Then \(\lambda (E)\) is not co-initial in \(\mathcal{Q}{\Vert }_{E}\), for some \(E\in \mathcal{I}\). So there is \(A\in \mathcal{Q}{\Vert }_{E}\) such that \(\lambda (E) \cap \overline{A}=\varnothing \), by homogeneity. Let \(w\in A\cap E\), and let \(F\in \mathsf {Lock}(\lambda ,w)\). Then \((\lambda (E), \lambda (E\cap F))\) is a reversal sequence that is not forcible, by lemma 3. So \(\lambda \) is not reversal-optimal.

\(\Leftarrow :\) Suppose \(\lambda \) is a patient solution and \(b=(\lambda (E_0), \lambda (E_1), \ldots , \lambda (E_n))\) is a reversal sequence. Let \(A_n\) be an element of \(\mathcal{Q}\) such that \(A_n\subseteq \lambda (E_n)\). By patience, there exists \(A_{n-1} \in \mathcal{Q}\), such that \(A_{n-1} \subseteq \lambda (E_{n-1})\) and \(A_{n-1} \prec A_n\). Also by patience, there is an answer \(A_{n-2} \subseteq \lambda (E_{n-2})\) such that \(A_{n-2} \prec A_{n-1}\), and so on. That yields a recipe for constructing a reversal sequence \(a=(A_0, A_1, \ldots , A_n)\), such that \(A_0 \prec A_1 \prec \cdots \prec A_n\) and \(a\le b\). Furthermore, it is easy to check that \(A_0 \cap \overline{ A_1 \cap \overline{\ldots \cap \overline{A_{n-1}\cap \overline{A_n}}}} = A_0\). By lemma 3, the reversal sequence a is forcible. Therefore b is forcible as well. It has been shown that every reversal sequence performed by a patient method is forcible. Therefore, patient solutions are reversal-optimal. \(\square \)

Proof (proposition 16)

By proposition 9, every natural, normal problem has a patient Ockham solution. By proposition 14, every patient Ockham solution is reversal-optimal. It remains to show that the method described in the proof of proposition 9 is cycle-free. Recall that the method is:

$$\begin{aligned} \lambda (E) = {\left\{ \begin{array}{ll} \bigcup {Simp}(\mathcal{Q}{\Vert }_{E}) &{}\,\,\, {\text { if }} {Simp}(\mathcal{Q}{\Vert }_{E}) \hbox { is co-initial in } \mathcal{Q}{\Vert }_{E}, \\ \bigcup \mathcal{Q}{\Vert }_{E} &{}\,\,\, {\text { otherwise}}. \\ \end{array}\right. } \end{aligned}$$

It is easy to see that \(\lambda \) is consistent. Suppose that \(c=(\lambda (E_0), \lambda (E_1), \lambda (E_2))\) is a cycle sequence. If \(\lambda (E_0) = \bigcup \mathcal{Q}{\Vert }_{E}\) then, by consistency, \(\lambda (E_1) \subseteq \lambda (E_0)\), and c is not a reversal sequence. So it must be that \(\lambda (E_0) = \bigcup {Simp}(\mathcal{Q}{\Vert }_{E_0})\). If c is a cycle sequence, \(\lambda (E_2) \subseteq \lambda (E_0) = \bigcup {Simp}(\mathcal{Q}{\Vert }_{E_0})\). Furthermore, by consistency, \(\lambda (E_2) \subseteq \mathcal{Q}{\Vert }_{E_2} \subseteq \mathcal{Q}{\Vert }_{E_1}\). By proposition 8, if \(\lambda (E_2) \subseteq \bigcup {Simp}(\mathcal{Q}{\Vert }_{E_0})\), and \(\lambda (E_2)\subseteq \mathcal{Q}{\Vert }_{E_1}\), then \(\lambda (E_2) \subseteq \bigcup {Simp}(\mathcal{Q}{\Vert }_{E_1})\). So c is not a reversal sequence. Contradiction. We have shown that \(\lambda \) performs no cycles of length three. Since every cycle sequence contains a cycle sub-sequence of length three, \(\lambda \) performs no cycles. \(\square \)

Proof (proposition 17)

Suppose that \(\lambda \) is reversal-optimal in \(\mathfrak {P}\). Then \(\lambda \) is a patient solution to \(\mathfrak {P}\), by proposition 14. By the proof of proposition 14, it follows that every reversal sequence produced by \(\lambda \) is forcible. Hence, if the simplicity order in \(\mathfrak {P}\) has no path of length \(\geqslant 2\), then \(\lambda \) performs no reversal sequence and is, therefore, reversal-admissible. Alternatively, let \(A_{0} \prec A_{1}\), for distinct \(A_{0}, A_{1} \in \mathcal{Q}\). Let \(\lambda '\) solve \(\mathfrak {P}\). It suffices for admissibility of \(\lambda \) to find information states \(E_{0}, E_{1}\) such that \(\lambda (E_{0}, E_{1}) = \lambda '(E_{0}, E_{1})\), for then \(\lambda '(E_{0}, E_{1}) \not < \lambda (E_{0}, E_{1})\). Choose world \(w_{0}\) in \(A_{0}\). Let information state \(F_{0}\) be locking for \(\lambda \) in \(w_{0}\) and let information state \(G_{0}\) be locking for \(\lambda '\) in \(w_{0}\). Thus, information state \(E_{0} = F_{0} \cap G_{0}\) is locking for both \(\lambda \) and \(\lambda '\) in \(w_{0}\). By the definition of \(\prec \), there exists world \(w_{1}\) in \(A_{1} \cap E_{0}\). Let \(F_{1}\) be locking for \(\lambda \) in \(w_{1}\) and let \(G_{1}\) be locking for \(\lambda '\) in \(w_{1}\). Then \(E_{0} \cap F_{1} \cap G_{1}\) is locking for both \(\lambda \) and \(\lambda '\) in \(w_{1}\). Let \(e = (E_{0}, E_{1})\). Then \(\lambda (e) = (A_{0}, A_{1}) = \lambda '(e)\), as desired. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kelly, K.T., Genin, K. & Lin, H. Realism, rhetoric, and reliability. Synthese 193, 1191–1223 (2016). https://doi.org/10.1007/s11229-015-0993-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11229-015-0993-9

Keywords

Navigation