Skip to main content
Log in

Probability in Two Deterministic Universes

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

How can probabilities make sense in a deterministic many-worlds theory? We address two facets of this problem: why should rational agents assign subjective probabilities to branching events, and why should branching events happen with relative frequencies matching their objective probabilities. To address the first question, we generalise the Deutsch–Wallace theorem to a wide class of many-world theories, and show that the subjective probabilities are given by a norm that depends on the dynamics of the theory: the 2-norm in the usual Many-Worlds interpretation of quantum mechanics, and the 1-norm in a classical many-worlds theory known as Kent’s universe. To address the second question, we show that if one takes the objective probability of an event to be the proportion of worlds in which this event is realised, then in most worlds the relative frequencies will approximate well the objective probabilities. This suggests that the task of determining the objective probabilities in a many-worlds theory reduces to the task of determining how to assign a measure to the worlds.

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

Similar content being viewed by others

Notes

  1. Dominance will only be needed to prove Theorem 1, that says that rational agents in a many-worlds theory assign subjective probabilities to the worlds. All other results follow without it, including Theorems 3 and 4, that say that rational agents in Kent’s universe and Many-Worlds bet according to the appropriate version of the Born rule.

  2. For concreteness, convergence is defined in the metric induced by the norm \(\Vert G\Vert := \Vert {\mathbf {c}}\Vert _3 + \Vert {\mathbf {r}}\Vert _1\).

  3. It is untenable in classical probability because in a given situation there are often several different plausible symmetries that give rise to different probability assignments. This problem does not arise here, as the symmetry at hand will be the one between the coefficients of the worlds, or the amplitudes of the quantum state.

  4. To the best of our knowledge there has been no attempt to calculate the number of worlds in an even remotely realistic model of a measurement.

  5. Note that Stern and Gerlach were not aware that they were measuring spin, rather they interpreted the experiment as a proof of Bohr–Sommerfelds spatial quantization hypothesis.

  6. One should take seriously the “hypothetical” here, as theories where the p-norm is preserved are rather pathological. As shown in Ref. [43], the only linear transformations that preserve the p-norm of all vectors for \(p\ne 1,2\) are permutations composed with phases. Here we get around this by only asking the transformation T to preserve the norm of the computational basis.

  7. In this example p must be of the form \(\log _2 n\) for integer \(n\ge 2\), but in general any real \(p\ge 1\) works.

  8. We talk about the roles objective probabilities play instead of the definition of objective probabilities because this is the best we can do. Unlike subjective probabilities, there is no widely accepted definition of objective probability that we could try to satisfy.

  9. One often hears a different story, that in the limit of an infinite number of trials the relative frequency is equal to the objective probability, full stop. This is simply mistaken, but given that the mathematics we present here are identical, those that want to insist on this mistake can do it equally well in many-world theories.

  10. Which does include the counting measure as a particular case, so we are by no means assuming that worlds cannot be counted.

  11. It should be clear that neither the regular nor the reverse Kent’s universe are realistic analogues of Many-Worlds: both of them require an exponential number of worlds, either in the end or in the beginning of the simulation.

  12. Sebens and Carroll claim, however, that an agent that assigns these probabilities can be Dutch-booked [46]: After the first measurement, but before the second, an agent that is ignorant of the outcome will assign probability 1 / 2 to being in each world, and will therefore accept a bet that pays 3€ in the world with outcome 1 and \(-3\)€ in the world with outcome 2. After the second measurement, the agent will assign probability 1 / 3 to being in each world, and will therefore accept a bet that pays \(-4\)€ in the world with a single outcome 1 and 2€ in the worlds with outcomes (2, 1) and (2, 2). If the agent accepts both bets, then in all worlds they will lose 1€.

    The problem with this argument is that the agent would not have accepted the first bet if they knew that they would be multiplied in the world with outcome 2: the assignment of probability 1 / 2 was a mistake, that the agent corrected after learning of the second multiplication. Given that they accepted the incorrect bet, however, the agent knows that accepting the second, fair, bet will cause them to be Dutch-booked, so they would reject it. Note, furthermore, that making two bets about the same situation using two different probabilities generically leads to Dutch-booking, even in single-world theories. Consider an agent that believes a coin to be fair, and therefore accepts a bet that pays 3€ if heads, and \(-3\)€ if tails. If the agent changes their mind, and decides instead that the coin has probability 2 / 3 of coming up tails, and accepts a bet that pays \(-4\)€ if heads and 2€ if tails, then the agent will lose 1€ independently of the result of the coin flip.

    This raises the question of how a realistic agent in Kent’s universe could ever assign probabilities, since they would depend on the whole tree of future branchings. The agent could simply recognize that it is not possible to determine the objective probabilities, and use only the part of the branching tree that they can foresee to calculate their subjective probabilities. Analogously, a grandparent could decide to divide their inheritance among their children weighted by how the number of grandchildren they begat, but refuse to speculate about how many children each grandchild will have.

References

  1. Everett, H.: “Relative state” formulation of quantum mechanics. Rev. Mod. Phys. 29, 454–462 (1957)

    Article  ADS  MathSciNet  Google Scholar 

  2. Finkelstein, D.: The logic of quantum physics. Trans. N. Y. Acad. Sci. 25, 621–637 (1963)

    Article  MathSciNet  Google Scholar 

  3. Hartle, J.B.: Quantum mechanics of individual systems. Am. J. Phys. 36, 704 (1968)

    Article  ADS  Google Scholar 

  4. DeWitt, B.S.: Quantum mechanics and reality. Phys. Today 23, 30 (1970)

    Article  Google Scholar 

  5. Graham, N.: The measurement of relative frequency. In: DeWitt, B., Graham, N. (eds.) The Many-Worlds Interpretation of Quantum Mechanics. Princeton University Press, Princeton (1973)

    Google Scholar 

  6. Farhi, E., Goldstone, J., Gutmann, S.: How probability arises in quantum mechanics. Ann. Phys. 192, 368–382 (1989)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Squires, E.J.: On an alleged “proof” of the quantum probability law. Phys. Lett. A 145, 67–68 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  8. Caves, C.M., Schack, R.: Properties of the frequency operator do not imply the quantum probability postulate. Ann. Phys. 315, 123–146 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Deutsch, D.: Quantum theory of probability and decisions. Proc. R. Soc. Lond. A 455, 3129 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Wallace, D.: Everettian rationality: defending Deutsch’s approach to probability in the Everett interpretation. Stud. Hist. Philos. Sci. B 34, 415–439 (2003)

    MathSciNet  MATH  Google Scholar 

  11. Wallace, D.: Quantum probability from subjective likelihood: improving on Deutsch’s proof of the probability rule. Stud. Hist. Philos. Sci. B 38, 311–332 (2007)

    MathSciNet  MATH  Google Scholar 

  12. Wallace, D.: A formal proof of the Born rule from decision-theoretic assumptions. In: Saunders, S., Barrett, J., Kent, A., Wallace, D. (eds.) Many Worlds? Everett, Quantum Theory & Reality. Oxford University Press, Oxford (2010)

    Google Scholar 

  13. Zurek, W.H.: Environment-assisted invariance, entanglement, and probabilities in quantum physics. Phys. Rev. Lett. 90, 120404 (2003)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Vaidman, L.: Probability in the many-worlds interpretation of quantum mechanics. In: Ben-Menahem, Y., Hemmo, M. (eds.) The Probable and the Improbable: Understanding Probability in Physics, Essays in Memory of Itamar Pitowsky. Springer, Berlin (2011)

    Google Scholar 

  15. Carroll, S.M., Sebens, C.T.: Many worlds, the Born rule, and self-locating uncertainty. In: Struppa, D.C., Tollaksen, J.M. (eds.) Quantum Theory: A Two-Time Success Story, p. 157. Springer, Milano (2014)

    Chapter  Google Scholar 

  16. Kent, A.: One world versus many: the inadequacy of Everettian accounts of evolution, probability, and scientific confirmation. In: Saunders, S., Barrett, J., Kent, A., Wallace, D. (eds.) Many Worlds? Everett, Quantum Theory & Reality. Oxford University Press, Oxford (2010)

    Google Scholar 

  17. Gill, R.D.: On an argument of David Deutsch. In: Schürmann, M., Franz, U. (eds.) Quantum Probability and Infinite Dimensional Analysis: From Foundations to Applications, pp. 277–292. World Scientific, Hackensack (2005)

    Chapter  Google Scholar 

  18. Baker, D.J.: Measurement outcomes and probability in Everettian quantum mechanics. Stud. Hist. Philos. Sci. B 38, 153–169 (2007)

    MathSciNet  MATH  Google Scholar 

  19. Albert, D.: Probability in the Everett picture. In: Saunders, S., Barrett, J., Kent, A., Wallace, D. (eds.) Many Worlds? Everett, Quantum Theory & Reality. Oxford University Press, Oxford (2010)

    Google Scholar 

  20. Dawid, R., Thébault, K.: Many worlds: decoherent or incoherent? Synthese 192, 1559–1580 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  21. Mandolesi, A.L.G.: Analysis of Wallace’s proof of the Born rule in Everettian quantum mechanics: formal aspects. Found. Phys. 48, 751–782 (2018)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Mandolesi, A.L.G.: Analysis of Wallace’s proof of the Born rule in Everettian quantum mechanics II: concepts and axioms. arXiv:1803.08762 [quant-ph]

  23. Wallace, D.: The Emergent Multiverse: Quantum Theory According to the Everett Interpretation. Oxford University Press, Oxford (2012)

    Book  MATH  Google Scholar 

  24. Joyce, J.: The Foundations of Causal Decision Theory. Cambridge University Press, Cambridge (2008)

    Google Scholar 

  25. Savage, L.: The Foundations of Statistics. Dover Publications, New York (1972)

    MATH  Google Scholar 

  26. van Fraassen, B.: Indifference: the symmetries of probability. In: van Fraassen, B. (ed.) Laws and Symmetry. Oxford University Press, Oxford (1989)

    Chapter  Google Scholar 

  27. Saunders, S.: Decoherence, relative states, and evolutionary adaptation. Found. Phys. 23, 1553–1585 (1993)

    Article  ADS  MathSciNet  Google Scholar 

  28. Wallace, D.: Everett and structure. Stud. Hist. Philos. Sci. B 34, 87–105 (2003)

    MathSciNet  MATH  Google Scholar 

  29. Zurek, W.H.: Decoherence, einselection, and the quantum origins of the classical. Rev. Mod. Phys. 75, 715–775 (2003)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Hartle, J.B.: The quasiclassical realms of this quantum universe. Found. Phys. 41, 982–1006 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Price, H.: Decisions, decisions, decisions: can Savage salvage Everettian probability? In: Saunders, S., Barrett, J., Kent, A., Wallace, D. (eds.) Many Worlds? Everett, Quantum Theory & Reality. Oxford University Press, Oxford (2010)

    Google Scholar 

  32. Gerlach, W., Stern, O.: Der experimentelle Nachweis der Richtungsquantelung im Magnetfeld. Z. Phys. 9, 349–352 (1922)

    Article  ADS  Google Scholar 

  33. Phipps, T.E., Stern, O.: Über die Einstellung der Richtungsquantelung. Z. Phys. 73, 185–191 (1932)

    Article  ADS  MATH  Google Scholar 

  34. Frisch, R., Segrè, E.: Über die Einstellung der Richtungsquantelung. II. Z. Phys. 80, 610–616 (1933)

    Article  ADS  Google Scholar 

  35. Sakurai, J.-J.: Modern Quantum Mechanics. Addison-Wesley, Boston (1993)

    Google Scholar 

  36. Feynman, R., Leighton, R., Sands, M.: The Feynman Lectures on Physics, vol. 3. Addison-Wesley, Boston (1977)

    MATH  Google Scholar 

  37. Landau, L., Lifshitz, E.: Quantum Mechanics: Non-relativistic Theory. Pergamon, Oxford (1965)

    MATH  Google Scholar 

  38. von Neumann, J.: Mathematische Grundlagen der Quantenmechanik. Springer, Berlin (1932)

    MATH  Google Scholar 

  39. Nielsen, M., Chuang, I.: Quantum Computation and Quantum Information. Cambridge University Press, New York (2000)

    MATH  Google Scholar 

  40. Peres, A.: Quantum Theory: Concepts and Methods. Springer, Dordrecht (2006)

    MATH  Google Scholar 

  41. Schrödinger, E.: An undulatory theory of the mechanics of atoms and molecules. Phys. Rev. 28, 1049–1070 (1926)

    Article  ADS  Google Scholar 

  42. Born, M.: Zur Quantenmechanik der Stoßvorgänge. Z. Phys. 37, 863–867 (1926)

    Article  ADS  MATH  Google Scholar 

  43. Aaronson, S.: Is quantum mechanics an island in theoryspace? arXiv:quant-ph/0401062

  44. Lewis, D.: A subjectivist’s guide to objective chance. In: Jeffrey, R.C. (ed.) Studies in Inductive Logic and Probability, vol. 2. University of California Press, Berkeley (1980)

    Google Scholar 

  45. Saunders, S.: Chance in the Everett interpretation. In: Saunders, S., Barrett, J., Kent, A., Wallace, D. (eds.) Many Worlds? Everett, Quantum Theory & Reality. Oxford University Press, Oxford (2010)

    Chapter  Google Scholar 

  46. Sebens, C.T., Carroll, S.M.: Self-locating uncertainty and the origin of probability in Everettian quantum mechanics. Br. J. Philos. Sci. 69, 25–74 (2018)

    MathSciNet  MATH  Google Scholar 

  47. Elga, A.: Defeating Dr. Evil with self-locating belief. Philos. Phenomenol. Res. 69, 383–396 (2004)

    Article  Google Scholar 

  48. Bohnenblust, F.: An axiomatic characterization of \(L_p\)-spaces. Duke Math. J. 6, 627–640 (1940)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Koenraad Audenaert, Časlav Brukner, Eric Cavalcanti, Fabio Costa, Daniel Süß, David Gross, Markus Heinrich, Philipp Höhn, Felipe M. Mora, Jacques Pienaar, Simon Saunders, and David Wallace for useful discussions. This work has been supported by the Excellence Initiative of the German Federal and State Governments (Grant ZUK 81).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mateus Araújo.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Consistent Fine-Graining

For which norms can the fine-graining argument work?

There are multiple ways to fine-grain a game with unequal coefficients into a symmetric game. For example, if \(\mu = \big \Vert \big (\Vert {\mathbf {1}}^{(n)}\Vert ,\Vert {\mathbf {1}}^{(m)}\Vert \big )\big \Vert \), where \({\mathbf {1}}^{(n)}\) is the vector with n ones, one can fine-grain the game

$$\begin{aligned} G = \begin{pmatrix} 1 &{}\quad r_1 \\ \mu &{}\quad r_2 \end{pmatrix} \end{aligned}$$
(73)

either by first taking it to

$$\begin{aligned} G' = \begin{pmatrix} 1 &{}\quad r_1 \\ \Vert {\mathbf {1}}^{(n)}\Vert &{}\quad r_2 \\ \Vert {\mathbf {1}}^{(m)}\Vert &{}\quad r_2 \end{pmatrix}, \end{aligned}$$
(74)

and then applying two more fine-grainings to take it to

$$\begin{aligned} G'' = \begin{pmatrix} 1 &{}\quad r_1 \\ {\mathbf {1}}^{(n)} &{}\quad {\mathbf {r}}_2^{(n)} \\ {\mathbf {1}}^{(m)} &{}\quad {\mathbf {r}}_2^{(m)} \end{pmatrix}, \end{aligned}$$
(75)

or by taking G directly to \(G''\), which will be possible only if \(\mu = \Vert {\mathbf {1}}^{(n+m)}\Vert \). We want all possible ways of fine-graining a game to give same result, so we demand the norm to be such that for all vectors \({\mathbf {v}}\) and \({\mathbf {w}}\) with disjoint support

$$\begin{aligned} \Vert {\mathbf {v}} + {\mathbf {w}}\Vert = \big \Vert \big (\Vert {\mathbf {v}}\Vert ,\Vert {\mathbf {w}}\Vert \big )\big \Vert . \end{aligned}$$
(76)

We also demand the norm to be permutation-invariant, as it seems unphysical to attribute meaning to the labelling of the vectors, and that \(\Vert (1,1)\Vert \ne 1\), because otherwise \(\Vert {\mathbf {1}}^{(n)}\Vert =1\) for all n, and it is therefore impossible to fine-grain any non-trivial game.

These conditions are enough to show that these norms must be equivalent to p-norms when restricted to vectors of rational numbers, as can be seen by adapting an argument by Bohnenblust [48]. We have then

Theorem 6

Let \(\Vert \cdot \Vert :{\mathbb {C}}^N \rightarrow {\mathbb {R}}\) be a permutation-invariant norm for \(N\ge 3\) such that \(\Vert (1,1)\Vert \ne 1\) and

$$\begin{aligned} \Vert {\mathbf {v}} + {\mathbf {w}}\Vert = \big \Vert \big (\Vert {\mathbf {v}}\Vert ,\Vert {\mathbf {w}}\Vert \big )\big \Vert \end{aligned}$$
(77)

for all vectors \({\mathbf {v}},{\mathbf {w}}\) with disjoint support. Then for any vector \({\mathbf {c}}\) such that the absolute value of its components is rational,

$$\begin{aligned} \Vert {\mathbf {c}}\Vert = \left( \sum _i |c_i|^p\right) ^{\frac{1}{p}}, \end{aligned}$$
(78)

for some real number \(p \ge 1\).

Proof

Let f be such that \(f(1):=\Vert 1\Vert \) and \(f(n+1) := \Vert (1,f(n))\Vert \). First note that \(f(1) = 1\), as \(\Vert 1\Vert = \Vert (1,0)\Vert = \Vert (\Vert (1,0)\Vert ,\Vert (0,0)\Vert )\Vert = \Vert (1,0)\Vert ^2\).

We need to show that f(n) is monotonous, and that \(f(n^k) = f^k(n)\). For the former, consider the identity \(2(f(n),0) = (f(n),1) + (f(n),-1)\) and take the norm of both sides. By the triangle inequality

$$\begin{aligned} 2f(n) \le \Vert (f(n),1)\Vert + \Vert (f(n),-1)\Vert = 2f(n+1). \end{aligned}$$
(79)

For the latter, first we show that \(f(n+m) = \Vert (f(m),f(n)\Vert \). Assume that it holds for some m. Then

$$\begin{aligned} f(n+1+m) = \Vert (f(m),f(n+1))\Vert = \Vert (f(m),1,f(n))\Vert = \Vert (f(m+1),f(n))\Vert .\nonumber \\ \end{aligned}$$
(80)

Since it holds for \(m=1\), by induction it holds for all m. Now assume that \(f(nm) = f(n)f(m)\) holds for some m. Then

$$\begin{aligned} f(n(m+1)) = f(nm+n) = \Vert (f(nm),f(n))\Vert = f(n)\Vert (f(m),1)\Vert = f(n)f(m+1),\nonumber \\ \end{aligned}$$
(81)

and therefore by induction this is true for all m, as it obviously holds for \(m=1\). This implies that \(f(n^k) = f^k(n)\).

Now let \(m,n\ge 2\) be some fixed integers, and h the integer such that for any positive integer k

$$\begin{aligned} m^h \le n^k < m^{h+1}. \end{aligned}$$
(82)

Applying \(f(\cdot )\) to these numbers, it follows that

$$\begin{aligned} h\log f(m) \le k \log f(n) \le (h+1) \log f(m), \end{aligned}$$
(83)

and using the fact that \(h \le k \frac{\log n}{\log m}\) and \(h > k \frac{\log n}{\log m}-1\) we have that

$$\begin{aligned} \frac{\log f(m)}{\log m}-\frac{\log f(m)}{k\log n} < \frac{\log f(n)}{\log n} \le \frac{\log f(m)}{\log m}+\frac{\log f(m)}{k\log n}. \end{aligned}$$
(84)

Taking the limit of k going to infinity lets us conclude that

$$\begin{aligned} \frac{\log f(m)}{\log m} = \frac{\log f(n)}{\log n}, \end{aligned}$$
(85)

which means that this fraction is a constant independent of n (and different than 0 as \(f(2)>1)\). Calling this constant 1 / p, we conclude that

$$\begin{aligned} f(n) = n^\frac{1}{p}. \end{aligned}$$

Now for any rational number m / n we have that

$$\begin{aligned} \Vert (1,m/n)\Vert = \frac{1}{n}\Vert (n,m)\Vert = \frac{1}{n}\Vert (f(n^p),f(m^p)\Vert = \frac{1}{n}f(n^p+m^p) = (1+(m/n)^p)^\frac{1}{p}, \end{aligned}$$

so by homogeneity \(\Vert (a,b)\Vert = (|a|^p + |b|^p)^{\frac{1}{p}}\) for any rationals |a| and |b|, and by induction for any vector \({\mathbf {c}}\) such that the absolute values of the components are rational numbers

$$\begin{aligned} \Vert {\mathbf {c}}\Vert = \left( \sum _i |c_i|^p\right) ^{\frac{1}{p}}. \end{aligned}$$
(86)

\(\square \)

If one furthermore assumes some regularity condition, then the result is valid for any complex vector.

Reducing Sequential Games to Simple Games

Consider the sequential game

$$\begin{aligned} G = \begin{pmatrix} c_1 &{}\quad r_1 \\ c_2 &{}\quad d_1 &{}\quad s_1 \\ &{}\quad d_2 &{}\quad s_2 \end{pmatrix}, \end{aligned}$$
(87)

where in the worlds with outcome 1 reward \(r_1\) is given, but in the worlds with outcome 2 the subgame \(H = \begin{pmatrix} d_1 &{} s_1 \\ d_2 &{} s_2 \end{pmatrix}\) is played. We want to reduce it to a simple game in both Kent’s universes and Many-Worlds.

In the regular Kent’s universe \(c_1\) worlds are created with reward \(r_1\), \(c_2d_1\) worlds are created with reward \(s_1\), and \(c_2d_2\) worlds are created with reward \(s_2\), so it seems natural to postulate that G is equivalent to the simple game

$$\begin{aligned} G' = \begin{pmatrix} c_1 &{}\quad r_1 \\ c_2d_1 &{}\quad s_1 \\ c_2d_2 &{}\quad s_2 \end{pmatrix}. \end{aligned}$$
(88)

In the reverse Kent’s universe, there are \(M(c_1+c_2)(d_1+d_2)\) worlds in the beginning. After the first branching, \(Mc_1(d_1+d_2)\) worlds are imprinted with outcome 1, and the remaining \(Mc_2(d_1+d_2)\) worlds are split again, with \(Mc_2d_1\) being imprinted with a further outcome 1, and \(Mc_2d_2\) with outcome 2. In the end there are \(Mc_1(d_1+d_2)\) with reward \(r_1\), \(Mc_2d_1\) worlds with reward \(s_1\), and \(Mc_2d_2\) worlds with reward \(s_2\), so it seems natural to postulate that G is equivalent to the simple game

$$\begin{aligned} G'' = \begin{pmatrix} c_1(d_1+d_2) &{}\quad r_1 \\ c_2d_1 &{}\quad s_1 \\ c_2d_2 &{}\quad s_2 \end{pmatrix}. \end{aligned}$$
(89)

Note that in the reverse Kent’s universe Substitution is satisfied: the value of the subgame H is

$$\begin{aligned} V(H) = \frac{1}{d_1+d_2}(d_1s_1+d_2s_2), \end{aligned}$$
(90)

and the value of G there is

$$\begin{aligned} V(G)&= \frac{1}{(c_1+c_2)(d_1+d_2)}\Big (c_1(d_1+d_2)r_1 + c_2(d_1s_1 + d_2s_2) \Big ) \end{aligned}$$
(91)
$$\begin{aligned}&= \frac{1}{c_1+c_2}(c_1r_1+c_2V(H)), \end{aligned}$$
(92)

which is equal to the value of \(\begin{pmatrix} c_1 &{} r_1 \\ c_2 &{} V(H) \end{pmatrix}\), as required. We shall not prove the general case, as that is quite straightforward.

In Many-Worlds, the game G is instantiated by making a measurement on the state \(| \psi \rangle = c_1| 1 \rangle + c_2| 2 \rangle ,\) giving reward \(r_1\) in the worlds with outcome 1, and in the worlds with outcome 2 doing a measurement on the state \(| \varphi \rangle = d_1| 1 \rangle + d_2| 2 \rangle ,\) finally giving rewards \(s_1\) and \(s_2\) in the worlds with the second outcomes 1 and 2. These measurements take the initial state \(| \psi \rangle | M_? \rangle | \varphi \rangle | D_? \rangle \) to the final state

$$\begin{aligned} | G \rangle = c_1| 1 \rangle | M_1 \rangle | \varphi \rangle | D_? \rangle | r_1 \rangle + c_2d_1| 2 \rangle | M_2 \rangle | 1 \rangle | D_1 \rangle | r_2 \rangle + c_2d_2| 2 \rangle | M_2 \rangle | 2 \rangle | D_2 \rangle | r_3 \rangle .\nonumber \\ \end{aligned}$$
(93)

An equivalent way to play this game is to make a joint measurement on the state

$$\begin{aligned} | \psi \rangle | \varphi \rangle = c_1d_1| 1 \rangle | 1 \rangle + c_1d_2| 1 \rangle | 2 \rangle + c_2d_1| 2 \rangle | 1 \rangle + c_2d_2| 2 \rangle | 2 \rangle , \end{aligned}$$
(94)

but in the worlds where the measurement on \(| \psi \rangle \) resulted in 1 apply the label ? to both outcomes of the measurement on \(| \varphi \rangle \), leading to the final state

$$\begin{aligned} | G' \rangle= & {} c_1d_1| 1 \rangle | M_1 \rangle | 1 \rangle | D_?' \rangle | r_1 \rangle + c_1d_2| 1 \rangle | M_1 \rangle | 2 \rangle | D_?'' \rangle | r_1 \rangle \nonumber \\&+ \, c_2d_1| 2 \rangle | M_2 \rangle | 1 \rangle | D_1 \rangle | s_1 \rangle + c_2d_2| 2 \rangle | M_2 \rangle | 2 \rangle | D_2 \rangle | s_2 \rangle \end{aligned}$$
(95)

where \(| D_?' \rangle \) and \(| D_?'' \rangle \) are measurements results physically distinct from \(| D_? \rangle \), but with the same label.

If one does not, however, coarse-grain the results (1, 1) and (1, 2) together, then this measurement procedure can be regarded as playing the simple game

$$\begin{aligned} G''' = \begin{pmatrix} c_1d_1 &{}\quad r_1 \\ c_1d_2 &{}\quad r_1 \\ c_2d_1 &{}\quad s_1 \\ c_2d_2 &{}\quad s_2 \end{pmatrix} \end{aligned}$$
(96)

instead. We postulate therefore that a rational agent in Many-Worlds should regard G and \(G'''\) as equivalent, or more formally that:

  • ReductionThe sequential game

    $$\begin{aligned} G = \begin{pmatrix} \alpha _1 &{}\quad r_1 \\ \vdots &{}\quad \vdots \\ \alpha _n &{}\quad \beta _1 &{}\quad s_1 \\ &{}\quad \vdots &{}\quad \vdots \\ &{}\quad \beta _m &{}\quad s_m \end{pmatrix} \end{aligned}$$
    (97)

    where subgame\((\varvec{\beta },{\mathbf {s}})\)is played in the worlds with outcomen, has the same value as the simple game

    $$\begin{aligned} G' = \begin{pmatrix} \alpha _1\varvec{\beta } &{}\quad r_1 \\ \vdots &{}\quad \vdots \\ \alpha _{n-1}\varvec{\beta } &{} r_{n-1} \\ \alpha _n\beta _1 &{}\quad s_1 \\ \vdots &{}\quad \vdots \\ \alpha _n\beta _{m-1} &{}\quad s_{m-1} \\ \alpha _n\beta _m &{}\quad s_m \end{pmatrix}. \end{aligned}$$
    (98)

This Reduction postulate suffices to prove Substitution as a theorem, as the value of the subgame \((\varvec{\beta },{\mathbf {s}})\) is

$$\begin{aligned} V(\varvec{\beta },{\mathbf {s}}) = \frac{1}{\Vert \varvec{\beta }\Vert ^p_p}\sum _{j=1}^m|\beta _j|^p, \end{aligned}$$
(99)

and the value of G is

$$\begin{aligned} V(G)= & {} \frac{1}{\Vert \varvec{\alpha }\Vert ^p_p\Vert \varvec{\beta }\Vert ^p_p}\left( \Vert \varvec{\beta }\Vert ^p_p\sum _{i=1}^{n-1}|\alpha _i|^pr_i + |\alpha _n|^p\sum _{j=1}^m|\beta _j|^p s_j\right) \end{aligned}$$
(100)
$$\begin{aligned}= & {} \frac{1}{\Vert \varvec{\alpha }\Vert ^p_p}\left( \sum _{i=1}^{n-1}|\alpha _i|^pr_i + |\alpha _n|^p V(\varvec{\beta },{\mathbf {s}})\right) , \end{aligned}$$
(101)

as required.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Araújo, M. Probability in Two Deterministic Universes. Found Phys 49, 202–231 (2019). https://doi.org/10.1007/s10701-019-00241-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-019-00241-7

Keywords

Navigation