Skip to main content
Log in

Population ethics in an infinite universe

  • Published:
Philosophical Studies Aims and scope Submit manuscript

If you wish to make an apple pie from scratch, you must first invent the universe.

—Carl Sagan.

Abstract

Population ethics studies the tradeoff between the total number of people who will ever live, and their quality of life. But widely accepted theories in modern cosmology say that spacetime is probably infinite. In this case, its population is also probably infinite, so the quantity/quality tradeoff of population ethics is no longer meaningful. Instead, we face the problem of how to ethically evaluate an infinite population of people dispersed throughout time and space. I argue that axiologies based on spatiotemporal Cesàro average utility are the most appropriate way to make this evaluation, because their unique properties make them superior to other axiologies that have been proposed.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. Here, I assume that consciousness and human-level intelligence are sufficient for a being to be worthy of moral consideration roughly comparable to that which we extend to other humans. I do not assume that human-level intelligence is necessary for a being to be worthy of moral consideration.

  2. For example: it is difficult to define a numerical measure that adequately aggregates the well-being of an infinite number of people (Bostrom, 2011). But a merely ordinal axiology (without a numerical measure) is inadequate for many purposes (Arntzenius, 2014). More problematically, there is a fundamental incompatibility between impartiality axioms (i.e. permutation-invariance) and Pareto axioms (i.e. monotonicity with respect to the welfare of the individuals). An axiology that satisfies a stronger form of one of these axioms must satisfy a weaker form of the other. See e.g. Diamond (1965), Svensson (1980), Chichilnisky and Heal (1997), Lauwers (1997, 1998, 2010, 2012), Vallentyne and Kagan (1997), Hamkins and Montero (2000), Fleurbaey and Michel (2003), Basu and Mitra (2003, 2007a, 2007b, 2007c), Sakai (2010a, 2010b, 2016), Lauwers and Vallentyne (2004), Asheim and Tungodden (2004), Banerjee (2006), Roemer and Suzumura (2007), Zame (2007), Asheim et al. (2010), Bostrom (2011), Dubey (2011), Dubey and Mitra (2011, 2014), Asheim and Zuber (2013, 2016), Pivato (2014, 2021, 2022, 2023), Jonsson and Voorneveld (2015, 2018), Askell (2018), Petri (2019), Jonsson (2020), Jonsson and Peterson (2020), Wilkinson (2021, 2023a, 2023b, 2023c), Asheim et al. (2022a, 2022b), and Li and Wakker (2023). See Asheim (2010) and Askell (2018) for reviews of some of this literature.

  3. (T) Formally, given a countably infinite, enumerated list of people with utilities \(u_1,u_2,u_3,\ldots\), the Cesàro average utility is the limit \(\displaystyle \lim \limits _{N{\rightarrow }{\infty }} \ \frac{u_1+u_2+u_3+\cdots +u_N}{N}\). Throughout the article, I will use footnotes flagged by (T) for “technical remarks” such as this one, for readers who want more precise mathematical formulations of some statements. These can be skipped without impeding understanding of the main text.

  4. Even there was someone in B with the same name, location and genetic code as someone in A, it is not clear that they would be “the same” person in any ethically relevant sense, if they grew up in wildly disparate worlds, led totally different lives, and developed wholly dissimilar personalities as a result.

  5. Here and throughout the paper, I use the term “utility” in an extremely broad sense. It could correspond to subjective hedonic state, or preference satisfaction, or quality of life according to some objective list of criteria, or some other measure of personal well-being.

  6. \({\mathcal { X}}\) is bounded in this section only for simplicity. In Section 5 we shall allow \({\mathcal { X}}\) to be much more general.

  7. (T) A metric is a function \(d:{\mathcal { S}}\times {\mathcal { S}}{{\longrightarrow }}{\mathbb {R}}_+\) such that (i) \(d(s,s)=0\) for all s in \({\mathcal { S}}\); (ii) \(d(r,s)=d(s,r)\) for all r and s in \({\mathcal { S}}\); and (iii) \(d(q,s)\le d(q,r)+d(r,s)\) for all q, r, and s in \({\mathcal { S}}\). For more information on the topological concepts used in this paper, please see Willard (2004) or any other introduction to topology.

  8. (T) Formally, \({\mathcal { N}}_c({\mathbf{ s}},r):=\{n\in {\mathcal { N}}\); \(d(s_n,c)<r\}\).

  9. (T) To be precise, the spatiotemporal Cesàro average utility of \(({\mathbf{ s}},{\mathbf{ x}})\) is \(\displaystyle \lim _{r{\rightarrow }{\infty }} \frac{1}{\left|{\mathcal { N}}_c({\mathbf{ s}},r)\right|} \sum _{n\in {\mathcal { N}}_c({\mathbf{ s}},r)} x_n\).

  10. (T) That is: \(({\mathbf{ s}},{\mathbf{ x}})\approx ({\mathbf{ s}}',{\mathbf{ y}})\) if both \(({\mathbf{ s}},{\mathbf{ x}})\succcurlyeq ({\mathbf{ s}}',{\mathbf{ y}})\) and \(({\mathbf{ s}},{\mathbf{ x}})\preccurlyeq ({\mathbf{ s}}',{\mathbf{ y}})\). Meanwhile \(({\mathbf{ s}},{\mathbf{ x}})\succ ({\mathbf{ s}}',{\mathbf{ y}})\) if \(({\mathbf{ s}},{\mathbf{ x}})\succcurlyeq ({\mathbf{ s}}',{\mathbf{ y}})\) but \(({\mathbf{ s}},{\mathbf{ x}})\not \approx ({\mathbf{ s}}',{\mathbf{ y}})\).

  11. (T) Formally: \(\displaystyle d_{\infty }({\mathbf{ x}},{\mathbf{ y}}):= \sup _{n\in {\mathcal { N}}} \ |x_n-y_n|\). This is a measure of the “distance” between \({\mathbf{ x}}\) and \({\mathbf{ y}}\).

  12. (T) To be precise: such that \(\lim \limits _{k{\rightarrow }{\infty }} d_{\infty }({\mathbf{ x}}^k,{\mathbf{ x}})=0\).

  13. A permutation is a one-to-one and onto function from \({\mathcal { N}}\) to itself; it “rearranges” the elements of \({\mathcal { N}}\). “Sufficiently large” will be made precise in Section 5.

  14. This stylised example assumes (unrealistically) that the future population of Earth is fixed, and corresponds precisely to the indices in the set \({\mathcal { I}}\).

  15. Pressman (2015) is a notable exception, who defends Average Utilitarianism against these objections.

  16. One could reformulate the Sadistic Conclusion in terms of proportions of certain kinds of people in an infinite population. This would say that a world where a \(1-\epsilon\) proportion of people have good lives and an \(\epsilon\) proportion have terrible lives is better than a world where only an \(\epsilon\) proportion have good lives and a \(1-\epsilon\) proportion of people have mediocre lives. But this is not obviously wrong, on reflection.

  17. One might think we could represent a finite population by adding to \({\mathcal { S}}\) an extra point “\({\infty }\)” that was “infinitely far away” from all the other points in \({\mathcal { S}}\), and then setting \(s_n={\infty }\) for all but finitely many n in \({\mathcal { N}}\). But the proof of the theorem would break down: it requires \({\widetilde{\mathbf{ x}}}\) to have infinitely many coordinates.

  18. I thank a referee for this suggestion.

  19. And presumably it would collapse into a black hole.

  20. (T) To be precise: the sum of the first 3 values is zero, as is the sum of the first 7 values, and the sum of the first 15 values. Indeed for any natural number n, the sum (and hence, the average) of the first \(2^n-1\) values is zero. However, the sum of the first 2 values is 1, the sum of the first 5 values is 2, the sum of the first 11 values is 4, and in general, for any natural number n, the sum of the first \(2^{n+1}+2^n-1\) values is \(2^n\). Thus, the average of the first \(2^{n+1}+2^n-1\) values is slightly greater than 1/3.

  21. For more on this debate, see Earman and Norton (1987), Earman (1989), Dorato (2000), Nerlich (2003) and Ladyman and Ross (2007), inter alia.

  22. To my knowledge, Arntzenius (2014) was the first to raise this problem in infinite-population ethics.

  23. A bijection from a set \({\mathcal { X}}\) to a set \({\mathcal { Y}}\) is a one-to-one, onto function from \({\mathcal { X}}\) to \({\mathcal { Y}}\). So for every element y in \({\mathcal { Y}}\), there is exactly one element x in \({\mathcal { X}}\) whose image is y.

  24. For instance, in classical physics, different reference frames correspond to different velocities. Thus, we can define the distance between two reference frames to be the magnitude of the relative velocity of one, relative to the other. More generally, in relativistic physics, we can define the distance between two reference frames to be the square root of the relative kinetic energy of one, relative to the other. For a more detailed explanation of these and other possible metrics, see Appendix B of Pivato (2023).

  25. That is: \({\mathcal { S}}={\mathbb {R}}^4\) and also \({\mathcal { O}}={\mathbb {R}}^4\).

  26. To address the concern about spacetime substantivalism versus relationalism raised in Sect. 3, we can stipulate that the mapping from \({\mathcal { O}}\) to \({\mathcal { S}}\) defined by each element of \(\Phi\) also depends on other parameters, such as the distribution of mass and energy. (Indeed, such a dependence is required by general relativity, as explained in Appendix B.) To avoid any further complexity, I shall not make this dependence explicit in the version I present here. But it can be incorporated without affecting any of what follows.

  27. (T) One could also require \({\mathcal { K}}_n\cap {\mathcal { K}}_m=\emptyset\) whenever \(n\ne m\). But this is not necessary for the results in this paper. \({\mathcal { K}}_n\) and \({\mathcal { K}}_m\) could overlap in cases of “branching” identities discussed by Parfit (1984).

  28. (T) Formally, \({\mathcal { S}}(c,r):= \{s\in {\mathcal { S}}\); \(d(c,s)<r\}\), \(\Phi (\psi ,r) := \{\phi \in \Phi\); \(\delta (\psi ,\phi )<r\}\), and \({\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r) := \{n\in {\mathcal { N}}\); \(\phi \left({\mathcal { K}}_n\right)\subseteq {\mathcal { S}}(c,r)\), for some \(\phi \in \Phi (\psi ,r)\}\).

  29. (T) Here are the formal statements of these conditions.    Mortality: \(\displaystyle {\mathcal { N}}=\bigcup _{r=1}^{\infty }{\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r)\). Local finitenes: \(\left|{\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r)\right|\) is finite for all \(r>0\). Subexponential growth: For any \(R>0\), \(\displaystyle \limsup _{r{\rightarrow }{\infty }} \frac{\left|{\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r+R)\right|}{\left|{\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r)\right|} = 1\).

  30. See Lemma 1 in Pivato (2023).

  31. (T) A metric space \({\mathcal { X}}\) is locally compact if every point has a neighbourhood that is compact. Roughly speaking, this means \({\mathcal { X}}\) is “finite-dimensional”. For example, the Euclidean space \({\mathbb {R}}^N\) is locally compact. Local compactness is not actually required for what follows, but it allows us to avoid certain technicalities.

  32. To the extent that personal identity is meaningful, one can assume it is also encoded in \({\mathcal { X}}\) (not in \({\mathcal { N}}\)). But the framework of this paper does not rely on any theory of personal identity (cf. Footnote 4).

  33. The requirement that \({\mathcal { B}}\) be locally compact corresponds to the idea that the property P can be specified relatively simply. Roughly speaking, it excludes highly contrived, unnatural properties, sensitive to microscopically precise details about a person’s life. But it allows \({\mathcal { B}}\) to be any open or closed subset of \({\mathcal { X}}\).

  34. (T) To be precise, the asymptotic frequency is defined: \(\displaystyle \lim _{r{\rightarrow }{\infty }} \frac{\#{\left\{ n\in {\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r) \; ; \; x_n\in {\mathcal { B}} \right\} }}{\left|{\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r)\right|}\). A social outcome \({\mathbf{ x}}\) is regular if this limit exists for every locally compact subset \({\mathcal { B}}\subseteq {\mathcal { X}}\).

  35. See Lemma 2 of Pivato (2023).

  36. (T) To be precise, spatiotemporal Cesàro average of u on \(({\varvec{\mathcal {K}}},{\mathbf{ x}})\) is \(\displaystyle \lim _{r{\rightarrow }{\infty }} \frac{1}{\left|{\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r)\right|} \sum _{n\in {\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r)} u(x_n)\).

  37. See Lemma 2 of Pivato (2023).

  38. Recall that the elements of \({\mathcal { N}}\) are just labels. There is no sense in which the label n in one world represents “the same person” as n represents in some other world (cf. Footnote 32). In light of this, A1 could be seen as a weak version of the Suppes-Sen Grading Principle, rather than the Pareto principle.

  39. (T) Formally: \(\displaystyle d_{\infty }({\mathbf{ x}},{\mathbf{ y}}):= \sup _{n\in {\mathcal { N}}} \ d(x_n,y_n)\). This is a measure of the “distance” between \({\mathbf{ x}}\) and \({\mathbf{ y}}\).

  40. I say “roughly” because this comparison is done with respect to all reference frames in \(\Psi\) that are within distance r of the central reference frame \(\psi\).

  41. This follows from Desideratum A4 below, along with Lemma 3 of Pivato (2023).

  42. Papers on infinite-population social welfare such as Lauwers (1998) call this a bounded permutation.

  43. (T) To be precise, \(\gamma\) is a Lévy permutation if \(\displaystyle \lim _{N{\rightarrow }{\infty }} \frac{\#\{n\in \{1,2,\ldots ,N\} ; \ \gamma (n)>N\}}{N} \ = \ 0.\)

  44. (T) In fact, one can even allow G to depend on n, as long as G(n) grows sublinearly. For example, if there is some constant K such that \(|\gamma (n) - n|\le K\cdot \sqrt{n}\) for all n in \({\mathbb {N}}\), then \(\gamma\) is a Lévy permutation.

  45. See Lemma 3 in Pivato (2023).

  46. The precise argument is more complicated, but also more general; see Proposition 4.1 of Pivato (2023). Also, this argument assumes that there is no ethically relevant sense in which a particular spacetime position in world A contains “the same person” as some position in world B (as does my formulation of Weak Pareto). This is because I am skeptical that such sameness-of-identity (or haecceity) between two different worlds is even meaningful (cf. Footnotes 4, 32 and 38). But some approaches to infinite-population ethics (e.g. Askell 2018) presuppose such haecceity, so they might not accept my conclusions regarding A vs. B.

  47. These four axiologies originally arose in the analysis of intergenerational justice, so they are defined for an infinite population of people arranged linearly in time. But they can be generalized to an infinite multidimensional spacetime, using the same “enumeration” trick that is used in this paper.

  48. Also, unlike spatiotemporal Cesàro average utilitarianism, all of these axiologies are actually incomplete binary relations, which are unable to make ethical comparisons between certain pairs of worlds.

  49. Similar views have been expressed by Parfit (1984, Appendix F), Cowen and Parfit (1992), Cowen (1992), and Bostrom (2011, §3.1), among others.

  50. I am not the first to observe that exponential discounting is incompatible with special relativity. Cowen (2007, §II.A, p.10) and Wilkinson (2023b) have made similar observations. Meanwhile, several economists have investigated the economic implications of relativistic time dilation — first facetiously (Krugman, 2010), and then more seriously (Angel, 2014; Carvalho and Gaspar, 2021). But this literature is mainly concerned with financial interest rates in a relativistic setting, rather than with ethical questions.

  51. Despite its name, the “Minkowski metric” is not a metric in the sense defined in Footnote 7.

  52. If we were concerned with individual decision-making, then an observer could exponentially discount according to proper time along her own trajectory through spacetime. This is because proper time along this trajectory determines her subjective perception of the passage of time. But the proper times experienced by observers on different spacetime trajectories generally do not agree. So this approach is totally unsuitable for defining an objective, observer-independent axiology.

  53. This manifests in the relativistic phenomena of time dilation and Lorentz-FitzGerald contraction.

  54. This holds even if spacetime is finite, as long as it is large enough for relativistic effects to be significant.

  55. (T) Formally, \({\mathcal { S}}(c,r):=\{s\in {\mathcal { S}}\); \(d(s,c)< r\}\).

  56. (T) For example, if \({\mathcal { S}}\) is four-dimensional Euclidean space, then \(V(r)=\frac{\pi ^2}{2} r^4\).

  57. (T) To be precise, the utility density is defined \(\displaystyle \lim _{r{\rightarrow }{\infty }} \ \frac{1}{V(r)} \sum _{n\in {\mathcal { N}}_c({\mathbf{ s}},r)} u(x_n)\).

  58. As noted in Sect. 3, Cesàro average utility is not vulnerable to this criticism.

  59. (T) Here, by “geometric structure” I mean a Lorentzian metric, which endows every tangent space of the manifold with an inner product like the Minkowski spacetime of Appendix A. By “entirely determined”, I am referring to the Einstein field equations, which relate this Lorentzian metric to the spatiotemporal distribution of matter and energy (described by the stress-energy tensor).

References

  • Angel, J. J. (2014). When finance meets physics: The impact of the speed of light on financial markets and their regulation. Financial Review, 49(2), 271–281.

    Article  Google Scholar 

  • Arntzenius, F. (2014). Utilitarianism, decision theory and eternity. Philosophical Perspectives, 28, 31–58.

    Article  Google Scholar 

  • Arrhenius, G., 2000. Future generations: A challenge for moral theory. Ph.D. thesis, Uppsala University.

  • Arrhenius, G., Population ethics. Oxford University Press (forthcoming), Oxford UK.

  • Arrhenius, G., Bykvist, K., Campbell, T., Finneron-Burns, E. (Eds.), 2022a. The Oxford Handbook of Population Ethics. Oxford University Press.

  • Arrhenius, G., Ryberg, J., Tännsjö, T., Summer 2022b. The repugnant conclusion. In: Zalta (2022). https://plato.stanford.edu/archives/sum2022/entries/repugnant-conclusion

  • Asheim, G. (2010). Intergenerational equity. Annual Review of Economics, 2(1), 197–222.

    Article  Google Scholar 

  • Asheim, G., Bossert, W., Sprumont, Y., & Suzumura, K. (2010). Infinite-horizon choice functions. Economical Theory, 43(1), 1–21.

    Article  Google Scholar 

  • Asheim, G., & Tungodden, B. (2004). Resolving distributional conflicts between generations. Economical Theory, 24(1), 221–230.

    Article  Google Scholar 

  • Asheim, G., & Zuber, S. (2013). A complete and strongly anonymous leximin relation on infinite streams. Social Choice and Welfare, 41(4), 819–834.

    Article  Google Scholar 

  • Asheim, G., & Zuber, S. (2016). Evaluating intergenerational risks. Journal of Mathematical Economics, 65, 104–117.

    Article  Google Scholar 

  • Asheim, G. B., Kamaga, K., Zuber, S., 2022a. Infinite population utilitarian criteria, CESifo Working Paper No. 9576.

  • Asheim, G. B., Kamaga, K., & Zuber, S. (2022). Maximal sensitivity under strong anonymity. Journal of Mathematical Economics, 103, 102768.

    Article  Google Scholar 

  • Askell, A., 2018. Pareto principles in infinite ethics. Ph.D. thesis, New York University.

  • Atsumi, H. (1965). Neoclassical growth and the efficient program of capital accumulation. Review of Economic Studies, 32, 127–136.

    Article  Google Scholar 

  • Banerjee, K. (2006). On the extension of the utilitarian and Suppes-Sen social welfare relations to infinite utility streams. Social Choice and Welfare, 27(2), 327–339.

    Article  Google Scholar 

  • Barrow, J. D., Davies, P., & Harper, C. L. (Eds.). (2004). Science and Ultimate Reality: From quantum to cosmos. Cambridge, UK: Cambridge University Press.

    Google Scholar 

  • Basu, K., & Mitra, T. (2003). Aggregating infinite utility streams with intergenerational equity: the impossibility of being Paretian. Econometrica, 71(5), 1557–1563.

    Article  Google Scholar 

  • Basu, K., Mitra, T., 2007a. On the existence of Paretian social welfare relations for infinite utility streams with extended anonymity. In: Roemer and Suzumura (2007), pp. 85–100.

  • Basu, K., Mitra, T., 2007b. Possibility theorems for aggregating infinite utility streams equitably. In: textitRoemer and Suzumura (2007), pp. 69–84.

  • Basu, K., & Mitra, T. (2007). Utilitarianism for infinite utility streams: A new welfare criterion and its axiomatic characterization. Journal of Economic Theory, 133(1), 350–373.

    Article  Google Scholar 

  • Blackorby, C., Bossert, W., & Donaldson, D. J. (2005). Population issues in social choice theory, welfare economics, and ethics. Cambridge University Press.

    Book  Google Scholar 

  • Bostrom, N. (2011). Infinite Ethics Analysis and Metaphysics, 10, 9–59.

    Google Scholar 

  • Carvalho, V. H., & Gaspar, R. M. (2021). Relativistic option pricing. International Journal of Financial Studies, 9(2), 4599.

    Article  Google Scholar 

  • Chichilnisky, G., & Heal, G. (1997). Social choice with infinite populations: Construction of a rule and impossibility results. Social Choice and Welfare, 14(2), 303–318.

    Article  Google Scholar 

  • Cowen, T., 1992. Consequentialism implies a zero rate of intergenerational discount. In: Fishkin and Laslett (1992), pp. 162–168.

  • Cowen, T. (2007). Caring about the distant future: Why it matters and what it means. University of Chicago Law Review, 74, 5–40.

    Google Scholar 

  • Cowen, T., Parfit, D., 1992. Against the social discount rate. In: Fishkin and Laslett (1992), pp. 144 –161.

  • DeWitt, B. S., 2004. The Everett interpretation of quantum mechanics, Ch. 10. In: Barrow et al. (2004), pp. 167–198.

  • DeWitt, B. S., & Graham, N. (Eds.). (1973). The Many Worlds Interpretation of Quantum Mechanics. Princeton University Press.

    Google Scholar 

  • Diamond, P. A. (1965). The evaluation of infinite utility streams. Econometrica, 4239, 170–177.

    Article  Google Scholar 

  • Dorato, M. (2000). Substantivalism, relationism, and structural spacetime realism. Foundations of physics, 30(10), 1605–1628.

    Article  Google Scholar 

  • Dubey, R. S. (2011). Fleurbaey-Michel conjecture on equitable weak Paretian social welfare order. Journal of Mathematical Economics, 47(4–5), 434–439.

    Article  Google Scholar 

  • Dubey, R. S., & Mitra, T. (2011). On equitable social welfare functions satisfying the weak Pareto axiom: a complete characterization. International Journal of Economic Theory, 7(3), 231–250.

    Article  Google Scholar 

  • Dubey, R. S., & Mitra, T. (2014). Combining monotonicity and strong equity: construction and representation of orders on infinite utility streams. Social Choice and Welfare, 43(3), 591–602.

    Article  Google Scholar 

  • Earman, J. (1989). World enough and spacetime: Absolute and relational theories of motion. MIT Press.

    Google Scholar 

  • Earman, J., & Norton, J. (1987). What price spacetime substantivalism? The hole story. The British Journal for the Philosophy of Science, 38(4), 515–525.

    Article  Google Scholar 

  • Ellis, G. F. R., & Brundrit, G. B. (1979). Life in the infinite universe. Quarterly Journal of the Royal Astronomical Society, 20, 37–41.

    Google Scholar 

  • Fishkin, J. S., Laslett, P. (Eds.), 1992. Justice Between Age Groups and Generations. Vol. 6 of Philosophy, Politics, and Society. Yale University Press.

  • Fleurbaey, M., & Michel, P. (2003). Intertemporal equity and the extension of the Ramsey criterion. Journal of Mathematical Economics, 39(7), 777–802.

    Article  Google Scholar 

  • Garriga, J., & Vilenkin, A. (2001). Many worlds in one. Physical Review D, 64(4), 043511.

    Article  Google Scholar 

  • Greaves, H. (2017). Population axiology. Philosophy. Compass, 12(11), 22366.

    Google Scholar 

  • Guth, A. H. (2000). Inflation and eternal inflation. Physics Reports, 333, 555–574.

    Article  Google Scholar 

  • Guth, A. H. (2001). Eternal inflation. Annals of the New York Academy of Sciences, 950, 66–82.

    Article  Google Scholar 

  • Guth, A. H. (2007). Eternal inflation and its implications. Journal of Physics A: Mathematical and Theoretical, 40(25), 6811.

    Article  Google Scholar 

  • Hamkins, J. D., & Montero, B. (2000). Utilitarianism in infinite worlds. Utilitas, 12(1), 91–96.

    Article  Google Scholar 

  • Jonsson, A., 2020. Infinite utility: time, counterparts and ultimate locations, (preprint).

  • Jonsson, A., & Peterson, M. (2020). Consequentialism in infinite worlds. Analysis, 80(2), 240–248.

    Article  Google Scholar 

  • Jonsson, A., & Voorneveld, M. (2015). Utilitarianism on infinite utility streams: Summable differences and finite averages. Economic Theory Bulletin, 3(1), 19–31.

    Article  Google Scholar 

  • Jonsson, A., & Voorneveld, M. (2018). The limit of discounted utilitarianism. Theoretical Economics, 13(1), 19–37.

    Article  Google Scholar 

  • Koopmans, T. C. (1960). Stationary ordinal utility and impatience. Econometrica, 28, 287–309.

    Article  Google Scholar 

  • Krugman, P. (2010). The theory of interstellar trade. Economic Inquiry, 48(4), 1119–1123.

    Article  Google Scholar 

  • Ladyman, J., & Ross, D. (2007). Every Thing must Go: Metaphysics Naturalized. Oxford, UK: Oxford University Press.

    Book  Google Scholar 

  • Lauwers, L. (1997). Topological aggregation, the case of an infinite population. Social Choice and Welfare, 14(2), 319–332.

    Article  Google Scholar 

  • Lauwers, L. (1998). Intertemporal objective functions: Strong Pareto versus anonymity. Mathematical Social Sciences, 35(1), 37–55.

    Article  Google Scholar 

  • Lauwers, L. (2010). Ordering infinite utility streams comes at the cost of a non-Ramsey set. Journal of Mathematics and Economics, 46(1), 32–37.

    Article  Google Scholar 

  • Lauwers, L. (2012). Intergenerational equity, efficiency, and constructibility. Economic Theory, 49(2), 227–242.

    Article  Google Scholar 

  • Lauwers, L., & Vallentyne, P. (2004). Infinite utilitarianism: more is always better. Economics & Philosophy, 20(2), 307–330.

    Article  Google Scholar 

  • Li, C., Wakker, P., 2023. A simple and very general axiomatization of average utility maximization for infinite streams, (working paper).

  • Linde, A. D. (1990). Particle Physics and Inflationary Cosmology. Harwood Academic.

    Book  Google Scholar 

  • Nerlich, G. (2003). Space-time substantivalism. The Oxford Handbook of Metaphysics (pp. 281–314). New York: Oxford University Press.

    Google Scholar 

  • Parfit, D. (1984). Reasons and Persons. Oxford University Press.

    Google Scholar 

  • Petri, H. (2019). Asymptotic properties of welfare relations. Economic Theory, 67(4), 853–874.

    Article  Google Scholar 

  • Pivato, M. (2014). Additive representation of separable preferences over infinite products. Theory and Decision, 77(1), 31–83.

    Article  Google Scholar 

  • Pivato, M. (2021). Intertemporal choice with continuity constraints. Mathematics of Operations Research, 46(3), 1203–1229.

    Article  Google Scholar 

  • Pivato, M. (2022). A characterization of Cesàro average utility. Journal of Economic Theory, 201, 105440.

    Article  Google Scholar 

  • Pivato, M., 2023. Cesàro average utilitarianism in relativistic spacetime. Social Choice and Welfare (to appear). https://doi.org/10.1007/s00355-023-01470-6

  • Pressman, M. (2015). A defence of average utilitarianism. Utilitas, 27(4), 389–424.

    Article  Google Scholar 

  • Ramsey, F. P. (1928). A mathematical theory of saving. The Economic Journal, 38(152), 543–559.

    Article  Google Scholar 

  • Roemer, J., & Suzumura, K. (Eds.). (2007). Intergenerational Equity and Sustainability. Palgrave Macmillan.

    Google Scholar 

  • Ryberg, J., Tännsjö, T. (Eds.), 2004. The Repugnant Conclusion: Essays on population ethics. Springer.

  • Sakai, T. (2010). A characterization and an impossibility of finite length anonymity for infinite generations. Journal of Mathematical Economics, 46(5), 877–883.

    Article  Google Scholar 

  • Sakai, T. (2010). Intergenerational equity and an explicit construction of welfare criteria. Social Choice and Welfare, 35(3), 393–414.

    Article  Google Scholar 

  • Sakai, T. (2016). Limit representations of intergenerational equity. Social Choice and Welfare, 47(2), 481–500.

    Article  Google Scholar 

  • Steinhardt, P. J., & Turok, N. (2002). A cyclic model of the universe. Science, 296(5572), 1436–1439.

    Article  Google Scholar 

  • Svensson, L.-G. (1980). Equity among generations. Econometrica, 658, 1251–1256.

    Article  Google Scholar 

  • Tegmark, M., 2004. Parallel universes, Ch. 21. In: Barrow et al. (2004), pp. 459–491.

  • Vallentyne, P., & Kagan, S. (1997). Infinite value and finitely additive value theory. The Journal of Philosophy, 94(1), 5–26.

    Article  Google Scholar 

  • von Weizsäcker, C. C. (1965). Existence of optimal programs of accumulation for an infinite time horizon. Review of Economic Studies, 32, 85–104.

    Article  Google Scholar 

  • Wilkinson, H. (2021). Infinite aggregation: Expanded addition. Philosophical Studies, 178(6), 1917–1949.

    Article  Google Scholar 

  • Wilkinson, H. (2023a). Infinite aggregation and risk. Australasian Journal of Philosophy, 101(2), 340–359.

    Article  Google Scholar 

  • Wilkinson, H., (2023b). Aggregation in an infinite, relativistic universe. Erkenntnis (to appear).

  • Wilkinson, H., (2023c). Chaos, add infinitum, (preprint).

  • Willard, S. (2004). General Topology. Dover Publications Inc.

    Google Scholar 

  • Zalta, E. N. (Ed.), Summer 2022. The Stanford Encyclopedia of Philosophy, Summer 2022 Edition. Metaphysics Research Lab, Stanford University.

  • Zame, W. R. (2007). Can intergenerational equity be operationalized? Theoretical Economics, 2, 187–202.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcus Pivato.

Additional information

This research supported by Labex MME-DII (ANR11-LBX-0023-01) and CHOp (ANR-17-CE26-0003). I am grateful to Geir Asheim, Jean Baccelli, Marc Fleurbaey, Adam Jonsson, Jean-François Laslier, Christian List, Bill Zwicker, and an anonymous reviewer for many excellent suggestions.

Appendices

Appendices

A Against exponentially discounted utility sums

In economics, the standard way to evaluate an infinite future stream of utility is by an exponentially discounted sum. Of course, discounting the interests of some people just because they have the misfortune to be born later than others can be challenged on purely moral grounds. In Ramsey’s (1928) famous words, it is “... a practice which is ethically indefensible and arises merely from the weakness of the imagination.”Footnote 49 Nevertheless, exponential discounting is ubiquitous in economics because of its ease and practicality. So it is worth explaining why this approach cannot work if we take modern physics seriously.Footnote 50

One problem is that exponential discounting is typically applied to moments in time, whereas we must also aggregate utilities across space. There are two natural responses: (i) confine attention only to the future light cone of the observer; and (ii) apply exponential discounting with respect to spatial as well as temporal distances. However, I shall now explain why neither solution is viable.

According to relativistic physics, no signal or other causal influence can travel faster than the speed of light. Consider an observer who is situated at some point o in spacetime. Her future light cone is the set of all points in spacetime that she can reach from o with a signal travelling at light speed or slower, as shown in Figure 7(a). This is the set of all locations in future spacetime where she can exert any causal influence —hence, arguably, the set of all locations relevant to a consequentialist moral evaluation of her present actions.

Fig. 7
figure 7

A sketch of relativistic spacetime. Heuristically, the vertical axis represents “time” in these figures, while the horizontal axis represents “space”. But this is only a heuristic: in relativistic physics there is no clear distinction between space and time. (a) Future and past light cones through the point o. The lines represent the paths of possible signals through spacetime. (b) Curves of constant distance from o in the Minkowski metric. Note that Minkowski distances are only well-defined inside the light cones

So it seems that she should just apply exponential discounting to events in her future light cone. To do this, she must assign a “time” coordinate to each point in her future light cone. But according to relativistic physics, there is no objective time coordinate. As shown in Figures 4 and 5 in Sect. 3, two observers with different velocities will assign different time coordinates to the future, and thus exponentially discount future utilities in different ways. So they will disagree in their ethical evaluations —even if they use the same discount factor. (See also the bottom row of Figure 6 in Sect. 3.)

There is a further problem. The main justification for exponential discounting (since Koopmans, 1960) is Stationarity: if the utility stream \((u_0,v_1,v_2,v_3,\ldots )\) is preferred to stream \((u_0,w_1,w_2,w_3,\ldots )\), then \((v_1,v_2,v_3,\ldots )\) should be preferred to \((w_1,w_2,w_3,\ldots )\). This axiom of “intertemporal consistency” prevents an observer from gratuitously reversing her preferences as time passes. But in relativistic spacetime, Stationarity is impossible: whenever the observer changes her velocity, her reference frame changes with it, in a manner which violates Stationarity even if she is an exponential discounter. So an observer can satisfy Stationarity only if she travels in a straight line at a constant speed, forever.

For the small velocity differences between different observers on planet Earth, and over the small spatiotemporal scale of everyday decisions, these interpersonal disagreements and intrapersonal deviations from Stationarity are negligible. But when extrapolated over the entire infinite future, they can be quite significant —even for small velocity differences. And for observers with large velocity differences (who may exist in the future), these disagreements and deviations can be significant even on small spatiotemporal scales.

A partial solution is a different form of discounting, which I shall call Minkowski discounting. This means that each observer exponentially discounts each spacetime location s in her future lightcone according to Minkowski metric distance from her current location to s. The Minkowski metric measures how much subjective time (or proper time) she would experience if she travelled from her current position to s in a straight line at a constant speed; see Figure 7(b).Footnote 51 The advantage of Minkowski discounting is that Minkowski distances are independent of the observer’s reference frame. So if Alice and Bob are both at the same spacetime position, they will Minkowski-discount events at s by the same amount, even if they have different velocities. So Alice and Bob will agree on their ethical evaluations. But the disadvantage of Minkowski discounting is that it violates Stationarity even for observers travelling at a constant speed. (This can be seen by shifting upwards the parabolas in Figure 7(b).) Likewise, two observers at different points in spacetime will Minkowski-discount the future differently, even if they have the same velocity.Footnote 52

In short, it should be clear that proposal (i) is unworkable. But now the problems with proposal (ii) should also be clear. In addition to being even more “ethically indefensible” than temporal discounting, spatial discounting faces the problem that in a relativistic spacetime, there is no observer-independent way to measure spatiotemporal distances. Different observers, travelling at different velocities, will have different reference frames, hence assign coordinates to spacetime in different ways, resulting in different measurements of distance.Footnote 53 Furthermore, the Minkowski metric is of no help. Although it assigns positive “distances” to points in the interior of an observer’s light cones, it assigns zero distance to any point on their boundaries (i.e. along the trajectory of photons) and assigns imaginary number distances to points outside the light cone. So using the Minkowski metric for spatial discounting would yield absurdities.

The upshot of this analysis is that exponential discounting is a nonstarter for any ethical theory which is intended to apply over large stretches of space and time.Footnote 54 This provides another reason (beyond the purely ethical argument) to seek a nondiscounted way of aggregating the utilities of people scattered across space and time.

B Against utility density

Recall the terminology of Sect. 2, where \({\mathcal { S}}\) is spacetime equipped with a metric d, and c is a “centre” point in \({\mathcal { S}}\). For any positive real number r, let \({\mathcal { S}}(c,r)\) be the ball of radius r around c in \({\mathcal { S}}\).Footnote 55 Let V(r) be the volume of this ball.Footnote 56 The set \({\mathcal { N}}_c({\mathbf{ s}},r)\) defined in Sect. 2 is just the set of people whose spatiotemporal locations are inside of \({\mathcal { S}}(c,r)\). As in Sect. 4.3, let \({\mathcal { X}}\) be a set of life outcomes. Let \(({\mathbf{ s}},{\mathbf{ x}})\) be a world, and let u be a function from \({\mathcal { X}}\) into \({\mathbb {R}}\), assigning a utility to every possible life outcome in \({\mathcal { X}}\). The utility density of this world is the limit, as r becomes very large, of the total utility of all people contained in a ball of radius r, divided by its volume.Footnote 57 (Bostrom (2011, §2.3) calls this value density.) As already noted in Sect. 1, an axiology based on utility density yields a population density version of the Repugnant Conclusion. Also, the definition of utility density assumes that there is an objective, observer-independent way to measure distances in spacetime —an assumption already refuted in Sect. 3 and Appendix A. But this axiology is also untenable for several other reasons. First, consider an infinite universe with only a finite number of people. In such a universe, the utility density would always be zero, regardless of the quality of people’s lives. Thus, a utility density axiology effectively requires an infinite population. Although I have earlier argued that an infinite population is the most likely world, it would be strange if our axiology depended on such a hypothesis to be nontrivial.Footnote 58 Alternately, suppose the population is infinite, but confined to a two-dimensional subset of three-dimensional space. In such a universe, the utility density must be zero, regardless of the quality of people’s lives. This is absurd.

Furthermore, our best scientific theories say that the universe will expand forever, so the spatial density of matter and energy is will decrease to zero over time. On the plausible assumption that all forms of life require some substrate of matter and energy, this means the density of people must also decrease to zero over time. Since the balls considered in the definition of utility density expand in time as well as in space, they will eventually be mostly occupied by the mostly empty spacetime of the far future; this again implies that the utility density must be zero, regardless of people’s wellbeing. This is absurd.

Of course, our theories may be wrong. If the mass-energy density of the universe is high enough, then its expansion will eventually halt and go into reverse, which would obviate the problem in the previous paragraph. But it would be very odd if the viability of our axiology depended upon whether the cosmos exceeds a certain critical density.

Fig. 8
figure 8

Left. A two-dimensional manifold equipped with a metric giving it the geometry of an open disk in a flat Euclidean plane, with finite area. Right. The same two-dimensional manifold equipped with a different metric, giving it the geometry of an unbounded hyperbolic plane, with infinite area. Every triangle in the picture has the same area. (Image credit: Anton Sherwood)

But this is just one symptom of a much deeper problem. The definition of utility density takes for granted that there is some objective way to define the “volume” of a region in spacetime. There is a natural notion of “volume” in familiar three-dimensional Euclidean space, and also in the four-dimensional Minkowski spacetime underlying special relativity (the Lebesgue measure). In both cases, it arises from the underlying geometric structure of the space (its inner product structure). However, in general relativity, spacetime is a four-dimensional differentiable manifold without any intrinsic geometric structure. The geometric structure of relativistic spacetime is entirely determined by its distribution of matter and energy.Footnote 59 Different mass-energy distributions yield different geometries, hence different measures of volume.

For a simple (non-relativistic) example, consider Figure 8. This shows a two-dimensional manifold (i.e. a surface) equipped with two different metrics. In the left-hand image, the manifold has a metric giving it the geometry of an open disk in a flat Euclidean plane. In the right-hand image, the same manifold has a different metric, giving it the geometry of a hyperbolic plane, which has infinite area. The relevant examples in general relativity are four-dimensional, and furthermore involve a Lorentzian metric. But the basic point remains the same: volume is determined by the geometry of the manifold, and in general relativity, this geometry is not fixed in advance.

This is problematic, because the volume V(r) plays a crucial role in the definition of utility density. There can be two universes, with exactly same people, each having exactly the same quality of life and occupying exactly the same “positions” on the underlying spacetime manifold (i.e. the same world \(({\mathbf{ s}},{\mathbf{ x}})\)), but with different distributions of matter and energy, leading to different geometries and hence different values for the utility density. Worse: a modification to \(({\mathbf{ s}},{\mathbf{ x}})\) which increases utility density in one universe might decrease utility density in the other universe. Hence whether or not this modification counts as an improvement would depend on the spatiotemporal distribution of (non-living) matter and energy. This is obviously absurd.

C Index of notation

\(\succcurlyeq\):

A weak order (i.e. complete, transitive binary relation) over worlds, representing an axiology. Its symmetric part is denoted \(\approx\), and its antisymmetric part is denoted \(\succ\).

\(\succcurlyeq _*\):

A weak order over individual life outcomes, induced by \(\succcurlyeq\). Its symmetric part is denoted \(\approx _*\), and its antisymmetric part is denoted \(\succ _*\).

c:

A point in spacetime \({\mathcal { S}}\) used as a “centre” for computing spatiotemporal Cesàro average utilities.

d:

A metric on \({\mathcal { X}}\), and also a metric on \({\mathcal { S}}\).

\(\delta\):

A metric on the set \(\Phi\) of reference frames; see §4.1.

\(d_{\infty }\):

The supremum metric induced by d on the set of social outcomes; see Footnotes 11 and 39.

\(\eta :{\mathcal { N}}{{\longrightarrow }}{\mathbb {N}}\):

An “enumeration” of people according to their locations in spacetime.

\({\varvec{\mathcal {K}}}=\{{\mathcal { K}}_n\}_{n\in {\mathcal { N}}}\):

A soma, indicating the spatiotemporal location and extension of every person. Here, \({\mathcal { K}}_n\subseteq {\mathcal { O}}\) for all \(n\in {\mathcal { N}}\); see §4.2.

\({\varvec{\mathcal {L}}}=\{{\mathcal { L}}_n\}_{n\in {\mathcal { N}}}\):

Another soma.

\({\mathcal { N}}\):

A countably infinite indexing set, used to label all the people who will ever live in the universe.

\({\mathcal { N}}_{c,\psi }({\varvec{\mathcal {K}}},r)\):

The set of all indices of people whose spacetime location appears to be contained in a ball of radius r around the center point c, with respect to some reference frame in a ball of radius r around the central reference frame \(\psi\); see Footnote 28 in §4.2.

\({\mathbb {N}}\):

\(:=\{1,2,3,\ldots \}\) The set of natural numbers.

\(n,m,\ldots\):

Generic elements of \({\mathcal { N}}\), or generic natural numbers.

\({\mathcal { O}}\):

A set representing objective spacetime; see §4.1.

\(\Phi\):

The set of possible reference frames; see §4.1.

\(\Phi (\psi ,r)\):

The set of all reference frames in a ball of radius r around the central reference frame \(\psi\); see Footnote 28 in §4.2.

\(\phi ,\phi ',\ldots\):

Generic reference frames; see §4.1.

\(\psi\):

A reference frame in \(\Phi\) used as a “central” reference frame for computing spatiotemporal Cesàro average utilities.

\({\mathbb {R}}\):

The set of all real numbers.

\({\mathbb {R}}_+\):

The set of all positive real numbers.

\(r,R,\ldots\):

Generic real numbers

\(r_{c,\psi }(n)\):

The smallest radius r such that person n appears to be contained in a ball of radius r around the center point c, with respect to some reference frame in a ball of radius r around the central reference frame \(\psi\).

\({\mathcal { S}}\):

A set representing subjectively perceived spacetime; see §4.1 and the start of §2.

\({\mathcal { S}}(c,r)\):

The set of all points in subjective spacetime \({\mathcal { S}}\) within distance r of the centre point c; see Footnote 28 in §4.2.

\(s,s'\ldots\):

Generic elements of \({\mathcal { S}}\).

\({\mathbf{ s}},{\mathbf{ s}}',\ldots\):

Generic spatiotemporal arrangements —that is, sequencess \((s_1,s_2,s_3,\ldots )\) that assign a spacetime location \(s_n\) in \({\mathcal { S}}\) to every index n in \({\mathcal { N}}\); see the start of §2.

\({\mathcal { W}}\):

The set of all nomologically possible worlds; see the end of §4.4.

\(u:{\mathcal { X}}{{\longrightarrow }}{\mathbb {R}}\):

A “utility” function. It assigns a numerical value to every possible life outcome, measuring the overall quality of that life for the purposes of ethical evaluation.

\({\mathcal { X}}\):

A locally compact metric space, representing the set of all possible individual life outcomes; see §4.3 and the start of §2.

\(x,y,z,\ldots\):

Generic life outcomes —i.e., elements of \({\mathcal { X}}\).

\({\mathbf{ x}},{\mathbf{ y}},{\mathbf{ z}},\ldots\):

Generic social outcomes.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pivato, M. Population ethics in an infinite universe. Philos Stud 180, 3383–3414 (2023). https://doi.org/10.1007/s11098-023-02014-5

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11098-023-02014-5

Navigation