Skip to main content
Log in

Optimal population and exhaustible resource constraints

  • Original Paper
  • Published:
Journal of Population Economics Aims and scope Submit manuscript

Abstract

A large literature considers the optimal size and growth rate of the human population, trading off the utility value of additional people with the costs of a larger population. In this literature, an important parameter is the social weight placed on population size; a standard result is that a planner with a larger weight on population chooses larger population levels and growth rates. We demonstrate that this result is conditionally overturned when an exhaustible resource constraint is introduced: if the discount rate is small enough, the optimal population today decreases with the welfare weight on population size. That is, a more total-utilitarian social planner could prefer a smaller population today than a more average-utilitarian social planner. We also present a numerical illustration applied to the case of climate change, where we show that under plausible real-world parameter values, our result matters for the direction and magnitude of optimal population policy.

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

Similar content being viewed by others

Notes

  1. As we will discuss later, an alternative approach that we do not study in this paper is to consider Pareto efficiency rather than social welfare maximization; Golosov et al. (2007) present a generalization of Pareto efficiency to a case with variable population.

  2. An important related paper in the population economics literature by Nerlove et al. (1986a) studies a similar, but distinct question. They consider whether parents’ procreative decisions would be inefficient in the face of two externalities: a fixed stock of land that makes a larger population less productive, and public goods that are cheaper for a larger population to support; they find that these externalities offset. An extension in Appendix 3, which brings their paper and ours closer together, models the technological progress as a public good that, as in their model, is more easily achieved by a larger population.

  3. We abstract from issues of inequality within generation by assuming that utility depends on consumption per person.

  4. If utility is negative, as would be the case if utility takes the CRRA form with a risk-aversion coefficient greater than 1, then Jones and Schoonbroodt (2010) show that α must be negative in any sensible social welfare function; a more total-utilitarian planner would then correspond to a smaller (more negative) value of α.

  5. Parfit (1984) notes that almost any important policy change will change the set of persons who are born, especially in future generations, but to a social planer that values utilities anonymously, for many policy changes the effect on the number of people may be small.

  6. An alternative approach to accounting for population is suggested by Blackorby et al. (1995): critical-level utilitarianism, where adding a new person to the population increases welfare only if their utility is above the critical value η.

  7. We thank Marc Fleurbaey for suggesting an analysis of such a specification.

  8. However, interest in the problem was not extinguished by this finding. Palivos and Yip (1993) claim to have found the opposite result: with CRRA utility and a coefficient of relative risk-aversion greater than one, they find that a larger exponent on population in the social welfare function leads to a smaller optimal population size; Boucekkine and Fabbri (2013) find a similar result. However, this does not imply that a more total-utilitarian planner would prefer a smaller population. As noted by Boucekkine and Fabbri (2013) and Boucekkine et al. (2014), a risk-aversion coefficient greater than one implies negative utility, which—if the utility of non-existence is assumed to be zero—implies that even the best possible life is worse in utility terms than not existing; if the exponent on population size is positive, the existence of persons is assumed to make society directly worse off. Jones and Schoonbroodt (2010) demonstrate that, in this case, an economically sensible social welfare function must feature a negative α, and it is easy to show that a higher weight on population then corresponds to a smaller (more negative) α. Thus, the finding of Palivos and Yip (1993) is useful because it confirms that this case is in fact entirely consistent with the standard result: a Benthamite planner, who places a higher weight on population, would choose a larger population.

  9. Recent surveys on the economics of climate change are provided by Pindyck (2013) and Stern (2013).

  10. Having an additional child imposes externalities upon other people, because that child will create additional emissions that impose costs on other people, but an emissions tax is powerless on this margin as long as the proceeds of the tax are redistributed to society: if each new child pays an expected amount of emissions taxes but receives an equal amount of lump-sum reimbursements, the tax on creating new emissions through fertility is zero. Thus, Harford (1997) finds that, whatever the level of the emissions tax, the externality generated by adding a person is the amount of the resources they and all of their descendants use up, including their greenhouse gas emissions.

  11. In the non-economic literature, Bradshaw and Brook (2014) argue that a reduction in the human population is not a quick fix for environmental problems, because reducing population would take too long; however, they do not compute or report any effects of population size on climate outcomes or social welfare. They appear to be considering massive reductions in population, as they speculate (but do not compute) that the optimal human population may be about 1 to 2 billion. Spears (2015) points out that these findings do not rule out that feasible reductions in population size or growth could importantly affect the risk of climate catastrophe. Neither of these papers compute an optimal population under any well-defined social welfare function, or even explicitly consider the social welfare implications of population size.

  12. They evaluate optimal policies for both average and total utilitarianism, and find that total utilitarianism leads to higher taxes on both children and emissions, and more emissions abatement, but they do not study the impact on the optimal population size.

  13. However, we do present an extension of our recursive dynasty model that includes physical capital in Appendix 3, to show that similar results apply in that case. We also abstract from the quantity-quality tradeoff examined by papers such as Baudin (2011).

  14. The role of population growth in weighting the importance of future periods under a TU SWF is discussed in the context of climate policy by Budolfson et al. (2017).

  15. Appendix 3 analyzes the robustness of our results to the weakening of these assumptions. An extension in which resource use per person is a choice variable is presented in Appendix 3, while Appendix 3 presents an extension in which the resource is renewable, and Appendix 3 contains one in which technology is endogenous to the size of the population.

  16. This role of technological progress is not a mere convenient modeling assumption: the interaction between environmental constraints and technological progress is central to the empirical understanding of twentieth century population growth by demographers and economists. For an overview, see Lam’s (2011) overview of how technological change, incentivized by prices in part as a result to population growth, shaped the use of limited natural resources by a growing population.

  17. A leading precursor to this model is presented by Cigno (1981), which gives the first characterization of a steady state with exhaustible resources and endogenous population assuming a constant saving ratio. Our present exogenous-utility framework abstracts away from savings; nevertheless, this model is an important precedent for moving away from Stiglitz’s (1974) prior assumption of a constant population growth rate. Instead, Cigno investigates the consequences of making population growth a function of consumption and capita per capita. More generally, this early literature considered carefully how to model the interaction of technological progress and population and consumption growth—we, too, depend on technological progress in our model, as the source of an incentive for the social planner to permit a larger population later.

  18. It is simple to confirm that the second-order conditions are satisfied; they are also satisfied in the other models in the body of the paper, and in all models in the appendices with two exceptions. As discussed in Appendices 3 and 4, there is the possibility of non-concavity in the case of our alternative specification for average utilitarianism, and in the model with capital. We simply assume the existence of an interior optimum in the former case, which is supported by simulation results with reasonable parameters, and we confirm numerically that the second-order condition is satisfied in our simulations for the latter case.

  19. If β γ < 1, the optimal population declines asymptotically to zero over time. In this case, a higher α leads to a higher starting population and a faster decline; the effect on current population is reversed, but the intuition is the same, as a more total-utilitarian planner wants to increase population more in the periods when it is already large.

  20. Note that faster technological improvement γ raises the optimal growth rate and lowers n 1; if people in the future will use less resources, the optimal population path allocates more resources to later periods when especially large populations can be supported. This effect is increasing in α: a more total-utilitarian planner reacts by lowering initial population and raising the growth rate by a greater amount. Therefore, \(\frac {dt^{*}}{d\gamma } > 0\): the crossing point of population paths moves later when technology improves faster. This result also holds numerically in our analysis of the dynasty and climate change models. We thank David de la Croix for the suggestion of analyzing the comparative statics with respect to γ.

  21. We continue to abstract from capital—except in the extension in Appendix 3—and to assume that generations do not overlap as in Golosov et al. (2007). Note however that our results hold with overlapping generations if only one generation is working at a time, or if production by each generation uses a unique technology.

  22. In the current specification, since income is constant, this could instead by interpreted as a time cost of κ percent per child; replacement fertility reduces net income to (1 − κ) percent of the maximum value. The distinction between time and monetary costs of fertility is beyond the scope of our analysis, but in our analysis of the variable resource use model in Appendix 3, we have confirmed that an altered specification of the model in which fertility costs are proportional to output does not qualitatively change the results from simulation. We thank David de la Croix for mentioning this issue.

  23. This is in the spirit of Nerlove et al.’s (1986a) model of a larger population contributing to a public good more inexpensively—here, technological change.

  24. This is comparable to the time periods studied in standard economic climate models such as Nordhaus’ DICE and RICE models, and the assumption that climate change eventually will be exogenously solved is similar to Nordhaus’ “backstop” technology.

  25. For example, if higher- α planners want a larger long-run population, or if technology continues to increase exponentially, expanding the population quickly after 2200 may be desirable, which would make steep population growth in the final decades of the 2100s beneficial; and yet it seems unrealistic to make our results depend on such unknowable very-long-term parameters.

  26. Of course, actual emissions are highly unequally distributed among people living today. Although population-reduction policies are typically discussed in the context of poorer countries, emissions per person are much higher in richer countries. Therefore, any actual policy implications of our results may be more relevant to richer countries where emissions are greater.

  27. By ignoring uncertainty, we do not have to model the outcome in the case of catastrophic impacts as in Weitzman (2009); in that sense, our analysis is very like Bohn and Stuart (2015) in assuming a level of emissions that must not be exceeded, though here that limit is intertemporal rather than period-by-period.

  28. In a somewhat different two-period setting where consumption is drawn from a fixed quantity of capital, Nerlove et al. (1985, 1986b) show that a more total-utilitarian planner prefers higher population growth, but in that setting the starting population is fixed, and a larger future population is made possible by reducing consumption per capita.

  29. Otherwise, if β is sufficiently close to one, then the optimal n 1 is infinitesimally small, as the planner is willing to delay population indefinitely.

  30. Indeed, in our simulations, we have found numerically extreme cases in which a higher α causes optimal population to rise slightly at first, then decline for many periods, and then rise again.

  31. However, Dasgupta (1998) also objects to the overall average utility specification, arguing that it is not time-consistent across generations. Thus, he argues that average utilitarianism as a philosophical concept is fundamentally flawed, an issue that we do not intend to get into; we simply assume that different planners could hold social welfare functions that differ according to the weight placed on population, whatever the philosophical issues underlying them.

  32. When dividing utility by population, we must use the discount rate β, as otherwise the term on the bottom is likely to be infinite; additionally, discounting on the top but not on the bottom would be inconsistent, and would imply that population should likely be made as large as possible in the near future, declining in the long run.

References

  • Abel G J, Barakat B, Samir K, Lutz W (2016) Meeting the sustainable development goals leads to lower world population growth. PNAS 113(50):14294–14299

    Article  Google Scholar 

  • Asheim G B, Buchholz W, Hartwick J M, Mitra T, Withagen C (2007) Constant savings rates and quasi-arithmetic population growth under exhaustible resource constraints. J Environ Econ Manag 53(2):213–229

    Article  Google Scholar 

  • Barro R J, Becker G S (1989) Fertility choice in a model of economic growth. Econometrica 57(2):481–501

    Article  Google Scholar 

  • Baudin T (2011) Family policies: what does the standard endogenous fertility model tell us? J Public Econ Theory 13(4):555–593

    Article  Google Scholar 

  • Becker G S, Barro R J (1988) A reformulation of the economic theory of fertility. Q J Econ 103(1):1–25

    Article  Google Scholar 

  • Blackorby C, Bossert W, Donaldson D (1995) Intertemporal population ethics: critical-level utilitarian principles. Econometrica 63(6):1303–1320

    Article  Google Scholar 

  • Bohn H, Stuart C (2015) Calculation of a population externality. Amer Econ J Econ Polic 7(2):61–87

    Article  Google Scholar 

  • Boucekkine R, Fabbri G (2013) Assessing Parfit’s repugnant conclusion within a canonical endogenous growth set-up. J Popul Econ 26(2):751–767

    Article  Google Scholar 

  • Boucekkine R, Fabbri G, Gozzi F (2014) Egalitarianism under population change: age structure does matter. J Math Econ 55:86–100

    Article  Google Scholar 

  • Bradshaw C J, Brook B W (2014) Human population reduction is not a quick fix for environmental problems. PNAS 111(46):16610–16615

    Article  Google Scholar 

  • Budolfson M, Dennig F, Fleurbaey M, Scovronick N, Siebert A, Spears D, Wagner F (2017) Optimal climate policy and the future of world economic development. World Bank Econ Rev, forthcoming

  • Cigno A (1981) Growth with exhaustible resources and endogenous population. Rev Econ Stud 48(2):281–287

    Article  Google Scholar 

  • Cole M A, Neumayer E (2004) Examining the impact of demographic factors on air pollution. Popul Environ 26(1):5–21

    Article  Google Scholar 

  • Conde-Ruiz J I, Giménez E L, Pérez-Nievas M (2010) Millian efficiency with endogenous fertility. Rev Econ Stud 77(1):154–187

    Article  Google Scholar 

  • Conly S (2015) One child. Do we have a right to more? Oxford University Press, New York

    Google Scholar 

  • Connelly M (2008) Fatal misconception: the struggle to control world population. Harvard University Press, Cambridge

    Google Scholar 

  • Dasgupta P (1969) On the concept of optimum population. Rev Econ Stud 36 (3):295–318

    Article  Google Scholar 

  • Dasgupta P (1995) The population problem: theory and evidence. J Econ Lit 33(4):1879–1902

    Google Scholar 

  • Dasgupta P (1998) Population, Consumption and Resources: Ethical Issues. Ecol Econ 24(2–3):139–152

    Article  Google Scholar 

  • Dasgupta S, Mitra T (1982) On some problems in the formulation of optimal population policies when resources are depletable. In: Eichhorn W, Henn R, Neumann K, Shephard R (eds) Economic theory of natural resources. Springer, pp 409–429

  • de la Croix D, Gosseries A (2009) Population policy through tradable procreation entitlements. Int Econ Rev 50(2):507–542

    Article  Google Scholar 

  • Dietz T, Rosa E A (1997) Effects of population and affluence on C O 2 emissions. PNAS 94(1):175–179

    Article  Google Scholar 

  • Gaffin S R, O’Neill B C (1997) Population and global warming with and without CO 2 Targets. Popul Environ 18(4):389–413

    Article  Google Scholar 

  • Golosov M, Jones L E, Tertilt M (2007) Efficiency with endogenous population growth. Econometrica 75(4):1039–1071

    Article  Google Scholar 

  • Harford J D (1997) Stock pollution, child-bearing externalities, and the social discount rate. J Environ Econ Manag 33(1):94–105

    Article  Google Scholar 

  • Harford J D (1998) The ultimate externality. Amer Econ Rev 88(1):260–265

    Google Scholar 

  • Jones L E, Schoonbroodt A (2010) Complements versus substitutes and trends in fertility choice in dynastic models. Int Econ Rev 51(3):671–699

    Article  Google Scholar 

  • Kelly D L, Kolstad C D (2001) Malthus and climate change: betting on a stable population. J Environ Econ Manag 41(2):135–161

    Article  Google Scholar 

  • Lam D (2011) How the world survived the population bomb: lessons from 50 years of extraordinary demographic history. Demography 48(4):1231–1262

    Article  Google Scholar 

  • Liddle B (2011) Consumption-driven environmental impact and age structure change in OECD countries: a cointegration-STIRPAT analysis. Demogr Res 24:749–770

    Article  Google Scholar 

  • Mitra T (1983) Limits on population growth under exhaustible resource constraints. Int Econ Rev 24(1):155–168

    Article  Google Scholar 

  • Nerlove M, Razin A, Sadka E (1982) Population size and the social welfare functions of Bentham and Mill. Econ Lett 10(1–2):61–64

    Article  Google Scholar 

  • Nerlove M, Razin A, Sadka E (1985) Population size: individual choice and social optima. Q J Econ 100(2):321–334

    Article  Google Scholar 

  • Nerlove M, Razin A, Sadka E (1986a) Endogenous population with public goods and malthusian fixed resources: efficiency or market failure. Int Econ Rev 27 (3):601–609

  • Nerlove M, Razin A, Sadka E (1986b) Some welfare theoretic implications of endogenous fertility. Int Econ Rev 27(1):3–31

  • Nordhaus W D (2008) A question of balance: weighing the options on global warming policies. Yale University Press, New Haven and London

    Google Scholar 

  • O’Neill B C, MacKellar F L, Lutz W (2005) Population and climate change. Cambridge University Press, New York

    Google Scholar 

  • O’Neill B C, Dalton M, Fuchs R, Jiang L, Pachauri S, Zigova K (2010) Global demographic trends and future carbon emissions. PNAS 107(41):17521–17526

    Article  Google Scholar 

  • O’Neill B C, Liddle B, Jiang L, Smith K R, Pachauri S, Dalton M, Fuchs R (2012) Demographic change and carbon dioxide emissions. Lancet 380 (9837):157–164

    Article  Google Scholar 

  • Palivos T, Yip C K (1993) Optimal population size and endogenous growth. Econ Lett 41(1):107–110

    Article  Google Scholar 

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

    Google Scholar 

  • Pindyck R S (2013) Climate change policy: what do the models tell us? J Econ Lit 51(3):860–872

    Article  Google Scholar 

  • Razin A, Ben-Zion U (1975) An intergenerational model of population growth. Amer Econ Rev 65(5):923–933

    Google Scholar 

  • Razin A, Yuen C -W (1995) Utilitarian tradeoff between population growth and income growth. J Popul Econ 8(1):81–87

    Article  Google Scholar 

  • Renström T I, Spataro L (2011) The optimum growth rate for population under critical-level utilitarianism. J Popul Econ 24(3):1181–1201

    Article  Google Scholar 

  • Robinson J A, Srinivasan T (1997) Long-term consequences of population growth: technological change, natural resources, and the environment. In: Rosenzweig M, Stark O (eds) Handbook of population and family economics, vol 1b. Elsevier, pp 1175–1298

  • Spears D (2015) Smaller human population in 2100 could importantly reduce the risk of climate catastrophe. PNAS 112(18):E2270

    Article  Google Scholar 

  • Spears D (2017) Making people happy or making happy people? Questionnaire-experimental studies of population ethics and policy. Soc Choice Welf 49(1):145–169

    Article  Google Scholar 

  • Stern N (2007) The economics of climate change: the Stern Review. Cambridge University Press, Cambridge and New York

    Book  Google Scholar 

  • Stern N (2013) The structure of economic modeling of the potential impacts of climate change: grafting gross underestimation of risk onto already narrow science models. J Econ Lit 51(3):838–859

    Article  Google Scholar 

  • Stiglitz J E (1974) Growth with exhaustible natural resources: the competitive economy. Rev Econ Stud 41 (Symposium on the Economics of Exhaustible Resources):139–152

    Article  Google Scholar 

  • Weitzman M L (2009) On modeling and interpreting the economics of catastrophic climate change. Rev Econ Stat 91(1):1–19

    Article  Google Scholar 

  • Wheeler D, Hammer D (2010) The economics of population policy for carbon emissions reduction in developing countries, working paper 229. Center for Global Development

Download references

Acknowledgments

We would like to thank Raouf Boucekkine, David de la Croix, Giorgio Fabbri, Marc Fleurbaey, and seminar participants at the University of Copenhagen and the Indian Statistical Institute—Delhi for their comments. We are especially grateful for the help and guidance provided by two anonymous referees of this journal and by the editor. Any errors or omissions are the responsibility of the authors.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nicholas Lawson.

Ethics declarations

This project received no funding. The authors declare that they have no conflict of interest. Although unrelated to this project, Dean Spears, in his capacity of Executive Director of r.i.c.e. (a 501(c)3 public charity), has received grants from the Bill and Melinda Gates Foundation, the NIH, USAID, and the IGC, and has been a paid consultant for the World Bank.

Additional information

Responsible editor: Alessandro Cigno

Appendices

Appendix A: Algebra and proofs

This appendix presents the algebra and proofs from our main model in Section 3 of the paper.

1.1 A.1 Simple two-period exogenous-utility model

Consider first the simplest possible version of the model, with only two relevant periods: t ∈{1, 2} (the population and utility in period 0 is exogenously fixed). The planner’s problem is:

$$\mathcal{L} = n_{1}^{\alpha}u_{1} + \beta n_{2}^{\alpha}u_{2} - \lambda\left( \frac{n_{1}}{T_{1}} + \frac{n_{2}}{T_{2}} - X\right). $$

Differentiating with respect to n 1 and n 2, we find:

$$\alpha n_{1}^{\alpha-1}u_{1} = \frac{\lambda}{T_{1}} $$
$$\beta\alpha n_{2}^{\alpha-1}u_{2} = \frac{\lambda}{T_{2}}. $$

Combining these expressions, and assuming that u 2 = u 1, we find:

$$\frac{n_{2}}{n_{1}} = \left( \beta\gamma\right)^{\frac{1}{1-\alpha}}. $$

Defining the growth factor as \(g \equiv \left (\beta \gamma \right )^{\frac {1}{1-\alpha }}\), and combining n 2 = g n 1 with the resource constraint:

$$ n_{1} = \frac{T_{2}X}{g + \gamma} $$
(4)
$$ n_{2} = \frac{gT_{2}X}{g + \gamma}. $$
(5)

These expressions depend on α only through g, leading to the following simple result:

Proposition 3

n 2 is increasing and n 1 is decreasing with α if and only if β γ > 1.

Proof 3

Differentiate (4) and (5) with respect to g:

$$\frac{dn_{1}}{dg} = \frac{-T_{2}X}{(g+\gamma)^{2}} < 0 $$
$$\frac{dn_{2}}{dg} = \frac{\gamma T_{2}X}{(g+\gamma)^{2}} > 0. $$

Since n 1and n 2depend on α only throughg, and \(\frac {dg}{d\alpha } > 0\)if andonly if β γ > 1,then \(\frac {dn_{1}}{d\alpha } < 0\)and\(\frac {dn_{2}}{d\alpha } > 0\)if andonly if β γ > 1. □

Thus, the optimal population today decreases with α, and the population tomorrow increases, whenever the optimal population growth is positive. When facing a resource constraint with technological improvement, total utilitarians prefer to have fewer people today so as to defer population to the future, when it will be less costly in terms of resource use and thus a larger population can be supported.Footnote 28 In a sense, having a higher α is like having a lower discount rate over the number of people: as long as population is increasing over time, the future is valued more highly when α is larger and the planner wants more people to be alive in that future.

1.2 A.2 Simple infinite-horizon exogenous-utility model

As noted in Section 3.1, in the exogenous-utility model, the lagrangian for the planner’s problem can be written as follows:

$$\mathcal{L} = \sum\limits_{t=0}^{\infty}{\beta^{t}n_{t}^{\alpha}u_{t}} - \lambda\left( \sum\limits_{t=1}^{\infty}{\frac{n_{t}}{T_{t}}} - X\right) $$

and differentiating with respect to n t and n t+1 gives:

$$\beta^{t}\alpha n_{t}^{\alpha-1}u_{t} = \frac{\lambda}{T_{t}} $$
$$\beta^{t+1}\alpha n_{t+1}^{\alpha-1}u_{t+1} = \frac{\lambda}{T_{t+1}}. $$

Combining these two expressions leads to an equation for optimal population growth:

$$\frac{n_{t+1}}{n_{t}} = \left( \frac{\beta\gamma u_{t+1}}{u_{t}}\right)^{\frac{1}{1-\alpha}}. $$

which, if we assume u t to be constant across time, simplifies to:

$$\frac{n_{t+1}}{n_{t}} = \left( \beta\gamma\right)^{\frac{1}{1-\alpha}} \equiv g. $$

Differentiating g with respect to α, we find:

$$\frac{dg}{d\alpha} = \ln\left( \beta\gamma\right)\left( \beta\gamma\right)^{\frac{1}{1-\alpha}}\left( \frac{1}{(1-\alpha)^{2}}\right) $$

and the only term that has an uncertain sign is the first one: if β γ > 1, and thus g > 1, then \(\frac {dg}{d\alpha } > 0\).

Using the equation for population in any period \(n_{t} = \left (\beta \gamma \right )^{\frac {t-1}{1-\alpha }}n_{1}\), and the resource constraint \({\sum }_{t=1}^{\infty }{\frac {n_{t}}{\gamma ^{t-1}T_{1}}} = X\), we can derive the starting value n 1:

$$n_{1}\sum\limits_{t=1}^{\infty}{\left( \beta\gamma^{\alpha}\right)^{\frac{t-1}{1-\alpha}}} = XT_{1}. $$

For this to give a positive solution for n 1, it must be the case that β γ α < 1.Footnote 29 Assuming this to be true, the solution for n 1 is:

$$n_{1} = \left[1 - \left( \beta\gamma^{\alpha}\right)^{\frac{1}{1-\alpha}}\right]XT_{1} $$

and this can be used to give the value for any n t :

$$n_{t} = \left( \beta\gamma\right)^{\frac{t-1}{1-\alpha}}\left[1 - \left( \beta\gamma^{\alpha}\right)^{\frac{1}{1-\alpha}}\right]XT_{1}. $$

Finally, we can differentiate n t with respect to α:

$$\frac{dn_{t}}{d\alpha} = \frac{XT_{1}\ln\left( \beta\gamma\right)\left( \beta\gamma\right)^{\frac{t-1}{1-\alpha}}}{(1-\alpha)^{2}}\left[(t-1)\left( 1 - \left( \beta\gamma^{\alpha}\right)^{\frac{1}{1-\alpha}}\right) - \left( \beta\gamma^{\alpha}\right)^{\frac{1}{1-\alpha}}\right]. $$

If we set this to zero, we find that the critical value of t is:

$$t^{*} = \frac{1}{1 - \left( \beta\gamma^{\alpha}\right)^{\frac{1}{1-\alpha}}}. $$

Notice that the first part of \(\frac {dn_{t}}{d\alpha }\)—the fraction prior to the square brackets—is always positive given the assumption that β γ > 1. The part within the square brackets is clearly increasing with t given that β γ α < 1, and thus we have that \(\frac {dn_{t}}{d\alpha } < 0\) for t < t and \(\frac {dn_{t}}{d\alpha } > 0\) if t > t , exactly as stated in Proposition 1.

1.3 A.3 Intermediate model

We want to prove the single-crossing property in the intermediate model; that is, the existence of a critical time t such that \(\frac {dn_{t}}{d\alpha } < 0\) if t < t and \(\frac {dn_{t}}{d\alpha } > 0\) if t > t .

The derivative of the planner’s problem with respect to n t is:

$$\frac{\partial\mathcal{L}}{\partial n_{t}} = \beta^{t}\alpha n_{t}^{\alpha-1}u(c(n_{t})) + \beta^{t}n_{t}^{\alpha}u^{\prime}(c(n_{t}))c^{\prime}(n_{t}) - \frac{\lambda}{T_{t}} = 0 $$

which leads immediately to:

$$ \left( \beta\gamma\right)^{t}n_{t}^{\alpha-1}\left[\alpha u(c(n_{t})) + n_{t}u^{\prime}(c(n_{t}))c^{\prime}(n_{t})\right] = \frac{\lambda}{T_{0}}. $$
(6)

We first prove that \(\frac {dn_{t}}{dt} > 0\), for which purpose we take the derivative with respect to t; after a bit of simplification, we have:

$$\begin{array}{@{}rcl@{}} &&\ln\left( \beta\gamma\right)n_{t}\left[\alpha u(c(n_{t})) + n_{t}u^{\prime}(c(n_{t}))c^{\prime}(n_{t})\right]\\ &&+ \left[(\alpha-1)\alpha u(c(n_{t})) + n_{t}\left( 2\alpha u^{\prime}(c(n_{t}))c^{\prime}(n_{t}) + n_{t}u^{\prime\prime}(c(n_{t}))c^{\prime}(n_{t})c^{\prime}(n_{t})\right.\right.\\ &&\quad\left.\left.+ n_{t}u^{\prime}(c(n_{t}))c^{\prime\prime}(n_{t})\right)\right]\frac{dn_{t}}{dt} = 0. \end{array} $$

All of the terms multiplying \(\frac {dn_{t}}{dt}\) are at least weakly negative (and the first two are strictly negative), while if β γ > 1, the first term is positive. Therefore, the derivative must be positive: \(\frac {dn_{t}}{dt} > 0\) if and only if β γ > 1.

Next, we take the derivative of the first-order condition with regard to α; whereas the multiplier λ is a constant with regard to t, it is not with regard to α: a change in α will necessitate a different λ to ensure that the resource constraint is binding. Therefore, we have:

$$\begin{array}{@{}rcl@{}} &&\left( \beta\gamma\right)^{t}\left[(\alpha-1)n_{t}^{\alpha-2}\left( \alpha u(c(n_{t})) + n_{t}u^{\prime}(c(n_{t}))c^{\prime}(n_{t})\right)\right.\\ &&\quad\quad\quad\left.+ n_{t}^{\alpha-1}\left( \alpha u^{\prime}(c(n_{t}))c^{\prime}(n_{t}) + u^{\prime}(c(n_{t}))c^{\prime}(n_{t})\right.\right.\\ &&\quad\quad\quad\left.\left. + n_{t}u^{\prime\prime}(c(n_{t}))c^{\prime}(n_{t})c^{\prime}(n_{t}) + n_{t}u^{\prime}(c(n_{t}))c^{\prime\prime}(n_{t})\right)\right]\frac{dn_{t}}{d\alpha}\\ &&+ \left( \beta\gamma\right)^{t}\left[\ln(n_{t})n_{t}^{\alpha-1}\left( \alpha u(c(n_{t})) + n_{t}u^{\prime}(c(n_{t}))c^{\prime}(n_{t})\right) + n_{t}^{\alpha-1}u(c(n_{t}))\right] = \frac{1}{T_{0}}\frac{d\lambda}{d\alpha}. \end{array} $$

Simplifying, the total derivative is:

$$\begin{array}{@{}rcl@{}} &&T_{0}\left( \beta\gamma\right)^{t}n_{t}^{\alpha-2}\left[(\alpha-1)\alpha u(c_{t}) + n_{t}\left( 2\alpha u^{\prime}(c_{t})c^{\prime}(n_{t}) + n_{t}u^{\prime\prime}(c_{t})c^{\prime}(n_{t})^{2}\right.\right.\\ &&\qquad\qquad\qquad\left.\left.+ n_{t}u^{\prime}(c_{t})c^{\prime\prime}(n_{t})\right)\right]\frac{dn_{t}}{d\alpha}\\ &&+ \lambda\ln(n_{t}) + T_{0}\left( \beta\gamma\right)^{t}n_{t}^{\alpha-1}u(c_{t}) = \frac{d\lambda}{d\alpha}. \end{array} $$

To further simplify notation, define g(t) as the terms multiplying \(\frac {dn_{t}}{d\alpha }\); as every term in g(t) is at least weakly negative, and the first two are strictly negative, we can conclude that g(t) < 0 for all values of t. Similarly, define \(f(t) \equiv \frac {d\lambda }{d\alpha } - \lambda \ln (n_{t}) - T_{0}\left (\beta \gamma \right )^{t}n_{t}^{\alpha -1}u(c_{t})\), and then we have:

$$\frac{dn_{t}}{d\alpha} = \frac{f(t)}{g(t)}. $$

The effect of α on n t thus comes down to the expression f(t), and how it changes over time. Because g(t) is always negative, proving single-crossing of f(t) would also prove single-crossing of \(\frac {dn_{t}}{d\alpha }\). We start by defining f(t) as a function of t only through n t , using (6) to substitute for \(T_{0}\left (\beta \gamma \right )^{t}n_{t}^{\alpha -1}\):

$$\begin{array}{@{}rcl@{}} f(t) &=& \frac{d\lambda}{d\alpha} - \lambda\ln(n_{t}) - \frac{\lambda u(c(n_{t}))}{\alpha u(c(n_{t})) + n_{t}u^{\prime}(c(n_{t}))c^{\prime}(n_{t})}\\ &=& \frac{d\lambda}{d\alpha} - \lambda\left[\ln(n_{t}) + \left( \alpha +\frac{n_{t}u^{\prime}(c((n_{t}))c^{\prime}(n_{t}))}{u(c(n_{t}))}\right)^{-1}\right]. \end{array} $$

Defining the derivative with respect to t as f (t), we now have \(f^{\prime }(t) = \frac {\partial f(t)}{\partial n_{t}}\frac {dn_{t}}{dt}\), and we have already shown that \(\frac {dn_{t}}{dt} > 0\) if β γ > 1. Therefore, consider \(\frac {\partial f(t)}{\partial n_{t}}\):

$$\begin{array}{@{}rcl@{}} \frac{\partial f(t)}{\partial n_{t}} \!\!&=&\!\! \!- \lambda\left[\frac{1}{n_{t}} \,-\, \left( \alpha \,+\, \frac{n_{t}u^{\prime}(c(n_{t}))c^{\prime}(n_{t})}{u(c(n_{t}))}\right)^{-2}\right.\\ &&\qquad\times\left.\left( \frac{\left( u^{\prime}(c(n_{t}))c^{\prime}(n_{t}) \,+\, n_{t}u^{\prime\prime}(c(n_{t}))c^{\prime}(n_{t})^{2} \,+\, n_{t}u^{\prime}(c(n_{t}))c^{\prime\prime}(n_{t})\right)u(c(n_{t})) \,-\, n_{t}u^{\prime}(c(n_{t}))^{2}c^{\prime}(n_{t})^{2}}{u(c(n_{t}))^{2}}\right)\right]. \end{array} $$

Every term in the numerator of the second line is negative, which means that \(\frac {\partial f(t)}{\partial n_{t}} < 0\). Therefore, f (t) < 0 if and only if β γ > 1.

Next, notice that \(\frac {dn_{t}}{d\alpha }\) cannot be either always positive or always negative, as long as the resource constraint is binding: if α increases, any increase in population at some point in time must be offset by a decrease at some other point in time. Therefore, f(t) cannot be always positive or always negative; given that f (t) < 0, f(t) must initially be positive and then become negative beyond some critical time t . Finally, we know that g(t) is always negative, and therefore \(\frac {dn_{t}}{d\alpha } = \frac {f(t)}{g(t)}\) must be negative for t < t and positive for t > t , if and only if β γ > 1. This is the result presented in Proposition 2.

1.4 A.4 Recursive dynasty model

Solving the planner’s problem with respect to n t , we find:

$$\beta^{t}\alpha n_{t}^{\alpha-1}u(c_{t}) + \beta^{t}n_{t}^{\alpha-2}u^{\prime}(c_{t})\kappa n_{t+1} - \beta^{t-1}n_{t-1}^{\alpha-1}u^{\prime}(c_{t-1})\kappa = \frac{\lambda}{T_{t}}. $$

Using the first-order condition for n t+1 to cancel out the λ, we then find:

$$\begin{array}{@{}rcl@{}} &&\beta\alpha n_{t}^{\alpha-1}u(c_{t}) + \beta n_{t}^{\alpha-2}u^{\prime}(c_{t})\kappa n_{t+1} - n_{t-1}^{\alpha-1}u^{\prime}(c_{t-1})\kappa\\ &=& \beta\gamma\left( \beta\alpha n_{t+1}^{\alpha-1}u(c_{t+1}) + \beta n_{t+1}^{\alpha-2}u^{\prime}(c_{t+1})\kappa n_{t+2} - n_{t}^{\alpha-1}u^{\prime}(c_{t})\kappa\right). \end{array} $$

and it is easy to rearrange this expression to find that, if a balanced growth path is optimal, g must be equal to \((\beta \gamma )^{\frac {1}{1-\alpha }}\).

Our goal is to use this equation to derive a similar result to that in Sections 3.1 or 3.2. For this purpose, we begin by considering the possibility of a balanced growth path, along which n t+1 = g n t . Given a fixed starting value of n 0 = 1, there is only one possible constant growth rate g that would exactly satisfy the resource constraint:

$$\begin{array}{@{}rcl@{}} \sum\limits_{t=0}^{\infty}{\frac{n_{t}}{T_{t}}} &=& X \rightarrow \sum\limits_{t=0}^{\infty}{\left( \frac{g}{\gamma}\right)^{t}} = XT_{0}\\ g^{*} &=& \left( \frac{XT_{0} - 1}{XT_{0}}\right)\gamma. \end{array} $$
(7)

These two conditions can only be satisfied at one particular value of α, specifically \(\alpha ^{*} = 1 - \frac {\ln (\beta \gamma )}{\ln (g^{*})}\). This is the only α for which a balanced-growth equilibrium exists; at any other value, implementing the feasible growth path g from Eq. 7 would make λ too small or too large to satisfy the first-order conditions.

Given the existence of the child-bearing cost κ, it is possible that having children could be sufficiently costly that desired population growth would be low and that the resource constraint would not be binding. Equation 3 holds regardless, but to ensure that the constraint is binding, we need \(\frac {\partial \mathcal {L}}{\partial n_{t}} > 0\) at λ = 0 on the balanced growth path, or:

$$\beta\alpha u(c) + \beta u^{\prime}(c)\kappa g - g^{1-\alpha}u^{\prime}(c)\kappa > 0. $$

Using the fact that g 1−α = β γ, this can be simplified to:

$$ \kappa < \frac{\alpha u(c)}{(\gamma - g)u^{\prime}(c)} $$
(8)

where c ≡ 1 − κ g is constant per-capita consumption. Greater intuition into condition (8) can be obtained if we assume that utility is linear in consumption, so that Eq. 8 simplifies to:

$$ \kappa < \frac{\alpha}{\gamma - (1-\alpha)g}. $$
(9)

This condition simply requires that the cost of raising children κ not be too large, and further note that it is quite a weak condition. γ is necessarily larger than (1 − α)g due to the assumption that β γ α < 1, and in most cases g will be of the same order of magnitude as γ; in fact, from Eq. 7 above, g approaches γ as X T 0 becomes large relative to one. If gγ, then Eq. 9 simplifies to \(\kappa < \frac {1}{g}\), which is exactly the condition required for consumption 1 − κ g to be positive. Equation 9 could only be significantly restrictive in a situation in which either β or X T 0 are very small, so that γ could be quite large and yet g could be close to 1.

Additionally, allowing for risk-aversion using a CRRA utility function further weakens this condition by raising utility relative to marginal utility. Let \(u(c) = \frac {c^{1-R}}{1-R}\) where R < 1 is the coefficient of relative risk-aversion; we restrict R to be less than one to ensure that utility remains positive. Then Eq. 8 becomes:

$$\kappa < \frac{\alpha c}{(\gamma - g)(1-R)} $$

and it is clear that the right-hand side is increasing in R, so that risk-aversion makes this a weaker condition by raising utility relative to marginal utility.

Now, suppose that α takes the value α defined above, the one that ensures that the balanced growth path described in Eq. 7 is optimal, and further assume that the value of κ is small enough that the resource constraint is binding. Consider the effect of a marginal increase in α. We will examine a case in which only two population values, n t and n t+1 for some t ≥ 1, are allowed to change from their balanced-growth-path values. We find the following condition for the effect of α on n t and n t+1:

Proposition 4

If α is initially at the value that ensures a balanced growth path and κ satisfies(8), and only n t and n t+1 are allowed to vary in response to a marginal change in α , \(\frac {dn_{t}}{d\alpha } < 0\) and \(\frac {dn_{t+1}}{d\alpha } > 0\) if and only if β γ > 1.

Proof 4

Totally differentiate (3) with respect to α, n t and n t+1:

$$\begin{array}{@{}rcl@{}} &&\left[\beta\alpha(\alpha-1)n_{t}^{\alpha-2}u(c_{t}) + 2\beta(\alpha-1)n_{t}^{\alpha-3}u^{\prime}(c_{t})\kappa n_{t+1}\right.\\ &&\quad\left.+ \beta n_{t}^{\alpha-4}u^{\prime\prime}(c_{t})\kappa^{2} n_{t+1}^{2} + n_{t-1}^{\alpha-2}u^{\prime\prime}(c_{t-1})\kappa^{2}\right]dn_{t}\\ &&-\left[\beta\alpha n_{t}^{\alpha-2}u^{\prime}(c_{t})\kappa + \beta n_{t}^{\alpha-3}u^{\prime\prime}(c_{t})\kappa^{2} n_{t+1} - \beta n_{t}^{\alpha-2}u^{\prime}(c_{t})\kappa\right]dn_{t+1}\\ &&+\left[\beta n_{t}^{\alpha-1}u(c_{t})(1 + \alpha\ln(n_{t})) + \beta n_{t}^{\alpha-2}u^{\prime}(c_{t})\kappa n_{t+1}\ln(n_{t})\right.\\ &&\quad\quad\left.- n_{t-1}^{\alpha-1}u^{\prime}(c_{t-1})\kappa\ln(n_{t-1})\right]d\alpha\\ &=& \beta\gamma\left[\beta\alpha(\alpha-1)n_{t+1}^{\alpha-2}u(c_{t+1}) + 2\beta(\alpha-1)n_{t+1}^{\alpha-3}u^{\prime}(c_{t+1})\kappa n_{t+2}\right.\\ &&\quad\quad\left.+ \beta n_{t+1}^{\alpha-4}u^{\prime\prime}(c_{t+1})\kappa^{2}n_{t+2}^{2} + n_{t}^{\alpha-2}u^{\prime\prime}(c_{t})\kappa^{2}\right]dn_{t+1}\\ &&-\beta\gamma\left[(\alpha-1)n_{t}^{\alpha-2}u^{\prime}(c_{t})\kappa + n_{t}^{\alpha-3}u^{\prime\prime}(c_{t})\kappa^{2}n_{t+1}\right]dn_{t}\\ &&+\beta\gamma\left[\beta n_{t+1}^{\alpha-1}u(c_{t+1})(1 + \alpha\ln(n_{t+1})) + \beta n_{t+1}^{\alpha-2}u^{\prime}(c_{t+1})\kappa n_{t+2}\ln(n_{t+1})\right.\\ &&\quad\quad\quad\left.- n_{t}^{\alpha-1}u^{\prime}(c_{t})\kappa\ln(n_{t})\right]d\alpha. \end{array} $$

The resource constraint requires that d n t+1 = −γ d n t ;we also impose a balanced growth path so that n t+1 = g n t , where\(g = \left (\beta \gamma \right )^{\frac {1}{1-\alpha }}\), and c t = c ≡ 1 − κ g. Thenwe rearrange to find:

$$\begin{array}{@{}rcl@{}} &&n_{t}^{\alpha-2}\left[\beta\alpha(\alpha-1)u(c)\left( 1 + \frac{\gamma}{g}\right) + 2\beta(\alpha-1)u^{\prime}(c)\kappa\left( 2\gamma + g\right)\right.\\ &&\quad\quad\quad\left.+ \beta u^{\prime\prime}(c)\kappa^{2}\left( g^{2} + \gamma^{2} + 4\gamma g\right)\right]dn_{t}\\ &=& \beta n_{t}^{\alpha-1}\ln(g)\left[\alpha u(c) + (g-\gamma)\kappa u^{\prime}(c)\right]d\alpha. \end{array} $$

Each of the terms in the square brackets on the left-hand side contains either (α − 1)or u (c), both of whichare negative, so it is immediately apparent that the left-hand side is some negative number multipliedby d n t .Therefore, \(\frac {dn_{t}}{d\alpha } < 0\)if and only if the square bracket on the right-hand side is positive; on the assumption that β γ > 1so that g > 1, weneed:

$$\kappa > \frac{\alpha u(c)}{(g-\gamma)u^{\prime}(c)} $$

which is exactly the condition for the resource constraint to be binding. □

If all the n t are allowed to vary with α, the problem becomes much more complicated; it is intuitive that the same result should typically hold, but since the overall pattern of effects could be quite complex, we are unable to extend our theoretical result to a more general case; similarly, we cannot extend the result analytically beyond the setting of a balanced growth path.Footnote 30

Appendix B: Comparison of α = 0 and α = 1

In this appendix, we present a special case of our model in which we can compare results with α = 0 and α = 1. In each of the models presented so far, at least one of these values of α leads to degenerate results, so we now consider a setting in which utility from consumption is quadratic. Additionally, to simplify the analysis to come, which will rely on continuity of derivatives with respect to time, we focus on a continuous-time setting:

$$u(t) = 1 + \psi_{1}n(t) - \psi_{2}n(t)^{2} $$

where ψ 1,ψ 2 > 0.

In this case, the planner’s problem is:

$$\mathcal{L} = {\int}_{t=0}^{\infty}{e^{-rt}n(t)^{\alpha}(1+\psi_{1}n(t)-\psi_{2}n(t)^{2})} - \lambda\left( {\int}_{t=0}^{\infty}{\frac{n(t)}{T(t)}} - X\right) $$

where r is the instantaneous discount rate, and the first-order condition for n(t) is:

$$e^{-rt}n(t)^{\alpha-1}\left[\alpha (1+\psi_{1}n(t)-\psi_{2}n(t)^{2}) + n(t)(\psi_{1}-2\psi_{2}n(t))\right] = \frac{\lambda}{T(t)}. $$

Simplifying, normalizing T(0) = 1, and writing T(t) = e pt where p is the technological growth rate (to maintain consistent notation), we arrive at:

$$e^{(p-r)t}n(t)^{\alpha-1}\left[\alpha + (1+\alpha)\psi_{1}n(t) - (2+\alpha)\psi_{2}n(t)^{2}\right] = \lambda. $$

We then differentiate with respect to t:

$$\begin{array}{@{}rcl@{}} &&(p\,-\,r)\lambda + e^{(p-r)t}n(t)^{\alpha-2}\left[(\alpha\,-\,1)\alpha \,+\, \alpha(1\,+\,\alpha)\psi_{1}n(t) \,-\, (1\,+\,\alpha)(2\,+\,\alpha)\psi_{2}n(t)^{2}\right]\\ &&\times\frac{dn(t)}{dt} = 0 \end{array} $$

and therefore we have the following results for α = 0 and α = 1:

$$\frac{dn(t)}{dt}|_{\alpha=0} = \frac{\lambda_{0}(p-r)e^{-(p-r)t}}{2\psi_{2}} $$
$$\frac{dn(t)}{dt}|_{\alpha=1} = \frac{\lambda_{1}(p-r)e^{-(p-r)t}}{6\psi_{2}n(t)-2\psi_{1}} $$

where λ α denotes the equilibrium value of λ for a given value of α.

We want to prove that, at any value of t for which n(t) is equal for α = 0 and α = 1, \(\frac {dn(t)}{dt}|_{\alpha =1} > \frac {dn(t)}{dt}|_{\alpha =0}\); if so, then given continuity, it is impossible for n(t)| α=1 to descend from above n(t)| α=0 to below n(t)| α=0 as t increases. If the resource constraint is binding in both cases—which we assume is the case— n(t)| α=1 cannot be always above or always below n(t)| α=0. Therefore, if \(\frac {dn(t)}{dt}|_{\alpha =1} > \frac {dn(t)}{dt}|_{\alpha =0}\) when n(t)| α=1 = n(t)| α=0, we will have proved a single-crossing property, in which n(t)| α=1 starts out below n(t)| α=0 and then rises above n(t)| α=0 after some critical t .

To prove this result, we need:

$$\frac{\lambda_{1}(p-r)e^{-(p-r)t}}{6\psi_{2}n(t)-2\psi_{1}} > \frac{\lambda_{0}(p-r)e^{-(p-r)t}}{2\psi_{2}}. $$

Assuming that p > r—which is analogous to our usual assumption that β γ > 1—and substituting for λ α , we need:

$$\frac{1 + 2\psi_{1}n(t) - 3\psi_{2}n(t)^{2}}{3\psi_{2}n(t)-\psi_{1}} > \frac{\psi_{1} - 2\psi_{2}n(t)}{\psi_{2}}. $$

We further assume that 3ψ 2 n(t) > ψ 1, which is implicitly a condition on the parameters of the utility function: the increase in utility with n(t) cannot be too steep. Then the necessary condition is:

$$\psi_{2} + 2\psi_{1}\psi_{2}n(t) - 3{\psi_{2}^{2}}n(t)^{2} > 3\psi_{1}\psi_{2}n(t) - {\psi_{1}^{2}} - 6{\psi_{2}^{2}}n(t)^{2} + 2\psi_{1}\psi_{2}n(t) $$

which simplifies to:

$${\psi_{1}^{2}} + \psi_{2} > 3\psi_{2}n(t)\left( \psi_{1} - \psi_{2}n(t)\right). $$

It is straightforward to prove that this final condition is always satisfied. The right-hand side is maximized at \(n(t) = \frac {\psi _{1}}{2\psi _{2}}\), at which value the right-hand side is equal to \(\frac {3{\psi _{1}^{2}}}{4}\), which is less than \({\psi _{1}^{2}} + \psi _{2}\). Therefore, as long as the parameters are such that 3ψ 2 n(t) > ψ 1 at the optimum, and as long as p > r, so that the optimal population is growing over time, there will be a critical time t , before which n(t) is higher for α = 0, and after which n(t) is higher for α = 1. When the planner is total-utilitarian rather than average-utilitarian, they will desire a population that starts out lower but rises higher in the long-run.

Appendix C: Extensions of main model

This appendix contains four extensions of the main models in the paper. First, we consider a case where the planner has a choice over the optimal level of emissions; we show that the optimal population equation takes the same form, and that the simulation results are very similar to those from the dynasty model in the main text. Second, we add capital to the dynasty model with variable resource use, and show that increasing substitutability between capital and exhaustible resources in production weakens our result, by moving to an earlier date the point at which the effect of α on population becomes positive, but does not overturn the result in the cases we consider. We then examine optimal population when the resource is renewable; we find that making the resource renew faster weakens our result and eventually overturns it. Finally, we show that if larger populations induce innovations that improve technology in the future, our result can again be overturned.

In each case, the new feature added to the model amounts to a weakening of the exhaustibility of the resource, either directly or indirectly. In this way, this appendix shows the limit of our result, and demonstrates that the essential feature required is the exhaustibility of the resource and the inability of society to find satisfactory substitutes.

1.1 C.1 Variable resource use in dynasty model

In this appendix, the recursive dynasty model from Section 3.3 is extended to include a choice over the amount of resource used up in each period. To be precise, the planner chooses e t , the amount of resource income obtained by each individual at time t, and the output function is linear in population size, so that per-capita income is f(e t ), where f (e) > 0 and f (e) < 0. Thus, the amount of X used up by each generation is \(\frac {n_{t}e_{t}}{T_{t}}\). The planner’s problem is defined by:

$$\mathcal{L} = {\sum}_{t=0}^{\infty}{\beta^{t}n_{t}^{\alpha}u\left( f(e_{t}) - \kappa\frac{n_{t+1}}{n_{t}}\right)} - \lambda\left( {\sum}_{t=1}^{\infty}{\frac{n_{t}e_{t}}{T_{t}}} - X\right). $$

where we continue to assume a constant growth rate γ for T. As noted in footnote 22, our numerical results are very similar if consumption is instead specified as \(f(e_{t})\left (1 - \kappa \frac {n_{t+1}}{n_{t}}\right )\), where fertility costs are consistent with a time cost interpretation; the crossing point t moves slightly earlier, but the qualitative results are the same.

The planner chooses e t and n t for all t ≥ 1, and we assume that both are fixed and normalized to one in period 0; differentiating with respect to e t and n t , and then using the derivatives for e t+1 and n t+1 to cancel out the λ, we find:

$$\begin{array}{@{}rcl@{}} \frac{e_{t+1}}{e_{t}}&&{}\left( \beta\alpha n_{t}^{\alpha-1}u(c_{t}) + \beta n_{t}^{\alpha-2}u^{\prime}(c_{t})\kappa n_{t+1} - n_{t-1}^{\alpha-1}u^{\prime}(c_{t-1})\kappa\right)\\ &&= \beta\gamma\left( \beta\alpha n_{t+1}^{\alpha-1}u(c_{t+1}) + \beta n_{t+1}^{\alpha-2}u^{\prime}(c_{t+1})\kappa n_{t+2} - n_{t}^{\alpha-1}u^{\prime}(c_{t})\kappa\right) \end{array} $$
$$n_{t}^{\alpha-1}u^{\prime}(c_{t})f^{\prime}(e_{t}) = \beta\gamma n_{t+1}^{\alpha-1}u^{\prime}(c_{t+1})f^{\prime}(e_{t+1}). $$

Here, the problem is even more complicated than in Section 3.3; a balanced growth path in which n t+1 = g n t and e t+1 = e t is feasible, and now for any value of α there is a constant e that satisfies the resource constraint, but there is no reason to think that any given combination of α and e would actually be optimal. Meanwhile, the analytical analysis is complicated by the fact that the optimal path of e t is not necessarily fixed.

Thus, we focus on simulations, using linear utility and the same parameter values as those presented in Fig. 2, but with e t chosen as well subject to a production function \(f(e_{t}) = 2e_{t} - {e_{t}^{2}}\). The latter implies that e t = 1, the implied value in Section 3.3, is the maximum feasible value, and thus that the choice of e t will be in [0, 1]. Figure 5 presents the optimal population and growth paths, exactly as in Fig. 2, and the result is also qualitatively identical: a planner with a higher α—a more total-utilitarian planner—would prefer a lower population for approximately the first 100 generations. The population growth rate is once again significantly higher in the first few periods with α = 0.4, as the population level is corrected to a higher path, and then growth settles down to a value slightly lower than that with α = 0.6.

Fig. 5
figure 5

Optimal population paths in dynasty model with variable resource use. Notes: All panels present results for α = {0.4,0.6}, from simulations in which β = 0.975, γ = 1.04, T 0 = 1, X = 100, κ = 0.1811, and f(e t ) = 2e t e t2. Panel (a) presents results from the first 140 periods, while (b) only presents the first 20 periods, as population growth rates remain flat after that point

Figure 6, meanwhile, presents the optimal paths for e t , which settle down very quickly, with a high- α planner preferring a lower resource use per period. This is intuitive, as a planner who places more value on population size relative to individual utility would prefer to restrict consumption in order to make a larger population possible. However, even with this effect, the main result of the paper holds: it is quite possible, even likely, that a planner with a higher welfare weight on population size would prefer a smaller population now and long into the future.

Fig. 6
figure 6

Optimal resource income paths in dynasty model. Notes: Results are presented for the first 140 periods for α = {0.4,0.6}, from simulations in which β = 0.975, γ = 1.04, T 0 = 1, X = 100, κ = 0.1811, and f(e t ) = 2e t e t2

1.2 C.2 Capital and exhaustible resources

We now add physical capital to the dynasty model, and allow the planner to choose both the level of resources used per period, e t , and the level of capital investment. If m t is the level of capital per person in period t and δ is the depreciation rate, investment per person in period t is m t+1 − (1 − δ)m t , which we assume can be made at a cost w(m t+1 − (1 − δ)m t )2. Then, specify output per person as \(f(e_{t},m_{t},n_{t}) = (2z_{t}(e_{t},m_{t}) - z_{t}(e_{t},m_{t})^{2})n_{t}^{\phi -1}\), where \(z_{t}(e_{t},m_{t}) = \left (ae_{t}^{\frac {\sigma -1}{\sigma }} + (1-a)m_{t}^{\frac {\sigma -1}{\sigma }}\right )^{\frac {\sigma }{\sigma -1}}\) is a CES aggregate of resource and capital with elasticity of substitution σ; we assume diminishing marginal returns to population by assuming that ϕ < 1 (this simply helps with ensuring convergence of the solution algorithm). Since \(2z_{t} - {z_{t}^{2}}\) takes a maximum at 1, this prevents the planner from accumulating ever-increasing amounts of economic resources per person. However, planners with different tastes may prefer different divisions of z t into e t and m t , and the more substitutable capital and the exhaustible resource are—that is, the larger is σ—the easier it will be to cut back on resource use and use capital instead, thus weakening the exhaustible resource constraint.

The planner’s problem is summarized in the following lagrangian:

$$\mathcal{L} = {\sum}_{t=0}^{\infty}{\beta^{t}n_{t}^{\alpha}u\left( f(e_{t},m_{t},n_{t}) - \kappa\frac{n_{t+1}}{n_{t}} - w\left( m_{t+1} - (1-\delta)m_{t}\right)^{2}\right)} - \lambda\left( {\sum}_{t=1}^{\infty}{\frac{n_{t}e_{t}}{T_{t}}} - X\right) $$

and we simplify the analysis by assuming linear utility, as in the simulations throughout the paper; then the first-order conditions are as follows:

$$\frac{\partial\mathcal{L}}{\partial n_{t}} = \beta^{t}n_{t}^{\alpha-2}\left( \alpha n_{t}c_{t} - (1-\delta)f(e_{t},m_{t},n_{t})n_{t} + \kappa n_{t+1}\right) - \beta^{t-1}n_{t-1}^{\alpha-1}\kappa - \lambda\frac{e_{t}}{T_{t}} = 0 $$
$$\frac{\partial\mathcal{L}}{\partial e_{t}} = \beta^{t}n_{t}^{\alpha}f_{e_{t}} - \lambda\frac{n_{t}}{T_{t}} = 0 $$
$$\frac{\partial\mathcal{L}}{\partial m_{t}} = \beta^{t}n_{t}^{\alpha}\left( f_{m_{t}} + 2(1-\delta)w(m_{t+1} - (1-\delta)m_{t})\right) - \beta^{t-1}n_{t-1}^{\alpha}2w(m_{t} - (1-\delta)m_{t-1}) = 0 $$

where \(f_{e_{t}}\) and \(f_{m_{t}}\) are partial derivatives of output with respect to e t and m t . If ϕ < 1, the second-order conditions are not guaranteed to be satisfied, but we have confirmed that they are satisfied in each of our numerical analyses to follow.

The first-order conditions are analytically quite complicated, but they can be simulated to illustrate their meaning. We use the standard calibration from before, except that X = 50 (and the simulations only run for 200 periods rather than 1000), and that we assume a = 0.5, w = 1, δ = 0.1, and ϕ = 0.9. Figure 7 presents optimal population paths for α = 0.4 and α = 0.6, as usual, but for four different values of σ, from 0.2 to 0.8.

Fig. 7
figure 7

Optimal population sizes in capital and resource model. Notes: All panels present results for the first 50 periods for α = {0.4,0.6}, from simulations in which β = 0.975, γ = 1.04, T 0 = 1, X = 50, κ = 0.1811, a = 0.5, w = 1, δ = 0.1, and ϕ = 0.9

We can see that, as σ increases, optimal populations generally increase, because easier substitution to capital means that the exhaustible resource constraint is less binding. More importantly for our purposes, however, the time at which the effect of α on optimal population becomes positive moves earlier. Higher- α planners have a greater incentive to switch from resource e t to capital m t , to loosen the exhaustible resource constraint and allow larger populations; we find that e t drops faster and to a lower level when α = 0.6, while m t increases faster and to a higher level with α = 0.6 in all but the σ = 0.2 case.

However, in the cases we study, our basic result continues to hold: the initial optimal population is lower for α = 0.6. The reason is that, in order to build up a large capital stock, the α = 0.6 planner must spend a lot on capital investment initially, which leaves few resources for fertility; as σ gets large, the optimal population actually drops initially to enable the planner to accumulate capital. At levels of σ above 0.8, we have had severe difficulties in finding a solution, as the planner wants to lower n 1 to a very low level—and at that level the cost of reproduction for period 2 is very high, making consumption in period 1 low and thus reducing the desire to have people in period 1. Thus, the problem appears to be unstable when substitution is sufficiently easy, but the reason is that high- α planners want to lower initial populations too much, so it is unlikely that a high- α planner will ever want a larger initial population than a lower- α planner.

Therefore, we conclude that allowing for substitution between capital and exhaustible resources can weaken our result, in that the crossing point at which \(\frac {dn_{t}}{d\alpha }\) becomes positive moves earlier. It is possible that sufficiently strong substitution could, under certain parameter configurations, completely overturn our result, but this does not appear to happen in the case we study.

1.3 C.3 Renewable resource

Next, we consider a case in which the resource is not exhaustible, but rather renews at a slow rate. Denoting the remaining stock of the resource at time t as X t , we assume that:

$$X_{t+1} = X_{t} - \frac{n_{t}}{T_{t}} + \eta(\bar{X}-X_{t}) $$

where T 1 = 1 and T t = γ t−1, and where we now use η to denote the rate of resource renewal, and where \(\bar {X}\) is the starting value of the resource stock. The resource renews itself in each period based on the amount existing at the beginning of the period, and grows at a faster rate when the resource is more depleted; one could easily think of alternative specifications for resource renewal, but this specification gives us a single parameter η that describes how renewable the resource is, which we can vary in the subsequent analysis.

We pursue our analysis in the exogenous-utility model. The planner’s problem is most easily specified as follows:

$$\max_{n_{t},X_{t}} W = {\sum}_{t=1}^{\infty}{\beta^{t-1}n_{t}^{\alpha}u_{t}} \hspace{0.2cm} s.t. \hspace{0.2cm} X_{t+1} = X_{t} - \frac{n_{t}}{T_{t}} + \eta(\bar{X}-X_{t}) \& \hspace{0.2cm} \frac{n_{t}}{T_{t}} \leq X_{t}, \forall t. $$

Thus, we think about the planner’s problem as choosing both n t and X t subject to the resource renewal equation and the feasibility constraint that states that no more resource can be used up in a period than exists at the start of the period. The lagrangian expression is:

$$\mathcal{L} = {\sum}_{t=1}^{\infty}{\left[\beta^{t-1}n_{t}^{\alpha}u_{t} - \lambda_{t}\left( X_{t+1} - X_{t} + \frac{n_{t}}{T_{t}} - \eta(\bar{X}-X_{t})\right) - \mu_{t}\left( \frac{n_{t}}{T_{t}} - X_{t}\right)\right]} $$

and taking the partial derivatives while assuming u t = 1 for all t:

$$\frac{\partial\mathcal{L}}{\partial n_{t}} = \beta^{t-1}\alpha n_{t}^{\alpha-1} - \frac{\lambda_{t}}{T_{t}} - \frac{\mu_{t}}{T_{t}} = 0 $$
$$\frac{\partial\mathcal{L}}{\partial X_{t}} = \lambda_{t}(1-\eta) - \lambda_{t-1} + \mu_{t} = 0. $$

Initially, the feasibility constraint will not be binding, unless the planner wishes to use up all of the available resource in the first period; we can check that the latter condition does not hold, and so we assume that μ t = 0 up to some time t at which the constraint begins to bind. This time t could be at infinity, or it could be a finite time. However, if t occurs in finite time, then once the feasibility constraint begins to bind, it must bind forever, and the resource stock must stay at the steady-state value of \(\theta \equiv \frac {\eta \bar {X}}{1+\eta }\), which is used up and renewed in every subsequent period.

To understand this, consider that allowing X t to increase from 𝜃 and converge to some higher value is clearly dominated by remaining at X t = 𝜃: the allowable resource usage is lower in every period in the former case, and thus population and welfare are also lower. Meanwhile, an oscillatory deviation from X t = 𝜃 is also welfare-reducing under the standard assumption that β γ α < 1. To prove this, suppose that we have reached a resource stock of 𝜃 at time t, and have to decide whether or not to remain at 𝜃 in period t + 1 and beyond. The baseline case is that we stay at X t+1 = X t+2 = 𝜃, and so forth to infinity. Against this, consider the welfare obtained from an increase in X t+1 to 𝜃 + 𝜖, followed by a return to 𝜃 for X t+2 and thereafter. In the former case, the population sizes are n t = 𝜃 γ t−1 and n t+1 = 𝜃 γ t, and the welfare over those two periods is:

$$W_{b} = \beta^{t-1}(\theta\gamma^{t-1})^{\alpha} + \beta^{t}(\theta\gamma^{t})^{\alpha}. $$

Meanwhile, in the oscillatory deviation, the population sizes are n t = (𝜃𝜖)γ t−1 and n t+1 = (𝜃 + (1 − η)𝜖)γ t, and the welfare over those two periods is:

$$W_{d} = \beta^{t-1}\left( (\theta-\epsilon)\gamma^{t-1}\right)^{\alpha} + \beta^{t}((\theta + (1-\eta)\epsilon)\gamma^{t})^{\alpha}. $$

Therefore, the net welfare gain from the oscillation is:

$${\Delta} W \equiv W_{d} - W_{b} = \left( \beta\gamma^{\alpha}\right)^{t-1}\left[(\theta-\epsilon)^{\alpha} - \theta^{\alpha}\right] + \left( \beta\gamma^{\alpha}\right)^{t}\left[(\theta + (1-\eta)\epsilon)^{\alpha} - \theta^{\alpha}\right]. $$

Now, assume that β γ α < 1 as before; then a necessary (but not sufficient) condition for the welfare gain to be positive is:

$$(\theta-\epsilon)^{\alpha} + (\theta + (1-\eta)\epsilon)^{\alpha} > 2\theta^{\alpha}. $$

However, Jensen’s inequality tells us that, since f(x) = x α is an increasing and concave function:

$$(\theta-\epsilon)^{\alpha} + (\theta + (1-\eta)\epsilon)^{\alpha} < (\theta-\epsilon)^{\alpha} + (\theta + \epsilon)^{\alpha} < 2\theta^{\alpha}. $$

Therefore, the oscillation must reduce welfare, and the only possible optimum is one in which X t remains at 𝜃 after t if t is reached in finite time.

Therefore, for the time up to t , we can solve the first-order conditions setting μ t = 0, which leads us to the following condition:

$$\frac{n_{t+1}}{n_{t}} = ((1-\eta)\beta\gamma)^{\frac{1}{1-\alpha}}. $$

If η = 0, then of course we are back in the standard exhaustible-resource world of Section 3, and the population growth rate is \((\beta \gamma )^{\frac {1}{1-\alpha }}\). However, faster resource renewal reduces the population growth rate, at least until the feasibility constraint binds; this may seem counter-intuitive, but the reason is that larger populations in early periods, by depleting the resource, now lead to more resource renewal in future periods, so there is an added benefit of a larger population at the beginning. Thus, the starting population level will increase with η, while the growth rate declines, so that later population values may increase or decrease.

The effect of α on the optimal population levels and growth rate is the same as before up to t , except that (1 − η) is added; if (1 − η)β γ > 1, so that population growth is positive, an increase in α raises the optimal population growth rate, and thus lowers the optimal starting population. However, this condition is less likely to be satisfied when η increases, so that an increase in resource renewability will eventually overturn our result: with sufficiently high η, population growth will be negative (though from a high value) until t is reached, and higher α will make that growth rate more negative, with a larger starting population.

This is illustrated in Fig. 8, which presents optimal population paths for the standard parameter values, with η = {0.001, 0.004, 0.008, 0.015}. In each case, t is reached in finite time, as can be seen in the final three figures. In the first three figures, the population growth rate increases with α, at least up to the point at which the constraint begins to bind, and the initial optimal population is lower when α = 0.6. When η = 0.015, population starts from a high level and then descends slowly, but faster and from a higher level when α = 0.6, exactly as predicted above.

Fig. 8
figure 8

Optimal population sizes with renewable resource. Notes: All panels present optimal population paths for α = {0.4,0.6}, from simulations in which β = 0.975, γ = 1.04, T 1 = 1, \(\bar {X} = 100\), and η = {0.001,0.004,0.008,0.015}. Panels (a) through (c) present the first 100 periods, while (d) presents the first 50 periods

1.4 C.4 Population-induced innovation

Finally, this appendix considers the case in which population at time t determines the technology level at time t + 1 (and perhaps beyond). This captures, in a reduced-form way, the idea that a larger population could provide more opportunities or impetus for innovation, and is one more way in which the exhaustibility of our resource could be weakened.

Making T t dependent on past population values adds considerably to the complexity of our model, so to illustrate the impact we focus on a simple 2-period version of our exogenous-utility model. We assume that T 1 = 1, but that T 2 is an increasing function of n 1: \(T_{2} = \gamma An_{1}^{\delta }\), where δ > 0, and where A is a constant with respect to α that we will vary with δ (as explained later) to prevent changes in δ from directly affecting the relative technological efficiencies.

We can then set up the maximization problem:

$$\mathcal{L} = n_{1}^{\alpha}u_{1} + \beta n_{2}^{\alpha}u_{2} - \lambda\left( n_{1} + \frac{n_{2}}{\gamma An_{1}^{\delta}} - X\right) $$

and if u 1 = u 2 = 1, the first-order conditions are:

$$\alpha n_{1}^{\alpha-1} = \lambda\left( 1 - \frac{\delta n_{2}}{\gamma An_{1}^{\delta+1}}\right) $$
$$\beta\alpha n_{2}^{\alpha-1} = \frac{\lambda}{\gamma An_{1}^{\delta}}. $$

Substituting for λ, we have:

$$n_{1}^{\alpha-1} = \beta\gamma An_{1}^{\delta}n_{2}^{\alpha-1}\left( 1 - \frac{\delta n_{2}}{\gamma An_{1}^{\delta+1}}\right) $$

along with the resource constraint, \(n_{1} + \frac {n_{2}}{\gamma An_{1}^{\delta }} = X\).

Even in a two-period setting, the resulting algebra is rather complicated, so as before we use numerical simulations to explore the effect of δ on our results. In Fig. 9, we present simulations for β = 0.975, γ = 1.04, and X = 1; for any value of α, we also set \(A = ({n_{1}^{0}})^{-\delta }\), where \({n_{1}^{0}}\) is the optimal n 1 when δ = 0 for that value of α. In this way, changing the value of δ will affect the technological parameter only by shifting n 1; there will be no direct effect whereby raising δ directly raises (if n 1 > 1) or lowers (if n 1 < 1) the value of \(\gamma n_{1}^{\delta }\).

Fig. 9
figure 9

Optimal population sizes and growth rate with induced innovation. Notes: Panel (a) presents the optimal population levels n 1 and n 2, while panel (b) presents the optimal growth rate \(\frac {n_{2}}{n_{1}}\), for α = {0.4,0.6}, from simulations in which β = 0.975, γ = 1.04, X = 1, and δ = [0,0.05]

At low values of δ, of course, our standard result holds: population growth is positive, and a higher α raises population growth further while lowering n 1. However, as δ increases, the same result obtains as in the previous appendix: population growth eventually falls below zero, and our result is flipped: a higher α means a larger initial population and a lower growth rate. As with renewable resources, induced innovation raises the value of large populations at early times, as they help to make the resource constraint less binding in the future. Therefore, a more total-utilitarian planner who wants to increase population when it is already large may choose to expand the population more in earlier periods.

Appendix D: An alternative specification for average utilitarianism

In our main analysis, we treat average utilitarianism as a generational average, applying our α parameter to the size of each generation. While we describe in Section 3 that this is the standard specification in the recent literature on the economics of population, it is not the universal standard: Nerlove et al. (1982, 1985, 1986b) all find that total utilitarians prefer larger populations in a setting in which average utilitarianism evaluates the average utility across all individuals over time. Furthermore, some philosophical objections have been raised against the generational average specification as a way of modelling average utilitarianism; for example, (Dasgupta 1998) argues that such a specification is ad hoc and lacks philosophical foundations.Footnote 31

Therefore, in this appendix, we demonstrate that similar results hold when we model average utilitarianism as an average across all time, rather than across a generation; we thank Marc Fleurbaey for suggesting this analysis. We will specify our general social welfare function as:

$$W = \frac{{\sum}_{t=0}^{\infty}{\beta^{t}n_{t}u_{t}}}{\left( {\sum}_{t=0}^{\infty}{\beta^{t}n_{t}}\right)^{1-\alpha}} $$

where α is once again the weight placed on population; if α = 0, the function expresses pure average utilitarianism, while if α = 1, the term on the bottom is equal to one and the social welfare function corresponds to total utilitarianism.Footnote 32 We will evaluate the effect of marginal changes to α on our results, but in each case the same conclusion would hold if considering the extreme cases of average (α = 0) and total (α = 1) utilitarianism.

To keep the algebra simple, we will only present here the two-period exogenous utility case; exactly analogous results follow in the infinite-horizon and dynasty models when considering changes in n t and n t+1, and are available upon request. The planner’s problem is:

$$\mathcal{L} = \frac{n_{1}u_{1} + \beta n_{2}u_{2}}{\left( n_{1} + \beta n_{2}\right)^{1-\alpha}} - \lambda\left( \frac{n_{1}}{T_{1}} + \frac{n_{2}}{T_{2}} - X\right). $$

However, we need to make one minor modification to the model to prevent knife’s-edge solutions. If u t is simply a constant (or if \(u_{t} = 1 - \kappa \frac {n_{t+1}}{n_{t}}\) in the dynasty model), then an interior solution can only exist if β γ = 1, and then any interior solution is valid. Instead, we assume that output features diminishing marginal returns with respect to population size, so that total output is equal to \(n_{t}^{\delta }\) for δ < 1, and we assume linear utility over per-capita consumption so that \(u_{t} = n_{t}^{\delta -1}\). Then the planner’s problem becomes:

$$\mathcal{L} = \frac{n_{1}^{\delta} + \beta n_{2}^{\delta}}{\left( n_{1} + \beta n_{2}\right)^{1-\alpha}} - \lambda\left( \frac{n_{1}}{T_{1}} + \frac{n_{2}}{T_{2}} - X\right). $$

Differentiating with respect to n 1 and n 2, we find:

$$\frac{\partial\mathcal{L}}{\partial n_{1}} = \frac{\delta n_{1}^{\delta-1}\left( n_{1} + \beta n_{2}\right) - (1-\alpha)\left( n_{1}^{\delta} + \beta n_{2}^{\delta}\right)}{\left( n_{1} + \beta n_{2}\right)^{2-\alpha}} - \frac{\lambda}{T_{1}} $$
$$\frac{\partial\mathcal{L}}{\partial n_{2}} = \frac{\beta\delta n_{2}^{\delta-1}\left( n_{1} + \beta n_{2}\right) - \beta(1-\alpha)\left( n_{1}^{\delta} + \beta n_{2}^{\delta}\right)}{\left( n_{1} + \beta n_{2}\right)^{2-\alpha}} - \frac{\lambda}{T_{2}}. $$

At the optimum, of course, both of these derivatives are equal to zero. Concavity cannot be guaranteed for all parameter values, but it holds in simulations with values similar to those used in Section 3; therefore, we assume concavity of the value function with respect to n 1 and n 2, and that thus an interior optimum exists.

However, the expressions are sufficiently complicated that we are not going to try to solve for n 1 and n 2; as it happens, we do not need to in order to sign the effect of α on n 1 and n 2. Suppose that we start from some \(\hat {\alpha }\); then we could solve the equations above for the optimal \(n_{1}\left (\hat {\alpha }\right )\) and \(n_{2}\left (\hat {\alpha }\right )\), and the associated \(\lambda \left (\hat {\alpha }\right )\). Now assume that α increases marginally to \(\tilde {\alpha } = \hat {\alpha } + \varepsilon \), where ε is vanishingly small. Further assume that we hold n 1 and n 2 fixed for now, and that λ adjusts with α to hold \(\frac {\partial \mathcal {L}}{\partial n_{2}} = 0\), so that λ takes the value \(\lambda _{2}(\tilde {\alpha })\) that would make the planner willing to hold n 2 unchanged. This, however, will not generally be the value of λ that will make \(\frac {\partial \mathcal {L}}{\partial n_{1}} = 0\), so we can evaluate whether the planner would like to increase or decrease n 1. If \(\frac {\partial \mathcal {L}}{\partial n_{2}} = 0\), then we know that:

$$\lambda_{2}(\tilde{\alpha}) = T_{2}\frac{\beta\delta n_{2}^{\delta-1}\left( n_{1} + \beta n_{2}\right) - \beta(1-\tilde{\alpha})\left( n_{1}^{\delta} + \beta n_{2}^{\delta}\right)}{\left( n_{1} + \beta n_{2}\right)^{2-\tilde{\alpha}}} $$

and assuming that T 2 = γ and T 1 = 1, this means:

$$\begin{array}{@{}rcl@{}} \frac{\partial\mathcal{L}}{\partial n_{1}} &=& \frac{\delta n_{1}^{\delta-1}\left( n_{1} + \beta n_{2}\right) - (1-\tilde{\alpha})\left( n_{1}^{\delta} + \beta n_{2}^{\delta}\right)}{\left( n_{1} + \beta n_{2}\right)^{2-\tilde{\alpha}}}\\ &&- \gamma\frac{\beta\delta n_{2}^{\delta-1}\left( n_{1} + \beta n_{2}\right) - \beta(1-\tilde{\alpha})\left( n_{1}^{\delta} + \beta n_{2}^{\delta}\right)}{\left( n_{1} + \beta n_{2}\right)^{2-\tilde{\alpha}}}\\ &=& \frac{\delta\left( n_{1} + \beta n_{2}\right)\left( n_{1}^{\delta-1} - \beta\gamma n_{2}^{\delta-1}\right) + (1-\tilde{\alpha})\left( n_{1}^{\delta} + \beta n_{2}^{\delta}\right)\left( \beta\gamma-1\right)}{\left( n_{1} + \beta n_{2}\right)^{2-\tilde{\alpha}}}. \end{array} $$

As stated, at \(\hat {\alpha }\), we know that \(\frac {\partial \mathcal {L}}{\partial n_{1}} = 0\) by definition, and thus the numerator must be equal to zero at \(\hat {\alpha }\). Increasing α affects the bottom of the fraction, but without changing the sign, whereas it also reduces the top as long as β γ > 1. Therefore, if β γ > 1, \(\frac {\partial \mathcal {L}}{\partial n_{1}}(\tilde {\alpha }) < 0\). If the equilibrium λ with \(\tilde {\alpha }\) was such that both \(\frac {\partial \mathcal {L}}{\partial n_{1}}\) and \(\frac {\partial \mathcal {L}}{\partial n_{2}}\) took the same sign at \(n_{1}(\hat {\alpha })\) and \(n_{2}(\hat {\alpha })\), there would be no allocation consistent with constrained optimization, as concavity would imply that both n 1 and n 2 should be moved in the same direction, which is not feasible given the resource constraint. Therefore, the optimal λ must be such that \(\frac {\partial \mathcal {L}}{\partial n_{1}}\) and \(\frac {\partial \mathcal {L}}{\partial n_{2}}\) take opposite signs at \(n_{1}(\hat {\alpha })\) and \(n_{2}(\hat {\alpha })\) given \(\tilde {\alpha }\), and given that both are decreasing in λ, this is only possible at a \(\lambda ^{*} < \lambda _{2}(\tilde {\alpha })\) at which \(\frac {\partial \mathcal {L}}{\partial n_{1}} < 0\) and \(\frac {\partial \mathcal {L}}{\partial n_{2}} > 0\). Then, by concavity, it must be that n 1 decreases with α and n 2 increases.

Thus, we again find the standard result from the paper: if β γ > 1, then a planner with a larger value of α—a more total-utilitarian planner—will want to reduce population today to allow for a larger population tomorrow. This proves that our result holds in this alternative specification for average utilitarianism, in the two-period exogenous-utility case; as stated earlier, exactly analogous results follow from the infinite-horizon and dynasty models when we consider changes in n t and n t+1, holding all other n fixed. In this specification, an increase in α means that increases in total discounted population are weighted more strongly, as they do not count as heavily in the denominator; therefore, the planner would like to increase population in the most efficient way, and if β γ > 1, technology is improving faster than discounting and it is best to increase the future population at the expense of today.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lawson, N., Spears, D. Optimal population and exhaustible resource constraints. J Popul Econ 31, 295–335 (2018). https://doi.org/10.1007/s00148-017-0665-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00148-017-0665-9

Keywords

JEL Classification

Navigation