Skip to main content
Log in

A bitter choice turned sweet: How acknowledging individuals’ concern at having a low relative income serves to align utilitarianism and egalitarianism

  • Regular Article
  • Published:
Journal of Evolutionary Economics Aims and scope Submit manuscript

An Erratum to this article was published on 09 June 2015

Abstract

When individuals’ utility is a convex combination of their income and their concern at having a low relative income (the weights attached to income and to the concern at having a low relative income sum up to one), the maximization of aggregate utility yields an equal income distribution. This alignment of utilitarianism and egalitarianism is obtained for any number of individuals, and for general utility functions that are convex combinations of a power function of income and the concern at having a low relative income. The alignment can also hold when the weights sum up to a number different than one.

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.

Similar content being viewed by others

Notes

  1. The inference that utilitarianism aligns with egalitarianism draws upon, but is not contingent on, the weights in the individuals’ utility functions necessarily summing up to 1. Under such an assumption, the utility functions have the characteristic that an individual’s weak taste for absolute income correlates with a strong distaste for low relative income (and vice versa). However, incorporation of the individuals’ distaste for low relative income can sustain the alignment of utilitarianism with egalitarianism even when the weights in the individuals’ utility functions do not sum up to 1, though additional conditions then need to be imposed. In Appendix B we provide the analysis of such case of preferences’ specification.

  2. In an Appendix, Stark (2013) attends to the origins of the concept of relative deprivation, elaborates on the measure, and provides more evidence that upward interpersonal comparisons affect significantly subjective wellbeing.

  3. To illustrate our notation, consider n = 3 and \(x\,=\,(x_{1}^{~} ,x_{2}^{~} ,x_{3}^{~} )\,=\,(2,1,1)\). For such an x, we have that \(x_{2}^{~} \,=\,x_{3}^{~} \!<\!x_{1}^{~} \). Then, x belongs to X Q , where Q is a permutation of the set {1,2,3} such that Q(1) = 2, Q(2) = 3, and Q(3) = 1; and it also belongs to X R , where R is a permutation of the set {1,2,3} such that R(1) = 3, R(2) = 2, and R(3) = 1.

  4. Note that given γ ∈ (0, 1), we have that \(\frac {\partial ^{2}F}{\partial {x_{l}^{2}}} \,=\,-\alpha _{l} \gamma (\gamma -1)x_{l}^{\gamma -2} >0\) for l ∈ {1,...,n} and that \(\frac {\partial ^{2}F}{\partial x_{l} \partial x_{k}} \!=0\) for kl, which implies that the Hessian matrix of F is positive definite.

  5. We have that \(\lim \limits _{\gamma \to 1} \gamma ^{\frac {1}{1-\gamma } }\,=\,\lim \limits _{\gamma \to 1} e^{\frac {\ln \gamma } {1-\gamma } }\,=\,1/e\) because from l’Hospital’s rule we obtain that \(\lim \limits _{\gamma \to 1} \frac {\ln \gamma } {1-\gamma } \,=\,-1\).

  6. For \(f(\gamma )\,=\,\gamma ^{\frac {1}{1-\gamma } }\) we have that \({f}^{\prime }(\gamma )\,=\,\frac {\gamma ^{\frac {\gamma } {1-\gamma } }}{(1-\gamma )^{2}}\left ({\ln \gamma \,-\,\frac {\gamma -1}{\gamma } } \right )\ge 0\) due to the fact that \(\ln x\ge \frac {x-1}{x}\) for any x > 0.

References

  • Andolfatto D (2002) A Theory of inalienable property rights. J of Political Econ 110:382–393

    Article  Google Scholar 

  • Aronsson T, Johansson-Stenman O (2008) When the Joneses’ consumption hurts: Optimal public good provision and nonlinear income taxation. J Publ Econ 92:986–997

    Article  Google Scholar 

  • Bentham J (1823) An introduction to the principles of morals and legislation. Clarendon Press, Oxford, 1907

    Google Scholar 

  • Blanchflower DG, Oswald AJ (2008) Hypertension and happiness across nations. J Health Econ 27:218–233

    Article  Google Scholar 

  • Blomquist SN (1993) Interdependent behavior and the effect of taxes. J Publ Econ 51:211–218

    Article  Google Scholar 

  • Boskin MJ, Sheshinski E (1978) Optimal redistributive taxation when individual welfare depends upon relative income. Quart J Econ 92:589–601

    Article  Google Scholar 

  • Callan MJ, Ellard JH, Shead N, Hodgins DC (2008) Gambling as a search for justice: Examining the role of personal relative deprivation in gambling urges and gambling behavior. Pers Soc Psychol Bull 34:1514–1529

    Article  Google Scholar 

  • Clark AE, Frijters P, Shields MA (2008) Relative income, happiness, and utility: An explanation for the Easterlin Paradox and other puzzles. J Econ Lit 46:95–144

    Article  Google Scholar 

  • Clark AE, Senik C (2010) Who compares to whom? The anatomy of income comparisons in Europe. Econ J 120:573–594

    Article  Google Scholar 

  • Dworkin R (2000) Sovereign virtue: The theory and practice of equality. Harvard University Press, Cambridge, MA

    Google Scholar 

  • Easterlin RA (1974) Does economic growth improve human lot? Some empirical evidence. In: David PA, Reder MW (eds) Nations and households in economic growth: Essays in honor of Moses Abramowitz. Academic Press, New York and London, pp 89–125

    Google Scholar 

  • Fan CS, Stark O (2011) A theory of migration as a response to occupational stigma. Int Econ Rev 52:549–571

    Article  Google Scholar 

  • Fliessbach K, Weber B, Trautner P, Dohmen T, Sunde U, Elger CE, Falk A (2007) Social comparison affects reward-related brain activity in the human ventral striatum. Science 318:1305–1308

    Article  Google Scholar 

  • Frey BS, Stutzer A (2002) Happiness and economics: How the economy and institutions affect human well-being. Princeton University Press, Princeton

    Google Scholar 

  • Haisley E, Mostafa R, Loewenstein G (2008) Subjective relative income and lottery ticket purchases. J Behav Decis Making 21:283–295

    Article  Google Scholar 

  • Hare RM (1981) Moral thinking: Its levels, method, and point. Clarendon Press, Oxford

    Book  Google Scholar 

  • Harsanyi JC (1977) Non-linear social welfare functions: A rejoinder to Professor Sen. In: Butts RE, Hintikka J (eds) Foundational problems in the special sciences. Reidel, Dordrecht, pp 294–295

    Google Scholar 

  • Luttmer EFP (2005) Neighbors as negatives: Relative earnings and well-being. Quart J Econ 120:963–1002

    Google Scholar 

  • Mill JS (1863) Utilitarianism. University of Chicago Press, Chicago, 1906

    Google Scholar 

  • Neumark D, Postlewaite A (1998) Relative income concerns and the rise in married women’s employment. J Publ Econ 70:157–183

    Article  Google Scholar 

  • Oswald AJ (1983) Altruism, jealousy and the theory of optimal non-linear taxation. J Publ Econ 20:77– 87

    Article  Google Scholar 

  • Rawls J (1999) A theory of justice. Harvard University Press, Cambridge, MA

    Google Scholar 

  • Roemer JE (1998) Equality of opportunity. Harvard University Press, Cambridge, MA

    Google Scholar 

  • Scanlon TM (1998) What do we owe to each other? Harvard University Press, Cambridge, MA

    Google Scholar 

  • Sen AK (1973) On economic inequality. Clarendon Press, Oxford

    Book  Google Scholar 

  • Sen AK (1980) Equality of what? In: McMurrin SM (ed) The Tanner lectures on human values, vol 1. Cambridge University Press, Cambridge, pp 197–220

    Google Scholar 

  • Sen AK (1982) Choice, welfare and measurement. Harvard University Press, Cambridge, MA

    Google Scholar 

  • Sorger G, Stark O (2013) Income distribution going awry: The reversal power of the concern for relative deprivation. J Econ Behav Organ 86:1–9

    Article  Google Scholar 

  • Stark O (1993) Nonmarket transfers and altruism. Europ Econ Rev 37:1413–1424

    Article  Google Scholar 

  • Stark O (2013) Stressful integration. Europ Econ Rev 63:1–9

    Article  Google Scholar 

  • Stark O, Hyll W (2011) On the economic architecture of the workplace: Repercussions of social comparisons among heterogeneous workers. J Lab Econ 29:349–375

    Article  Google Scholar 

  • Stark O, Taylor JE (1991) Migration incentives, migration types: The role of relative deprivation. Econ J 101:1163–1178

    Article  Google Scholar 

  • Takahashi H, Kato M, Matsuura M, Mobbs D, Suhara T, Okubo Y (2009) When your gain is my pain and your pain is my gain: Neural correlates of envy and schadenfreude. Science 323:937–939

    Article  Google Scholar 

  • Tullock G (1975) On mathematics as decoration. In: Chisholm RK (ed) Papers in economic criticism. Memphis University Press, Memphis, pp 22–23

    Google Scholar 

  • Wendner R, Goulder LH (2008) Status effects, public goods provision, and excess burden. J Publ Econ 92:1968–1985

    Article  Google Scholar 

  • Zizzo DJ, Oswald AJ (2001) Are people willing to pay to reduce others’ incomes? Ann Econ Stat 63-64:39–65

    Google Scholar 

Download references

Acknowledgments

We are indebted to an Associate Editor for thoughtful advice and insightful comments, to a referee, and to Luigi Orsenigo for guidance and encouragement.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Oded Stark.

Appendices

Appendix A: Proof of Proposition 1

With n individuals, there are n! possible orderings of incomes (permutations of the set {1,2,...,n}). Assuming a specific ordering \({x_{1}} \!\le \! {x_{2}} \!\le \! ... \!\le \!x_{n} \), we proceed to show that a corresponding result holds independently of the values of \(\alpha _{1}^{~} ,...,\alpha _{n} \). By symmetry of the maximized function with respect to incomes and income weights, the result obtained below holds for other orderings of incomes.

Without loss of generality, and normalizing the sum of incomes if necessary, we assume that this sum is \(\displaystyle \sum \limits _{i=1}^{n} {x_{i}} \,=\,1\). The set of admissible distributions \(D\!\equiv \! \!\left \{ {x\in \mathbf {R}_{\ge 0}^{n} \text {: } x_{1}^{~} \le x_{2}^{~} \le ... \le x_{n} ,\displaystyle \sum \limits _{i=1}^{n} {x_{i}} =1} \right \}\) is convex (in fact, it is a convex polytope) and has a finite number of extremal points of the form

$$x^{k}=(\underbrace{0,0,...,0}_{n-k},\underbrace{1/k,1/k,...,1/k}_{k}),\,\, k\in \{ {1,2,...,n} \}, $$

where we have that x n = (1/n, 1/n,...,1/n) = x .

Over the closed set D, the function S W F is linear and therefore its maximum on D is attained in one of the extremal points x k. We have that

$$S{\kern-.5pt}W{}\!F(x^{k})=\underbrace{\frac{1}{k}\sum\limits_{i=n-k+1}^{n} {\alpha_{i}}}_{i{\kern-.5pt}n{\kern-.5pt}c{\kern-.5pt}o{\kern-.5pt}m{\kern-.5pt}e}\,\!-{\kern-.7pt}\!\underbrace{\frac{1}{n}\sum\limits_{i=1}^{n-k} {(1-\alpha_{i} )}}_{r{\kern-.5pt}e{\kern-.5pt}l{\kern-.5pt}a{\kern-.5pt}t{\kern-.5pt}i{\kern-.5pt}v{\kern-.5pt}e\,\,{\kern.5pt}d{\kern-.5pt}e{\kern-.9pt}p{\kern-.5pt}r{\kern-.5pt}i{\kern-.5pt}v{\kern-.5pt}a{\kern-.5pt}t{\kern-.5pt}i{\kern-.5pt}o{\kern-.5pt}n}, $$

namely, compared to x n, x k for kn represents taking away income from k individuals and distributing that income equally between the remainder nk individuals.

A distribution x n constitutes a unique maximum if and only if S W F(x n) >S W F(x k) for any kn. To see this, note that

$$\frac{n-k}{n}=\left( {\frac{1}{k}-\frac{1}{n}} \right)k>\left( {\frac{1}{k}-\frac{1}{n}} \right)\sum\limits_{i=n-k+1}^{n} {\alpha_{i}} , $$
$$\frac{1}{n}\sum\limits_{i=n-k+1}^{n} {\alpha_{i}} >\frac{1}{k}\sum\limits_{i=n-k+1}^{n} {\alpha_{i}} -\frac{n-k}{n}, $$
$$\frac{1}{n}\sum\limits_{i=n-k+1}^{n} {\alpha_{i}} +\frac{1}{n}\sum\limits_{i=1}^{n-k} {\alpha_{i}} >\frac{1}{k}\sum\limits_{i=n-k+1}^{n} {\alpha_{i}} -\frac{n-k}{n}+\frac{1}{n}\sum\limits_{i=1}^{n-k} {\alpha_{i}} , $$
$$\frac{1}{n}\sum\limits_{i=1}^{n} {\alpha_{i}} >\frac{1}{k}\sum\limits_{i=n-k+1}^{n} {\alpha_{i}} +\frac{1}{n}\left[ {\sum\limits_{i=1}^{n-k} {\alpha_{i}} -(n-k)} \right]. $$

We have

$$\sum\limits_{i=1}^{n-k} {\alpha_{i}} -(n-k)=\sum\limits_{i=1}^{n-k} {(\alpha_{i} -\text{1)}} =-\sum\limits_{i=1}^{n-k} {(1-\alpha_{i} \text{)}} . $$

Consequently, we get

$$ S{\kern-.5pt}W{}\!F(x^{n})=\frac{1}{n}\sum\limits_{i=1}^{n} {\alpha_{i}} >\frac{1}{k}\sum\limits_{i=n-k+1}^{n} {\alpha_{i}} -\frac{1}{n}\sum\limits_{i=1}^{n-k} {(1-\alpha_{i} )} =S{\kern-.5pt}W{}\!F(x^{k}) $$
(A1)

for any kn, which completes the proof that x n = x is the maximum. □

Appendix B: A loosened assumption on the weights in the individuals’ preferences

In this Appendix we attend to a model of preferences in which there are two (positive) parameters in the individual’s utility function rather than one: γ i > 0 weight to be placed on individual’s i own income, and β i > 0 weight to be placed on low relative income, that is, the individual’s utility function is

$$ u_{i} (x)=\gamma_{i} x_{i} -\beta_{i} R\!D_{i} (x). $$
(B1)

When this extra degree of freedom is allowed, we obtain the result that, as we increase the β weights, loosely speaking, the optimal utilitarian solution moves toward egalitarianism or, more correctly, social welfare decreases as we move away from egalitarianism.

Taking first as an example two individuals, and using the same notation as in the beginning of Section 2, for the case of x i < x j , when income is transferred from individual j to individual i, without changing the hierarchy of the two income earners, the marginal increase in i’s utility is γ i + β i , whereas the marginal decrease in j’s utility is γ j . Unlike in the case of two parameters \(\alpha _{1}^{~} ,\alpha _{2}^{~} \) in which the marginal gain of individual i, which is equal to 1, is greater than the marginal loss of individual j, which is equal to α j ∈ (0, 1), here the marginal gain of i will be greater than the marginal loss of j if γ i + β i > γ j , which translates into the condition that the marginal disutility of low relative income of the poorer individual, β i , has to be larger than the difference between the two individuals’ marginal utilities of own income, γ j γ i . From analyzing in a similar manner the case of x i > x j , we get an analogous condition on the marginal increase in social welfare, namely β j > γ i γ j .

Summing up: when the weights in the utility function of an individual do not sum up to 1, then for equality of incomes to be the optimal utilitarian solution, that is, for \(x_{1}^{\ast } \,=\,x_{2}^{\ast } \,=\,\frac {A}{2}\) to obtain, we need to have that \(\beta _{1}^{~} \!\!>\!\!\gamma _{2}^{~} \!-{\kern -.5pt}\gamma _{1}^{~} \) and that \(\beta _{2}^{~} \!>\!\gamma _{1}^{~}\! -\gamma _{2}^{~} \).

This extension to a setting of the weights not summing up to one can be carried further and be applied to the case of more than two individuals. Let the utility of the i-th, i ∈ {1,...,n}, individual in a population with income vector x = (x 1,...,x n ) be described by (B1). In the following proposition we state and prove a condition on the weights γ i , β i in the population which, if satisfied, yields that the egalitarian income distribution maximizes utilitarian social welfare.

Proposition B1

If

$$ \gamma_{j}^{~} +\beta_{j}^{~} >\gamma_{l}^{\phantom{\frac{1}{2}}} \!\!+\frac{n-2}{n}\beta_{l} $$
(B2)

for all j ≠l, then the egalitarian income distribution maximizes social welfare.

Proof 3

The problem of the utilitarian social planner under a budget constraint A > 0 is

$$\begin{array}{@{}rcl@{}} &&\max SW\!F(x)\\ &&\text{for} x\in \mathbf{R}_{\ge 0}^{n} \text{s.t.} \sum\limits_{i=1}^{n} {x_{i}} = A, \end{array} $$
(B3)

where \(S{\kern -.5pt}W{}\!F(x)\!\,=\,\!\sum \limits _{i=1}^{n} {u_{i} (x)} \). With n individuals, there are n! possible orderings of incomes (permutations of the set {1,2,...,n}). In what follows, we divide the maximization problem (4) into n! sub-problems corresponding each to a given ordering of incomes, and we show that in each sub-problem the egalitarian income distribution x = (A/n,A/n,...,A/n) is optimal.

Specifically, we divide the set \(X\!\!\!\,=\,\!\!\!\left \{ {x\!\in \!\mathbf {R}_{\ge 0}^{n} :\displaystyle \sum \limits _{i=1}^{n} {x_{i}} {\kern -.5pt} ={\kern -.5pt} A} \right \}\) into sets X P ,\( X \!\,=\,\underset {P}{\bigcup } X_{P}\), where the summation is over all possible orderings of incomes. For xX P , the income of individual P(i) is the i-th lowest in the population, namely \(x_{P(1)}^{~} \le x_{P(2)}^{~} \!\le \! ...\!\le \! x_{P(n)}^{~} \); and \(X_{P} \,=\,\{x\in X:x_{P(k)}^{~} <x_{P(l)}^{~} \Rightarrow k<l\}\).Footnote 3 Obviously, x X P for every P. For the ordering P, the maximization problem is

$$ \max\limits_{x\in X_{P}} S{\kern-.5pt}W{}\!F(x). $$
(B4)

For xX P the social welfare function can be rewritten as

$$S{\kern-.5pt}W{}\!F(x)= \sum\limits_{i=1}^{n} {a_{P(i)}^{~} x_{P(i)}^{~}} , $$

where

$$ a_{P(i)}^{~} =\gamma_{P(i)}^{~} +\frac{(n-i)\beta_{P(i)}} {n}-\frac{1}{n}\sum\limits_{j=1}^{i-1} {\beta_{P(j)}} . $$
(B5)

Consider the income distribution xX P which is not egalitarian, namely xx , and a set of individuals whose income is the lowest in the population under distribution x. Obviously, individual P(1) belongs to this set. We introduce \(i_{x} \,=\,\max \{i\!:\! x_{P(1)}^{~} \,=\,x_{\!P(i)}^{~} \}\). The individuals P(1),...,P(i x ) have the same income, the lowest in the population. Given that xx , we have that P(i x ) < n, and that \(x_{P(i_{x} )}^{~} \!\!<\!\!x_{P(i_{x} +1)}^{~} \). We construct an income distribution x X P such that x x in the following way: for ii x + 1, \({x}^{\prime }_{\!P(i)} \,=\,\frac {1}{i_{x} +1}\displaystyle \sum \limits _{j=1}^{i_{x} +1} {x_{P(j)}^{\!}} \,=\,\frac {i_{x} x_{P(i_{x})}^{~} +x_{P(i_{x} +1)}^{~}}{i_{x} +1}\), and for i > i x + 1, \({x}^{\prime }_{\!P(i)} \,=\,x_{P(i)}^{~} \). We thus obtain that

$$\begin{array}{@{}rcl@{}} &&{} S{\kern-.5pt}W{}\!F({x}^{\prime})-S{\kern-.5pt}W{}\!F(x) \\ &&{} =\!\sum\limits_{j=1}^{i_{x}} {a_{\!P(j)}^{~}\! \left( {\frac{i_{x} x_{\!P(i_{x} )}^{~} \!{\kern.5pt}+x_{P(i_{x} +1)}^{~}} {i_{x} \,+\,1}\,-\,x_{\!P(i_{x} )}^{~}} \!\right)}\!+a_{\!P(i_{x} +1)}^{~}\! \left( \frac{i_{x} x_{\!P(i_{x} )}^{~} \!{\kern.5pt}+x_{\!P(i_{x} +1)}^{~}} {i_{x} \,+\,1} \,-\,x_{\!P(i_{x} +1)}^{~}\! \right) \\ &&{} =\frac{x_{P(i_{x} +1)}^{~} -x_{P(i_{x} )}^{~}} {i_{x} +1}\left( {\sum\limits_{j=1}^{i_{x}} {a_{P(j)}^{~}} -i_{x} a_{P(i_{x} +1)}^{~} } \right) . \end{array} $$
(B6)

Next, from (3) applied to j ∈ {1,...,i} and to l =i+1, we obtain by summation that

$$\sum\limits_{j=1}^{i} {(\gamma_{P(j)}^{~} +\beta_{P(j)}^{~} )} >i\gamma_{P(i+1)}^{~} +\frac{i(n-2)\beta_{P(i+1)}} {n}, $$

which implies

$$ \sum\limits_{j=1}^{i} {(\gamma_{P(j)}^{~} +\beta_{P(j)} )} >i\gamma_{P(i+1)}^{~} +\frac{i(n-i-1)\beta_{P(i+1)}} {n}. $$
(B7)

Recalling (B5), because

$$\sum\limits_{j=1}^{i} {a_{P(j)}^{~}} =\sum\limits_{j=1}^{i} {\gamma_{P(j)}^{~}} +\frac{n-i}{n}\sum\limits_{j=1}^{i} {\beta_{P(j)}} , $$

and because

$$ia_{P(i+1)}^{~} =i\gamma_{P(i+1)}^{~} +\frac{i(n-i-1)\beta_{P(i+1)}} {n}-\frac{i}{n}\sum\limits_{j=1}^{i} {\beta_{P(j)}} , $$

we get from (B7) that

$$ \sum\limits_{j=1}^{i} {a_{P(j)}^{~}} >ia_{P(i+1)}^{~} $$
(B8)

for all i∈ {1,...,n−1}.

Thus, by joining (B6) and (B8), we have that S W F (x )−S W F (x) > 0. Consequently, distribution x such that xx cannot be a solution to (B4). Because the set X P is compact, S W F(x) attains its maximum over this set. Therefore, distribution x is the solution to (B4) for any ordering P, and x is the unique maximum of the social welfare function over the set X. □

Appendix C: Proof of Proposition 2

Akin to the proof of Proposition 1, and again without loss of generality, we assume a specific ordering of the incomes, \(x_{1}^{~} \!\le \! x_{2}^{~} \!\le ... \!\le x_{n} \), and show that the result holds independently of the values of α 1,...,α n . The symmetry of the maximized function with respect to incomes and income weights implies that the result obtained below holds for other orderings of incomes. For the assumed ordering, the maximization problem of the utilitarian social planner is

$$\begin{array}{@{}rcl@{}} &&\max\limits_{x\in {\Omega}} \left\{ {\sum\limits_{i=1}^{n} {\left[ {\alpha_{i} x_{i}^{\gamma} -\frac{1-\alpha_{i}} {n}\sum\limits_{j=i+1}^{n} {\left( {x_{j} -x_{i}} \right)}} \right]}} \right\} \\ &&\text{s.t. } {\Omega} \text{=}\left\{ {\sum\limits_{i=1}^{n} {x_{i}} =A;\, x_{1}^{~} \ge 0;\, x_{i} \ge x_{i-1}^{~} \text{ for } i=2,...,n} \right\} . \end{array} $$

We denote \(F(x)\,=\,-\!\displaystyle \sum \limits _{i=1}^{n} {\left [ {\alpha _{i} x_{i}^{\gamma } \!-\frac {1-\alpha _{i}} {n}\displaystyle \sum \limits _{j=i+1}^{n} {\left ({x_{j} \,-\,x_{i}} \right )}} \right ]} \), \(h(x)\,=\,\!\displaystyle \sum \limits _{i=1}^{n} {x_{i} \,-\,A} \), and \(g_{i} (x)=x_{i-1}^{~} -x_{i} \) for all i∈ {2,...,n}, \(g_{1}^{~} (x)=-x_{1}^{~} \). Transforming the maximization problem to a minimization problem,

$$\begin{array}{@{}rcl@{}} &&\min F(x) \\ &&\text{s.t. } h(x)=0,\, g_{i} (x)\le 0\,\text{ for } \,i\!\!\in \!\{1,...,n\}, \end{array} $$

we obtain the Lagrangian

$$\begin{array}{@{}rcl@{}} L&=&F(x)+\sum\limits_{i=1}^{n} {\mu_{i} g_{i} (x)+\lambda h(x)} \\ &=&-\sum\limits_{i=1}^{n} {\left[ {\alpha_{i} x_{i}^{\gamma} -\frac{1-\alpha_{i}} {n}\sum\limits_{j=i+1}^{n} {\left( {x_{j} -x_{i}} \right)}} \right]} +\sum\limits_{i=1}^{n} {\mu_{i} g_{i} (x)+\lambda h(x)} . \\ \end{array} $$

We have that

$$\frac{\partial L}{\partial x_{1}^{~}} =-\alpha_{1}^{~} \gamma x_{1}^{\gamma -1} -(n-1)\frac{1-\alpha_{1}} {n}-\mu_{1}^{~} +\mu_{2}^{~} +\lambda , $$

that

$$\frac{\partial L}{\partial x_{n}} =-\alpha_{n} \gamma x_{n}^{\gamma -1} +\sum\limits_{j=1}^{n-1} {\frac{1-\alpha_{j}} {n}} -\mu_{n} +\lambda , $$

and that

$$\frac{\partial L}{\partial x_{l}^{~}} =-\alpha_{l}^{~} \gamma x_{l}^{\gamma -1} +\sum\limits_{j=1}^{l-1} {\frac{1-\alpha_{j}} {n}} -(n-l)\frac{1-\alpha_{l}^{~}} {n}-\mu_{l}^{~} +\mu_{l+1}^{~} +\lambda $$

for l ∈ {2,...,n−1}.

Because the maximized function F is convex,Footnote 4 and the feasible set D is described by affine constraints, the following set of the Karush-Kuhn-Tucker conditions is both sufficient and necessary for x to be a unique minimum:

$$\begin{array}{@{}rcl@{}} &&\frac{\partial L}{\partial x_{l}^{~}} (x)=0,\, l\in \left\{ {1,...,n} \right\} \!, \\ &&\mu_{l}^{~} g_{l}^{~} (x)=0,\, \mu_{l}^{~} \ge 0,\text{} g_{l}^{~} (x)\le 0,\, l\in \{1,...,n\} , \\ &&h(x)=0 . \end{array} $$

We now show that point x =(A/n,A/n,...,A/n) satisfies the preceding conditions and that, therefore, it constitutes the global minimum of this problem.

We note that because \(x_{1}^{\ast } =A/n>0\), we set \(\mu _{1}^{~} =0\), and hence the considered problem reduces to a set of linear equations in \(\mu _{2}^{~} ,...,\mu _{n} ,\lambda \). We denote

$$C_{l} =\left\{ \begin{array}{ll} -\alpha_{1}^{~} \gamma A^{\gamma -1}n^{1-\gamma} -(n-1)\frac{1-\alpha_{1}^{~}}{n} & \qquad l=1, \\ -\alpha_{n} \gamma A^{\gamma -1}n^{1-\gamma} +\displaystyle\sum\limits_{j=1}^{n-1} \frac{1-\alpha_{j}}{n} & \qquad l=n,\\ -\alpha_{l}^{~} \gamma A^{\gamma -1}n^{1-\gamma} +\displaystyle\sum\limits_{j=1}^{l-1} \frac{1-\alpha_{j}} {n}-(n-l)\frac{1-\alpha_{l}^{~}}{n} & \qquad \text{otherwise.} \end{array} \right. $$

Thus, we obtain

$$\left\{\begin{array}{ll} C_{1} =-\mu_{2}^{~} -\lambda &\qquad l=1, \\ C_{n} =\mu_{n} -\lambda &\qquad l=n, \\ C_{l} =\mu_{l}^{~} -\mu_{l+1}^{~} -\lambda &\qquad \text{otherwise,}\\ \end{array}\right. $$

which can be written as

$$C=\mathrm{M}v, $$

where \(C=(C_{l} )_{l=1}^{n}, v=\left ({\lambda ,\mu _{2}^{~},...,\mu _{n}} \right )\), and

$$\mathrm{M}=\left[ \begin{array}{ccccccc} {-1} & {-1} & 0 & {...} & {...} & {...} & 0 \\ {-1} & 1 & {-1} & 0 & {...} & {...} & 0 \\ {-1} & 0 & 1 & {-1} & 0 & {...} & 0 \\ {...} & {...} & {...} & {...} & {...} & {...} & {...} \\ {-1} & 0 & {...} & {...} & 0 & 1 & {-1} \\ {-1} & 0 & {...} & {...} & {...} & 0 & 1 \\ \end{array}\right]. $$

It can easily be verified that

$$\mathrm{M}^{-1}=\left[ \begin{array}{ccccccc} {-\frac{1}{n}} & {-\frac{1}{n}} & {-\frac{1}{n}} & {-\frac{1}{n}} & {...} & {-\frac{1}{n}} & {-\frac{1}{n}} \\ [12pt] {-\frac{n-1}{n}} & {\frac{1}{n}} & {\frac{1}{n}} & {\frac{1}{n}} & {...} & {\frac{1}{n}} & {\frac{1}{n}} \\ [12pt] {-\frac{n-2}{n}} & {-\frac{n-2}{n}} & {\frac{2}{n}} & {\frac{2}{n}} & {...} & {\frac{2}{n}} & {\frac{2}{n}}\\ {...} & {...} & {...} & {...} & {...} & {...} & {...} \\ {-\frac{1}{n}} & {-\frac{1}{n}} & {-\frac{1}{n}} & {-\frac{1}{n}} & {...} & {-\frac{1}{n}} & {\frac{n-1}{n}} \\ \end{array}\right]. $$

Then,

$$\lambda =-\frac{1}{n}\sum\limits_{l=1}^{n} {C_{l}} , $$

and for i∈{1,2,...,n−1}

$$\mu_{i+1} =-\frac{n-i}{n}\sum\limits_{l=1}^{i} {C_{l}} +\frac{i}{n}\sum\limits_{l=i+1}^{n} {C_{l}} \equiv f_{i+1} (\alpha_{1} ,...,\alpha_{n} ). $$

Obviously the functions f i + 1 are linear and, moreover, for i∈{1,2,...,n − 1}, we have that f i+1(1,...,1) = 0. We calculate for j = l

$$\frac{\partial C_{l}} {\partial \alpha_{l}^{~}} =-\gamma A^{\gamma -1}n^{1-\gamma} +\frac{n-l}{n}, $$

for j <l

$$\frac{\partial C_{l}} {\partial \alpha_{j}} =-\frac{1}{n}, $$

and for j > l

$$\frac{\partial C_{l}} {\partial \alpha_{j}} =0. $$

Therefore, for j > i we obtain

$$\frac{\partial f_{i+1}} {\partial \alpha_{j}} =\frac{i}{n}\sum\limits_{l=j}^{n} {\frac{\partial C_{l}} {\partial \alpha _{j}} } =\frac{i}{n}\left( {\frac{\partial C_{j}} {\partial \alpha_{j}} +\sum\limits_{l=j+1}^{n} {\frac{\partial C_{l}} {\partial \alpha_{j}} }} \right)=-i\gamma A^{\gamma -1}n^{-\gamma} <0, $$

whereas for ji, we have

$$\begin{array}{@{}rcl@{}} &&\frac{\partial f_{i+1}} {\partial \alpha_{j}} =-\frac{n-i}{n}\left( {\frac{\partial C_{j}} {\partial \alpha_{j}} +\sum\limits_{l=j+1}^{i} {\frac{\partial C_{l}} {\partial \alpha_{j}} }} \right)+\frac{i}{n}\sum\limits_{l=i+1}^{n} {\frac{\partial C_{l}} {\partial \alpha_{j}} } \\ &&=-\frac{n-i}{n}\left( {-\gamma A^{\gamma -1}n^{1-\gamma} +\frac{n-j}{n}-\frac{i-j}{n}} \right)-\frac{i(n-i)}{n^{2}} \\ &&=\frac{n-i}{n}\left( {\gamma A^{\gamma -1}n^{1-\gamma} -1} \right). \end{array} $$

From the assumption that An/e we get that \(A\ge n\gamma ^{\frac {1}{1-\gamma } }\), where this last inequality is due to the fact that \(\lim \limits _{\gamma \to 1} \gamma ^{\frac {1}{1-\gamma } }\,=\,1/e\),Footnote 5 and that \(\gamma ^{\frac {1}{1-\gamma } }\) as a function of γ is non- decreasing.Footnote 6 Thus, we have that \(\left ({\gamma A^{\gamma -1}n^{1-\gamma } \,-\,1} \right )\!\le \left [ {\gamma \! \left ({n\gamma ^{\frac {1}{1-\gamma } }} \right )^{\gamma -1}n^{1-\gamma } \,-\,1} \right ]\) =0. Therefore, \(\frac {\partial f_{i+1}} {\partial \alpha _{j}} \!\le \! 0\) for any i ∈ {1,2,...,n−1} and j = {1,...,n}. Thus, because f i+1(1,...,1) = 0, for any set of weights \((\alpha _{1}^{~} ,...,\alpha _{n} )\!\!\in \!\! (0, 1)^{n}\) we obtain that f i+1 \( (\alpha _{1}^{~} ,...,\alpha _{n} )\ge 0\) for i ∈ {1,2,...,n−1}.

In sum, we obtained a set of non-negative multipliers \(\mu _{2}^{~} ,...,\mu _{n} \) for inequality constraints \(g_{2}^{~} ,...,g_{n} \) which are satisfied by x with equality. This completes the proof that x is the global maximum of the considered problem. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Stark, O., Jakubek, M. & Kobus, M. A bitter choice turned sweet: How acknowledging individuals’ concern at having a low relative income serves to align utilitarianism and egalitarianism. J Evol Econ 25, 541–557 (2015). https://doi.org/10.1007/s00191-015-0396-6

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00191-015-0396-6

Keywords

JEL Classification

Navigation