Skip to main content

Advertisement

Log in

Stochastic dominance and thick-tailed wealth distributions

  • Published:
Journal of Economics Aims and scope Submit manuscript

Abstract

Right-skewed and thick-tailed wealth distributions have been documented as an empirical regularity across space and time. A key mechanism for explaining these distributional features is proportional random growth. We investigate the comparative statics of a well-defined class of random growth models when allowing for stochastically ordered shifts in the wealth return process. An order-contingent monotone comparative statics property is identified, according to which pure increases in risk (e.g. higher volatility of capital returns) foster top wealth concentration whereas first-order stochastically dominated shifts in the return process (induced by e.g. proportional capital income taxation) rather lower inequality at the upper end of the distribution. Our analysis points to the potentially ambiguous effects on top wealth inequality of introducing or modifying capital income tax treatments in the presence of stochastic returns.

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. A distribution is said to be (i) right-skewed when it displays an asymmetrically long upper tail, and (ii) thick-tailed if the upper tail decays as a power law. A formal definition is provided in the “Appendix”.

  2. Stochastic returns—on either wealth or labor, or both—are key to determining the emergence of thick-tailed stationary distributions, see e.g. D’Amato and Di Pietro (2014).

  3. Comparative statics predictions such as monotonicity properties are of first-order importance in economics (e.g. Milgrom and Shannon 1994; Athey 2002).

  4. This literature is too broad to be fully referenced here. Davidson and Duclos (2000) provide a thorough survey of the topic.

  5. Following the standard notation in probability theory, we denote random variables with upper case letters, whereas particular realizations of such variables are written in corresponding lower case letters.

  6. As shown in Brandt (1986), for any initial condition \(W_0\) there exists a unique stationary limit law of (1), which is that of \(W=A_0+\sum _{t=1}^{\infty }B_{-t}\prod _{j=0}^{t-1}A_{-j}\), provided \(E[\log |A_0|]<0\) and \(E[\log ^{+}|B_0|]<\infty \), where \(x^{+}:=\max \left\{ x,0 \right\} \) for \(x \in \mathbb {R}\).

  7. While not needed for our analysis, slightly stricter regularity conditions are required in Kesten (1973) and Goldie (1991). A more formal presentation of the basic results from Kesten (1973) is reported in the “Appendix”. Extensions of this set of results to generalized Kesten processes with Markovian coefficients have been obtained in Saporta (2005) and Benhabib et al. (2011).

  8. The state variable \(w_t\) can be thought of as cash-on-hand, i.e. the resources available to the agent for consumption and saving, after the realization of current income.

  9. Notice that the characterization of mean-preserving second-order stochastic dominance change in distribution does not mandate a monotone function \(k(\cdot )\), as the proof of the equivalence stated in part (ii) only relies on the second-order derivative of the latter.

  10. That is, households care about post return bequests and therefore take into account the uncertainty inherent in the random process for wealth returns when determining their optimal savings.

  11. This occurs, for instance, when capital income taxation is so strong as to force the support of the after-tax rates of returns to lie in the unit interval. In this case, for any finite \(\mu \) the condition \(E[A^\mu ]=1\) does not hold, and by convention \(\mu =+\infty \).

  12. Details on the computation are reported in the “Appendix”.

References

  • Acemoglu D, Robinson JA (2015) The rise and decline of general laws of capitalism. J Econ Perspect 29:3–28

    Article  Google Scholar 

  • Alvaredo F, Atkinson AB, Morelli S (2016) The challenge of measuring UK wealth inequality in the 2000s. Fisc Stud 37:13–33

    Article  Google Scholar 

  • Athey S (2002) Monotone comparative statics under uncertainty. Q J Econ 117:187–223

    Article  Google Scholar 

  • Atkinson A (1970) The measurement of inequality. J Econ Theory 2:244–263

    Article  Google Scholar 

  • Atkinson A, Bourguignon F (1989) The design of direct taxation and family benefits. J Publ Econ 41:3–29

    Article  Google Scholar 

  • Becker GS, Tomes N (1979) An equilibrium theory of the distribution of income and intergenerational mobility. J Polit Econ 87:1153–1189

    Article  Google Scholar 

  • Benhabib J, Bisin A, Zhu S (2011) The distribution of wealth and fiscal policy in economies with finitely-lived agents. Econometrica 79:123–157

    Article  Google Scholar 

  • Benhabib J, Bisin A, Zhu S (2015) The wealth distribution in Bewley economies with capital income risk. J Econ Theory 159:489–515

    Article  Google Scholar 

  • Benhabib J, Bisin A (2016) Skewed wealth distributions: theory and empirics. NBER Working Paper No. 21924

  • Bossmann M, Kleiber C, Wälde K (2007) Bequests, taxation and the distribution of wealth in a general equilibrium model. J Publ Econ 91:1247–1271

    Article  Google Scholar 

  • Brandt A (1986) The stochastic equation \(Y_{n+1} = A_{n}Y_{n} + B_{n}\) with stationary coefficients. Adv Appl Probab 18:211–220

    Article  Google Scholar 

  • Cagetti M, De Nardi M (2009) Estate taxation, entrepreneurship, and wealth. Am Econ Rev 99(1):85–111

    Article  Google Scholar 

  • Castaneda A, Diaz-Gimenez J, Rios-Rull JV (2003) Accounting for the US earnings and wealth inequality. J Polit Econ 111:818–857

    Article  Google Scholar 

  • Champernowne DG (1953) A model of income distribution. Econ J 63:318–351

    Article  Google Scholar 

  • D’Amato M, Di Pietro C (2014) Occupational mobility and wealth evolution in a model of educational investment with credit market imperfections. J Econ Inequal 12:73–98

    Article  Google Scholar 

  • Davidson R, Duclos J-Y (2000) Statistical inference for stochastic dominance and for the measurement of poverty and inequality. Econometrica 68:1435–1464

    Article  Google Scholar 

  • Frechet M (1939) Sur les Formules de Repartition des Revenus. Revue del Institut International de Statistique 7:32–38

    Article  Google Scholar 

  • Gabaix X (2009) Power laws in economics and finance. Ann Rev Econ 1:255–293

    Article  Google Scholar 

  • Goldie CM (1991) Implicit renewal theory and tails of solutions of random equations. Ann Appl Probab 1:126–166

    Article  Google Scholar 

  • Kesten H (1973) Random difference equations and renewal theory for products of random matrices. Acta Math 131:207–248

    Article  Google Scholar 

  • Klass OS, Biham O, Levy M, Malcai O, Solomon S (2007) The Forbes 400, the Pareto power-law and efficient markets. Eur Phys J B Condens Matter Complex Syst 55:143–147

    Article  Google Scholar 

  • Meyer J (2014) The theory of risk and risk aversion. In: Machina M, Viscusi W (eds) Handbook of the economics of risk and uncertainty. North Holland Press, London

    Google Scholar 

  • Milgrom P, Shannon C (1994) Monotone comparative statics. Econometrica 62:157–180

    Article  Google Scholar 

  • Piketty T (2003) Income inequality in France, 1901–1998. J Political Econ 111:1004–1042

    Article  Google Scholar 

  • Piketty T (2014) Capital in the twenty-first century. Harvard University Press, Cambridge

    Book  Google Scholar 

  • Piketty T, Saez E (2003) Income inequality in the United States, 1913–1998. Q J Econ 118:1–39

    Article  Google Scholar 

  • Saporta B (2005) Tail of the stationary solution of the stochastic equation \(Y_{n+1} = a_{n}Y_{n} + b_{n}\) With Markovian coefficients. Stoch Process Appl 115:1954–1978

    Article  Google Scholar 

  • Wolff E (1987) Estimates of household wealth inequality in the US, 1962–1983. Rev Income Wealth 33:231–256

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to Giacomo Corneo (Editor) and one anonymous reviewer for many valuable comments and suggestions, which have helped to substantially improve the paper. We also wish to thank Marcello D’Amato, Tullio Jappelli, Thomas Kneib, Salvatore Morelli, Marco Pagano, Holger Strulik and seminar conference participants in Göttingen, Halle-Saale (17th IWH-CIREQ-GW Macroeconometric Workshop) and Naples (CSEF) for insightful discussions. The usual disclaimer applies.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marco M. Sorge.

Appendix

Appendix

Thick-tailed distributions

Let L(x) be a real-valued function. Then L(x) is said to vary regularly with index \(\mu \in (0, \infty )\) if

$$\begin{aligned} \lim _{x \rightarrow \infty } \frac{L(tx)}{L(x)}=t^{-\mu },\quad \forall t>0 \end{aligned}$$

A probability distribution with a differentiable CDF F(x) is a power-law with tail index \(\mu \) if \(L(x):=\left( 1-F(x) \right) \) is regularly varying with index \(\mu >0\). The distribution F(x) is then said to be thick-tailed with tail index \(\mu \).

Regularity conditions for Kesten processes (Kesten 1973).

Consider the stochastic Eq. (1) where \((A_t, B_t) \in \mathbb {R}^2_+\) are i.i.d. random variables. For some \(\mu >1\), let the the following hold:

$$\begin{aligned} E[A^\mu ]=1;\quad E\left[ (A)^\mu \max \left\{ ln(A), 0 \right\} \right] <\infty ;\quad E[B]\in (0, \infty ) \end{aligned}$$

Also, assume \(\left( B/(1-A) \right) \) is non-degenerate and the distribution of ln(A) is non-lattice (i.e. its support is not contained in \(\lambda \mathbb {Z}\) for some \(\lambda \)). Then there exist constants \((c_{-}, c_+) \)—at least one positive—such that for \(x \rightarrow \infty \):

$$\begin{aligned} \begin{aligned} x^{\mu } Pr(W>x) \quad \rightarrow \quad c_+ \\ x^{\mu } Pr(W<-x) \quad \rightarrow \quad c_{-} \end{aligned} \end{aligned}$$

where W is the stationary solution to (1), i.e. \(W\overset{d}{=} A W+B\). The solution to (1) then converges in probability to W as \(t \rightarrow \infty \).

Proof of Proposition 1

Recall that we restrict attention to sequences \((A_t)\) and \((A_t')\) complying with Assumption 1, and therefore associated with tail indexes \(\mu _A >1\) and \(\mu _{A'}>1\), respectively. Let a denote the generic realization of the random variable A. We start noticing that the function \(g: a \mapsto a^{\mu }\) is strictly increasing and strictly convex for all \(a \in \mathbb {R}_+\).

  1. i)

    Consider \(F_{A} \succcurlyeq _{1} F_{A'}\) with \(F_{A} \ne F_{A'}\). Then from Theorem 1, part (i) we have

    $$\begin{aligned} 1=E[(A')^{\mu _{A'}}]=\int _{\mathbb S} (a')^{\mu _{A'}} d F_{A'} < \int _{\mathbb S} (a)^{\mu _{A'}} d F_{A} \end{aligned}$$

    Suppose that \(\mu _{A'} \le \mu _{A}\). Then by Holder’s inequality we have

    $$\begin{aligned} \int _{\mathbb S}(a)^{\mu _{A'}}dF_{A} \le \left[ \int _{\mathbb S}\left( (a)^{\mu _{A'}} \right) ^{\frac{\mu _A}{\mu _{A'}}}dF_{A} \right] ^{\frac{\mu _{A'}}{\mu _A}}=\left[ \int _{\mathbb S} (a)^{\mu _A}dF_{A} \right] ^{\frac{\mu _{A'}}{\mu _A}}=1 \end{aligned}$$

    which is a contradiction. Hence it must be \(\mu _{A'}> \mu _A\).

  2. ii)

    Consider \(F_{A} \succcurlyeq _{2} F_{A'}\) with \(\int _{\mathbb S} a dF_{a} = \int _{\mathbb S} a' dF_{A'}\) \(F_{A} \ne F_{A'}\) From Theorem 1, part (ii) we have

    $$\begin{aligned} 1=E[(A)^{\mu _A}]=\int _{\mathbb S} (a)^{\mu _A} d F_{a} < \int _{\mathbb S} (a')^{\mu _A} d F_{A'}=E[(A')^{\mu _A}] \end{aligned}$$

    and the second assertion follows by applying again Holder’s inequality.

\(\square \)

Proof of Corollary 1

Since the composite function \(g: z \mapsto (\phi (z))^{\mu }\)—with \(\mu >1\)—is strictly increasing if and only if \(d \phi /d z >0\) for all z, and strictly convex if and only if

$$\begin{aligned} (\mu -1) \phi (z)^{\mu -2} \left( d \phi / d z \right) ^2+\phi (z)^{\mu -1}(d^2 \phi / d z^2) >0\quad \forall z \end{aligned}$$

which is equivalent to requiring

$$\begin{aligned} \frac{d^2 \phi }{d z^2}>-(\mu -1)\frac{\left( d \phi / d z \right) ^2}{\phi (z)}=\Lambda (z)\quad \forall z \end{aligned}$$
(11)

both points follow readily from Proposition 1. \(\square \)

Proof of Proposition 2

In the following, we consider the transformation \(R'=(1-\tau (R))R\) of the random variable \(R \sim F_R\) (a copy of \(R_t\)), with \(F_{R'}=F_{(1-\tau (R))R}\).

  1. i)

    The assertion follows immediately if the mapping \((1-\tau (r_t))r_t\) is monotone in \(r_t\) (and hence invertible). More generally, let \(g: \mathbb {R}_+ \mapsto \mathbb {R}_+\) be a function such that for any \(x \in \mathbb {R}_{+}\) it holds \(g(x) \le x\). Let also M denote the set of all random variables \(X: \Omega \mapsto \mathbb {R}_+\) defined on a properly filtered probability space \((\Omega , \Sigma , P)\). Then there always exists a function \(g^*: M \mapsto M\) such that for any \(\omega \in \Omega \) the function \(g^*(X): \Omega \rightarrow \mathbb {R}_+\) is given by \(g^*(X)(\omega )=g(X(\omega ))\).

    Consider now a random variable X with density \(f_X\), and define \(\Gamma _{y}:=\left\{ x= X(\omega ): g(x)\le y \right\} \) for any real y. By construction \(y\le x\) for all x belonging to the border of \(\Gamma _y\), i.e for all \(x\in \partial \Gamma _y:=\left\{ x: g(x)= y \right\} \). Let \(\underline{x}=\min \partial \Gamma _y \), then it holds

    $$\begin{aligned} Pr(g^*(X)<y) = \int _{\Gamma _{y}}f(x) dx \ge \int _{0}^{\underline{x}} f(x) dx \ge \int _{0}^{y} f(x) dx = Pr(X< y) \end{aligned}$$
    (12)

    Since the after-tax rate of return \(r'_t=(1-\tau (r_t))r_t\) satisfies the restriction \(r'_t \le r_t\), it plays the role of the \(g^*(\cdot )\) function in the above. Using the first and the last terms in (12), we can therefore write \(Pr(R_t<y) \le Pr(R'_t <y)\) for any real y, from which it follows \(F_R \succcurlyeq _{1} F_{R'}\).

  2. ii)

    Let \(\tau (r_t)=\tau \in (0,1)\), such that \(r'_t=(1-\tau )r_t\). By the previous point \(F_R\) first-order stochastically dominates \(F_{R'}\). However, since the imposition of a capital income tax distorts households’ saving decisions, a behavioral response arises—i.e. a change in the share \(\psi (\theta , \delta , F_R)\)—whose strength depends on the underlying probability distribution shift. Hence, Corollary 1 cannot be applied. In order to resort to Proposition 1, we need to show that \(F_R \succcurlyeq _{1} F_{R'}\) implies \(F_{A(R)} \succcurlyeq _{1} F_{A(R')}\). Under CRRA preferences, the effective random return on wealth (i.e. the multiplicative coefficient of the wealth accumulation equation) is

    $$\begin{aligned} A(X)=\frac{X}{1+\left( \theta E[X^{1-\delta }] \right) ^{-\frac{1}{\delta }}},\quad X \in \left\{ R, R' \right\} \end{aligned}$$
    (13)

\(\square \)

Let \(R'=(1-\tau )R\). Suppose, by contradiction, that there exists some \(r \in \mathbb {S} \subset \mathbb {R}_+\) such that

$$\begin{aligned} Pr(A(R)<r)>Pr(A(R')<r) \end{aligned}$$

which is equivalent to requiring

$$\begin{aligned} r \cdot \left[ 1+\left( \theta E[R^{1-\delta }] \right) ^{-\frac{1}{\delta }} \right] > \frac{r}{1-\tau } \cdot \left[ 1+\left( \theta E[((1-\tau )R)^{1-\delta }] \right) ^{-\frac{1}{\delta }} \right] \end{aligned}$$
(14)

or

$$\begin{aligned} 1+\left( \theta E[R^{1-\delta }] \right) ^{-\frac{1}{\delta }} > \frac{1}{1-\tau } +\frac{1}{(1-\tau )^{1/\delta }}\left( \theta E[R^{1-\delta }] \right) ^{-\frac{1}{\delta }} \end{aligned}$$
(15)

Since \(\tau \in (0,1)\), by simple algebra a contradiction is obtained. Hence, for all \(r \in \mathbb {S}\) it must be that \(Pr(A(R)<r)<Pr(A(R')<r)\), or \(F_{A(R)} \succcurlyeq _{1} F_{A(R')}\).

Proof of Corollary 2

We first show that, for any tax schedule \(\tau (\cdot )\), it holds \(F_{A(R)} \succcurlyeq _{1} F_{A(R')}\) if \(\delta \le 1\) (sufficiency). This is trivially true in the knife-edge case \(\delta =1\), as the imposition of a capital income tax does not change the fraction of current wealth devoted to savings, Corollary 1 part i) therefore applies.

Consider now the case \(\delta <1\). Since \(g: z \mapsto (z)^{1-\delta }\) is an increasing function, \(R \succcurlyeq _{1} R'\) implies \(E[R^{1-\delta }]>E[R'^{1-\delta }]\). Recalling the saving decision rule (9), this in turn implies \(\psi (\theta , \delta , F_R)>\psi (\theta , \delta , F_{R'})\). According to the invariance property of the stochastic dominance ordering, if \(R \succcurlyeq _{1} R'\) then \(c\cdot R \succcurlyeq _{1} c\cdot R' \) for any constant \(c>0\). The assertion then readily follows from noticing that \(R' \succcurlyeq _{1} \frac{\psi (\theta , \delta , F_{R'})}{\psi (\theta , \delta , F_R)} R'\).

To prove the necessity part, we show that, when \(\delta >1,\) there exist tax schedules \(\tau (\cdot )\) for which \(F_R \succcurlyeq _{1} F_{R'}\) does not imply \(F_{A(R)} \succcurlyeq _{1} F_{A(R')}\). To this end, for a fixed (realized) rate of return \(\underline{r} \in \mathbb {S}\), let us consider a progressive capital income tax schedule which takes a simple two-bracket form:

$$\begin{aligned} \tau (r)= \left\{ \begin{array}{rl} \tau _{min}, &{}\text{ if } \, r < \underline{r} \\ \\ \tau _{max}, &{}\text{ if } \, r \ge \underline{r}\\ \end{array} \right. \end{aligned}$$
(16)

where \(0\le \tau _{min}< \tau _{max}<1\). For all random variables \(R \in \mathbb {S}\) the following stochastic dominance ordering holds:

$$\begin{aligned} (1-\tau _{min}) R \succcurlyeq _1(1-\tau (R)) R \succcurlyeq _1 (1-\tau _{max}) R \end{aligned}$$
(17)

Let us now consider the random variable defined by (13). For \(r < \underline{r}\) we have

$$\begin{aligned} Pr(A((1-\tau (r)) R)< r) = Pr\left( R < r \left( \frac{1+\left( \theta E[((1-\tau (r)) R)^{1-\delta }] \right) ^{-\frac{1}{\delta }}}{1-\tau _{min}}\right) \right) \end{aligned}$$
(18)

and

$$\begin{aligned} Pr(A((1-\tau _{min}) R)< r) = Pr\left( R < r \left( \frac{1+\left( \theta E[((1-\tau _{min} R)^{1-\delta }] \right) ^{-\frac{1}{\delta }}}{1-\tau _{min}}\right) \right) \end{aligned}$$
(19)

For \(\delta > 1\), we get \(E[((1-\tau (r)) R)^{1-\delta }] > E[((1-\tau _{min}) R)^{1-\delta }]\), from which it follows

$$\begin{aligned} \left( \frac{1+\left( \theta E[((1-\tau (r)) R)^{1-\delta }] \right) ^{-\frac{1}{\delta }}}{1-\tau _{min}}\right) < \left( \frac{1+\left( \theta E[((1-\tau _{min} R)^{1-\delta }] \right) ^{-\frac{1}{\delta }}}{1-\tau _{min}}\right) \end{aligned}$$
(20)

Hence, it must be that \(Pr(A((1-\tau _{min}) R)< r) > Pr(A((1-\tau (r)) R) < r)\), and the first-order stochastic dominance ordering between the random variables \((1-\tau _{min}) R\) and \((1-\tau (R)) R\) is not inherited by the random variables \(A((1-\tau _{min}) R)\) and \(A((1-\tau (r)) R)\). \(\square \)

Piece-wise linear taxation.

The CDF of the after-tax return on wealth \((1-\tau )r_t\) under the flat rate \(\tau \)—with the pre-tax \(r_t \sim U[0, \bar{r}]\) – simply reads as \(F(r)=\frac{r}{\bar{r}(1-\tau )}\), with mean \(r_{M}:=E[(1-\tau )r_t]=(1-\tau )\frac{\bar{r}}{2}\). Any mean-preserving spread on \(r_t\)—call it \(r'_t\), with associated CDF \(F'(\cdot )\)—obtained through a change in the tax system, must be such that \(\int _{0}^{\bar{r}(1-\tau )}r'_t dF'=r_{M}\) and \(\int _0^{y} F(r)dr \le \int _0^{y} F'(r)dr\) for all \(y \le \bar{r}(1-\tau )\).

A straightforward instance of such risk transformation is a piece-wise linear tax scheme according to which (i) a lower flat rate \(\tau '<\tau \) is imposed on all taxpayers, and (ii) a fixed tax T (fulfilling some conditions) is also raised on higher-than-average returns on wealth \(r > r_M\). Specifically, let \(1-\tau '=\frac{1-\tau }{\gamma }\) for a given \(\gamma \in (0,1)\) and T such that \(T<r_M\) (all taxpayers gaining \(r>r_M\) on their wealth can afford paying this lump sum) and \(T=(\tau -\tau ')\bar{r}>0\) (the lump sum is strictly positive). Combining these requirement yields \(T=\frac{1-\gamma }{\gamma }(1-\tau )\bar{r}\) with \(\gamma >\frac{2}{3}\). It is easy to see that the CDF of the after-tax rates of return on wealth in the presence of the new flat rate \(\tau '\) writes as

$$\begin{aligned} F'(r) = \left\{ \begin{array}{ll} \frac{r}{\bar{r} (1-\tau ')} &{} \quad \text{ if } \quad r \le r_M \\ \\ \frac{r}{\bar{r} (1-\tau ')} +(1-\gamma ) &{} \quad \text{ if } \quad r>r_{M} \end{array} \right. \end{aligned}$$
(21)

and hence the associated mean \(E[r'_t]\) is equal to \(r_M\). Since \(F(0)=F'(0)=0\) and \(F\left( (1-\tau )\bar{r} \right) =F'\left( (1-\tau )\bar{r}\right) =1\), by the equality of the means we have

$$\begin{aligned} \int _{0}^{(1-\tau ) \bar{r}}\left( F(r)-F'(r) \right) dr= & {} \left[ - \int _0^{(1-\tau ) \bar{r}} r (f( r ) - f'( r ) ) d r \right] \\&+\,\left[ F\left( (1-\tau )\bar{r} \right) -F'\left( (1-\tau )\bar{r} \right) \right] (1-\tau )\bar{r}=0 \end{aligned}$$

from which it follows \(\int _0^{y} F(r)dr \le \int _0^{y} F'(r)dr\) for all y. Hence, \(F'\) constitutes a mean-preserving second-order dominant change in risk.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Di Pietro, C., Sorge, M.M. Stochastic dominance and thick-tailed wealth distributions. J Econ 123, 141–159 (2018). https://doi.org/10.1007/s00712-017-0544-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00712-017-0544-6

Keywords

JEL Classification

Navigation