Skip to main content

Advertisement

Log in

Bequests, estate taxes, and wealth distributions

  • Research Article
  • Published:
Economic Theory Aims and scope Submit manuscript

Abstract

Bossmann et al. (J Public Econ 91:1247–1271, 2007) found that estate taxes reduce the long-run wealth inequality. This result contrasts with the findings of the previous literature with idiosyncratic labor efficiency risk. We use a decomposition technique, developed by Davies (J Labor Econ 4:538–559, 1986), to reinvestigate the impact of estate taxes on the long-run wealth inequality. We find that the different results of estate taxes are due to the different redistribution effects.

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

Similar content being viewed by others

Notes

  1. Simulations of Davies and Kuhn (1991) show that estate taxes reduce wealth inequality in the short run, even though they increase inequality in the long run.

  2. Mino and Nakamoto (2016) studied wealth inequality in an economy of consumption externalities and heterogeneous preferences.

  3. Studies of incomplete-market heterogeneous agents models, such as Aiyagari (1994), Castaneda et al. (2003) and De Nardi (2004), Benhabib et al. (2015), and De Nardi and Yang (2016), incorporate precautionary savings motives into their models. They solve agent’s policy functions numerically and simulate the stationary wealth distribution. Benhabib et al. (2011) also found agent’s policy functions explicitly. They use idiosyncratic investment risk to generate the observed fat tail of the wealth distribution in the USA.

  4. Since our model has linear policy functions, wealth distribution does not influence the aggregate economy. Algan et al. (2011) built a model in which wealth redistribution can influence the aggregate output. Antunes et al. (2015) investigated the feedback of wealth distribution on the aggregate economy.

  5. Our model has a simple demographic structure. Modeling a more complicated demographic structure Mierau and Turnovsky (2014) studied the relationship between demography and wealth inequality.

  6. We use \(\{x_{t}\}\) to represent a sequence in this paper.

  7. Note that we do not need to assume that \(\hbox {Var}(l_{t})<\infty \).

  8. Here \(=_{st}\) denotes equality in distribution.

  9. We state this intuition formally in Lemma 2 of “Appendix A.4”.

  10. For the mathematical definition of the Lorenz curve, \(L_{X}(p)\), see Gastwirth (1971).

  11. For inequality measures, our main reference book is Shaked and Shanthikumar (2010). A good reference of Lorenz dominance is Arnold (1987).

  12. See pages 68–69 of Marshall and Olkin (2007).

  13. The linear process is

    $$\begin{aligned} l_{t+1}=\bar{l}+v(l_{t}-\bar{l})+\varepsilon _{t+1} \end{aligned}$$

    where \(\bar{l}=1\) and \(0<v<1\). \(\{\varepsilon _{t}\}\) is i.i.d. with a zero mean, a finite variance, and a lower bound sufficient to keep \( l_{t+1}>0 \). This process is used in Davies (1986) and Davies and Kuhn (1991). Solon (1992) and Zimmerman (1992) used different datasets in the USA to study the intergenerational mobility and found that the elasticity of child’s earnings with respect to parent’s earnings is about 0.4.

  14. See comments after Proposition 1 about the increase in mean wealth caused by bequest motives.

  15. There is one minor difference between our result and that of Bossmann et al. (2007). We can only show \(\hbox {CV}(a_{\infty }^{A})\ge \hbox {CV}(a_{\infty }^{B})\), while they show \(\hbox {CV}(a_{\infty }^{A})>\hbox {CV}(a_{\infty }^{B})\).

  16. This intuition comes from the mathematical result that \(X+a\) Lorenz dominates \(X+b\) for any nonnegative random variable X with a finite positive mean and \(a>b>0\) [see Theorem 3.A.25 of Shaked and Shanthikumar (2010)]. Thus, \(X+a\) is more equal than \(X+b\).

  17. For an individual with before-tax wealth x, the effective average tax rate is

    $$\begin{aligned} \frac{x-\left[ \left( 1-\zeta \right) x+\zeta E(X)\right] }{x}=\zeta \left[ 1-\frac{E(X)}{x}\right] , \end{aligned}$$

    which is increasing in x.

  18. See Fellman (1976) for a study on the effect of progressive taxes on income distributions.

  19. In a simulation exercise Bossmann et al. (2007) found that the estate tax reduces the Gini coefficient of the long-run wealth distribution. Our theoretical result of Theorem 2 supports the simulation results of the Gini coefficient in Bossmann et al. (2007).

  20. Zhu (2017) introduced idiosyncratic investment risk into the Becker–Tomes model and found that the inheritance effect of the estate tax reduces the long-run wealth inequality in the model with sufficiently volatile idiosyncratic investment risk.

  21. We assume that young agents live together with their parents.

  22. The negative root of Eq. (18) cannot be the equilibrium interest rate in the economy with housing. The agent has two ways of holding assets from the young period to the old period, savings and housing. We assume that housing does not depreciate. And the housing price does not change in the stationary equilibrium. Thus, the return of housing is positive. There exist arbitrages if the interest rate of saving is negative.

  23. The Gini coefficient of the earnings distribution is 0.33. The Gini coefficients of the long-run wealth distribution in Table 4 are larger than this number since there exists a life cycle pattern of savings.

  24. In the equilibrium r could be negative. Since saving is the only way to bring wealth to the next period, even if r is negative, the agent still saves.

  25. For two random variables X and Y, X is smaller than Y in the convex order, denoted by \(X\preceq _{cx}Y\), if and only if

    $$\begin{aligned} E[\phi (X)]\le E[\phi (Y)] \end{aligned}$$

    for all convex functions \(\phi : \mathbb {R} \rightarrow \mathbb {R} \), provided the expectations exist. For more properties of the convex order, see Shaked and Shanthikumar (2010).

  26. We abuse notations a little bit. We use \(a_{t}\) instead of \(a_{t}^{B}\) without confusions.

  27. Let X be a random variable with a finite mean. \(E(X)\preceq _{cx}X\) can be established by applying Jensen’s Inequality and the definition of the convex order.

  28. \(X\preceq _{cx}Y\) implies \(bX\preceq _{cx}bY\) for any \(b\in \mathbb {R} \). Note that \(\phi (bx)\) is a convex function of \(x\in \mathbb {R} \) if \(\phi (x)\) is a convex function of \(x\in \mathbb {R} \).

  29. After we solve the general equilibrium in our benchmark model with “joy of giving” bequest motives, the estate tax does not affect the prices of r and w when utility functions are logarithmic. Thus, the estate tax does not have a general equilibrium effect. However, the estate tax does affect the prices of r and w in a model with altruistic bequest motives even for logarithmic utility functions. Then, the estate tax does have a general equilibrium effect. Thus, a model with altruistic bequest motives and endogenous prices of r and w is not comparable to our benchmark model.

    We then decide to follow the studies of Becker and Tomes (1979) and Davies (1986) to assume that the prices of r and w are exogenous. Thus, we can concentrate on the inheritance effect and the redistribution effect of the estate tax.

  30. Note that

    $$\begin{aligned} c_{12}=\frac{1}{\left[ (1+r)\left( 1-\zeta \right) \right] ^{-1}+\left[ (1+r)\left( 1-\zeta \right) \right] ^{-\frac{1}{\eta }}\chi ^{-\frac{1}{\eta }}}. \end{aligned}$$

References

  • Aiyagari, S.R.: Uninsured idiosyncratic risk and aggregate saving. Q. J. Econ. 109, 659–684 (1994)

    Article  Google Scholar 

  • Algan, A., Challe, E., Ragot, X.: Incomplete markets and the output-inflation tradeoff. Econ. Theory 46, 55–84 (2011)

    Article  Google Scholar 

  • Antunes, A., Cavalcanti, T., Villamil, A.: The effects of credit subsidies on development. Econ. Theory 58, 1–30 (2015)

    Article  Google Scholar 

  • Arnold, B.: Majorization and the Lorenz Order: A Brief Introduction. Springer, Berlin (1987)

    Book  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Chatterjee, S.: Transitional dynamics and the distribution of wealth in a neoclassical growth model. J. Public Econ. 54, 97–119 (1994)

    Article  Google Scholar 

  • Davies, J.: Does redistribution reduce inequality? J. Labor Econ. 4, 538–559 (1986)

    Article  Google Scholar 

  • Davies, J., Kuhn, P.: A dynamic model of redistribution, inheritance, and inequality. Can. J. Econ. 24, 324–344 (1991)

    Article  Google Scholar 

  • De Nardi, M.: Wealth inequality and intergenerational links. Rev. Econ. Stud. 71, 743–768 (2004)

    Article  Google Scholar 

  • De Nardi, M., Yang, F.: Wealth inequality, family background, and estate taxation. J. Monet. Econ. 77, 130–145 (2016)

    Article  Google Scholar 

  • Fellman, J.: The effect of transformations on Lorenz curves. Econometrica 44, 823–824 (1976)

    Article  Google Scholar 

  • Gajdos, T., Weymark, J.: Introduction to inequality and risk. J. Econ. Theory 147, 1313–1330 (2012)

    Article  Google Scholar 

  • Gale, W., Perozek, M.: Do estate taxes reduce saving? In: Gale, W., Hines Jr., J., Slemrod, J. (eds.) Rethinking Estate and Gift Taxation, pp. 216–257. Brookings Institution Press, Washington (2001)

    Chapter  Google Scholar 

  • Gastwirth, J.: A general definition of the Lorenz curve. Econometrica 39, 1037–1039 (1971)

    Article  Google Scholar 

  • Kopczuk, W.: Taxation of intergenerational transfers and wealth. In: Auerbach, A., Chetty, R., Feldstein, M., Saez, E. (eds.) Handbook of Public Economics, pp. 329–390. Elsevier, Amsterdam (2013)

    Chapter  Google Scholar 

  • Marshall, A., Olkin, I.: Life Distributions. Springer, New York (2007)

    Google Scholar 

  • Mierau, J., Turnovsky, S.: Demography, growth, and inequality. Econ. Theory 55, 29–68 (2014)

    Article  Google Scholar 

  • Mino, K., Nakamoto, Y.: Heterogeneous conformism and wealth distribution in a neoclassical growth model. Econ. Theory 62, 689–717 (2016)

    Article  Google Scholar 

  • Pestieau, P., Thilbault, E.: Love the children or money: reflections on debt neutrality and estate taxation. Econ. Theory 50, 31–57 (2012)

    Article  Google Scholar 

  • Rothschild, M., Stiglitz, J.: Some further results on the measurement of inequality. J. Econ. Theory 6, 188–204 (1973)

    Article  Google Scholar 

  • Shaked, M., Shanthikumar, G.: Stochastic Orders. Springer, New York (2010)

    Google Scholar 

  • Solon, G.: Intergenerational income mobility in the United States. Am. Econ. Rev. 82, 393–408 (1992)

    Google Scholar 

  • Zhu, S.: A Becker–Tomes model with investment risk. Mimeo, Beihang University (2017)

  • Zilcha, I.: Intergenerational transfers, production and income distribution. J. Public Econ. 87, 489–513 (2003)

    Article  Google Scholar 

  • Zimmerman, D.: Regression toward mediocrity in economic stature. Am. Econ. Rev. 82, 409–429 (1992)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shenghao Zhu.

Additional information

A previous version of this paper was circulated as “Intergenerational links, taxation, and wealth distributions.” We thank Daniel Barczyk, Jess Benhabib, Alberto Bisin, Yongheng Deng, Aditya Goenka, Tomoo Kikuchi, Christian Kleiber, Haoming Liu, Baochun Peng, Ariell Reshef, Thomas Sargent, Klaus W älde, C.C. Yang, Ting Zeng, and Jie Zhang. Shenghao Zhu acknowledges the financial support from Advanced Innovation Center for Big Data-Based Precision Medicine, Beihang University.

A Appendix

A Appendix

1.1 A.1 Proof of Proposition 1

Proof

Letting \(K_{t+1}=K_{t}=K\) in Eq. (11) we have

$$\begin{aligned} K=\left( \frac{1-\alpha +\varphi \alpha }{1+\tilde{\beta }^{-\frac{1}{\eta } }-\varphi (1-\delta )}A\right) ^{\frac{1}{1-\alpha }}, \end{aligned}$$
(A.1)

where \(\tilde{\beta }=\beta \left[ 1+\chi ^{\frac{1}{\eta }}(1-\zeta )^{\frac{ 1-\eta }{\eta }}\right] ^{\eta }(1+r)^{1-\eta }\) and \(r=\alpha AK^{\alpha -1}-\delta \).

Plugging Eq. (A.1) into \(r=\alpha AK^{\alpha -1}-\delta \) we have

$$\begin{aligned} \frac{r+\delta }{\alpha }=\frac{1+\tilde{\beta }^{-\frac{1}{\eta }}-\varphi (1-\delta )}{1-\alpha +\varphi \alpha }. \end{aligned}$$
(A.2)

Plugging \(\tilde{\beta }=\beta \left[ 1+\chi ^{\frac{1}{\eta }}(1-\zeta )^{ \frac{1-\eta }{\eta }}\right] ^{\eta }(1+r)^{1-\eta }=\frac{\beta }{ (1-\varphi )^{\eta }}(1+r)^{1-\eta }\) into Eq. (A.2) we have

$$\begin{aligned} \frac{1-\alpha }{\alpha }(r+\delta )+\varphi (1+r)-(1-\varphi )\beta ^{- \frac{1}{\eta }}(1+r)^{1-\frac{1}{\eta }}=1. \end{aligned}$$

We show Proposition 1 in two cases:

Case (i) \(\eta >1\)

Note that \(0<\varphi <1\). Define

$$\begin{aligned} h(\varphi ,r)=\frac{1-\alpha }{\alpha }(r+\delta )+\varphi (1+r)-(1-\varphi )\beta ^{-\frac{1}{\eta }}(1+r)^{1-\frac{1}{\eta }}. \end{aligned}$$

The equilibrium r is determined by

$$\begin{aligned} h(\varphi ,r)=1. \end{aligned}$$

Note that \(h(\varphi ,r)\) is a continuous function of r, with

$$\begin{aligned} h(\varphi ,-\delta )=\varphi (1-\delta )-(1-\varphi )\beta ^{-\frac{1}{\eta } }(1-\delta )^{1-\frac{1}{\eta }}<\varphi (1-\delta )<1 \end{aligned}$$

and

$$\begin{aligned} \lim _{r\rightarrow \infty }h(\varphi ,r)=\infty \end{aligned}$$

Also \(h_{22}(\varphi ,r)=\left( 1-\frac{1}{\eta }\right) \frac{1}{\eta } (1-\varphi )\beta ^{-\frac{1}{\eta }}(1+r)^{-\frac{1}{\eta }-1}>0\) due to \( \eta >1\). Thus, \(h(\varphi ,r)\) is a strictly convex function of r. Therefore, there must exist a unique equilibrium \(r>-\delta \).Footnote 24

Note that \(h(\varphi ,r)\) is strictly increasing in \(\varphi \). For \(\varphi _{1}<\varphi _{2}<1\), suppose that

$$\begin{aligned} h(\varphi _{1},r_{1})=1\text { and }h(\varphi _{2},r_{2})=1. \end{aligned}$$

We have

$$\begin{aligned} h(\varphi _{2},r_{1})>h(\varphi _{1},r_{1})=1. \end{aligned}$$

Thus, \(r_{2}<r_{1}\) since \(h(\varphi _{2},-\delta )<1\) and \(h(\varphi _{2},r)\) is a continuous function of r. A higher \(\chi \) implies a higher \(\varphi \) . Thus, a higher \(\chi \) implies a lower r and a higher K.

Case (ii) \(\eta =1\)

In this case \(\tilde{\beta }=\beta (1+\chi )\) and \(\varphi =\frac{1}{1+\frac{1 }{\chi }}\), Eq. (A.1) implies

$$\begin{aligned} K= & {} \left( \frac{1-\alpha +\chi }{1+\frac{1}{\beta }+\delta \chi }A\right) ^{\frac{1}{1-\alpha }} \\= & {} \left( \left[ \frac{1}{\delta }-\frac{\frac{1}{\delta }\left( 1+\frac{1}{ \beta }\right) -(1-\alpha )}{1+\frac{1}{\beta }+\delta \chi }\right] A\right) ^{\frac{1}{1-\alpha }}. \end{aligned}$$

Thus, a higher \(\chi \) implies a higher K. \(\square \)

1.2 A.2 Proof of Proposition 2

Proof

Obviously \(c_{4}\ge 0\). From Eq. (A.2) we have

$$\begin{aligned} 1+r=\frac{\left( 1+\tilde{\beta }^{-\frac{1}{\eta }}\right) \alpha +(1-\delta )(1-\alpha )}{(1-\alpha )+\varphi \alpha } \end{aligned}$$

Thus,

$$\begin{aligned} c_{4}=\frac{\left( 1-\zeta \right) \varphi (1+r)}{1+\tilde{\beta }^{-\frac{1}{ \eta }}}=\left( 1-\zeta \right) \frac{\alpha +\frac{1-\delta }{1+\tilde{\beta }^{-\frac{1}{\eta }}}(1-\alpha )}{\alpha +\frac{1}{\varphi }(1-\alpha )}<1 \end{aligned}$$

since \(\tilde{\beta }=\beta \left[ 1+\chi ^{\frac{1}{\eta }}(1-\zeta )^{\frac{ 1-\eta }{\eta }}\right] ^{\eta }(1+r)^{1-\eta }>0\) and \(0<\varphi <1\). \(\square \)

1.3 A.3 Proof of Proposition 3

Proof

From Eq. (13) we have

$$\begin{aligned} a_{t+1}=c_{3}l_{t}+c_{4}a_{t}+c_{5}, \end{aligned}$$

where \(c_{3}=\frac{1}{1+\tilde{\beta }^{-\frac{1}{\eta }}}w\), \(c_{4}=\frac{ \left( 1-\zeta \right) \varphi (1+r)}{1+\tilde{\beta }^{-\frac{1}{\eta }}}\), and \(c_{5}=\frac{\zeta \varphi (1+r)}{1+\tilde{\beta }^{-\frac{1}{\eta }}}K\). Let \(B_{t}=c_{5}+c_{3}l_{t}\). We have

$$\begin{aligned} a_{t+1}=c_{4}a_{t}+B_{t}. \end{aligned}$$
(A.3)

Note that \(\{B_{t}\}\) is stationary and ergodic since \(\{l_{t}\}\) is stationary and ergodic by Assumption 1. We have \(-\infty \le \log c_{4}<0\). Also \(E(B_{t})=c_{5}+c_{3}<\infty \), since \(E(l_{t})=1\) by Assumption 2. Thus, \(E(\log B_{t})^{+}\le E(B_{t})<\infty \). By Theorem 1 of Brandt (1986) we know that \(a_{t}\) converges to \(\sum _{j=0}^{\infty }c_{4}^{j}B_{t-j-1}\) almost surely as t approaches infinity. Thus, we have

$$\begin{aligned} a_{t}\rightarrow _{st}\sum _{j=0}^{\infty }c_{4}^{j}B_{t-j-1}\text { as } t\rightarrow \infty . \end{aligned}$$

Since \(\{B_{t}\}\) is stationary, we know that the sequence of \( (B_{t-1},B_{t-2}, \ldots ,B_{t-j-1}, \ldots )\) has the same distribution as the sequence of \((B_{0},B_{1}, \ldots ,B_{s}, \ldots )\). Thus, we have

$$\begin{aligned} \sum _{j=0}^{\infty }c_{4}^{j}B_{t-j-1}=_{st}\sum _{s=0}^{\infty }c_{4}^{s}B_{s},\text { }\forall t\in \mathbb {Z} . \end{aligned}$$

Let

$$\begin{aligned} a_{\infty }=_{st}\sum _{s=0}^{\infty }c_{4}^{s}B_{s}=_{st}c_{3}\sum _{s=0}^{\infty }c_{4}^{s}l_{s}+\frac{c_{5}}{ 1-c_{4}}. \end{aligned}$$

Thus, we know that

$$\begin{aligned} a_{t}\rightarrow _{st}a_{\infty }\text { as }t\rightarrow \infty . \end{aligned}$$

\(\square \)

1.4 A.4 Proof of Theorem 1

Proof

Theorem 3.A.36 of Shaked and Shanthikumar (2010) shows that

Lemma 2

Let \(X_{1}\), \(X_{2}\), \(\ldots \), \(X_{n}\) and Y be \(n+1\) random variables. If \(X_{i}\preceq _{cx}Y\), \(i=1\), 2, \(\ldots \), n, then

$$\begin{aligned} \sum _{i=1}^{n}a_{i}X_{i}\preceq _{cx}Y, \end{aligned}$$

whenever \(a_{i}\ge 0\), \(i=1\), 2, \(\ldots \), n, and \( \sum _{i=1}^{n}a_{i}=1\).Footnote 25

Theorem 3.A.10 of Shaked and Shanthikumar (2010) states that

Lemma 3

Let X and Y be two nonnegative random variables with equal means. Then, \(X\preceq _{cx}Y\) if and only if \(L_{X}(p)\ge L_{Y}(p)\) for all \(p\in [0,1]\).

Note that \(a_{\infty }^{A}\) has the same Lorenz curve as \(l_{1}\). We only need to show that \(a_{\infty }^{B}\succeq _{L}l_{1}\).

In economy B, pick \(a_{1}=\frac{c_{3}}{1-c_{4}}\).Footnote 26 Thus,

$$\begin{aligned} a_{1}\preceq _{cx}\frac{c_{3}}{1-c_{4}}l_{1} \end{aligned}$$

since \(a_{1}=E\left( \frac{c_{3}}{1-c_{4}}l_{1}\right) \).Footnote 27

Suppose that

$$\begin{aligned} a_{t}\preceq _{cx}\frac{c_{3}}{1-c_{4}}l_{1}\text {.} \end{aligned}$$

Thus, \(\frac{1-c_{4}}{c_{3}}a_{t}\preceq _{cx}l_{1}\).Footnote 28

And

$$\begin{aligned} a_{t+1}= & {} c_{3}l_{t}+c_{4}a_{t} \\= & {} \frac{c_{3}}{1-c_{4}}\left( (1-c_{4})l_{t}+c_{4}\frac{1-c_{4}}{c_{3}} a_{t}\right) . \end{aligned}$$

Note that \((1-c_{4})l_{t}+c_{4}\frac{1-c_{4}}{c_{3}}a_{t}\) is a weighted average of \(l_{t}\) and \(\frac{1-c_{4}}{c_{3}}a_{t}\). For \(\forall t\ge 1\), \( l_{t}\) and \(l_{1}\) have the same distribution. We have \(l_{t}\preceq _{cx}l_{1}\), \(\forall t\ge 1\). By Lemma 2 we have

$$\begin{aligned} (1-c_{4})l_{t}+c_{4}\frac{1-c_{4}}{c_{3}}a_{t}\preceq _{cx}l_{1}. \end{aligned}$$

Thus,

$$\begin{aligned} a_{t+1}\preceq _{cx}\frac{c_{3}}{1-c_{4}}l_{1}. \end{aligned}$$

By mathematical induction we have

$$\begin{aligned} a_{t}\preceq _{cx}\frac{c_{3}}{1-c_{4}}l_{1}, \ \ \forall t\ge 1. \end{aligned}$$

Since \(a_{t}\rightarrow _{st}a_{\infty }^{B}\) as t approaches infinity, we have

$$\begin{aligned} a_{\infty }^{B}\preceq _{cx}\frac{c_{3}}{1-c_{4}}l_{1}, \end{aligned}$$

by part (c) of Theorem 3.A.12 of Shaked and Shanthikumar (2010). By Lemma 3 we have \(a_{\infty }^{B}\succeq _{L}\frac{c_{3}}{1-c_{4}} l_{1} \) since \(E\left( a_{\infty }^{B}\right) =E\left( \frac{c_{3}}{1-c_{4}} l_{1}\right) =\frac{c_{3}}{1-c_{4}}\). Thus, \(a_{\infty }^{B}\succeq _{L}l_{1} \). \(\square \)

1.5 A.5 Proof of Lemma 1

Proof

Let

$$\begin{aligned} g(x)=(1-\zeta )x+\zeta E\left( X\right) , \ \ x\in [0,+\infty ) \end{aligned}$$

and

$$\begin{aligned} h(x)=(1-\hat{\zeta })x+\hat{\zeta }E\left( X\right) , \ \ x\in [0,+\infty ) \end{aligned}$$

Note that \(g(\cdot )\) and \(h(\cdot )\) are nonnegative increasing functions defined on \([0,+\infty )\), since \(0\le \hat{\zeta }\le \zeta <1\). Also \( g(x)>0\) and \(h(x)>0\) for \(x>0\). Note that \(\frac{h(x)}{g(x)}\) is increasing in \(x\in (0,+\infty )\), since

$$\begin{aligned} \frac{h(x)}{g(x)}= & {} \frac{(1-\hat{\zeta })x+\hat{\zeta }E\left( X\right) }{ (1-\zeta )x+\zeta E\left( X\right) } \\= & {} \frac{1-\hat{\zeta }}{1-\zeta }\left[ 1-\frac{\zeta -\hat{\zeta }}{\left( 1- \hat{\zeta }\right) \left( 1-\zeta \right) }\frac{E(X)}{x+\frac{\zeta }{ 1-\zeta }E(X)}\right] . \end{aligned}$$

By Theorem 3.A.26 of Shaked and Shanthikumar (2010) we have \(g(X)\succeq _{L}h(X)\), i.e., \((1-\zeta )X+\zeta E\left( X\right) \succeq _{L}(1-\hat{\zeta })X+\hat{\zeta }E\left( X\right) \). By Lemma 3 we have \((1-\zeta )X+\zeta E\left( X\right) \preceq _{cx}(1-\hat{\zeta })X+\hat{\zeta }E\left( X\right) \) since \(E\left[ (1-\zeta )X+\zeta E\left( X\right) \right] =E(X)=E \left[ (1-\hat{\zeta })X+\hat{\zeta }E\left( X\right) \right] \). \(\square \)

1.6 A.6 Proof of Theorem 2

Proof

Note that \(a_{\infty }^{\zeta }\) is the stationary distribution of the stochastic process \(\{a_{t}^{\zeta }\}\) which is generated by

$$\begin{aligned} a_{t+1}^{\zeta }=c_{6}l_{t}+c_{7}\left[ (1-\zeta )a_{t}^{\zeta }+\zeta \bar{K }\right] \end{aligned}$$

and a given \(a_{1}^{\zeta }\). And \(a_{\infty }^{\hat{\zeta }}\) is the stationary distribution of the stochastic process \(\{a_{t}^{\hat{\zeta }}\}\) which is generated by

$$\begin{aligned} a_{t+1}^{\hat{\zeta }}=c_{6}l_{t}+c_{7}\left[ (1-\hat{\zeta })a_{t}^{\hat{\zeta }}+\hat{\zeta }\bar{K}\right] \end{aligned}$$

and a given \(a_{1}^{\hat{\zeta }}\).

Let \(a_{1}^{\zeta }=_{st}a_{1}^{\hat{\zeta }}\). Thus, \(a_{1}^{\zeta }\preceq _{cx}a_{1}^{\hat{\zeta }}\) by the definition of the convex order.

Now suppose that \(a_{t}^{\zeta }\preceq _{cx}a_{t}^{\hat{\zeta }}\). By Lemma 1 we have

$$\begin{aligned} (1-\zeta )a_{t}^{\zeta }+\zeta \bar{K}\preceq _{cx}(1-\hat{\zeta } )a_{t}^{\zeta }+\hat{\zeta }\bar{K} \end{aligned}$$

since \(E\left( a_{t}^{\zeta }\right) =\bar{K}\).

By Corollary 3.A.22 of Shaked and Shanthikumar (2010) we have \((1-\hat{\zeta } )a_{t}^{\zeta }\preceq _{cx}(1-\hat{\zeta })a_{t}^{\hat{\zeta }}\) since \( \left( 1-\hat{\zeta }\right) \) is independent of \(a_{t}^{\zeta }\) and \(a_{t}^{ \hat{\zeta }}\). By Part (d) of Theorem 3.A.12 of Shaked and Shanthikumar (2010) we have

$$\begin{aligned} (1-\hat{\zeta })a_{t}^{\zeta }+\hat{\zeta }\bar{K}\preceq _{cx}(1-\hat{\zeta } )a_{t}^{\hat{\zeta }}+\hat{\zeta }\bar{K}, \end{aligned}$$

since \(\hat{\zeta }\bar{K}\) is independent of \((1-\hat{\zeta })a_{t}^{\zeta }\) and \((1-\hat{\zeta })a_{t}^{\hat{\zeta }}\). By the transitivity of the convex order we have

$$\begin{aligned} (1-\zeta )a_{t}^{\zeta }+\zeta \bar{K}\preceq _{cx}(1-\hat{\zeta })a_{t}^{ \hat{\zeta }}+\hat{\zeta }\bar{K}. \end{aligned}$$

Thus, we have \(c_{7}\left[ (1-\zeta )a_{t}^{\zeta }+\zeta \bar{K}\right] \preceq _{cx}c_{7}\left[ (1-\hat{\zeta })a_{t}^{\hat{\zeta }}+\hat{\zeta }\bar{K }\right] \) by the property of the convex order in Footnote 28. Note that \( c_{6}l_{t}\) and \(c_{7}\left[ (1-\zeta )a_{t}^{\zeta }+\zeta \bar{K}\right] \) are independent. And \(c_{6}l_{t}\) and \(c_{7}\left[ (1-\hat{\zeta })a_{t}^{ \hat{\zeta }}+\hat{\zeta }\bar{K}\right] \) are independent. Thus, by part (d) of Theorem 3.A.12 of Shaked and Shanthikumar (2010), we have

$$\begin{aligned} c_{6}l_{t}+c_{7}\left[ (1-\zeta )a_{t}^{\zeta }+\zeta \bar{K}\right] \preceq _{cx}c_{6}l_{t}+c_{7}\left[ (1-\hat{\zeta })a_{t}^{\hat{\zeta }}+\hat{\zeta } \bar{K}\right] . \end{aligned}$$

Thus, we have

$$\begin{aligned} a_{t+1}^{\zeta }\preceq _{cx}a_{t+1}^{\hat{\zeta }}. \end{aligned}$$

By mathematical induction we have

$$\begin{aligned} a_{t}^{\zeta }\preceq _{cx}a_{t}^{\hat{\zeta }},\text { }\forall t\ge 1. \end{aligned}$$

Since \(a_{t}^{\zeta }\rightarrow _{st}a_{\infty }^{\zeta }\) and \(a_{t}^{\hat{ \zeta }}\rightarrow _{st}a_{\infty }^{\hat{\zeta }}\) as t approaches infinity, we have

$$\begin{aligned} a_{\infty }^{\zeta }\preceq _{cx}a_{\infty }^{\hat{\zeta }}, \end{aligned}$$

by part (c) of Theorem 3.A.12 of Shaked and Shanthikumar (2010). By Lemma 3 we have

$$\begin{aligned} a_{\infty }^{\zeta }\succeq _{L}a_{\infty }^{\hat{\zeta }}, \end{aligned}$$

since \(E\left( a_{\infty }^{\zeta }\right) =E\left( a_{\infty }^{\hat{\zeta } }\right) =\bar{K}\). \(\square \)

1.7 A.7 An alternative setup of the model

Here we investigate an alternative setup of our benchmark model. The main difference is that \(b_{t}\) in our benchmark model is the before-tax bequest. In this alternative setup, \(b_{t}\) is the after-tax bequest.

The agent’s problem is

$$\begin{aligned}&\max _{c_{t}^{y},s_{t},c_{t+1}^{o},b_{t+1}}\log c_{t}^{y}+\beta \left( \log c_{t+1}^{o}+\chi \log b_{t+1}\right) \\&\quad \hbox {s.t.} \ \ c_{t}^{y}+s_{t} =w_{t}l_{t}+b_{t}+g_{t}, \\&\quad c_{t+1}^{o}+\left( 1+\zeta \right) b_{t+1} =(1+r_{t+1})s_{t}. \end{aligned}$$

The agent’s optimal policy functions are

$$\begin{aligned} c_{t+1}^{o}= & {} \frac{1}{1+\chi }(1+r_{t+1})s_{t},\\ b_{t+1}= & {} \frac{\chi }{\left( 1+\chi \right) \left( 1+\zeta \right) } (1+r_{t+1})s_{t},\\ c_{t}^{y}= & {} \frac{1}{1+\beta \left( 1+\chi \right) }\left( w_{t}l_{t}+b_{t}+g_{t}\right) , \end{aligned}$$

and

$$\begin{aligned} \quad \, s_{t}=\frac{\beta \left( 1+\chi \right) }{1+\beta \left( 1+\chi \right) } \left( w_{t}l_{t}+b_{t}+g_{t}\right) . \end{aligned}$$

From the government’s budget constraint we have

$$\begin{aligned} g_{t}=\zeta \int b_{t}di, \end{aligned}$$

where \(\int di\) denotes the aggregation of old agents.

Thus, the aggregate capital follows

$$\begin{aligned} K_{t+1}= & {} \int s_{t}di \\= & {} \frac{\beta \left( 1+\chi \right) }{1+\beta \left( 1+\chi \right) }\left[ w_{t}+\frac{\chi }{1+\chi }(1+r_{t})K_{t}\right] , \end{aligned}$$

where \(w_{t}=(1-\alpha )AK_{t}^{\alpha }\) and \(r_{t}=\alpha AK_{t}^{\alpha -1}-\delta \).

In the steady-state aggregate economy we have \(K_{t+1}=K_{t}=\bar{K}\). Thus, we have

$$\begin{aligned} \bar{K}=\left( \frac{1-\alpha +\chi }{1+\frac{1}{\beta }+\delta \chi } A\right) ^{\frac{1}{1-\alpha }}. \end{aligned}$$

The estate tax does not affect the aggregate capital. Then, \(\bar{w} =(1-\alpha )A\left( \bar{K}\right) ^{\alpha }\) and \(\bar{r}=\alpha A\left( \bar{K}\right) ^{\alpha -1}-\delta \).

Let \(a_{t+1}=s_{t}\). From the agent’s policy functions we have the individual wealth accumulation equation,

$$\begin{aligned} a_{t+1}=c_{6}l_{t}+c_{7}\left[ \frac{1}{1+\zeta }a_{t}+\frac{\zeta }{1+\zeta }\bar{K}\right] , \end{aligned}$$
(21)

with \(c_{6}=\frac{1}{1+\frac{1}{\beta \left( 1+\chi \right) }}\bar{w}\) and \( c_{7}=\frac{1}{\left( 1+\frac{1}{\beta \left( 1+\chi \right) }\right) \left( 1+\frac{1}{\chi }\right) }\left( 1+\bar{r}\right) \). Both \(c_{6}\) and \(c_{7}\) do not depend on the estate tax \(\zeta \).

From Eq. (21) we have the long-run wealth distribution,

$$\begin{aligned} a_{\infty }=_{st}\sum _{s=0}^{\infty }\left( \tilde{c}_{8}\right) ^{s}\left( c_{6}l_{s}+c_{7}\frac{\zeta }{1+\zeta }\bar{K}\right) , \end{aligned}$$

where \(\tilde{c}_{8}=c_{7}\frac{1}{1+\zeta }\). Comparing Eqs. (21 ) and (15) we find that all the theoretical results of the long-run wealth inequality in the benchmark model still hold in this alternative setup.

1.8 A.8 The Becker–Tomes model

Here we briefly review some main results of Becker–Tomes models by Becker and Tomes (1979) and Davies (1986). As in Becker and Tomes (1979) and Davies (1986), we assume that each agent only lives for one period. At the end of the period, the agent dies and gives birth to one child. The prices of r and w are exogenous and constant. Davies (1986) explained that the aim of using exogenous prices of r and w in his paper is exactly close to the general equilibrium effect of the estate tax.Footnote 29

As in Becker and Tomes (1979) and Davies (1986) we assume that the agent can correctly anticipate the labor efficiency of his child. The agent’s problem is

$$\begin{aligned}&\max _{c_{t},b_{t+1},I_{t+1}}\frac{c_{t}^{1-\eta }-1}{1-\eta }+\chi \frac{ I_{t+1}^{1-\eta }-1}{1-\eta }\\&\quad \hbox {s.t.} \ \ c_{t}+b_{t+1}=I_{t},\\&\quad I_{t+1}=wl_{t+1}+(1+r)\left( 1-\zeta \right) b_{t+1}+g, \end{aligned}$$

where \(I_{t+1}\) is the total wealth of the child. The agent’s optimal policy functions are

$$\begin{aligned} c_{t}= & {} \frac{1}{1+\left[ (1+r)\left( 1-\zeta \right) \right] ^{\frac{1-\eta }{ \eta }}\chi ^{\frac{1}{\eta }}}\left( I_{t}+\frac{wl_{t+1}+g}{(1+r)\left( 1-\zeta \right) }\right) ,\\ b_{t+1}= & {} \frac{1}{1+\left[ (1+r)\left( 1-\zeta \right) \right] ^{\frac{\eta -1 }{\eta }}\chi ^{-\frac{1}{\eta }}}I_{t}-\frac{1}{1+\left[ (1+r)\left( 1-\zeta \right) \right] ^{\frac{1-\eta }{\eta }}\chi ^{\frac{1}{\eta }}} \frac{wl_{t+1}+g}{(1+r)\left( 1-\zeta \right) }, \end{aligned}$$

and

$$\begin{aligned} I_{t+1}=\frac{(1+r)\left( 1-\zeta \right) }{1+\left[ (1+r)\left( 1-\zeta \right) \right] ^{\frac{\eta -1}{\eta }}\chi ^{-\frac{1}{\eta }}}\left( I_{t}+\frac{wl_{t+1}+g}{(1+r)\left( 1-\zeta \right) }\right) . \end{aligned}$$
(22)

From the government’s budget constraint we have

$$\begin{aligned} g=\zeta (1+r)\int b_{t}di, \end{aligned}$$

where \(\int di\) denotes the aggregation of young agents.

In the steady-state aggregate economy we have \(\int I_{t+1}di=\int I_{t}di= \bar{I}\). Thus, we have

$$\begin{aligned} \bar{I}=\int I_{t}di=w+(1+r)\left( 1-\zeta \right) \int b_{t}di+g=w+\frac{g}{ \zeta }. \end{aligned}$$
(23)

From Eq. (22) we have

$$\begin{aligned} \bar{I}=\frac{(1+r)\left( 1-\zeta \right) }{1+\left[ (1+r)\left( 1-\zeta \right) \right] ^{\frac{\eta -1}{\eta }}\chi ^{-\frac{1}{\eta }}}\left( \bar{ I}+\frac{w+g}{(1+r)\left( 1-\zeta \right) }\right) . \end{aligned}$$
(24)

Combining Eqs. (23) and (24) we have

$$\begin{aligned} \bar{I}=\frac{1}{1-(1+r)\left( 1-\left[ (1+r)\left( 1-\zeta \right) \right] ^{-\frac{1}{\eta }}\chi ^{-\frac{1}{\eta }}\right) }w, \end{aligned}$$

and

$$\begin{aligned} g=mw, \end{aligned}$$

where \(m=\frac{\zeta }{\frac{1}{(1+r)\left( 1-\left[ (1+r)\left( 1-\zeta \right) \right] ^{-\frac{1}{\eta }}\chi ^{-\frac{1}{\eta }}\right) }-1}\).

From Eq. (22) we have the individual wealth accumulation equation,

$$\begin{aligned} I_{t+1}=c_{12}I_{t}+\frac{1}{1+\left[ (1+r)\left( 1-\zeta \right) \right] ^{ \frac{\eta -1}{\eta }}\chi ^{-\frac{1}{\eta }}}\left( wl_{t+1}+g\right) , \end{aligned}$$
(25)

where \(c_{12}=\frac{(1+r)\left( 1-\zeta \right) }{1+\left[ (1+r)\left( 1-\zeta \right) \right] ^{\frac{\eta -1}{\eta }}\chi ^{-\frac{1}{\eta }}}\).

For simplicity we assume that \(\{l_{t}\}\) is i.i.d. Thus, we have

$$\begin{aligned} \hbox {Var}(I_{t})=\frac{c_{12}^{2}\hbox {Var}(l_{t})}{\left( 1-c_{12}^{2}\right) \left[ (1+r)\left( 1-\zeta \right) \right] ^{2}}w^{2}. \end{aligned}$$

The impact of \(\zeta \) on \(c_{12}\) in the individual wealth accumulation Eq. (25) represents the inheritance effect of the estate tax on the stationary wealth distribution. The higher the estate tax \(\zeta \), the lower the \(c_{12}\).Footnote 30 Thus, the inheritance effect of the estate tax increases the long-run wealth inequality. The impact of \(\zeta \) on the lump-sum transfer g in the individual wealth accumulation Eq. (25) represents the redistribution effect of the estate tax.

We can calculate the coefficient of variation,

$$\begin{aligned}&\hbox {CV}(I_{t}) \\= & {} \frac{\sqrt{\hbox {Var}(I_{t})}}{\bar{I}} \\= & {} \frac{1-(1+r)\left( 1-\left[ (1+r)\left( 1-\zeta \right) \right] ^{-\frac{ 1}{\eta }}\chi ^{-\frac{1}{\eta }}\right) }{(1+r)\left( 1-\zeta \right) } \frac{c_{12}}{\sqrt{1-c_{12}^{2}}}\sqrt{\hbox {Var}(l_{t})}. \end{aligned}$$

To illustrate the impact of the estate tax \(\zeta \) on the lump-sum transfer g and that of the estate tax \(\zeta \) on the long-run wealth inequality, we implement a simple calibration exercise. We pick \(\eta =2\), \(\chi =0.8\), \( r=2\), and \(w=1\). We assume that one generation lasts for 30 years. Thus, \( r=1\) corresponds to the annual interest rate of \(3.7\%\). We increase the estate tax \(\zeta \) from 0.1 to 0.5. Figure 2 shows that the higher the estate tax \(\zeta \), the lower the g. Thus, the redistribution effect of the estate tax increases the long-run wealth inequality.

Fig. 2
figure 2

The impact of the estate tax on the transfer

We also investigate the net effect of the estate tax on the long-run wealth inequality. We assume that \(l_{t}\sim U[0,2]\). Thus, \(E(l_{t})=1\) and \( \hbox {Var}(l_{t})=\frac{2}{3}\). Figure 3 shows that the higher the estate tax \( \zeta \), the higher the CV of the long-run wealth inequality.

Fig. 3
figure 3

The impact of the estate tax on the CV

Figure 4 shows that the higher the estate tax \(\zeta \), the higher the Gini coefficient of the long-run wealth inequality. Figures 3 and 4 show that the estate tax increases the long-run wealth inequality. This result is reasonable since both the inheritance effect and the redistribution effect of the estate tax increase the long-run wealth inequality.

Fig. 4
figure 4

The impact of the estate tax on the Gini coefficient

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wan, J., Zhu, S. Bequests, estate taxes, and wealth distributions. Econ Theory 67, 179–210 (2019). https://doi.org/10.1007/s00199-017-1091-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-017-1091-7

Keywords

JEL Classification

Navigation