Skip to main content

Advertisement

Log in

A Becker–Tomes model with investment risk

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

Abstract

Recent studies find that sufficiently volatile idiosyncratic investment risk plays an important role in generating wealth inequality. I introduce idiosyncratic investment risk into the Becker and Tomes (J Polit Econ 87:1153–1189, 1979) model and find an explicit expression of the stationary wealth distribution in this simple model. This explicit expression brings us new insights of how bequest motives and estate taxes influence wealth distributions. I find that inheritance increases wealth inequality in models with idiosyncratic investment risk through exaggerating labor earnings uncertainty, while inheritance decreases wealth inequality in the Becker and Tomes (1979) model through mitigating labor earnings uncertainty. This causes estate taxes to have different impacts on wealth inequality in my model and the Becker and Tomes (1979) model.

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. This large class of studies includes, for example, Quadrini (2000), Cagetti and De Nardi (2006, 2009), Benhabib et al. (2011), Panousi (2012), Shourideh (2014), Achdou et al. (2015), and Benhabib et al. (2015b).

  2. Since my 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.

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

  4. The tax scheme in Pestieau and Possen (1979) has the form

    $$\begin{aligned} S_{A}=pS_{B}^{c}, \end{aligned}$$

    where \(p\ge 1\), \(0\le c<1\). \(S_{A}\) represents the after-tax estate, and \( S_{B}\) the before-tax estate. The lower the value of c, the greater the degree of progressivity of the tax. The constant p is the instrument through which the government returns the tax revenues to the economy.

  5. The mechanisms generating stationary distributions in the two models are also different. Pestieau and Possen (1979) study a progressive property or estate tax. They use the concavity to generate a stationary wealth distribution. And the stationary distribution is lognormal. In this paper, I study a flat estate tax and use a Kesten process to generate a stationary distribution with a Pareto tail.

  6. Benhabib et al. (2015b) generate a stationary wealth distribution with a fat tail in an infinite-horizon model, and their model permits agents to have the precautionary savings motive.

  7. See Benhabib et al. (2015a) for careful quantitative analyses of the Benhabib et al. (2011) model.

  8. To isolate the redistribution effect from their mechanism, Benhabib et al. (2011) intentionally assume that the government wastes collected revenues and does not redistribute them.

  9. In a working paper version, Piketty and Saez (2012, 2013) use a multiplicative random coefficient to generate the fat tail of the wealth distribution.

  10. Modeling a more complicated demographic structure Mierau and Turnovsky (2014) studied the relationship between demography and wealth inequality.

  11. I use \(\{x_{t}\}\) to denote a sequence.

  12. A Markov process \(\left\{ x_{t}\right\} \) is irreducible if there exists a measure \(\varphi \) such that whenever \(\varphi (A)>0\) the process \(\left\{ x_{t}\right\} \) enters the set A in finite time with a positive probability. See page 82 of Meyn and Tweedie (2009).

  13. Davies (1986) uses a mean-reverting process,

    $$\begin{aligned} H_{t}=(1-\omega )\hat{H}+\omega H_{t-1}+\varepsilon _{t}, \end{aligned}$$

    with \(0<\omega <1\). And \(\hat{H}\) is the long-run mean of \(H_{t}\). He assumes that \(\varepsilon _{t}\) is independent of \(H_{t-1}\) and has a zero mean and a constant variance. To ensure \(H_{t}>0\) he assumes that \( \varepsilon _{t}\) is strictly bounded from below by \(-(1-\omega )\hat{H}\). If we furthermore assume that \(\varepsilon _{t}\) is bounded from above and have a continuous density function on its support, such a process \(\{H_{t}\}\) satisfies Assumptions 1 and 2 of my model.

  14. Angeletos (2007) studies the impact of idiosyncratic investment risk on the aggregate capital stock in a neoclassical growth model.

  15. The proof of Theorem 1 could be way simplified if we assume that both \(\{H_{t}\}\) and \(\left\{ \tilde{R}_{t}\right\} \) are i.i.d. along generations.

  16. I relax this assumption in Sect. 7.

  17. I introduce an exogenous economic growth rate into the economy and extend the benchmark model in Sect. 6.1.

  18. Zipf’s law refers to a distribution with an asymptotic Pareto tail of an exponent close to 1.

  19. Let \(F_{X}(x)\) and \(F_{Y}(x)\) be the distribution functions of random variables X and Y, respectively. X first-order stochastically dominates Y, denoted as \(X\succeq _\mathrm{FSD}Y\), if, and only if,

    $$\begin{aligned} F_{X}(x)\le F_{Y}(x) \end{aligned}$$

    for all \(x\in \mathbb {R} \). See page 2 of Müller and Stoyan (2002).

  20. Wan and Zhu (2017) apply this decomposition technique and find that different formulations of bequest motives affect the redistribution effect (the transfer effect) of estate taxes. See discussions in Sect. 5.2.

  21. For a nonnegative random variable Y with a finite positive mean and a constant \(c>0\), Y and cY have the same Lorenz curve, i.e., a Lorenz curve satisfies the scale invariance axiom.

  22. 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 \(X+b\).

  23. See more discussions about the redistribution effect in Sect. 5.2.

  24. See comments in the last paragraph of page 547 of Davies (1986).

  25. See footnote 15 of Davies (1986).

  26. The proofs of the results in this section are quite long. I put them in the online technical appendix (Zhu 2018).

  27. Note that Assumption 4\(^{\prime \prime \prime }\) implies that \((1-b)\chi E\left( \tilde{R}_{t+1}\right) <1\).

  28. See also Benhabib and Zhu (2008).

  29. \( \mathbb {Z} \) denotes the set of integers.

References

  • Achdou, Y., Han, J., Lasry, J., Lions, P., Moll, B.: Heterogeneous Agent Models in Continuous Time. Princeton University, Mimeo (2015)

    Google Scholar 

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

    Article  Google Scholar 

  • Angeletos, G.: Uninsured idiosyncratic investment risk and aggregate saving. Rev. Econ. Dyn. 10, 1–30 (2007)

    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 

  • Aoki, S., Nirei, M.: Zipf’s law, Pareto law, and the evolution of top incomes in the United States. Am. Econ. J. Macroecon. 9, 36–71 (2017)

    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., Zhu, S.: Age, luck, and inheritance. NBER working paper no. 14128 (2008)

  • 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., Luo, M.: Wealth distribution and social mobility in the US: a quantitative approach. NBER working paper no. 21721 (2015a)

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

    Article  Google Scholar 

  • Bossmann, M., Kleiber, C., Wälde, 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)

    Google Scholar 

  • Cagetti, M., De Nardi, M.: Entrepreneurship, frictions, and wealth. J. Polit. Econ. 114, 835–870 (2006)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Cao, D., Luo, W.: Persistent heterogeneous returns and top end wealth inequality. Rev. Econ. Dyn. 26, 301–326 (2017)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Farhi, E., Werning, I.: Progressive estate taxation. Q. J. Econ. 125, 635–673 (2010)

    Article  Google Scholar 

  • Gabaix, X.: Zipf’s law for cities: an explanation. Q. J. Econ. 114, 739–767 (1999)

    Article  Google Scholar 

  • Kaymak, B., Poschke, M.: The evolution of wealth inequality over half a century: the role of taxes, transfers and technology. J. Monet. Econ. 77, 1–25 (2016)

    Article  Google Scholar 

  • Klass, O., Biham, O., Levy, M., Malcai, O., Solomon, S.: The Forbes 400 and the Pareto wealth distribution. Econ. Lett. 90, 290–295 (2006)

    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 

  • Loève, M.: Probability Theory I, 4th edn. Springer, Berlin (1977)

    Book  Google Scholar 

  • Meyn, S., Tweedie, R.: Markov Chains and Stochastic Stability, 2nd edn. Cambridge University Press, Cambridge (2009)

    Book  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 

  • Mirek, M.: Heavy tail phenomenon and convergence to stable laws for iterated Lipschitz maps. Probab. Theory Relat. Fields 151, 705–734 (2011)

    Article  Google Scholar 

  • Müller, A., Stoyan, D.: Comparison Methods for Stochastic Models and Risks. Wiley, Hoboken (2002)

    Google Scholar 

  • Ok, E.: Probability Theory with Economic Applications. New York University, Mimeo (2016)

    Google Scholar 

  • Panousi, V.: Capital Taxation with Entrepreneurial Risk. Federal Reserve Board, Mimeo (2012)

    Google Scholar 

  • Panousi, V., Reis, C.: Optimal Capital Taxation with Idiosyncratic Investment Risk. Federal Reserve Board, Mimeo (2015)

    Google Scholar 

  • Pestieau, P., Possen, U.: A model of wealth distribution. Econometrica 47, 761–772 (1979)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Piketty, T.: Capital in the Twenty-First Century. The Belknap Press of Harvard University Press, Cambridge (2014)

    Book  Google Scholar 

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

    Article  Google Scholar 

  • Piketty, T., Saez, E.: A theory of optimal capital taxation. NBER working paper no. 17989 (2012)

  • Piketty, T., Saez, E.: A theory of optimal inheritance taxation. Econometrica 81, 1851–1886 (2013)

    Article  Google Scholar 

  • Quadrini, V.: Entrepreneurship, saving, and social mobility. Rev. Econ. Dyn. 3, 1–40 (2000)

    Article  Google Scholar 

  • Rointershtein, A.: One-dimensional linear recursions with Markov-dependent coefficients. Ann. Appl. Probab. 17, 572–608 (2007)

    Article  Google Scholar 

  • Royden, H.L.: Real Analysis, 3rd edn. Prentice-Hall Inc., Upper Saddle River (1988)

    Google Scholar 

  • Saez, E., Zucman, G.: Wealth inequality in the United States since 1913: evidence from capitalized income tax data. Q. J. Econ. 131, 519–578 (2016)

    Article  Google Scholar 

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

    Google Scholar 

  • Shourideh, A.: Optimal Taxation of Wealthy Individuals. University of Pennsylvnia, Mimeo (2014)

    Google Scholar 

  • Wan, J., Zhu, S.: Bequests, estate taxes, and wealth distributions. Econ. Theory (2017). https://doi.org/10.1007/s00199-017-1091-7

  • Zhu, S.: Technical appendix of unobservable \(\tilde{R}_{t+1}\): Appendix A.9 of “A Becker–Tomes model with investment risk.” University of International Business and Economics, Online Appendix (2018)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shenghao Zhu.

Additional information

I would like to thank Jess Benhabib, Alberto Bisin, Basant Kapur, Tomoo Kikuchi, Vincenzo Quadrini, Danny Quah, Thomas Sargent, Jing Wan, C.C. Yang, Hanqin Zhang, Jie Zhang, and seminar participants at ECINEQ 2013 and SAET 2013.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 155 KB)

Appendix

Appendix

1.1 The tail of the stationary wealth distribution

In the steady state of the aggregate economy, Eq. (5) implies

$$\begin{aligned} L_{t+1}=d_{t+1}L_{t}+\eta _{t+1}, \end{aligned}$$
(A.1)

where

$$\begin{aligned} d_{t+1}=\frac{\tilde{R}_{t+1}(1-b)}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{ \frac{\gamma -1}{\gamma }}\chi ^{-\frac{1}{\gamma }}}, \end{aligned}$$
(A.2)

and

$$\begin{aligned} \eta _{t+1}=\frac{1}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{\frac{\gamma -1}{ \gamma }}\chi ^{-\frac{1}{\gamma }}}\left( H_{t+1}+G\right) . \end{aligned}$$
(A.3)

In order to investigate the stationary wealth distribution with serially correlated \(\{H_{t}\}\) and \(\{\tilde{R}_{t}\}\), we need the following definition.

Definition 2

Let \((\ell ,\mathcal {F})\) be a measurable space and let \(\left\{ x_{n}\right\} \) be a stationary Markov process with transition kernel \( Q(x,\cdot )\) defined on it. A Markov-modulated process (MMP) associated with \(\left\{ x_{n}\right\} \) is a stationary Markov process \(\left\{ (x_{n},\zeta _{n})\right\} \) defined on a product space \((\ell \times \Upsilon ,\mathcal {F}\otimes \Xi )\), whose transitions depend only on the position of \(x_{n}\). That is, for any \(n\ge 0\), \(A\in \mathcal {F}\), \(B\in \Xi \),

$$\begin{aligned} Pr \left( x_{n}\in A,\zeta _{n}\in B\mid \sigma \left( (x_{i},\zeta _{i}):i<n\right) \right) =\int _{A}Q(x,dy)\Gamma (x,y,B)|_{x=x_{n-1}}, \end{aligned}$$

where \(\Gamma (x,y,\cdot )=Pr (\zeta _{1}\in \cdot \mid x_{0}=0,x_{1}=y)\) is a kernel on \((\ell \times \ell \times \Xi )\).

Lemma 1

Let

$$\begin{aligned} m(x)=\log E(d_{t+1})^{x}. \end{aligned}$$

Then m(x) is a convex function of \(x>0\).

Proof

See page 158 of Loève (1977). \(\square \)

Lemma 2

m(x) is a continuous function of \(x>0\).

Proof

By Proposition 17 of Chapter 5 in Royden (1988), Lemma 1 implies Lemma 2. \(\square \)

Proof of Theorem 1

Note that the process \(\left\{ (H_{t},v_{t})\right\} \), where \(v_{t}=\left( d_{t},\eta _{t}\right) \), is a Markov-modulated process associated with \(\{H_{t}\}\).

In order to apply Theorem 1.5 of Rointershtein (2007) to the process \( \{L_{t}\}\), we will verify (A1)–(A7) of Assumption 1.2 in Rointershtein (2007).

(A1) is obviously satisfied since the Borel sigma-algebra is countably generated.

By Assumption 1, \(\{H_{t}\}\) is irreducible. Thus (A2) is satisfied.

By Assumption 2, we have

$$\begin{aligned} Pr \left( H_{t+1}\le h\mid H_{t}=x\right) =\int _{0}^{h}f(x,y)\hbox {d}y=\int _{0}^{h}\bar{H}f(x,y)\frac{1}{\bar{H}}\hbox {d}y, \end{aligned}$$

for \(h\in (0,\bar{H})\). Let \(\mu ^{Leb}\) be the Lebesgue measure. We construct a probability measure \(\lambda \) on \((0,\bar{H})\), such that \( \lambda (A)=\frac{1}{\bar{H}}\mu ^{Leb}(A)\) for any Borel set A. Since f(xy) is uniformly bounded above on \((0,\bar{H})\times (0,\bar{H})\), \(\bar{ H}f(x,y)\) is also uniformly bounded above on \((0,\bar{H})\times (0,\bar{H})\) . Thus the family of functions \(\{\bar{H}f(x,\cdot ):(0,\bar{H})\rightarrow [0,\infty )\}_{x\in (0,\bar{H})}\) is uniformly integrable with respect to the measure \(\lambda \). Then (A3) is satisfied for \(m_{1}=1\) and the measure \(\lambda \) we construct.

From Assumption 2, we know that \(Pr (H_{t+1}\in (0,\bar{H}))=1\). And from Assumption 4 we know that \(Pr (\tilde{R}_{t+1}\in [\) Ṟ,\(\bar{ R}])=1\). Thus from Eq. (A.3) we know that there exists an \( \bar{\eta }>0\) such that \(Pr (\eta _{t+1}<\bar{\eta })=1\). Thus (A4) is satisfied.

From Assumption 4, we know that \(Pr (\tilde{R}_{t+1}\in [\) Ṟ, \(\bar{R}])=1\). Thus \(d_{t+1}\) is bounded, since \(d_{t+1}\) is a continuous function of \(\tilde{R}_{t+1}\) [see Eq. (A.2)]. We also know that \(d_{t+1}\) is bounded away from zero since Ṟ\(>0\). Thus there exists an \( c_{\rho }>1\) such that \(Pr (\frac{1}{c_{\rho }}<d_{t+1}<c_{\rho })=1\). Thus (A5) is satisfied.

For \(x>0\), we have

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{1}{n}\log E\left[ \prod \limits _{i=1}^{n-1} \left( d_{i}\right) ^{x}\right] =\log E(d_{t+1})^{x}=m(x), \end{aligned}$$

since, by Assumption 3, \(\left\{ \tilde{R}_{t}\right\} \) is i.i.d. along generations. From Lemmas 1 and 2, we know that m(x) is convex and continuous. From Assumption 5, we have \(E\left( d_{t+1}\right) <1\). Thus \(m(1)<0\). By Assumption 6, \(E(d_{t+1})^{2}>1\). Thus \(m(2)>0\). Then we know that there exists a unique \(\mu \in (1,2)\) such that \(m(\mu )=0\), i.e.

$$\begin{aligned} E(d_{t+1})^{\mu }=1. \end{aligned}$$

Thus (A6) is satisfied.

By Assumption 4, \(\tilde{R}_{t+1}\) has a probability density function \( l(\cdot )\) on [Ṟ,\(\bar{R}]\). Thus the distribution of \(\log d_{t+1}\) is nonarithmetic. Thus (A7) is satisfied.

We have verified (A1)–(A7) of Assumption 1.2 in Rointershtein (2007). Applying Theorem 1.5 of Rointershtein (2007) to the process \(\{L_{t}\}\), we have

$$\begin{aligned} \lim _{x\rightarrow \infty }\frac{Pr (L_{\infty }>x)}{x^{-\mu }}=c, \end{aligned}$$

with \(c>0\). \(\square \)

1.2 Proof of Theorem 2

Proof

Let \(\mu ^{\prime }\in (1,2)\) solves

$$\begin{aligned} E(d_{t+1}^{\prime })^{\mu ^{\prime }}=1. \end{aligned}$$

Since \(d_{t+1}\succeq _\mathrm{FSD}d_{t+1}^{\prime }\) and \(f(d)=(d)^{\mu ^{\prime }} \) is an increasing function of d, applying Theorem 1.2.8 of Müller and Stoyan (2002), we have

$$\begin{aligned} E(d_{t+1})^{\mu ^{\prime }}\ge E(d_{t+1}^{\prime })^{\mu ^{\prime }}=1. \end{aligned}$$

Thus

$$\begin{aligned} \log E(d_{t+1})^{\mu ^{\prime }}\ge 0. \end{aligned}$$

By Assumption 5, we have \(\log E(d_{t+1})<0\). From Lemmas 1 and 2, we know that \(\log E(d_{t+1})^{x}\) is convex and continuous in x. Thus there exists a \(\mu >1\) such that \(\log E\left( d_{t+1}\right) ^{\mu }=0 \), and \(\mu \le \mu ^{\prime }\). \(\square \)

1.3 Proof of Proposition 2

Proof

Suppose that \(\chi >\chi ^{\prime }\). Thus by Eq. (A.2) we know that

$$\begin{aligned} d_{t+1}=\frac{\tilde{R}_{t+1}(1-b)}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{ \frac{\gamma -1}{\gamma }}\chi ^{-\frac{1}{\gamma }}}>d_{t+1}^{\prime }= \frac{\tilde{R}_{t+1}(1-b)}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{\frac{ \gamma -1}{\gamma }}\left( \chi ^{\prime }\right) ^{-\frac{1}{\gamma }}}. \end{aligned}$$

For \(x\in \mathbb {R} \), \(d_{t+1}\le x\) implies \(d_{t+1}^{\prime }\le x\). Thus we have

$$\begin{aligned} Pr (d_{t+1}\le x)\le Pr (d_{t+1}^{\prime }\le x), \end{aligned}$$

for \(x\in \mathbb {R} \). Then we know that \(d_{t+1}\succeq _\mathrm{FSD}d_{t+1}^{\prime }\). Applying Theorem 2, we know that the Pareto exponent \(\mu \) of the wealth distribution under \(\chi \) is smaller than under \(\chi ^{\prime }\). \(\square \)

1.4 Proof of Proposition 4

Proof

By Eq. (A.2), we have

$$\begin{aligned} d_{t+1}= & {} \frac{\tilde{R}_{t+1}(1-b)}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{\frac{\gamma -1}{\gamma }}\chi ^{-\frac{1}{\gamma }}} \\= & {} \frac{1}{\left[ \tilde{R}_{t+1}(1-b)\right] ^{-1}+\left[ \tilde{R} _{t+1}(1-b)\right] ^{-\frac{1}{\gamma }}\chi ^{-\frac{1}{\gamma }}}. \end{aligned}$$

Thus \(d_{t+1}\) decreases with b.

Suppose that \(b<b^{\prime }\). Thus we have \(d_{t+1}>d_{t+1}^{\prime }\). For \( x\in \mathbb {R} \), \(d_{t+1}\le x\) implies \(d_{t+1}^{\prime }\le x\). Thus we have

$$\begin{aligned} Pr (d_{t+1}\le x)\le Pr (d_{t+1}^{\prime }\le x), \end{aligned}$$

for \(x\in \mathbb {R} \). Then we know that \(d_{t+1}\succeq _\mathrm{FSD}d_{t+1}^{\prime }\). Applying Theorem 2, we know that the Pareto exponent \(\mu \) of the wealth distribution under b is smaller than under \(b^{\prime }\). \(\square \)

1.5 Proof of Proposition 5

Proof

Suppose that \(g<g^{\prime }\). Thus we have \(\frac{d_{t+1}}{g}>\frac{ d_{t+1}}{g^{\prime }}\). For \(x\in \mathbb {R} \), \(\frac{d_{t+1}}{g}\le x\) implies \(\frac{d_{t+1}}{g^{\prime }}\le x\). Thus we have

$$\begin{aligned} Pr \left( \frac{d_{t+1}}{g}\le x\right) \le Pr \left( \frac{d_{t+1}}{ g^{\prime }}\le x\right) , \end{aligned}$$

for \(x\in \mathbb {R} \). Then we know that \(\frac{d_{t+1}}{g}\succeq _\mathrm{FSD}\frac{d_{t+1}}{ g^{\prime }}\). Applying Theorem 2 to the process \(\left\{ \hat{L} _{t}\right\} \), we know that the Pareto exponent \(\mu \) of the wealth distribution under g is smaller than under \(g^{\prime }\). \(\square \)

1.6 Serially correlated \(\left\{ \tilde{R}_{t}\right\} \)

We introduce a stationary Markov process \(\left\{ x_{t}\right\} \) into the model such that the process \(\left\{ (x_{t},\psi _{t})\right\} \), where \( \psi _{t}=\left( \tilde{R}_{t},H_{t}\right) \), is a Markov-modulated process associated with \(\{x_{t}\}\). Thus the Markov process \(\left\{ x_{t}\right\} \) is the underlying process.

Assumption 1\(^{\prime }\).\(\{x_{t}\}\) is on the measurable space \(( \mathbb {R} ,\mathbf {B})\), where \(\mathbf {B}\) is the Borel sigma-algebra.

Assumption 2\(^{\prime }\).\(\{x_{t}\}\) is irreducible.

Assumption 3\(^{\prime }\). Let \(Q(x,\cdot )\) be the transition kernel of \(\{x_{t}\}\). There exist a probability measure \(\lambda \) on \(( \mathbb {R} ,\mathbf {B})\), a number \(m_{1}\in \mathbb {N} \), and a measurable density kernel \(f(x,y): \mathbb {R} ^{2}\rightarrow [0,\infty )\) such that

$$\begin{aligned} Q^{m_{1}}(x,A)=\int _{A}f(x,y)\lambda (dy), \end{aligned}$$

and the family of functions \(\{f(x,\cdot ): \mathbb {R} \rightarrow [0,\infty )\}_{x\in \mathbb {R} }\) is uniformly integrable with respect to the measure \(\lambda \).

Assumption 4\(^{\prime }\).\(H_{t}\in (0,\bar{H})\).

Assumption 5\(^{\prime }\).\(\tilde{R}_{t}\in [\)\( ,\bar{R}]\) with Ṟ\(>0\).

Assumption 6\(^{\prime }\). Let \(\Lambda (x)=\lim \sup _{n\rightarrow \infty }\frac{1}{n}\log E\left[ \prod \nolimits _{i=1}^{n-1} \left( d_{i}\right) ^{x}\right] \). There exist \(\mu _{1}>1\) and \(\mu _{2}>1\) such that \(\Lambda (\mu _{1})\ge 0\) and \(\Lambda (\mu _{2})<0\).

Assumption 7\(^{\prime }\). There do not exist a constant \( \alpha >0\) and a measurable function \(\beta : \mathbb {R} \rightarrow [0,\alpha )\) such thatFootnote 29

$$\begin{aligned} Pr \left( \log (d_{1})\in \beta (x_{0})-\beta (x_{1})+\alpha \mathbb {Z} \right) =1. \end{aligned}$$

Proof of Proposition 7

The process \(\left\{ (x_{t},v_{t})\right\} \), where \(v_{t}=\left( d_{t},\eta _{t}\right) \), is a Markov-modulated process associated with \(\{x_{t}\}\) since the process \( \left\{ (x_{t},\psi _{t})\right\} \), where \(\psi _{t}=\left( \tilde{R} _{t},H_{t}\right) \), is a Markov-modulated process associated with \( \{x_{t}\} \).

In order to apply Theorem 1.5 of Rointershtein (2007) to the process \( \{L_{t}\}\), we will verify (A1)–(A7) of Assumption 1.2 in Rointershtein (2007).

(A1) is obviously satisfied since the Borel sigma-algebra is countably generated.

By Assumptions 2\(^{\prime }\) and 3\(^{\prime }\), (A2) and (A3) are satisfied.

From Assumption 4\(^{\prime }\), we know that \(Pr (H_{t+1}\in (0,\bar{H}))=1\). And from Assumption 5\(^{\prime }\) we know that \(Pr (\tilde{R}_{t+1}\in [\) Ṟ,\(\bar{R}])=1\). Thus from Eq. (A.3) we know that there exists an \(\bar{\eta }>0\) such that \(Pr (\eta _{t+1}<\bar{\eta } )=1 \). Thus (A4) is satisfied.

From Assumption 5\(^{\prime }\), we know that \(Pr (\tilde{R}_{t+1}\in [\) Ṟ, \(\bar{R}])=1\). Thus \(d_{t+1}\) is bounded, since \(d_{t+1}\) is a continuous function of \(\tilde{R}_{t+1}\) [see Eq. (A.2)]. We also know that \(d_{t+1}\) is bounded away from zero since Ṟ\(>0\). Thus there exists an \(c_{\rho }>1\) such that \(Pr (\frac{1}{c_{\rho }}<d_{t+1}<c_{\rho })=1\). Thus (A5) is satisfied.

By Assumptions 6\(^{\prime }\) and 7\(^{\prime }\), (A6) and (A7) are satisfied.

We have verified (A1)–(A7) of Assumption 1.2 in Rointershtein (2007). By Lemma 2.3 of Rointershtein (2007) we know that, for \(x>0\), the following limit exists,

$$\begin{aligned} \Lambda (x)=\lim _{n\rightarrow \infty }\frac{1}{n}\log E\left[ \prod \limits _{i=1}^{n-1}\left( d_{i}\right) ^{x}\right] . \end{aligned}$$

We then show the following lemma.

Lemma 3

\(\Lambda (x)\) is a convex function of \(x>0\).

Proof

Note that \(\log E\left[ \prod \nolimits _{i=1}^{n-1}\left( d_{i}\right) ^{x}\right] =\log E\left[ \left( \prod \nolimits _{i=1}^{n-1}d_{i}\right) ^{x} \right] \). Viewing \(\left( \prod \nolimits _{i=1}^{n-1}d_{i}\right) \) as a random variable, we know, from Lemma 1, that \(\log E\left[ \left( \prod \nolimits _{i=1}^{n-1}d_{i}\right) ^{x}\right] \) is a convex function of \( x>0\). Thus \(\frac{1}{n}\log E\left[ \prod \nolimits _{i=1}^{n-1}\left( d_{i}\right) ^{x}\right] \) is a convex function of \(x>0\). Then we know that \( \Lambda (x)\) is a convex function of \(x>0\). \(\square \)

Thus we have Lemma 4 as a corollary to Lemma 3.

Lemma 4

\(\Lambda (x)\) is a continuous function of \(x>0\).

Proof

By Proposition 17 of Chapter 5 in Royden (1988), Lemma 3 implies Lemma 4. \(\square \)

Lemmas 3 and 4 imply that \(\Lambda (x)\) is convex and continuous. From Assumption 6\(^{\prime }\), we know that there exist \(\mu _{1}>1\) and \(\mu _{2}>1\) such that \(\Lambda (\mu _{1})\ge 0\) and \(\Lambda (\mu _{2})<0\). Thus there exists a unique \(\mu >1\) such that

$$\begin{aligned} \Lambda (\mu )=0. \end{aligned}$$

Applying Theorem 1.5 of Rointershtein (2007) to the process \(\{L_{t}\}\), we have

$$\begin{aligned} \lim _{x\rightarrow \infty }\frac{Pr (L_{\infty }>x)}{x^{-\mu }}=c, \end{aligned}$$

with \(c>0\). \(\square \)

1.7 Proof of Proposition 8

Proof

Suppose that \(\chi >\chi ^{\prime }\). Thus by Eq. (A.2) we know that

$$\begin{aligned} d_{t+1}=\frac{\tilde{R}_{t+1}(1-b)}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{ \frac{\gamma -1}{\gamma }}\chi ^{-\frac{1}{\gamma }}}>d_{t+1}^{\prime }= \frac{\tilde{R}_{t+1}(1-b)}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{\frac{ \gamma -1}{\gamma }}\left( \chi ^{\prime }\right) ^{-\frac{1}{\gamma }}}. \end{aligned}$$

Let

$$\begin{aligned} \Lambda _{1}(x)=\lim _{n\rightarrow \infty }\frac{1}{n}\log E\left[ \prod \limits _{i=1}^{n-1}\left( d_{i}^{\prime }\right) ^{x}\right] , \end{aligned}$$

for \(x>0\). Suppose that \(\Lambda _{1}(\mu ^{\prime })=0\). Thus we have

$$\begin{aligned} \Lambda (\mu ^{\prime })=\lim _{n\rightarrow \infty }\frac{1}{n}\log E\left[ \prod \limits _{i=1}^{n-1}\left( d_{i}\right) ^{\mu ^{\prime }}\right] \ge \lim _{n\rightarrow \infty }\frac{1}{n}\log E\left[ \prod \limits _{i=1}^{n-1} \left( d_{i}^{\prime }\right) ^{\mu ^{\prime }}\right] =0. \end{aligned}$$

From Assumption 6\(^{\prime }\), we know that there exists \(\mu _{2}>1\) such that \(\Lambda (\mu _{2})<0\). Then we know that there exists a \(\mu >1\) such that \(\Lambda (\mu )=0\), and \(\mu \le \mu ^{\prime }\). \(\square \)

1.8 Proof of Proposition 9

Proof

I apply Theorem 1.8 of Mirek (2011) to prove Proposition 9 . I need to verify Assumptions 1.6 and 1.7 of Mirek (2011).

Let \(\vartheta =\left( \tilde{R},H\right) \) and

$$\begin{aligned} \psi _{\vartheta }(L)=\left\{ \begin{array}{ll} H-(1-b)\tilde{R}\theta L+G, &{} \text { if }L\le \phi \\ dL+\eta , &{} \hbox {otherwise} \end{array} \right. . \end{aligned}$$

The process \(\left\{ L_{t}\right\} \) is generated by \(L_{t+1}=\psi _{\vartheta }(L_{t})\). Thus \(\psi _{\vartheta }(L)\) is Lipschitz continuous.

Verification of Assumption 1.6 of Mirek (2011). For every \(z>0\), let

$$\begin{aligned} \psi _{\vartheta ,z}(L)=z\psi _{\vartheta }\left( \frac{1}{z}L\right) . \end{aligned}$$

\(\psi _{\vartheta ,z}\) are called dilatations of \(\psi _{\vartheta }\). Let

$$\begin{aligned} \bar{\psi }_{\vartheta }(L)=\lim _{z\rightarrow 0}\psi _{\vartheta ,z}(L). \end{aligned}$$

Thus we have

$$\begin{aligned} \bar{\psi }_{\vartheta }(L)=\lim _{z\rightarrow 0}\psi _{\vartheta ,z}(L)=\lim _{z\rightarrow 0}\left[ z\psi _{\vartheta }\left( \frac{1}{z} L\right) \right] =dL,\text { for }\forall L\ge 0. \end{aligned}$$

Since \(\psi _{\vartheta }(L)\) is piecewise linear, It is easy to find a random variable \(N_{\vartheta }\) with bounded support such that

$$\begin{aligned} |\psi _{\vartheta }(L)-dL|\le N_{\vartheta },\text { for }\forall L\ge 0. \end{aligned}$$

Assumption 1.6 of Mirek (2011) is satisfied.

Verification of Assumption 1.7 of Mirek (2011). As for Assumption 1.7 of Mirek (2011), condition (H3) is satisfied since

$$\begin{aligned} d_{t+1}=\frac{\tilde{R}_{t+1}(1-b)}{1+\left[ \tilde{R}_{t+1}(1-b)\right] ^{ \frac{\gamma -1}{\gamma }}\chi ^{-\frac{1}{\gamma }}}, \end{aligned}$$

by Eq. (A.2). \(\left\{ \tilde{R}_{t}\right\} \) is i.i.d. along time and the support of \(\tilde{R}_{t}\) is closed.

The law of \(\log d\) is nonarithmetic since \(\tilde{R}_{t}\) has a probability density function \(l(\cdot )\) on [Ṟ,\(\bar{R}]\) by Assumption 3\(^{\prime \prime }\). Thus (H4) in Assumption 1.7 of Mirek (2011) is satisfied.

Let \(m(x)=\log E(d)^{x}\). From Assumption 4\(^{\prime \prime }\), we have \( E\left( d\right) <1\). Thus \(m(1)<0\). By Assumption 5\(^{\prime \prime }\), \( E(d)^{2}>1\). Thus \(m(2)>0\). From Lemmas 1 and 2, we know that m(x) is convex and continuous. Thus there exists a unique \(\mu \in (1,2)\) such that \(m(\mu )=0\), i.e.

$$\begin{aligned} E(d)^{\mu }=1. \end{aligned}$$

We also know that \(E\left( d^{\mu }|\log d|\right) <\infty \), since d is bounded.

\(E\left[ \left( N_{\vartheta }\right) ^{\mu }\right] <\infty \), since \( N_{\vartheta }\) is bounded.

Assumption 1.7 of Mirek (2011) is satisfied.

Applying Theorem 1.8 of Mirek (2011), we find that the stationary distribution of the process \(\left\{ L_{t}\right\} \), \(L_{\infty }\), has an asymptotic Pareto tail, i.e.

$$\begin{aligned} \lim _{x\rightarrow \infty }\frac{Pr (L_{\infty }>x)}{x^{-\mu }}=c, \end{aligned}$$

with \(c>0\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhu, S. A Becker–Tomes model with investment risk. Econ Theory 67, 951–981 (2019). https://doi.org/10.1007/s00199-018-1103-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-018-1103-2

Keywords

JEL Classification

Navigation