Skip to main content
Log in

Asset prices in segmented and integrated markets

  • Published:
Finance and Stochastics Aims and scope Submit manuscript

Abstract

This paper evaluates the effect of market integration on prices and welfare, in a model where two Lucas trees grow in separate regions with similar investors. We find equilibrium asset price dynamics and welfare both in segmentation, when each region holds its own asset and consumes its dividend, and in integration, when both regions trade both assets and consume both dividends. Integration always increases welfare. Asset prices may increase or decrease, depending on the time of integration, but decrease on average. Cross-asset correlation is zero or negative before integration, but significantly positive afterwards, explaining some effects commonly associated with financialisation.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. As Foerster and Karolyi [26] surmise, “International asset pricing models suggest that when investors realise that barriers to investments are to be removed, expected returns should decrease as prices are bid up on the expectation of the removal of these barriers”. For partial equilibrium models of international asset pricing, see also Solnik [41, 42], Stulz [45, 44], Adler and Dumas [2].

  2. It is trivial to obtain uncorrelated returns with power utility by considering two segmented markets with independent Lucas trees. However, in such a model, one market eventually overtakes the other, violating cointegration.

  3. With logarithmic preferences, see the integrated market of Cochrane et al. [16] and the partially integrated markets of Pavlova and Rigobon [38], Bhamra et al. [7]. For power utility, see Martin [33].

  4. This assumption is also a reasonable approximation to the statistical properties of dividends across large economic areas. For example, in the period 1991–2018, the correlation among dividend growth between large-capitalisation stocks in the United States, Europe and Japan were 0.25 (US–EU), −0.07 (US–JP) and −0.11 (EU–JP), as estimated from the annual dividends paid by the index funds VFINX, VEUREX and VPACX which track their respective markets.

  5. Under Assumption 3.1, the process \(X\) exists globally and never reaches zero or one, i.e., we have \(P[X_{t}\in (0,1)\), \(\forall t\ge 0 ]=1\). See for example Borodin and Salminen [9, No. 6, Chap. II.1].

  6. The case \(\gamma =1\) corresponds to logarithmic utility, which leads to simpler but slightly different calculations, for brevity not included here.

  7. Here the maximisation takes place over \(c\in \mathcal{C}\) and \(\phi = (\phi ^{(1)}, \phi ^{(2)}) \in \overline{\mathcal{P}}\), as the integrated market includes two assets. See Appendix A.1 for details.

References

  1. Abramowitz, M., Stegun, I.A.: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. Dover, New York (1965)

    MATH  Google Scholar 

  2. Adler, M., Dumas, B.: International portfolio choice and corporation finance: a synthesis. Journal of Finance 38, 925–984 (1983)

    Article  Google Scholar 

  3. Banner, A.D., Fernholz, R., Karatzas, I.: Atlas models of equity markets. Ann. Appl. Probab. 15, 2296–2330 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  4. Basak, S., Cuoco, D.: An equilibrium model with restricted stock market participation. Rev. Financ. Stud. 11, 309–341 (1998)

    Article  Google Scholar 

  5. Basak, S., Pavlova, A.: A model of financialization of commodities. Journal of Finance 71, 1511–1556 (2016)

    Article  Google Scholar 

  6. Bekaert, G., Harvey, C.R.: Emerging equity market volatility. J. Financ. Econ. 43, 29–77 (1997)

    Article  Google Scholar 

  7. Bhamra, H.S., Coeurdacier, N., Guibaud, S.: A dynamic equilibrium model of imperfectly integrated financial markets. J. Econ. Theory 154, 490–542 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bhardwaj, G., Gorton, G., Rouwenhorst, G.: Facts and fantasies about commodity futures ten years later. Tech. rep., National Bureau of Economic Research (2015). Available online at https://www.nber.org/papers/w21243.pdf

  9. Borodin, A.N., Salminen, P.: Handbook of Brownian Motion – Facts and Formulae. Birkhäuser, Basel (2012)

    MATH  Google Scholar 

  10. Brown, G., Sarkozy, N.: Oil prices need government supervision. Wall St. J. (December 9, 2009). Available online at https://www.wsj.com/articles/SB124701217125708963

  11. Buraschi, A., Trojani, F., Vedolin, A.: When uncertainty blows in the orchard: comovement and equilibrium volatility risk premia. Journal of Finance 69, 101–137 (2014)

    Article  Google Scholar 

  12. Carmona, R.: Financialization of the commodities markets: a non-technical introduction. In: Aïd, R., et al. (eds.) Commodities, Energy and Environmental Finance, pp. 3–37. Springer, Berlin (2015)

    Chapter  Google Scholar 

  13. Chabakauri, G.: Dynamic equilibrium with two stocks, heterogeneous investors, and portfolio constraints. Rev. Financ. Stud. 26, 3104–3141 (2013)

    Article  Google Scholar 

  14. Chan, P., Sircar, R., Stein, M.V.: A feedback model for the financialization of commodity markets. SIAM J. Financ. Math. 6, 870–899 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  15. Cochrane, J.H.: The dog that did not bark: a defense of return predictability. Rev. Financ. Stud. 21, 1533–1575 (2008)

    Article  Google Scholar 

  16. Cochrane, J.H., Longstaff, F.A., Santa-Clara, P.: Two trees. Rev. Financ. Stud. 21, 347–385 (2007)

    Article  Google Scholar 

  17. Cont, R., Wagalath, L.: Institutional investors and the dependence structure of asset returns. Int. J. Theor. Appl. Finance 19, 1650010 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  18. Cuchiero, C.: Polynomial processes in stochastic portfolio theory. In: Stochastic Processes and Their Applications, vol. 129, pp. 1829–1872 (2019)

    MATH  Google Scholar 

  19. Dumas, B.: Dynamic equilibrium and the real exchange rate in a spatially separated world. Rev. Financ. Stud. 5, 153–180 (1992)

    Article  Google Scholar 

  20. Dumas, B., Harvey, C.R., Ruiz, P.: Are correlations of stock returns justified by subsequent changes in national outputs? J. Int. Money Financ. 22, 777–811 (2003)

    Article  Google Scholar 

  21. Farhi, E., Gabaix, X.: Rare disasters and exchange rates. Q. J. Econ. 131, 1–52 (2015)

    Article  MATH  Google Scholar 

  22. Fernholz, R., Karatzas, I.: Relative arbitrage in volatility-stabilized markets. Ann. Finance 1, 149–177 (2005)

    Article  MATH  Google Scholar 

  23. Filipović, D., Larsson, M.: Polynomial diffusions and applications in finance. Finance Stoch. 20, 931–972 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  24. Filipović, D., Larsson, M.: Polynomial jump-diffusion models. Stoch. Syst. 10, 71–97 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  25. Filipović, D., Larsson, M., Trolle, A.B.: Linear-rational term structure models. Journal of Finance 72, 655–704 (2017)

    Article  Google Scholar 

  26. Foerster, S.R., Karolyi, G.A.: The effects of market segmentation and investor recognition on asset prices: evidence from foreign stocks listing in the United States. Journal of Finance 54, 981–1013 (1999)

    Article  Google Scholar 

  27. Goetzmann, W.N., Li, L., Rouwenhorst, K.G.: Long-term global market correlations. J. Bus. 78, 1–38 (2005)

    Article  Google Scholar 

  28. Gordon, M.J., Shapiro, E.: Capital equipment analysis: the required rate of profit. Manag. Sci. 3, 102–110 (1956)

    Article  Google Scholar 

  29. Hansen, S.L.: Cross-sectional asset pricing with heterogeneous preferences and beliefs. J. Econ. Dyn. Control 58, 125–151 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  30. Hurd, T.R., Kuznetsov, A.: Explicit formulas for Laplace transforms of stochastic integrals. Markov Process. Relat. Fields 14, 277–290 (2008)

    MathSciNet  MATH  Google Scholar 

  31. Karatzas, I., Lehoczky, J.P., Shreve, S.E., Xu, G.L.: Martingale and duality methods for utility maximization in an incomplete market. SIAM J. Control Optim. 29, 702–730 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  32. Lucas, R.E. Jr.: Asset prices in an exchange economy. Econometrica 46, 1429–1445 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  33. Martin, I.: The Lucas orchard. Econometrica 81, 55–111 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  34. Masters, M.W.: Testimony before the committee on homeland security and governmental affairs. US Senate, Washington, May 20 (2008). Available online at https://www.hsgac.senate.gov/imo/media/doc/052008Masters.pdf

  35. Menzly, L., Santos, T., Veronesi, P.: Understanding predictability. J. Polit. Econ. 112, 1–47 (2004)

    Article  Google Scholar 

  36. Merton, R.C.: A simple model of capital market equilibrium with incomplete information. Journal of Finance 42, 483–510 (1987)

    Article  Google Scholar 

  37. Pal, S.: Analysis of market weights under volatility-stabilized market models. Ann. Appl. Probab. 21, 1180–1213 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  38. Pavlova, A., Rigobon, R.: Asset prices and exchange rates. Rev. Financ. Stud. 20, 1139–1180 (2007)

    Article  Google Scholar 

  39. Santos, T., Veronesi, P.: Labor income and predictable stock returns. Rev. Financ. Stud. 19, 1–44 (2006)

    Article  Google Scholar 

  40. Santos, T., Veronesi, P.: Habit formation, the cross section of stock returns and the cash-flow risk puzzle. J. Financ. Econ. 98, 385–413 (2010)

    Article  Google Scholar 

  41. Solnik, B.H.: An equilibrium model of the international capital market. J. Econ. Theory 8, 500–524 (1974)

    Article  MathSciNet  Google Scholar 

  42. Solnik, B.H.: The international pricing of risk: an empirical investigation of the world capital market structure. Journal of Finance 29, 365–378 (1974)

    Article  Google Scholar 

  43. Solnik, B.H., Boucrelle, C., Le Fur, , Y.: International market correlation and volatility. Financ. Anal. J. 52(5), 17–34 (1996)

    Article  Google Scholar 

  44. Stulz, R.M.: A model of international asset pricing. J. Financ. Econ. 9, 383–406 (1981)

    Article  Google Scholar 

  45. Stulz, R.M.: On the effects of barriers to international investment. Journal of Finance 36, 923–934 (1981)

    Article  Google Scholar 

  46. Tang, K., Xiong, W.: Index investment and the financialization of commodities. Financ. Anal. J. 68(5), 54–74 (2012)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

For helpful comments, we thank Michael Brennan, Peter Bank, Jan Kallsen, Kasper Larsen, Andrea Roncoroni, Matthew Spiegel, Andrea Vedolin and seminar participants at Bar-Ilan University, Bȩdlewo, CEMA, CIRM, ESSEC, Econometric Society Meeting, MACSI, University of Pisa, Rice University, Shanghai Jiao Tong, UCC and UCD Smurfit. We are very grateful to the anonymous referee and the Associate Editor, who helped to improve the paper and to better place it in context. Special thanks to Silvia Gaspari for excellent research assistance.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paolo Guasoni.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

P. Guasoni is partially supported by the ERC (279582), NSF (DMS-1412529), and SFI (16/IA/4443 and 16/SPP/3347).

Appendix: Proofs

Appendix: Proofs

1.1 A.1 Preliminaries

The next proof relies on the definition of consumption and investment policies \(\mathcal{C}\) and \(\mathcal{P}\), respectively, which combine into admissible strategies.

Definition A.1

In all markets (segmented or integrated), the set \(\mathcal{C}\) of consumption policies consists of all adapted processes \((c_{t})_{t\ge 0}\) with \(c_{t}>0\) a.s. for all \(t\ge 0\) and \(\mathbb{E}[\int ^{T}_{0} c_{t} dt]<\infty \) for all \(T >0\). In each segmented market, the set of investment policies \(\mathcal{P}^{i}\), \(i=1,2\), consists of all predictable processes \((\phi _{t})_{t\ge 0}\) such that \(\mathbb{E}[\int ^{T}_{0} \phi _{t}^{2} d \langle P^{(i)} \rangle _{t}]< \infty \) for all \(T>0\), where \(P^{(i)}\) is the respective asset price. In the integrated market, the set \(\overline{\mathcal{P}}\) of investment policies consists of \(\mathbb{R}^{2}\)-valued, predictable processes \((\phi _{t})_{t\ge 0}\) such that \(\mathbb{E}[\langle \int ^{\cdot }_{0} \phi \cdot dP\rangle _{T}]< \infty \) for all \(T>0\). The set \(\mathcal{A}\) of admissible strategies consists of all pairs \((c,\phi )\) such that \(c\in \mathcal{C}\), \(\phi \in \mathcal{P}^{(i)}\) for each segmented market and \(\phi \in \overline{\mathcal{P}}\) for the integrated market, and the corresponding wealth in (3.6) or (3.7) satisfies \(Y_{t}\ge 0\) a.s. for all \(t\ge 0\).

Definition 3.3 of an equilibrium leads to the familiar representation of prices as discounted cash flows and the safe rate as minus the growth rate of the stochastic discount factor. Although the result is often taken for granted and can be informally derived from a perturbation argument, it is not guaranteed to hold in general and the literature does offer counterexamples; see Basak and Cuoco [4, Remark 4 and Corollary 3] and the discussion in Karatzas et al. [31]. Thus we offer a proof that applies to the model considered here.

Proposition A.2

In the segmented markets, the equilibrium asset prices are

$$\begin{aligned} P^{(i)}_{t} = \mathbb{E}_{t}\bigg[ \int ^{\infty }_{t} \frac{M^{(i)}_{s}}{M^{(i)}_{t}} D^{(i)}_{s} ds \bigg], \qquad \textit{where } M^{(i)}_{t} = e^{-\beta t}(D^{(i)}_{t})^{-\gamma }, \end{aligned}$$

while equilibrium safe rates \(r^{(i)}_{t}\)are identified by the local martingale condition for \((M^{(i)}_{t} e^{\int _{0}^{t} r^{(i)}_{s} ds})_{t\geq 0}\). Likewise, in the integrated market, the equilibrium asset prices are

$$\begin{aligned} \bar{P}^{(i)}_{t} = \mathbb{E}_{t}\bigg[ \int ^{\infty }_{t} \frac{\bar{M}_{s}}{\bar{M}_{t}} D^{(i)}_{s} ds \bigg], \qquad \textit{where } \bar{M}_{t} = e^{-\beta t}(D^{(1)}_{t} + D^{(2)}_{t})^{- \gamma } . \end{aligned}$$

Proof

Note that if \(Y_{0}=P^{(i)}_{0}\), \(c_{t}=D^{(i)}_{t}\) and \(\phi _{t}\equiv 1\), then (3.6) implies that \(Y_{t}=P^{(i)}_{t}\) for all \(t \geq 0\). Fix \(i=1,2\), \(t_{0}>0\), \(\vartheta > 0\) and \(K>1\). Define the stopping time

$$ \tau _{0}:=(t_{0}+\vartheta )\wedge \inf \bigg\{ t>t_{0} \,: \, P^{(i)}_{t} \leq \frac{1}{K} P^{(i)}_{t_{0}} \text{ and } D^{(i)}_{t}\geq K D^{(i)}_{t_{0}} \bigg\} . $$

For any \(\delta \in (-\vartheta , \vartheta )\), consider an event \(A \in \mathcal{F}_{t_{0}}\) on which we adopt an alternative strategy, where from time \(t_{0}\) to \(\tau _{0}\) the consumption changes from \(D^{(i)}_{t}\) to \({c^{\delta }_{t}=(1-\frac{\delta }{\vartheta })D^{(i)}_{t}}\) with the difference in consumption invested in the risky asset, so that the number of shares changes from 1 to \({\phi ^{\delta }_{t}:=1+ \frac{\delta }{\vartheta }\Delta _{t}}\) for some process \((\Delta _{t})_{t\geq 0}\). Thus the corresponding wealth on \(A\) becomes

$$ {Y^{\delta }_{t}=\left (1+ \frac{\delta }{\vartheta }\Delta _{t}\right )P^{(i)}_{t}}. $$

To satisfy the budget constraints (3.6), \((\Delta _{t})\) satisfies on \(A\)

$$\begin{aligned} d\bigg(1+\frac{\delta }{\vartheta }\Delta _{t}\bigg)P^{(i)}_{t} &= r^{(i)}_{t} \bigg(\Big(1+\frac{\delta }{\vartheta }\Delta _{t}\Big)P^{(i)}_{t} - \Big(1+\frac{\delta }{\vartheta }\Delta _{t}\Big) P^{(i)}_{t}\bigg) dt \\ &\phantom{=:}+ \bigg(1+ \frac{\delta }{\vartheta } \Delta _{t}\bigg) ({dP^{(i)}_{t}}+D^{(i)}_{t} dt )- \bigg(1-\frac{\delta }{\vartheta }\bigg)D^{(i)}_{t} dt, \end{aligned}$$

therefore

$$\begin{aligned} \frac{\delta }{\vartheta } P^{(i)}_{t} d\Delta _{t}+ \frac{\delta }{\vartheta } d \langle \Delta , P^{(i)} \rangle _{t} = \bigg(\frac{\delta }{\vartheta }+\frac{\delta }{\vartheta }\Delta _{t} \bigg) D^{(i)}_{t} dt \end{aligned}$$

and so

$$\begin{aligned} d\Delta _{t} = (1+\Delta _{t} )\frac{D^{(i)}_{t}}{P^{(i)}_{t}}dt . \end{aligned}$$

Hence we get on \(A\) that

Δ t = t 0 t e s t D u ( i ) P u ( i ) d u D s ( i ) P s ( i ) dsϑ K 2 e ϑ K 2 D t 0 ( i ) P t 0 ( i ) D t 0 ( i ) P t 0 ( i ) on  t 0 , τ 0 .

After \(\tau _{0}\), the investor holds \(\phi ^{\delta }_{t}\equiv 1+\frac{\delta }{\vartheta }\Delta _{\tau _{0}}\) unit of risky asset and consumes \({c^{\delta }_{t}=(1+ \frac{\delta }{\vartheta } \Delta _{\tau _{0}})D^{(i)}_{t}}\). To ensure that consumption remains positive, assume also that

$$ {\delta >-\frac{1}{K^{2}} e^{-\vartheta K^{2} \frac{D^{(i)}_{t_{0}}}{P^{(i)}_{t_{0}}}} \frac{P^{(i)}_{t_{0}}}{D^{(i)}_{t_{0}}}} . $$

The change in expected utility from \((c_{t},\phi _{t})_{t\geq 0}\) to \((c^{\delta }_{t},\phi ^{\delta }_{t})_{t\geq 0}\) is thus

$$\begin{aligned} \Delta ^{\delta }J&= \mathbb{E}\bigg[1_{A}\bigg(\int ^{\tau _{0}}_{t_{0}} \! \frac{ e^{-\beta t} ( (1-\frac{\delta }{\vartheta } )D^{(i)}_{t} )^{1-\gamma }}{1-\gamma } dt +\!\int _{\tau _{0}}^{+\infty } \! \frac{ e^{-\beta t} ( (1+\frac{\delta }{\vartheta }\Delta _{\tau _{0}} )D^{(i)}_{t} )^{1-\gamma }}{1-\gamma } dt \\ &\phantom{=:} \qquad \qquad -\int _{t_{0}}^{+\infty } \frac{ e^{-\beta t} (D^{(i)}_{t} )^{1-\gamma }}{1-\gamma } dt\bigg) \bigg]. \end{aligned}$$

Because \(\frac{x^{1-\gamma }}{1-\gamma }\) is concave, we have \(y^{-\gamma }(y-x) \leq \frac{y^{1-\gamma }}{1-\gamma }- \frac{x^{1-\gamma }}{1-\gamma }\) for any \(x,y>0\), whence

$$\begin{aligned} \Delta ^{\delta }J\geq \mathbb{E}\bigg[1_{A}\bigg( -&\frac{\delta }{\vartheta } \int ^{\tau _{0}}_{t_{0}} e^{-\beta t}D^{(i)}_{t} \bigg(\Big(1- \frac{\delta }{\vartheta }\Big)D^{(i)}_{t}\bigg)^{-\gamma } dt \\ +& \frac{\delta }{\vartheta }\Delta _{\tau _{0}}\int _{\tau _{0}}^{+ \infty } D^{(i)}_{t} e^{-\beta t}\bigg(\Big(1+\frac{\delta }{\vartheta } \Delta _{\tau _{0}}\Big)D^{(i)}_{t}\bigg)^{-\gamma } dt\bigg)\bigg]. \end{aligned}$$

As \((c_{t},\phi _{t})_{t\geq 0}\) is optimal, it follows that \(\lim _{\delta \downarrow 0}\frac{ \Delta ^{\delta }J}{\delta }\leq 0\) and \(\lim _{\delta \uparrow 0}\frac{ \Delta ^{\delta }J}{\delta }\geq 0\), whence

$$\begin{aligned} \lim _{\delta \rightarrow 0} \frac{1}{\vartheta } \mathbb{E}\bigg[1_{A} \bigg(&-\int ^{\tau _{0}}_{t_{0}} e^{-\beta t}D^{(i)}_{t} \bigg(\Big(1- \frac{\delta }{\vartheta }\Big)D^{(i)}_{t}\bigg)^{-\gamma } dt \\ & +\Delta _{\tau _{0}}\int _{\tau _{0}}^{+\infty } D^{(i)}_{t} e^{- \beta t}\bigg(\Big(1+\frac{\delta }{\vartheta }\Delta _{\tau _{0}}\Big)D^{(i)}_{t} \bigg)^{-\gamma } dt\bigg)\bigg] = 0 \end{aligned}$$

and therefore

$$ \frac{\mathbb{E} [1_{A} (-\int ^{\tau _{0}}_{t_{0}} e^{-\beta t}D^{(i)}_{t} (D^{(i)}_{t} )^{-\gamma } dt+\Delta _{\tau _{0}}\int _{\tau _{0}}^{+\infty } D^{(i)}_{t} e^{-\beta t} (D^{(i)}_{t} )^{-\gamma } dt ) ]}{\vartheta }= 0. $$
(A.1)

Because (A.1) holds for all \(\vartheta \neq 0\),

$$\begin{aligned} \lim _{\vartheta \downarrow 0} \frac{\mathbb{E} [1_{A} (-\int ^{\tau _{0}}_{t_{0}} e^{-\beta t}D^{(i)}_{t} (D^{(i)}_{t} )^{-\gamma } dt+\Delta _{\tau _{0}}\int _{\tau _{0}}^{+\infty } D^{(i)}_{t} e^{-\beta t} (D^{(i)}_{t} )^{-\gamma } dt ) ]}{\vartheta } = 0. \end{aligned}$$
(A.2)

Note that

$$\begin{aligned} \lim _{\vartheta \downarrow 0} \frac{\int ^{\tau _{0}}_{t_{0}} e^{-\beta t}D^{(i)}_{t} (D^{(i)}_{t} )^{-\gamma } dt}{\vartheta } =&e^{- \beta t_{0}} (D^{(i)}_{t_{0}} )^{1-\gamma }, \\ \lim _{\vartheta \downarrow 0} \int _{\tau _{0}}^{+\infty } D^{(i)}_{t} e^{-\beta t} (D^{(i)}_{t} )^{-\gamma } dt =& \int _{t_{0}}^{+\infty } D^{(i)}_{t} e^{-\beta t} (D^{(i)}_{t} )^{-\gamma } dt \end{aligned}$$

and

$$ \lim _{\vartheta \downarrow 0}\frac{\Delta _{\tau _{0}}}{\vartheta }= \lim _{\vartheta \downarrow 0}\frac{1}{\vartheta }\int ^{\tau _{0}}_{t_{0}} e^{\int _{s}^{\tau _{0}} \frac{D^{(i)}_{u}}{P^{(i)}_{u}} du} \frac{D^{(i)}_{s}}{P^{(i)}_{s}} ds= \frac{D^{(i)}_{t_{0}}}{P^{(i)}_{t_{0}}}. $$

Thus an application of the dominated convergence theorem to (A.2) yields

$$ \mathbb{E}\bigg[1_{A}\bigg(-e^{-\beta t_{0}} (D^{(i)}_{t_{0}} )^{1- \gamma }+\frac{D^{(i)}_{t_{0}}}{P^{(i)}_{t_{0}}}\mathbb{E}_{t_{0}} \Big[\int _{t_{0}}^{+\infty } D^{(i)}_{t} e^{-\beta t} (D^{(i)}_{t} )^{- \gamma } dt\Big]\bigg)\bigg]= 0. $$

As \(A \in \mathcal{F}_{t_{0}}\) is arbitrary, it follows that

$$ P^{(i)}_{t_{0}}= \mathbb{E}_{t_{0}}\bigg[\int _{t_{0}}^{+\infty } D^{(i)}_{t} e^{-\beta (t-t_{0})}\bigg(\frac{D^{(i)}_{t}}{D^{(i)}_{t_{0}}}\bigg)^{- \gamma } dt\bigg]. $$
(A.3)

To identify the safe rate \(r^{(i)}\), fix \(i \in \{1,2\}\), \(t_{1}>t_{0}>0\), \(\vartheta \in (0, t_{1}-t_{0})\) and \({K>1}\), and define stopping times

$$\begin{aligned} \eta _{K}&:=\inf \bigg\{ t>t_{0} \,:\, e^{\int ^{t}_{t_{0}} \vert r^{(i)}_{u} \vert du} \geq K \text{ and } P^{(i)}_{t} \leq \frac{1}{K} P^{(i)}_{t_{0}} \bigg\} , \\ \tau _{0}&:= (t_{0}+\vartheta )\wedge \eta _{K}\wedge \inf \bigg\{ t>t_{0} : D^{(i)}_{t} \geq K D^{(i)}_{t_{0}} \text{ and} \\ &\qquad \qquad \qquad \qquad \qquad \qquad \quad \frac{1}{P^{(i)}_{t} } \int ^{t}_{t_{0}} e^{\int ^{t}_{s} \vert r^{(i)}_{u} \vert du}D_{s}^{(i)}ds \geq \frac{K}{2} \bigg\} . \end{aligned}$$

For any \(0 < |\delta |<\frac{\vartheta }{K}\), further define another stopping time

$$ \tau _{1}:= t_{1}\wedge \eta _{K}\wedge \inf \bigg\{ t>\tau _{0} \,:\, \bigg\vert \frac{\delta }{\vartheta }\frac{1}{P^{(i)}_{t}}\int ^{\tau _{0}}_{t_{0}} e^{\int ^{t}_{s} r^{(i)}_{u} du}D_{s}^{(i)}ds\bigg\vert \geq \frac{1}{2} \bigg\} . $$

Consider an event \(A \in \mathcal{F}_{t_{0}}\) on which we adopt an alternative strategy in which, as before, from \(t_{0}\) to \(\tau _{0}\) the consumption changes from \(D^{(i)}_{t}\) to \({c^{\delta }_{t}=(1-\frac{\delta }{\vartheta })D^{(i)}_{t}}\), while the difference now is invested in the safe asset. Thus the corresponding wealth on \(A\) is \({Y^{\delta }_{t}=P^{(i)}_{t}+\frac{\delta }{\vartheta }\Delta _{t}}\), where \({\Delta _{t}=\int ^{t}_{t_{0}} e^{\int ^{t}_{s} r^{(i)}_{u} du}D_{s}^{(i)} ds}\) on \(A\). From time \(\tau _{0}\) to \(\tau _{1}\), the strategy remains \((c^{\delta }_{t},\phi ^{\delta }_{t})=(D^{(i)}_{t},1)\), and thus wealth on \(A\) becomes \({Y^{\delta }_{t}=P^{(i)}_{t}+\frac{\delta }{\vartheta }\Delta _{t}}\), where \({\Delta _{t}=\int ^{\tau _{0}}_{t_{0}} e^{\int ^{t}_{s} r^{(i)}_{u} du} D^{(i)}_{s} ds}\) on \(A\). After \(\tau _{1}\), the lump sum \({\frac{\delta }{\vartheta }\Delta _{\tau _{1}}}\) is used to buy \({\frac{\delta }{\vartheta } \frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}}}\) units of stock so that the number of shares on \(A\) becomes \({\phi ^{\delta }_{t}=1+\frac{\delta }{\vartheta } \frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}}}\) and the consumption rate is \({c^{\delta }_{t}=(1+ \frac{\delta }{\vartheta } \frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}})D^{(i)}_{t}}\) for \(t \geq \tau _{1}\) on \(A\). Then the change in expected utility from \((c_{t},\phi _{t})_{t\geq 0}\) to \((c^{\delta }_{t},\phi ^{\delta }_{t})_{t\geq 0}\) is

$$\begin{aligned} \Delta ^{\delta }J = \mathbb{E}\bigg[1_{A}\bigg(&\int ^{\tau _{0}}_{t_{0}} \frac{ e^{-\beta t} ( (1-\frac{\delta }{\vartheta } )D^{(i)}_{t} )^{1-\gamma }}{1-\gamma } dt \\ & +\int _{\tau _{1}}^{+\infty } \frac{ e^{-\beta t} ( (1+\frac{\delta }{\vartheta }\frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}} )D^{(i)}_{t} )^{1-\gamma }}{1-\gamma } dt \!-\!\int _{t_{0}}^{+\infty } \frac{ e^{-\beta t} (D^{(i)}_{t} )^{1-\gamma }}{1-\gamma } dt\!\bigg) \bigg]. \end{aligned}$$

Again, concavity implies that \(y^{-\gamma }(y-x) \leq \frac{y^{1-\gamma }}{1-\gamma }- \frac{x^{1-\gamma }}{1-\gamma } \text{ for any } x\), \(y>0\); thus

$$\begin{aligned} \Delta ^{\delta }J \geq \mathbb{E}\bigg[1_{A}\bigg(&- \frac{\delta }{\vartheta }\int ^{\tau _{0}}_{t_{0}} e^{-\beta t} D^{(i)}_{t} \bigg(\Big(1-\frac{\delta }{\vartheta }\Big)D^{(i)}_{t}\bigg)^{-\gamma } dt \\ & +\frac{\delta }{\vartheta } \frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}}\int _{\tau _{1}}^{+ \infty } e^{-\beta t}D^{(i)}_{t}\bigg(\Big(1+\frac{\delta }{\vartheta } \frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}}\Big)D^{(i)}_{t} \bigg)^{-\gamma } dt\bigg)\bigg]. \end{aligned}$$

As \((c_{t},\phi _{t})_{t\geq 0}\) is optimal, we have \(\lim _{\delta \downarrow 0}\frac{ \Delta ^{\delta }J}{\delta }\leq 0\) and \(\lim _{\delta \uparrow 0}\frac{ \Delta ^{\delta }J}{\delta }\geq 0\) and thus

$$\begin{aligned} \lim _{\delta \rightarrow 0} \frac{1}{\vartheta }\mathbb{E}\bigg[1_{A} \bigg(&-\int ^{\tau _{0}}_{t_{0}} e^{-\beta t} D^{(i)}_{t} \bigg( \Big(1-\frac{\delta }{\vartheta }\Big)D^{(i)}_{t}\bigg)^{-\gamma } dt \\ &+\frac{\Delta _{\tau _{1}}}{ P^{(i)}_{\tau _{1}}}\int _{\tau _{1}}^{+ \infty } D^{(i)}_{t} e^{-\beta t}\bigg(\Big(1+\frac{\delta }{\vartheta } \frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}}\Big)D^{(i)}_{t} \bigg)^{-\gamma } dt\bigg)\bigg] = 0. \end{aligned}$$

Because \(\lim _{\delta \rightarrow 0}\tau _{1}=\tau _{2}:=\min \{t_{1},\eta _{K} \}\) almost surely and \(\frac{\Delta _{\tau _{1}}}{P^{(i)}_{\tau _{1}}} \leq (\tau _{1}-t_{0})K^{3} \frac{D^{(i)}_{t_{0}}}{P^{(i)}_{t_{0}}}\), dominated convergence yields

$$ \lim _{\vartheta \downarrow 0} \frac{\mathbb{E} [1_{A} (-\int ^{\tau _{0}}_{t_{0}} e^{-\beta t} (D^{(i)}_{t} )^{1-\gamma } dt+\frac{\Delta _{\tau _{2}}}{ P^{(i)}_{\tau _{2}}}\int _{\tau _{2}}^{+\infty } e^{-\beta t} (D^{(i)}_{t} )^{1-\gamma } dt ) ]}{\vartheta }= 0. $$
(A.4)

Recalling that \({\Delta _{\tau _{2}}=\int ^{\tau _{0}}_{t_{0}} e^{\int ^{\tau _{2}}_{s} r^{(i)}_{u} du} D^{(i)}_{s} ds}\), it follows that

$$\begin{aligned} \lim _{\vartheta \downarrow 0} \frac{\Delta _{\tau _{2}}}{\vartheta } =&e^{ \int ^{\tau _{2}}_{t_{0}} r^{(i)}_{u} du} D^{(i)}_{t_{0}} . \end{aligned}$$

Applying the dominated convergence theorem to (A.4), (A.3) implies that

$$\begin{aligned} &\mathbb{E}\bigg[1_{A}\bigg(-e^{-\beta t_{0}} (D^{(i)}_{t_{0}} )^{1- \gamma }+ \frac{e^{\int ^{\tau _{2}}_{t_{0}} r^{(i)}_{u} du} D^{(i)}_{t_{0}} }{P^{(i)}_{\tau _{2}}} \mathbb{E}_{\tau _{2}}\Big[\int _{\tau _{2}}^{+\infty } D^{(i)}_{t} e^{- \beta t} (D^{(i)}_{t} )^{-\gamma } dt\Big]\bigg)\bigg] \\ &= e^{-\beta t_{0}}\mathbb{E}\big[1_{A} D^{(i)}_{t_{0}}\big(- (D^{(i)}_{t_{0}} )^{-\gamma }+e^{\int ^{\tau _{2}}_{t_{0}} r^{(i)}_{u} du} e^{-\beta ( \tau _{2}-t_{0})} (D^{(i)}_{\tau _{2}} )^{-\gamma }\big)\big] =0 . \end{aligned}$$

As \(A \in \mathcal{F}_{t_{0}}\) is arbitrary, we get

$$\begin{aligned} e^{\int ^{t_{0}}_{0} (r^{(i)}_{u}-\beta ) du} (D^{(i)}_{t_{0}} )^{- \gamma }= \mathbb{E}_{t_{0}} \big[e^{\int ^{\min \{t_{1},\eta _{K} \}}_{0} (r^{(i)}_{u}-\beta ) du} (D^{(i)}_{\min \{t_{1},\eta _{K} \}} )^{- \gamma } \big], \end{aligned}$$

which implies that \(( e^{\int ^{t}_{0} (r^{(i)}_{u}-\beta ) du} (D^{(i)}_{t})^{-\gamma })_{t \geq 0}\) is a local martingale. □

1.2 A.2 Proof of Theorem 4.1

The proof focuses on region 1, as the argument for region 2 is analogous.

(i) Recalling that \(e^{\int ^{t}_{0} r^{(1)}_{s} ds}M^{(1)}_{t}=e^{\int ^{t}_{0} (r^{(1)}_{s}- \beta ) ds} c_{t}^{\gamma }=e^{\int ^{t}_{0} (r^{(1)}_{s}-\beta ) ds}(D_{t}^{(1)})^{ \gamma }\) and applying Itô’s lemma, it follows that

$$\begin{aligned} &d \big(e^{\int ^{t}_{0} (r^{(1)}_{s}-\beta ) ds} (D_{t}^{(1)} )^{- \gamma }\big) \\ &= (r^{(1)}_{t}-\beta ) e^{\int ^{t}_{0} (r^{(1)}_{s}-\beta ) ds} (D_{t}^{(1)} )^{-\gamma } dt \\ &\phantom{=:}- \gamma e^{\int ^{t}_{0} (r^{(1)}_{s}-\beta ) ds} (D^{(1)}_{t} )^{- \gamma -1} \big((\mu -\kappa ) D^{(1)}_{t}+\kappa w_{1} D_{t} \big)dt \\ &\phantom{=:}+\frac{\gamma (1+\gamma )\sigma ^{2}}{2}e^{\int ^{t}_{0} (r^{(1)}_{s}- \beta ) ds} (D^{(1)}_{t} )^{-\gamma -1}D_{t} dt+ dL_{t} \\ &= e^{\int ^{t}_{0} (r^{(1)}_{s}-\beta ) ds} (D^{(1)}_{t} )^{-\gamma } \\ &\phantom{=:}\times \bigg(r^{(1)}_{t}-\beta -\gamma (\mu -\kappa )-\gamma \kappa w_{1} \frac{D_{t}}{D^{(1)}_{t}}+\frac{\gamma (\gamma +1) \sigma ^{2}}{2} \frac{D_{t}}{D^{(1)}_{t}} \bigg) dt+dL_{t} \end{aligned}$$

for some local martingale \(L\) (a stochastic integral with respect to Brownian motion, whose expression is inconsequential). Thus the equilibrium safe rate is

$$ {r^{(1)}_{t}=\beta +\gamma (\mu -\kappa )+\gamma \mu w_{1} \frac{D_{t}}{D^{(1)}_{t}}-\frac{\gamma (\gamma +1) \sigma ^{2}}{2} \frac{D_{t}}{D^{(1)}_{t}}} . $$

Write the price of asset 1 in terms of \(X_{t}\) and \(D_{t}\) as

$$\begin{aligned} P^{(1)}_{t}&= \mathbb{E}_{t}\bigg[\int ^{\infty }_{t} e^{-\beta (s-t)} \frac{ (D_{s}^{(1)} )^{1-\gamma }}{ (D_{t}^{(1)} )^{-\gamma }} ds\bigg] \\ &=\mathbb{E}_{t}\bigg[\int ^{\infty }_{t} e^{-\beta (s-t)} \frac{ D_{s}^{1-\gamma }}{ D_{t}^{-\gamma }} \frac{X_{s}^{1-\gamma }}{X_{t}^{-\gamma }} ds\bigg] \\ &= \int ^{\infty }_{t} e^{-\beta (s-t)} \mathbb{E}_{t}\bigg[ \frac{ D_{s}^{1-\gamma }}{ D_{t}^{-\gamma }} \frac{X_{s}^{1-\gamma }}{X_{t}^{-\gamma }}\bigg] ds \\ &= \int ^{\infty }_{t} e^{-\beta (s-t)} \mathbb{E}_{t}\bigg[ \frac{ D_{s}^{1-\gamma }}{D_{t}^{-\gamma }}\bigg] \mathbb{E}_{t}\bigg[ \frac{X_{s}^{1-\gamma }}{X_{t}^{-\gamma }}\bigg] ds \\ &= D_{t} X_{t}^{\gamma } \int ^{\infty }_{t} e^{-\beta (s-t)} \mathbb{E}_{t} \bigg[\bigg(\frac{D_{s}}{D_{t}}\bigg)^{1-\gamma }\bigg]\mathbb{E}_{t} [X_{s}^{1- \gamma } ] ds , \end{aligned}$$
(A.5)

where the fourth equality follows by the independence of \((B^{D}_{t})_{t\geq 0}\) and \((B^{X}_{t})_{t\geq 0}\). As \((D_{t})_{t\geq 0}\) is a geometric Brownian motion, it satisfies

$$ \mathbb{E}\bigg[\bigg(\frac{D_{s}}{D_{t}}\bigg)^{1-\gamma } \bigg]= \mathbb{E} \big[e^{(1-\gamma ) (\mu -\frac{\sigma ^{2}}{2} )(s-t)+(1- \gamma )\sigma (B^{D}_{s}-B^{D}_{t} )} \big] =e^{(1-\gamma ) (\mu - \frac{\gamma \sigma ^{2}}{2} )(s-t)}. $$

Then (A.5) becomes

$$\begin{aligned} P^{(1)}_{t} =&D_{t} X_{t}^{\gamma }\int ^{\infty }_{t} e^{-\theta (s-t)} \mathbb{E}_{t} [X_{s}^{1-\gamma } ] ds =D_{t} X_{t}^{\gamma }\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} e^{-\theta (s-t)} X_{s}^{1-\gamma } ds \bigg] \\ =& D_{t} X_{t}^{\gamma }\mathbb{E}\bigg[\int ^{\infty }_{t} e^{-\theta (s-t)} X_{s}^{1-\gamma } ds \bigg\vert X_{t} \bigg]= D_{t} X_{t}^{\gamma } f^{(1)}(X_{t}), \end{aligned}$$

where the third equality follows by the Markov property of \((X_{t})_{t\geq 0}\) and

$$\begin{aligned} f^{(1)}(x) :=\mathbb{E}\bigg[\int ^{\infty }_{t} e^{-\theta (s-t)} X_{s}^{1- \gamma } ds \bigg\vert X_{t}=x \bigg] =\mathbb{E}\bigg[\int ^{\infty }_{0} e^{-\theta s} X_{s}^{1-\gamma } ds \bigg\vert X_{0}=x\bigg]. \end{aligned}$$

Note that the second equality holds true because of the Markov property.

(ii) For the welfare, recall that

$$ \mathcal{U}^{(1)}_{t} =\mathbb{E}_{t}\bigg[\int ^{\infty }_{t} e^{- \beta (s-t)}\frac{c_{s}^{1-\gamma }}{1-\gamma } ds\bigg]=\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} e^{-\beta (s-t)} \frac{ (D_{s}^{(1)} )^{1-\gamma }}{1-\gamma } ds\bigg]. $$

Comparing this expression to the formula for \(P_{t}^{(1)}\) in (A.5), note that

$$\begin{aligned} \mathcal{U}^{(1)}_{t} &= \frac{1}{1-\gamma } (D^{(1)}_{t} )^{-\gamma } P^{(1)}_{t}= \frac{1}{1-\gamma } (D^{(1)}_{t} )^{1-\gamma }X_{t}^{\gamma -1}f^{(1)}(X_{t}) \\ &= \frac{1}{1-\gamma } D_{t}^{1-\gamma }f^{(1)}(X_{t}). \end{aligned}$$

(iii) As \(P^{(1)}_{t}=D_{t} X_{t}^{\gamma }f^{(1)}(X_{t})\), where \(f^{(1)}\) is defined in (4.1), Itô’s formula yields

$$\begin{aligned} &dP^{(1)}_{t} \\ &= X_{t}^{\gamma }f^{(1)}(X_{t}) dD_{t}+ D_{t} X_{t}^{\gamma -1}\big( \gamma f^{(1)}(X_{t})+X_{t} (f^{(1)} )'(X_{t}) \big)dX_{t} \\ &\quad {} +X_{t}^{\gamma -1}\big(\gamma f^{(1)}(X_{t})+X_{t} (f^{(1)} )'(X_{t}) \big) d \langle D,X \rangle _{t} \\ &\quad {} +\frac{1}{2} D_{t} X_{t}^{\gamma -2}\big(\gamma (\gamma -1) f^{(1)}(X_{t})+2 \gamma X_{t} (f^{(1)} )'(X_{t}) +X_{t}^{2} (f^{(1)} )''(X_{t}) \big) d \langle X \rangle _{t} \\ &=D_{t} X_{t}^{\gamma -1} \bigg(\Big(\mu X_{t} f^{(1)}(X_{t})+\kappa (w_{1}-X_{t}) \big(\gamma f^{(1)}(X_{t})+X_{t} (f^{(1)} )'(X_{t}) \big) \\ & \qquad \qquad \ \ \,\quad {}+\frac{\sigma ^{2}}{2}(1-X_{t})\big( \gamma (\gamma -1) f^{(1)}(X_{t})+2\gamma X_{t} (f^{(1)} )'(X_{t}) \\ &\phantom{=:} \qquad \qquad \qquad \qquad \qquad \,\ \ {} +X_{t}^{2} (f^{(1)} )''(X_{t}) \big)\Big) dt +\sigma X_{t} f^{(1)}(X_{t}) dB^{D}_{t} \\ & \phantom{=:}\qquad \quad \quad {} +\sigma \sqrt{X_{t}(1-X_{t})}\big(\gamma f^{(1)}(X_{t})+X_{t} (f^{(1)} )'(X_{t}) \big) dB^{X}_{t}\bigg) \\ &= \sigma D_{t} X_{t}^{\gamma }f^{(1)} ( X_{t} )dB^{D}_{t} + A^{1}_{t} dt \\ & \phantom{=:}+ \sigma D_{t} X_{t}^{\gamma -1}\big(\gamma f^{(1)}(X_{t})+X_{t}({f^{(1)}})'(X_{t}) \big)\sqrt{X_{t} (1-X_{t})}dB^{X}_{t} \\ &=\sigma D_{t} X_{t}^{\gamma }f^{(1)} ( X_{t} )dB^{D}_{t} + \sigma D_{t} X_{t}^{\gamma -1}h^{(1)} ( X_{t} )\sqrt{X_{t} (1-X_{t})}dB^{X}_{t} + A^{1}_{t} dt \end{aligned}$$
(A.6)

for an adapted process \(A^{1}\). An analogous computation for region 2 yields

$$\begin{aligned} &dP^{(2)}_{t} \\ &= \sigma D_{t} (1-X_{t})^{\gamma }f^{(2)} ( X_{t} ) dB^{D}_{t} + A^{2}_{t} dt \\ &\phantom{=:}+ \sigma D_{t} (1-X_{t})^{\gamma -1} \big(-\gamma f^{(2)}(X_{t})+(1-X_{t})({f^{(2)}})'(X_{t}) \big)\sqrt{X_{t} (1-X_{t})}dB^{X}_{t} \\ &= \sigma D_{t} (1-X_{t})^{\gamma }f^{(2)} ( X_{t} ) dB^{D}_{t} \\ & \phantom{=:}+\sigma D_{t} (1-X_{t})^{\gamma -1}h^{(2)} (X_{t} )\sqrt{X_{t} (1-X_{t})} dB^{X}_{t}+ A^{2}_{t} dt \end{aligned}$$

for another adapted process \(A^{2}\). Hence

$$\begin{aligned} \frac{d \langle P^{(1)} \rangle _{t}}{dt} &= \sigma ^{2} D^{2}_{t} X_{t}^{2( \gamma -1)} \Big( \big(X_{t} f^{(1)}(X_{t})\big)^{2} +\big(h^{(1)}(X_{t}) \big)^{2} X_{t}(1-X_{t})\Big), \\ \frac{d \langle P^{(2)} \rangle _{t}}{dt}& = \sigma ^{2} D^{2}_{t} (1-X_{t})^{2( \gamma -1)} \\ &\phantom{=:}\times \Big( \big((1-X_{t}) f^{(2)}(X_{t})\big)^{2} +\big(h^{(2)}(X_{t}) \big)^{2} X_{t}(1-X_{t})\Big), \\ \frac{d \langle P^{(1)}, P^{(2)} \rangle _{t}}{dt} &= \sigma ^{2} D^{2}_{t} X_{t}^{\gamma }(1-X_{t})^{\gamma } \big( f^{(1)}(X_{t}) f^{(2)}(X_{t}) +h^{(1)}(X_{t})h^{(2)}(X_{t}) \big), \end{aligned}$$

and the claim follows by (4.5).

(iv) Again, the proof focuses on region 1, as the argument for region 2 is analogous. Following the derivation of (A.6), we get

$$\begin{aligned} &\frac{dP^{(1)}_{t}}{P^{(1)}_{t}} - \sigma dB^{D}_{t} - \sigma \sqrt{ \frac{1-X_{t}}{X_{t}}}\bigg(\gamma +X_{t} \frac{ (f^{(1)} )'(X_{t})}{f^{(1)}(X_{t})} \bigg) dB^{X}_{t} \\ &= \bigg(\mu +\kappa \frac{w_{1}-X_{t}}{X_{t}}\Big(\gamma +X_{t} \frac{ (f^{(1)} )'(X_{t})}{f^{(1)}(X_{t})} \Big) \\ &\qquad {}+ \frac{\sigma ^{2}}{2}\frac{1-X_{t}}{X_{t}}\Big( \gamma (\gamma -1) +2\gamma X_{t} \frac{ (f^{(1)} )'(X_{t})}{f^{(1)}(X_{t})}+X_{t}^{2} \frac{ (f^{(1)} )''(X_{t})}{f^{(1)}(X_{t})} \Big)\bigg) \\ &=: \mu ^{(1)}_{t}-\frac{D^{(1)}_{t}}{P^{(1)}_{t}} dt \end{aligned}$$

which proves the claim. □

Next, under Assumption 4.2, closed-form expressions for the \(f^{(i)}\) in terms of hypergeometric functions follow (cf. the formulas derived by Hurd and Kuznetsov [30] for related functionals of the Jacobi process).

Proposition A.3

Let Assumptions 3.4and 4.2hold. Then

$$\begin{aligned} f^{(1)}(x) &= \frac{2}{\sigma ^{2} \omega ^{(1)}}\bigg( \frac{\Gamma ^{(1)}_{1}}{1+c_{1}-\gamma } F^{(1)}_{1}(x) G^{(1)}_{1}(1)+ \frac{\Gamma ^{(1)}_{2}}{1+c_{1}-\gamma } x^{2-\gamma } F^{(1)}_{2}(x) G^{(1)}_{1}(x) \\ &\qquad \qquad \quad {}+ \frac{\Gamma ^{(1)}_{2}}{2-\gamma } F^{(1)}_{1}(x) \big(G^{(1)}_{2}(1)-x^{2-\gamma } G^{(1)}_{2}(x)\big) \bigg), \\ f^{(2)}(x) &= \frac{2}{\sigma ^{2} \omega ^{(2)}}\bigg( \frac{\Gamma ^{(2)}_{1}}{1+c_{2}-\gamma } F^{(2)}_{1}(1-x) G^{(2)}_{1}(1) \\ &\qquad \qquad \quad {}+ \frac{\Gamma ^{(2)}_{2} (1-x)^{2-\gamma }}{1+c_{2}-\gamma } F^{(2)}_{2}(1-x) G^{(2)}_{1}(1-x) \\ &\qquad \qquad \quad {}+ \frac{\Gamma ^{(2)}_{2}}{2-\gamma } F^{(2)}_{1}(1-x) \big(G^{(2)}_{2}(1)-(1-x)^{2-\gamma } G^{(2)}_{2}(1-x)\big) \bigg), \end{aligned}$$

where

$$\begin{aligned} F^{(i)}_{1}(x)&:={}_{2}F_{1}(a,b;c_{i};x), \\ G^{(i)}_{1}(x)&:={}_{3}F_{2}(c_{i}-a,c_{i}-b,1+c_{i}-\gamma ;c_{i},2+c_{i}- \gamma ;x), \\ F^{(i)}_{2}(x)&:={}_{2}F_{1}(b+1-c_{i},a+1-c_{i};2-c_{i};x), \\ G^{(i)}_{2}(x)&:={}_{3}F_{2}(1-b,1-a,2-\gamma ;2-c_{i},3-\gamma ;x), \\ a&:= \frac{- (1-\frac{2\kappa }{\sigma ^{2}} )+\sqrt{ (1-\frac{2\kappa }{\sigma ^{2}} )^{2}-\frac{8\theta }{\sigma ^{2}}}}{2}, \\ b&:= \frac{- (1-\frac{2\kappa }{\sigma ^{2}} )-\sqrt{ (1-\frac{2\kappa }{\sigma ^{2}} )^{2}-\frac{8\theta }{\sigma ^{2}}}}{2}, \\ c_{i}& :=\frac{2\kappa w_{i} }{\sigma ^{2}}, \end{aligned}$$

\(\Gamma ^{(i)}_{1}:= \frac{\Gamma (a+b+1-c_{i})\Gamma (1-c_{i})}{\Gamma (a+1-c_{i})\Gamma (b+1-c_{i})}\), \(\Gamma ^{(i)}_{2}:= \frac{\Gamma (a+b+1-c_{i})\Gamma (c_{i}-1)}{\Gamma (a)\Gamma (b)}\), and the Wronskian constant \(\omega ^{(i)}\)is

$$ \omega ^{(i)}:=\big(\varphi ^{(i)} (y) (F_{1}^{(i)} )' (y)- (\varphi ^{(i)} )'(y)F_{1}^{(i)}(y)\big)(1-y)^{\frac{2\kappa (1-w_{i})}{\sigma ^{2}}} y^{ \frac{2\kappa w_{i} }{\sigma ^{2}}} $$

with \(\varphi ^{(i)}(x)={}_{2}F_{1}(a,b;a+b+1-c_{i};1-x)\)and the hypergeometric functions \({}_{2}F_{1}\)and \({}_{3}F_{2}\)given by

$$\begin{aligned} {}_{2}F_{1}(a,b;c;x)&:=\sum ^{\infty }_{n=0} \frac{(a)_{n} (b)_{n}}{(c)_{n} }\frac{x^{n}}{n!}, \\ {}_{3}F_{2}(a,b,p;c,q;x)&:=\sum ^{\infty }_{n=0} \frac{(a)_{n} (b)_{n} (p)_{n}}{(c)_{n} (q)_{n}}\frac{x^{n}}{n!}, \end{aligned}$$

where \((p)_{n}:=p (p+1) \cdots (p+n-2) (p+n-1)\).

Proof

The proof focuses on region 1, as the formulae for region 2 follow by replacing \(x\) with \(1-x\) and \(w_{1}\) with \(w_{2}\). First, we can characterise the function \(f^{(1)}\) in terms of the density function and speed measure of the process \(X\): we have

$$\begin{aligned} \mathbb{E}\bigg[\int ^{\infty }_{0} e^{-\theta s} X_{s}^{1-\gamma } ds \bigg\vert X_{0}=x\bigg] =&\int ^{\infty }_{0} e^{-\theta s} \bigg( \int ^{1}_{0} y^{1-\gamma } p(s;x,y) m(y) dy \bigg) ds \\ =& \int ^{1}_{0} y^{1-\gamma } \bigg(\int ^{\infty }_{0} e^{-\theta s} p(s;x,y) ds \bigg) m(y) dy, \end{aligned}$$
(A.7)

where \(p(s;x,y)\) is the transition density of the Markov process \(X_{t}\) at \(y\), conditionally on \(X_{0}=x\), with respect to the speed measure \(m(dy)=m(y)dy\). The representation of the Green function for scalar diffusions (e.g. Borodin and Salminen [9, II.10–11]) yields \({m(x)=\frac{2}{\sigma ^{2}}x^{c_{1}-1}(1-x)^{a+b-c_{1}}}\) and

$$\begin{aligned} \int ^{\infty }_{0} e^{-\theta s} p(s;x,y) ds= \textstyle\begin{cases} \frac{1}{\omega ^{(1)}} F_{1}^{(1)}(x)\varphi ^{(1)}(y),&\quad x\leq y, \\ \frac{1}{\omega ^{(1)}} F_{1}^{(1)}(y)\varphi ^{(1)}(x), &\quad x \geq y, \end{cases}\displaystyle \end{aligned}$$

where \(F_{1}^{(1)}\) and \(\varphi ^{(1)}\) are the fundamental solutions of the ordinary differential equation

$$ (x-x^{2} )g''(x)+\frac{2\mu }{\sigma ^{2}} (w_{1}-x) g'(x)= \frac{2\theta }{\sigma ^{2}} g(x) $$

with the respective boundary conditions

$$ \textstyle\begin{array}{rlrl} F_{1}^{(1)}(0+)&>0, \qquad &(F_{1}^{(1)} )'(0+) s(0+)&=0, \\ \varphi ^{(1)}(0+)&=+\infty , \qquad &(\varphi ^{(1)} )'(0+) s(0+)&>- \infty , \\ \varphi ^{(1)}(1-)&>0, \qquad &(\varphi ^{(1)} )'(1-) s(1-)&=0, \\ F_{1}^{(1)}(1-)&=+\infty , \qquad & (F_{1}^{(1)} )'(1-) s(1-)&>- \infty , \end{array} $$

where \(s(x):=\frac{2}{\sigma ^{2}}x^{c_{1}-1}(1-x)^{a+b-c_{1}}=x(1-x)m(x)\). Thus (A.7) further simplifies to

$$\begin{aligned} &\mathbb{E}\bigg[\int ^{\infty }_{0} e^{-\theta s} X_{s}^{1-\gamma } ds \bigg\vert X_{0}=x\bigg] \\ &= \frac{2}{\sigma ^{2} \omega ^{(1)}} \bigg({}_{2}F_{1}(a,b;a+b+1-c_{1};1-x) \\ &\phantom{=}\qquad \qquad {} \times \int ^{x}_{0} {}_{2}F_{1}(a,b;c_{1};y) (1-y)^{a+b-c_{1}} y^{c_{1}-\gamma } dy \\ &\phantom{=}\qquad \qquad {} + {}_{2}F_{1}(a,b;c_{1};x) \\ &\phantom{=}\qquad \qquad \quad {}\times \int ^{1}_{x} {}_{2}F_{1}(a,b;a+b+1-c_{1};1-y) (1-y)^{a+b-c_{1}} y^{c_{1}-\gamma } dy\bigg). \end{aligned}$$

Note that Assumption 4.2 ensures that the first integral on the right-hand side converges. Applying the identities for hypergeometric functions in Abramowitz and Stegun [1, Chap. 15.3], the expression for \(f^{(1)}\) simplifies to

$$\begin{aligned} &\mathbb{E}\bigg[\int ^{\infty }_{0} e^{-\theta s} X_{s}^{1-\gamma } ds \bigg\vert X_{0}=x\bigg] \\ &= \frac{2}{\sigma ^{2} \omega ^{(1)}}\bigg(\Gamma _{1}^{(1)} {}_{2}F_{1}(a,b;c_{1};x) \int ^{x}_{0} {}_{2}F_{1}(a,b;c_{1};y) (1-y)^{a+b-c_{1}} y^{c_{1}- \gamma } dy \\ &\qquad \qquad \quad {} + \Gamma _{2}^{(1)} x^{1-c_{1}} {}_{2}F_{1}(b+1-c_{1},a+1-c_{1};2-c_{1};x) \\ &\qquad \qquad \qquad {} \times \int ^{x}_{0} {}_{2}F_{1}(a,b;c_{1};y) (1-y)^{a+b-c_{1}} y^{c_{1}-\gamma } dy \\ &\qquad \qquad \quad {}+ \Gamma _{1}^{(1)} {}_{2}F_{1}(a,b;c_{1};x) \int ^{1}_{x} {}_{2}F_{1}(a,b;c_{1};y) (1-y)^{a+b-c_{1}} y^{c_{1}- \gamma } dy \\ &\qquad \qquad \quad {} + \Gamma _{2}^{(1)} {}_{2}F_{1}(a,b;c_{1};x) \\ &\qquad \qquad \qquad {} \times \int ^{1}_{x} {}_{2}F_{1}(b+1-c_{1},a+1-c_{1};2-c_{1};y) (1-y)^{a+b-c_{1}} y^{1-\gamma } dy\bigg) \\ & = \frac{2}{\sigma ^{2} \omega ^{(1)}}\bigg(\Gamma _{1}^{(1)} {}_{2}F_{1}(a,b;c_{1};x) \int ^{1}_{0} {}_{2}F_{1}(c_{1}-a,c_{1}-b;c_{1};y) y^{c_{1}-\gamma } dy \\ & \qquad \qquad \quad {} + \Gamma _{2}^{(1)} x^{1-c_{1}} {}_{2}F_{1}(b+1-c_{1},a+1-c_{1};2-c_{1};x) \\ &\qquad \qquad \qquad {} \times \int ^{x}_{0} {}_{2}F_{1}(c_{1}-a,c_{1}-b;c_{1};y) y^{c-\gamma } dy \\ & \qquad \qquad \quad {} + \Gamma _{2}^{(1)} {}_{2}F_{1}(a,b;c_{1};x) \int ^{1}_{x} {}_{2}F_{1}(1-b,1-a;2-c_{1};y) y^{1-\gamma } dy\bigg) \\ & = \frac{2}{\sigma ^{2} \omega ^{(1)}}\bigg( \frac{\Gamma _{1}^{(1)} {}_{2}F_{1}(a,b;c_{1};x)}{1+c_{1}-\gamma } {}_{3}F_{2}(c_{1}-a,c_{1}-b,1+c_{1}- \gamma ;c_{1},2+c_{1}-\gamma ;1) \\ &\qquad \qquad \quad {} + \Gamma _{2}^{(1)} \frac{x^{2-\gamma }}{1+c_{1}-\gamma } {}_{2}F_{1}(b+1-c_{1},a+1-c_{1};2-c_{1};x) \\ & \qquad \qquad \qquad {} \times {}_{3}F_{2}(c_{1}-a,c_{1}-b,1+c_{1}- \gamma ;c_{1},2+c_{1}-\gamma ;x) \\ &\qquad \qquad \quad {} + \Gamma _{2}^{(1)}\frac{1}{2-\gamma } {}_{2}F_{1}(a,b;c_{1};x) \\ &\qquad \qquad \qquad {} \times \Big({}_{3}F_{2}(1-b,1-a,2- \gamma ;2-c_{1},3-\gamma ;1) \\ & \qquad \qquad \qquad \qquad {} - x^{2-\gamma } {}_{3}F_{2}(1-b,1-a,2- \gamma ;2-c_{1},3-\gamma ;x) \Big)\Bigg), \end{aligned}$$

which yields the claim for \(f^{(1)}\). □

1.3 A.3 Proof of Theorem 4.3

(i) Note that

$$\begin{aligned} \bar{P}^{(1)}_{t} =& \mathbb{E}_{t}\bigg[\int ^{\infty }_{t} e^{-\beta (s-t)} \bigg(\frac{D_{s}}{D_{t}}\bigg)^{1-\gamma } \frac{D_{t}}{D_{s}} D^{(1)}_{s} ds\bigg] \\ =&D_{t} \, \mathbb{E}_{t}\bigg[\int ^{\infty }_{t} e^{-\beta (s-t)} \bigg(\frac{D_{s}}{D_{t}}\bigg)^{1-\gamma } X_{s} ds\bigg] \\ =& D_{t}\int ^{\infty }_{t} e^{-\beta (s-t)} \mathbb{E}_{t}\bigg[ \bigg(\frac{D_{s}}{D_{t}}\bigg)^{1-\gamma }\bigg] \mathbb{E}_{t} [X_{s} ] ds \\ =&D_{t}\int ^{\infty }_{t} e^{-\theta (s-t)} \mathbb{E}_{t} [X_{s} ] ds \\ =&D_{t}\int ^{\infty }_{t} e^{-\theta (s-t)} \mathbb{E} [X_{s}\vert X_{t} ] ds =D_{t}f(X_{t}), \end{aligned}$$

where the fifth equality follows by the Markov property of \((X_{t})_{t \geq 0}\) and \(f\) is defined as

$$ f(x) :=\int ^{\infty }_{t} e^{-\theta (s-t)} \mathbb{E}\left [X_{s} \vert X_{t}=x\right ] ds= \int ^{\infty }_{0} e^{-\theta s} \mathbb{E} \left [X_{s} \vert X_{0}=x\right ] ds, $$
(A.8)

where the second equality again uses the Markov property. Integrating (3.2) and taking the expectation yields \(\mathbb{E}[X_{t}\vert X_{0}=x]=e^{-\kappa t} x+w_{1}(1-e^{-\kappa t})\). Therefore, (A.8) further simplifies to

$$ f(x) =\int ^{\infty }_{0} e^{-\theta s} \big(e^{-\kappa s} x+w_{1} (1-e^{- \kappa s} )\big) ds=\frac{1}{\theta +\kappa }x+ \frac{\kappa }{(\theta +\kappa )\theta }w_{1} , $$

whence the price of asset 1 is

$$\begin{aligned} \bar{P}^{(1)}_{t}&= D_{t} f(X_{t}) \\ &= \frac{1}{\theta +\kappa }X_{t} D_{t} + \frac{\kappa }{(\theta +\kappa )\theta }w_{1} D_{t} \\ &= \frac{1}{\theta +\kappa } D^{(1)}_{t}+ \frac{\kappa w_{1}}{(\theta +\kappa )\theta } (D^{(1)}_{t}+D^{(2)}_{t} ) \\ &=\frac{1}{\theta }\bigg(\frac{\theta + \kappa w_{1}}{\theta +\kappa } D_{t}^{(1)} + \frac{\kappa w_{1}}{\theta +\kappa } D_{t}^{(2)}\bigg), \end{aligned}$$

and the price of asset 2 follows analogously.

As \(e^{\int ^{t}_{0} {\bar{r}_{s}} ds}\bar{M}_{t}=e^{\int ^{t}_{0} ({ \bar{r_{s}}}-\beta ) ds}D_{t}^{-\gamma }\), \(t \geq 0\), is a local martingale and

$$\begin{aligned} &d \big(e^{\int ^{t}_{0} (\bar{r_{s}}-\beta ) ds}D_{t}^{-\gamma } \big) \\ &=\left (\bar{r_{t}}-\beta \right ) e^{\int ^{t}_{0} (\bar{r_{s}}- \beta ) ds} D_{t}^{-\gamma } dt \\ &\phantom{=:}- \gamma e^{\int ^{t}_{0} (\bar{r_{s}}-\beta ) ds} D_{t}^{-\gamma } \bigg(\mu +\frac{(- \gamma -1) \sigma ^{2}}{2}\bigg)dt+ dL_{t} \\ &= e^{\int ^{t}_{0} (\bar{r_{s}}-\beta ) ds} D_{t}^{-\gamma } \bigg( \bar{r_{t}}-\beta -\gamma \Big(\mu -\frac{(\gamma +1)\sigma ^{2}}{2} \Big)\bigg) dt+ dL_{t}, \end{aligned}$$

for some local martingale \(L\), it follows that \(\bar{r_{t}}=\beta +\gamma \mu - \frac{\gamma (\gamma +1)\sigma ^{2}}{2}\).

(ii) For the welfare, note that

$$\begin{aligned} \overline{\mathcal{U}}_{t} =& \mathbb{E}_{t}\bigg[\int ^{\infty }_{t} e^{- \beta (s-t)}\frac{c_{s}^{1-\gamma }}{1-\gamma } ds\bigg] =\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} e^{-\beta (s-t)} \frac{ D_{s}^{1-\gamma }}{1-\gamma } ds\bigg] \\ =&\frac{D_{t}^{1-\gamma }}{1-\gamma }\mathbb{E}_{t}\bigg[\int ^{\infty }_{t} e^{-\beta (s-t)} \left (\frac{D_{s}}{D_{t}}\right )^{1-\gamma } ds \bigg] \\ =&\frac{D_{t}^{1-\gamma }}{1-\gamma }\int ^{\infty }_{t} e^{-\beta (s-t)} \mathbb{E}_{t}\bigg[\bigg(\frac{D_{s}}{D_{t}}\bigg)^{1-\gamma } \bigg] ds \\ =&\frac{D_{t}^{1-\gamma }}{1-\gamma }\int ^{\infty }_{t} e^{-\theta (s-t)} ds =\frac{D_{t}^{1-\gamma }}{1-\gamma }\frac{1}{\theta }. \end{aligned}$$

(iii) In the integrated market, recalling that \(a_{0}\) and \(b_{i}\) are defined in Theorem 4.3, we have \(\bar{P}^{(1)}_{t}=a_{0} D_{t} X_{t}+b_{1} D_{t}\) and \(\bar{P}^{(2)}_{t}=a_{0} D_{t} (1-X_{t})+b_{2} D_{t}\). By Itô’s formula,

$$\begin{aligned} d\bar{P}^{(1)}_{t}&= \big(\mu (a_{0} X_{t}+b_{1}) D_{t} +a_{0} \kappa (w_{1}-X_{t}) D_{t}\big)dt \\ &\phantom{=:}+\sigma (a_{0} X_{t}+b_{1}) D_{t} dB^{D}_{t}+\sigma a_{0} D_{t} \sqrt{X_{t}(1-X_{t})} dB^{X}_{t}, \end{aligned}$$
(A.9)
$$\begin{aligned} d\bar{P}^{(2)}_{t}&= \Big(\mu \big(a_{0} (1-X_{t})+b_{2}\big) D_{t} -a_{0} \kappa (w_{1}-X_{t}) D_{t}\Big)dt \\ &\phantom{=:}+\sigma \big(a_{0} (1-X_{t})+b_{2}\big) D_{t} dB^{D}_{t}-\sigma a_{0} D_{t} \sqrt{X_{t}(1-X_{t})} dB^{X}_{t}, \end{aligned}$$
(A.10)

whence

$$\begin{aligned} \frac{d \langle \bar{P}^{(1)} \rangle _{t}}{dt}&=\sigma ^{2} D^{2}_{t} \big( (a_{0} X_{t}+b_{1} )^{2}+a_{0}^{2} X_{t}(1-X_{t})\big), \\ \frac{d \langle \bar{P}^{(2)} \rangle _{t}}{dt}&=\sigma ^{2} D^{2}_{t} \Big( \big(a_{0} (1-X_{t})+b_{2} \big)^{2}+a_{0}^{2} X_{t}(1-X_{t}) \Big), \\ \frac{d \langle \bar{P}^{(1)}, \bar{P}^{(2)} \rangle _{t}}{dt}&=\sigma ^{2} D^{2}_{t} \Big( (a_{0} X_{t}+b_{1} )\big(a_{0} (1-X_{t})+b_{2}\big)-a_{0}^{2} X_{t}(1-X_{t})\Big), \end{aligned}$$

and the claim follows by (4.5).

(iv) Equations (A.9) and (A.10) yield

$$\begin{aligned} \frac{d\bar{P}^{(1)}_{t}}{\bar{P}^{(1)}_{t}} =& \bigg(\mu + \frac{a_{0} \kappa (w_{1}-X_{t})}{a_{0} X_{t}+b_{1}} \bigg)dt+\sigma dB^{D}_{t}+ \sigma \frac{a_{0} \sqrt{X_{t}(1-X_{t})}}{a_{0} X_{t}+b_{1}} dB^{X}_{t}, \\ \frac{d\bar{P}^{(2)}_{t}}{\bar{P}^{(2)}_{t}} =& \bigg(\mu - \frac{a_{0} \kappa (w_{1}-X_{t})}{a_{0} (1-X_{t})+b_{2}} \bigg)dt+ \sigma dB^{D}_{t}-\sigma \frac{a_{0} \sqrt{X_{t}(1-X_{t})}}{a_{0} (1-X_{t})+b_{2}} dB^{X}_{t}. \end{aligned}$$

From Theorem 4.3 (iii), recall that \(a_{0}=\frac{1}{\theta +\kappa }\), \(b_{1}=\frac{\kappa w_{1}}{\theta (\theta +\kappa )}\) and \(b_{2}=\frac{\kappa w_{2}}{\theta (\theta +\kappa )}\). Thus

$$\begin{aligned} a_{0} \kappa w_{1}&= \frac{\kappa w_{1}}{\theta +\kappa }=\theta b_{1}, \qquad a_{0} \kappa (1-w_{1})= \frac{\kappa w_{2}}{\theta +\kappa }= \theta b_{2}, \\ 1-a_{0} \kappa &= \frac{\theta }{\theta +\kappa } = \theta a_{0}. \end{aligned}$$

Therefore we get

$$\begin{aligned} a_{0} \kappa (w_{1}-X_{t})+X_{t}&= a_{0} \kappa w_{1} +X_{t}(1-a_{0} \kappa )= \theta (b_{1}+a_{0} X_{t}), \\ -a_{0} \kappa (w_{1}-X_{t})+1-X_{t}&= \theta a_{0}+ \theta b_{2} - X_{t} \theta a_{0} = \theta \big(a_{0}(1-X_{t})+b_{2}\big), \end{aligned}$$

so that the expected returns are

$$\begin{aligned} \bar{\mu }^{(1)}_{t} =&\mu + \frac{a_{0} \kappa (w_{1}-X_{t})}{a_{0} X_{t}+b_{1}}+ \frac{X_{t}}{a_{0} X_{t}+b_{1}}=\mu +\theta , \\ \bar{\mu }^{(2)}_{t} =& \mu - \frac{a_{0} \kappa (w_{1}-X_{t})}{a_{0} (1-X_{t})+b_{2}}+ \frac{1-X_{t}}{a_{0} (1-X_{t})+b_{2}}=\mu +\theta . \end{aligned}$$

 □

1.4 A.4 Proof of Proposition 5.1

The proof focuses on region 1, as region 2 is analogous.

In the segmented market, the pricing formula in Proposition A.3 yields

$$\begin{aligned} &\lim _{T\rightarrow \infty }\mathbb{E}\bigg[ \frac{P_{T}^{(1)}}{D_{T}^{(1)}}\bigg] \\ & = \lim _{T \rightarrow \infty }\mathbb{E} [X_{T}^{\gamma -1}f^{(1)}(X_{T}) ]= \int ^{1}_{0} x^{\gamma -1}f^{(1)}(x) m(x) dx \\ &= \int ^{1}_{0} \frac{2 x^{c_{1}+\gamma -2} (1-x)^{a+b-c_{1}}}{\sigma ^{2} \omega ^{(1)} B(a+b-c_{1}+1,c_{1})} \\ &\phantom{=:} \qquad \times \bigg(\Gamma _{1}^{(1)} {}_{2}F_{1}(a,b;c_{1};x)\int ^{1}_{0} {}_{2}F_{1}(c_{1}-a,c_{1}-b;c_{1};y) y^{c_{1}-\gamma } dy \\ &\phantom{=:}\qquad \qquad + \Gamma _{2}^{(1)} x^{1-c_{1}} {}_{2}F_{1}(b+1-c_{1},a+1-c_{1};2-c_{1};x) \\ &\phantom{=:}\qquad \qquad \quad \times \int ^{x}_{0} {}_{2}F_{1}(c_{1}-a,c_{1}-b;c_{1};y) y^{c_{1}-\gamma } dy \\ & \phantom{=:}\qquad \qquad + \Gamma _{2}^{(1)} {}_{2}F_{1}(a,b;c_{1};x) \int ^{1}_{x} {}_{2}F_{1}(1-b,1-a;2-c_{1};y) y^{1-\gamma } dy\bigg) dx \\ & = \frac{2}{\sigma ^{2} \omega ^{(1)} B(a+b-c_{1}+1,c_{1})} \\ &\phantom{=:} \times \bigg(\Gamma _{1}^{(1)} \int ^{1}_{0} {}_{2}F_{1}(c_{1}-a,c_{1}-b;c_{1};x) x^{c_{1}+\gamma -2} dx \\ & \phantom{=:}\qquad \times \int ^{1}_{0} {}_{2}F_{1}(c_{1}-a,c_{1}-b;c_{1};y) y^{c_{1}- \gamma } dy \\ & \phantom{=:}\qquad + \Gamma _{2}^{(1)} \int ^{1}_{0} x^{\gamma -1} {}_{2}F_{1}(1-b,1-a;2-c_{1};x) \\ & \phantom{=:}\qquad \qquad \qquad {} \times \Big(\int ^{x}_{0} {}_{2}F_{1}(c_{1}-a,c_{1}-b;c_{1};y) y^{c_{1}-\gamma } dy \Big) dx \\ & \phantom{=:}\qquad + \Gamma _{2}^{(1)} \int ^{1}_{0} y^{1-\gamma } {}_{2}F_{1}(1-b,1-a;2-c_{1};y) \\ & \phantom{=:} \qquad \qquad \qquad {} \times \Big( \int ^{y}_{0} x^{c_{1}+ \gamma -2} {}_{2}F_{1}(a,b;c_{1};x) dx\Big)dy\bigg) \\ & = \frac{2}{\sigma ^{2} \omega ^{(1)} B(a+b-c_{1}+1,c_{1})} \\ &\phantom{=:}\times \bigg(\Gamma _{1}^{(1)} \frac{G_{1}^{(1)}(1)}{c_{1}+1-\gamma } \frac{G_{3}^{(1)}(1)}{c_{1}+\gamma -1} \\ &\quad \quad \quad \,\,+ \Gamma _{2}^{(1)} \int ^{1}_{0} x^{c_{1}} F_{3}^{(1)}(x) \Big(\frac{G_{1}^{(1)}(x)}{c_{1}+1-\gamma }+ \frac{G_{3}^{(1)}(x)}{c_{1}+\gamma -1}\Big) dx\bigg). \end{aligned}$$

On the other hand, in the integrated market, Theorem 4.3 implies that

$$\begin{aligned} &\lim _{T\rightarrow \infty }\mathbb{E}\bigg[ \frac{\bar{P}_{T}^{(1)}}{D_{T}^{(1)}}\bigg] \\ &=\lim _{T\rightarrow \infty }\mathbb{E}\left [ \frac{1}{\theta +\kappa }\left (1+\frac{\kappa w_{1}}{\theta } \frac{1}{X_{T}}\right )\right ]= \frac{1}{\theta +\kappa }\bigg(1+ \frac{\kappa w_{1}}{\theta }\int ^{1}_{0} \frac{1}{x} m(x) dx\bigg) \\ &= \frac{1}{\theta +\kappa }\bigg(1+\frac{\kappa w_{1}}{\theta }\int ^{1}_{0} \frac{ x^{c_{1}-2} (1-x)^{a+b-c_{1}}}{ B(a+b-c_{1}+1,c_{1})} dx \bigg)=\frac{1}{\theta +\kappa }\left (1+\frac{\kappa w_{1}}{\theta } \frac{a+b}{c_{1}-1}\right ) \\ &=\frac{1}{\theta +\kappa }\bigg(1+\frac{\kappa w_{1}}{\theta } \frac{2\frac{\kappa }{\sigma ^{2}}-1}{2\frac{\kappa w_{1}}{\sigma ^{2}}-1} \bigg), \end{aligned}$$

where \(m(x):=\frac{ x^{c_{1}-1} (1-x)^{a+b-c_{1}}}{ B(a+b-c_{1}+1,c_{1})}\).  □

1.5 A.5 Proof of Proposition 5.2

The real exchange rate is the ratio between the stochastic discount factors in the two countries, i.e., \(p_{t}=\frac{M_{t}^{(1)}}{M_{t}^{(2)}}\). Under segmentation, Proposition A.2 yields that \(p_{t}= \frac{e^{-\beta t} (D^{(1)}_{t})^{-\gamma }}{e^{-\beta t} (D^{(2)}_{t})^{-\gamma }}\), and Proposition 5.2 follows by recalling that \(D^{(1)}_{t}=D_{t} X_{t}\) and \(D^{(2)}_{t}=D_{t} (1-X_{t})\).  □

1.6 A.6 Proof of Proposition 6.1

Because of the market clearing conditions (6.1)–(6.3) and the obtained optimal consumption (6.4), the stochastic discount factor \(\bar{M}_{t}\) solves

$$ D_{t}=c^{(1)}_{t}+c^{(2)}_{t}= (y^{(1)}e^{\beta _{1} t} \bar{M}_{t} )^{- \frac{1}{\gamma }} + (y^{(2)}e^{\beta _{2} t} \bar{M}_{t} )^{- \frac{1}{2\gamma }}. $$

Solving this quadratic equation gives

$$\begin{aligned} \bar{M}_{t}^{-\frac{1}{2\gamma }} =& \frac{- (y^{(2)}e^{\beta _{2} t} )^{-\frac{1}{2\gamma }}+\sqrt{ (y^{(2)}e^{\beta _{2} t} )^{-\frac{1}{\gamma }}+4D_{t} (y^{(1)}e^{\beta _{1} t} )^{-\frac{1}{\gamma }} }}{2 (y^{(1)}e^{\beta _{1} t} )^{-\frac{1}{\gamma }}} \\ =& \frac{ (y^{(2)} )^{-\frac{1}{2\gamma }}}{2 (y^{(1)} )^{-\frac{1}{\gamma }}}e^{ (\frac{\beta _{1}}{\gamma }-\frac{\beta _{2}}{2\gamma } )t}\bigg(-1+ \sqrt{1+ \frac{4 (y^{(1)} )^{-\frac{1}{\gamma }}}{ (y^{(2)} )^{-\frac{1}{\gamma }}} D_{t} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }t}}\bigg). \end{aligned}$$

With the stochastic discount factor \(\bar{M}_{t}\), the wealth process \(Y^{(i)}\) of agent \(i\) is

$$\begin{aligned} \bar{M}_{t} Y_{t}^{(i)}= \mathbb{E}_{t} \left [\int ^{\infty }_{t} \bar{M}_{u} c_{u}^{(i)} du \right ]. \end{aligned}$$

Thus the ratio of wealths becomes

$$\begin{aligned} \frac{Y_{t}^{(2)}}{Y_{t}^{(1)}} =& \frac{\mathbb{E}_{t} [\int ^{\infty }_{t} \bar{M}_{u} c_{u}^{(2)} du ]}{\mathbb{E}_{t} [\int ^{\infty }_{t} \bar{M}_{u} c_{u}^{(1)} du ]}= \frac{ (y^{(2)} )^{-\frac{1}{2\gamma }}}{ (y^{(1)} )^{-\frac{1}{\gamma }}} \frac{\mathbb{E}_{t} [\int ^{\infty }_{t} \bar{M}_{u}^{1-\frac{1}{2\gamma }} e^{-\frac{\beta _{2}}{2\gamma }u} du ]}{\mathbb{E}_{t} [\int ^{\infty }_{t} \bar{M}_{u}^{1-\frac{1}{\gamma }} e^{-\frac{\beta _{1}}{\gamma }u} du ]} \\ =& 2 \frac{\mathbb{E}_{t} [\int ^{\infty }_{t} (-1+\sqrt{1+K D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u}} )^{1-2\gamma } e^{ (\frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1-\gamma ) )u} du ]}{\mathbb{E}_{t} [\int ^{\infty }_{t} (-1+\sqrt{1+K D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u}} )^{2-2\gamma } e^{ (\frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1-\gamma ) )u} du ]}, \qquad \quad \end{aligned}$$
(A.11)

where \(K\) is a positive constant depending on \(y^{(i)}\) and \(\gamma \) only.

To establish an upper bound (respectively, lower bound) for the expectation in the numerator (respectively, denominator), consider first the case of \(\mu \!-\!\frac{1}{2}\sigma ^{2}\!+\!\frac{\beta _{2}\!-\!\beta _{1}}{\gamma }>\!0\). On the one hand,

$$\begin{aligned} &\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \bigg(-1+\sqrt{1+K D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u}}\bigg)^{1-2\gamma } e^{ ( \frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1- \gamma ) )u} du \bigg] \\ &\leq \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \bigg( \frac{2+\sqrt{K D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u}}}{K D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u}} \bigg)^{2\gamma -1} e^{ (\frac{\beta _{1}}{\gamma }(1-2\gamma )- \frac{\beta _{2}}{\gamma }(1-\gamma ) )u} du \bigg] \\ &\leq 2^{2\gamma -2}\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \Big( 2^{2 \gamma -1} \big(K D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u} \big)^{1-2\gamma } + \big(K D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u} \big)^{\frac{1-2\gamma }{2}} \Big) \\ & \qquad \qquad \qquad \qquad {}\times e^{ ( \frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1- \gamma ) )u} du \bigg] \\ &= \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \big( K_{1} e^{-\beta _{2} u} D_{u}^{1-2\gamma } +K_{2} e^{ (\frac{\beta _{1}}{2\gamma }(1-2\gamma )- \frac{\beta _{2}}{2\gamma } )u} D_{u}^{\frac{1-2\gamma }{2}} \big) du \bigg] \\ &= K_{1} D_{t}^{1-2\gamma } \int ^{\infty }_{t} e^{-\beta _{2} u} e^{ ((1-2 \gamma )\mu - \gamma (1-2\gamma ) \sigma ^{2} )(u-t)} du \\ &\phantom{=:}+ K_{2} D_{t}^{\frac{1-2\gamma }{2}} \int ^{\infty }_{t} e^{ ( \frac{\beta _{1}}{2\gamma }(1-2\gamma )-\frac{\beta _{2}}{2\gamma } )u} e^{ (\frac{1-2\gamma }{2}\mu - \frac{1}{2} \frac{1-2\gamma }{2} \frac{1+2\gamma }{2} \sigma ^{2} )(u-t)} du \\ &= K_{1} e^{-\beta _{2} t} D_{t}^{1-2\gamma }+ K_{2} e^{ ( \frac{\beta _{1}}{2\gamma }(1-2\gamma )-\frac{\beta _{2}}{2\gamma } )t} D_{t}^{ \frac{1-2\gamma }{2}}, \end{aligned}$$
(A.12)

where \(K_{1}\) and \(K_{2}\) are some positive constants independent of \(t\). Above, the first inequality follows from \((\sqrt{1+x}-1)^{-\delta }=(\frac{\sqrt{1+x}+1}{x})^{\delta }\leq ( \frac{2+\sqrt{x}}{x})^{\delta }\) for all \(x>0\), \(\delta \geq 0\), and the second inequality from Jensen’s inequality \((x+y)^{\delta } \leq 2^{\delta -1 }(x^{\delta }+y^{\delta })\) for all \(x,y \geq 0\), \(\delta \geq 1\). On the other hand,

$$\begin{aligned} &\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \bigg(-1+\sqrt{1+K D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u}}\bigg)^{2-2\gamma } e^{ ( \frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1- \gamma ) )u} du \bigg] \\ &\geq \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \big(K D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u}\big)^{1-\gamma } e^{ ( \frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1- \gamma ) )u} du \bigg] \\ &= K^{1-\gamma }\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} e^{-\beta _{1} u} D_{u}^{1-\gamma } du \bigg] = K_{3} e^{-\beta _{1} t} D_{t}^{1-\gamma }, \end{aligned}$$
(A.13)

where \(K_{3}\) is a positive constant independent of \(t\) and the inequality follows from \(-1+\sqrt{1+x} \leq \sqrt{x}\). With (A.12) and (A.13), (A.11) becomes

$$\begin{aligned} \frac{Y_{t}^{(2)}}{Y_{t}^{(1)}}&\leq 2 \frac{K_{1} e^{-\beta _{2} t} D_{t}^{1-2\gamma }+ K_{2} e^{ (\frac{\beta _{1}}{2\gamma }(1-2\gamma )-\frac{\beta _{2}}{2\gamma } )t} D_{t}^{\frac{1-2\gamma }{2}}}{K_{3} e^{-\beta _{1} t} D_{t}^{1-\gamma }} \\ &= K_{1} e^{(\beta _{1}-\beta _{2})t}D_{t}^{-\gamma }+K_{2} e^{ \frac{\beta _{1}-\beta _{2}}{2\gamma }t} D_{t}^{-\frac{1}{2}} \\ &= K_{1} D_{0}^{-\gamma } e^{-\gamma (\mu -\frac{1}{2}\sigma ^{2}+ \frac{\beta _{2}-\beta _{1}}{\gamma } )t-\gamma \sigma B^{D}_{t}}+K_{2} D_{0}^{-\frac{1}{2}} e^{-\frac{1}{2} (\mu -\frac{1}{2}\sigma ^{2}+ \frac{\beta _{2}-\beta _{1}}{\gamma } )t-\frac{1}{2} \sigma B^{D}_{t}}, \end{aligned}$$

which tends to zero as \(t \rightarrow \infty \) because \(\mu -\frac{1}{2}\sigma ^{2}+\frac{\beta _{2}-\beta _{1}}{\gamma }>0\).

Now consider the case of \(\mu -\frac{1}{2}\sigma ^{2}+\frac{\beta _{2}-\beta _{1}}{\gamma }<0\). On the one hand,

$$\begin{aligned} &\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \bigg(-1+\sqrt{1+K D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u}}\bigg)^{2-2\gamma } e^{ ( \frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1- \gamma ) )u} du \bigg] \\ &\leq \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \bigg( \frac{2+\sqrt{K D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u}}}{K D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u}} \bigg)^{2\gamma -2} e^{ (\frac{\beta _{1}}{\gamma }(1-2\gamma )- \frac{\beta _{2}}{\gamma }(1-\gamma ) )u} du \bigg] \\ &\leq \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \Big( K_{1} \big( D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u}\big)^{2-2\gamma } + K_{2} \big( D_{u} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }u}\big)^{1-\gamma } \Big) \\ & \phantom{=:}\qquad \qquad {} \times e^{ (\frac{\beta _{1}}{\gamma }(1-2 \gamma )-\frac{\beta _{2}}{\gamma }(1-\gamma ) )u} du \bigg] \\ &= \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \big( K_{1} e^{ ( \frac{\beta _{2}}{\gamma }(1-\gamma )-\frac{\beta _{1}}{\gamma } )u} D_{u}^{2-2 \gamma } + K_{2} e^{-\beta _{1} u} D_{u}^{1-\gamma }\big) du \bigg] \\ &= K_{1} e^{ (\frac{\beta _{2}}{\gamma }(1-\gamma )- \frac{\beta _{1}}{\gamma } )t} D_{t}^{2-2\gamma } + K_{2} e^{-\beta _{1} t} D_{t}^{1-\gamma }, \end{aligned}$$

where \(K_{1}\) and \(K_{2}\) are positive constants independent of \(t\). On the other hand,

$$\begin{aligned} &\mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \bigg(-1+\sqrt{1+K D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u}}\bigg)^{1-2\gamma } e^{ ( \frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1- \gamma ) )u} du \bigg] \\ &\geq \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} \big(K D_{u} e^{ \frac{\beta _{2}-\beta _{1}}{\gamma }u}\big)^{1-2\gamma } e^{ ( \frac{\beta _{1}}{\gamma }(1-2\gamma )-\frac{\beta _{2}}{\gamma }(1- \gamma ) )u} du \bigg] \\ &= K^{1-2\gamma } \mathbb{E}_{t} \bigg[\int ^{\infty }_{t} e^{-\beta _{2} u} D_{u}^{1-2\gamma } du \bigg] = K_{3} e^{-\beta _{2} t} D_{t}^{1-2 \gamma }, \end{aligned}$$

where \(K_{3}\) is a positive constant independent of \(t\) and the inequality holds due to \(-1+\sqrt{1+x}= \frac{x}{1+\sqrt{1+x}} \leq x\) and the fact that \(1-2\gamma <0\). Then (A.11) becomes

$$\begin{aligned} \frac{Y_{t}^{(2)}}{Y_{t}^{(1)}} \geq & 2 \frac{K_{3} e^{-\beta _{2} t} D_{t}^{1-2\gamma }}{K_{1} e^{ (\frac{\beta _{2}}{\gamma }(1-\gamma )-\frac{\beta _{1}}{\gamma } )t} D_{t}^{2-2\gamma } + K_{2} e^{-\beta _{1} t} D_{t}^{1-\gamma }} \\ =& \frac{1}{K_{1} e^{\frac{\beta _{2}-\beta _{1}}{\gamma }t}D_{t}+K_{2} e^{(\beta _{2}-\beta _{1})t} D_{t}^{\gamma }} \\ =& \frac{1}{K_{1} D_{0} e^{ (\mu -\frac{1}{2}\sigma ^{2}+\frac{\beta _{2}-\beta _{1}}{\gamma } )t+ \sigma B^{D}_{t}}+K_{2} D_{0}^{\gamma } e^{\gamma (\mu -\frac{1}{2}\sigma ^{2}+\frac{\beta _{2}-\beta _{1}}{\gamma } )t+\gamma \sigma B^{D}_{t}}}, \end{aligned}$$

which tends to \(+\infty \) as \(t \rightarrow \infty \) because \(\mu -\frac{1}{2}\sigma ^{2}+\frac{\beta _{2}-\beta _{1}}{\gamma }<0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guasoni, P., Wong, K.C. Asset prices in segmented and integrated markets. Finance Stoch 24, 939–980 (2020). https://doi.org/10.1007/s00780-020-00433-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00780-020-00433-4

Keywords

Mathematics Subject Classification (2010)

JEL Classification

Navigation