Skip to main content
Log in

Fraction-Degree Reference Dependent Stochastic Dominance

  • Published:
Methodology and Computing in Applied Probability Aims and scope Submit manuscript

Abstract

For addressing the Allis-type anomalies, a fractional degree reference dependent stochastic dominance rule is developed which is a generalization of the integer degree reference dependent stochastic dominance rules. This new rule can effectively explain why the risk comparison does not satisfy translational invariance and scaling invariance in some cases. The rule also has a good property that it is compatible with the endowment effect of risk. This rule can help risk-averse but not absolute risk-averse decision makers to compare risks relative to reference points. We present some tractable equivalent integral conditions for the fractional degree reference dependent stochastic dominance rule, as well as some practical applications for the rule in economics and finance.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  • Bawa S (1982) Research bibliography stochastic dominance: A research bibliography. Manage Sci 28:698–712

    Article  MathSciNet  Google Scholar 

  • Bell DE (1985) Disappointment in Decision Making under Uncertainty. Oper Res 33(1):1–27

    Article  MathSciNet  Google Scholar 

  • Bell DE (1988) The value of pre-decision side bets for utility maximizers. Manage Sci 34:797–800

    Article  Google Scholar 

  • Bi HW, Zhu W (2019) The non-integer higher-order Stochastic dominance. Oper Res Lett 47(2):77–82

    Article  MathSciNet  Google Scholar 

  • Chateauneuf A, Cohen M, Meilijson I (2005) More pessimism than greediness: a characterization of monotone risk aversion in the rank-dependent expected utility model. Econ Theor 25:649–667

    Article  MathSciNet  Google Scholar 

  • Denuit M, Müller A (2002) Smooth generators of integral stochastic orders. Ann Appl Probab 12:1174–1184

    Article  MathSciNet  Google Scholar 

  • Delquié P, Cillo A (2006) Disappointment without prior expectation: a unifying perspective on decision under risk. J Risk Uncertainty 33:197–215

    Article  Google Scholar 

  • Dentcheva D, Ruszczyński A (2003) Optimization with stochastic dominance constraints. SIAM J Optim 14:548–566

    Article  MathSciNet  Google Scholar 

  • Dentcheva D, Ruszczyński A (2006) Portfolio optimization with stochastic dominance constraints. J Bank Finance 30:433–451

    Article  Google Scholar 

  • Dittmann I, Maug E, Spalt O (2010) Sticks or carrots? Optimal CEO compensation when managers are loss averse. J Finance 65(6):2015–2050

    Article  Google Scholar 

  • Fishburn PC (1964) Decision and Value Theory. John Wiley & Sons, New York

    MATH  Google Scholar 

  • Friedman M, Savage LJ (1948) The utility analysis of choices involving risk. J Polit Econ 56:279–304

    Article  Google Scholar 

  • Guo D, Hu Y, Wang S, Zhao L (2016) Comparing risks with reference points: A stochastic dominance approach. Insurance Math Econom 70:105–116

    Article  MathSciNet  Google Scholar 

  • Gollier C (2016) Stochastic volatility implies fourth-degree risk dominance: Applications to asset pricing. TSE Working Papers

  • Hanoch G, Levy H (1969) The efficiency analysis of choices involving risk. Rev Econ Stud 36:335–346

    Article  Google Scholar 

  • Huang R, Tzeng L, Zhao L (2019) Fractional Degree Stochastic Dominance. Management Science forthcoming

  • Kahnemann D, Tversky A (1979) Prospect Theory: An Analysis of Decision under Risk. Econometrica 47(2):263–291

    Article  MathSciNet  Google Scholar 

  • Köszegi B, Rabin M (2006) A Model of Reference-Dependent Preferences. Quart J Econ 121(4):1133–65

    MATH  Google Scholar 

  • Köszegi B, Rabin M (2007) Reference-dependent risk attitudes. Amer Econ Rev 97:1047–1073

    Article  Google Scholar 

  • Levy H (1992) Stochastic dominance and expected utility: Survey and analysis. Manage Sci 38:555–593

    Article  Google Scholar 

  • Machina M (1982) Expected utility analysis without the independence axiom. Econometrica 50:277–323

    Article  MathSciNet  Google Scholar 

  • Mao T, Hu T (2018) Fractional stochastic dominance. Technical Report. University of Science and Technology of China, Hefei

    Google Scholar 

  • Markowitz H (1952) The utility of wealth. J Polit Econ 60:151–158

    Article  Google Scholar 

  • Müller A, Stoyan D (2002) Comparison Methods for Stochastic Models and Risks. John Wiley & Sons, Chichester, UK

    MATH  Google Scholar 

  • Müller A, Scarsini M, Tsetlin I, Winkler R (2017) Between first-and second-Order stochatic dominance. Manage Sci 63:2933–2974

    Article  Google Scholar 

  • Post T, Kopa M (2016) Portfolio choice based on third-degree stochastic dominance. Manage Sci 63(10):3381–3392

    Article  Google Scholar 

  • Quiggin J (1993) Generalized Expected Utility Theory: the Rank-Dependent Model. Kluwer Academic, Boston

    Book  Google Scholar 

  • Rabin M (2000) Risk Aversion and Expected-Utility Theory: A Calibration Theorem. Econometrica 68(5):1281–92

    Article  Google Scholar 

  • Richard TH, Johnson EJ (1990) Gambling with the House Money and Trying to Break Even: The Effects of Prior Outcomes on Risky Choice. Manage Sci 36(6):643–60

    Google Scholar 

  • Rothschild M, Stiglitz JE (1970) Increasing risk. I. A definition. J Econ Theo 2:225–243

  • Segal U (1987) Some remarks on Quiggin’s anticipated utility. J Econ Behav Org 8:145–154

    Article  Google Scholar 

  • Sprenger C (2015) An endowment effect for risk: Experimental tests of stochastic reference points. J Polit Econ 123:1456–1499

    Article  Google Scholar 

  • Sugden R (2003) Reference-Dependent Subjective Expected Utility. J Econ Theo 11(2):172–191

    Article  MathSciNet  Google Scholar 

  • Tversky A, Kahneman D (1991) Loss aversion in risk-less choice: A reference dependent model. Quart J Econ 106:1039–1061

    Article  Google Scholar 

  • von Neumann J, Morgenstern O (1947) Theory of games and economic behavior, 2nd edn. Princeton University Press, Princeton, NJ

    MATH  Google Scholar 

  • Yang J, Han S, Zhang X, Zhou D (2019) Optimization with fractional stochastic dominance constraints. Technical Report. Zhejiang Sci-Tech University, Hangzhou

    Google Scholar 

Download references

Acknowledgements

J. Yang was supported by the NNSF of China (No. 11701518). S. Han was supported by the NNSF of China (No. 12071436).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shuguang Han.

Additional information

Publisher’s Note

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

Appendices

Appendices

1.1 Proof of Theorem 1

Proof

\({\mathrm{[1]} \Rightarrow \mathrm{[2]}}\). Since \(U_{\gamma }^{*}\) is invariant under translations, any \(u\in U_{\gamma }^{*}\) can be approximated by a sequence of functions \(\{u_{n}\in U_{\gamma }, n=1,2,\cdots \}\) as in the proof of Theorem 2.1 in Denuit and Müller (2002). From this result it follows that \({\mathrm{[1]}}\) implies \({\mathrm{[2]}}\).

\({\mathrm{[2]} \Rightarrow \mathrm{[3]}}\). For a fixed \(t\in \mathfrak {R}\), we define the consumption utility function u(xt) with the following right derivative:

$$u'(x;t)=\left\{ \begin{array}{ll} \gamma , &{} x\le t~\text{ and }~ G(x)\le F(x), \\ 1, &{} x\le t~\text{ and }~ G(x)>F(x),\\ 0, &{} x>t. \end{array} \right.$$

Obviously, \(u(x;t)\in U_{\gamma }^{*}\). Due to the integration by part, it holds that

$$\begin{aligned}&E\left[ v(Y,r,u(x;t))\right] -E\left[ v(X,r,u(x;t)\right] \\&~~~=\int _{-\infty }^{\infty }u'(x;t)\left( F(x)-G(x)\right) dx+\eta \lambda \int _{-\infty }^{r}u'(x;t)\left( F(x)-G(x)\right) dx\\&~~~~~~~+\eta \int _{r}^{\infty }u'(x;t)\left( F(x)-G(x)\right) dx\\&~~~=\left( 1+\eta \right) \left[ \int _{-\infty }^{\infty }u'(x;t)\left( F(x)-G(x)\right) f_{\eta , \lambda }(x;r)dx\right] \\&~~~=\left( 1+\eta \right) \left[ \int _{-\infty }^{t}\left[ \gamma \left( F(x)-G(x)\right) _{+}-\left( (G(x)-F(x)\right) _{+}\right] f_{\eta , \lambda }(x;r)dx\right] . \end{aligned}$$

Hence, for all \(t\in \mathfrak {R}\), \(\eta ^{*}>0\) and \(\lambda ^{*}> 1\), \(E\left[ v(Y,r,u(x;t))\right] \ge E\left[ v(X,r,u(x;t))\right]\) implies that

$$\int _{-\infty }^{t}\left[ \gamma \left( F(x)-G(x)\right) _{+}-\left( G(x)-F(x)\right) _{+}\right] f_{\eta ^{*}, \lambda ^{*}}(x;r)dx\ge 0.$$

\({\mathrm{[3]} \Rightarrow \mathrm{[1]}}\). Let \(u \in U_{\gamma }\). Without loss of generality we can assume

$$R: =\sup _{x\in \mathfrak {R}}u'(x)\in (0,\infty ).$$

For any fixed \(n\ge 2\), define \(\varepsilon _n=2^{-n}\) and K as the largest integer k for which

$$R(1-k\varepsilon _n)\ge \inf _{x\in \mathfrak {R}}u'(x),$$

and define a partition of a real line into intervals \([x_{k}, x_{k+1}]\) as follows: let \(x_0=-\infty\), \(x_{K+1}=\infty\) and

$$x_{k}=\sup \left\{ x:u'(x)\ge R(1-k\varepsilon _n)\right\} ,\qquad k=1,...,K .$$

Then we define

$$m_{k}=\sup \left\{ u'(x):x_{k-1}< x \le x_{k}\right\} =R\left[ 1-(k-1)\varepsilon _n\right] .$$

It follows that

$$\gamma (m_k-R\varepsilon _n)\le u'(x)\le m_k, \ \text{ for } x \in (x_{k-1}, x_k], \ k=1,\ldots , K+1.$$

This implies that for all \(0\le k\le K\),

$$\begin{aligned}&\sum _{k=0}^{K}\int _{x_{k}}^{x_{k+1}}\left[ F(x)-G(x)\right] _{+}u'(x)f_{\eta ,\lambda }(x;r)dx\\&~~~\ge \gamma \sum _{k=0}^{K} \left( m_k-R\varepsilon _n\right) \int _{x_{k}}^{x_{k+1}}\left[ F(x)-G(x)\right] _{+}f_{\eta ,\lambda }(x;r)dx. \end{aligned}$$

and

$$\begin{aligned}&\sum _{k=0}^{K}\int _{x_{k}}^{x_{k+1}}\left[ G(x)-F(x)\right] _{+}u'(x)f_{\eta ,\lambda }(x;r)dx\\&~~~\le \sum _{k=0}^{K}m_k\int _{x_{k}}^{x_{k+1}}\left[ G(x)-F(x)\right] _{+}f_{\eta ,\lambda }(x;r)dx. \end{aligned}$$

Let

$$T_k=\int _{x_k}^{x_{k+1}}\left[ \gamma \left( F(x)-G(x)\right) _{+}-\left( G(x)-F(x)\right) _{+}\right] f_{\eta ,\lambda }(x;r)dx,$$

and

$$c_k=\int _{x_k}^{x_{k+1}}\left( F(x)-G(x)\right) _{+}f_{\eta ,\lambda }(x;r)dx.$$

Thus,

$$\begin{aligned} {E[v(Y,r,u)]-E[v(X,r,u)]}= & {} \sum _{k=0}^{K}\int _{x_{k}}^{x_{k+1}}\left[ F(x)-G(x)\right] _{+}u'(x)f_{\eta ,\lambda }(x;r)dx\\&~-\sum _{k=0}^{K}\int _{x_{k}}^{x_{k+1}}\left[ G(x)-F(x)\right] _{+}u'(x)f_{\eta ,\lambda }(x;r)dx\\\ge & {} \sum _{k=0}^{K}m_kT_k-\gamma R \varepsilon _n \sum _{k=0}^{K}c_k. \end{aligned}$$

Set

$$A(x,r)=\frac{f_{\eta ,\lambda }(x,r)}{f_{\eta ^{*},\lambda ^{*}}(x,r)}=\left\{ \begin{array}{ll} \frac{\frac{1+\eta \lambda }{1+\eta }}{\frac{1+\eta ^*\lambda ^{*}}{1+\eta ^{*}}}, &{} x\le r \\ 1, &{} x>r. \end{array} \right.$$

Note that for all \(k=0, \ldots , K+1\),

$$\begin{aligned}&\sum _{i=0}^{k}T_i=\int _{-\infty }^{x_{k+1}}\left[ \gamma \left( F(x)-G(x)\right) _{+}-\left( G(x)-F(x)\right) _{+}\right] f_{\eta ^{*}, \lambda ^{*}}(x,r) A(x,r)dx \end{aligned}$$

and

$$\int _{-\infty }^{x_{k+1}}\left[ \gamma \left( F(x)-G(x)\right) _{+}-\left( G(x)-F(x)\right) _{+}\right] f_{\eta ^{*}, \lambda ^{*}}(x,r)dx\ge 0,$$

and A(xr) is positive and noningcreasing, which implies that \(\sum _{i=0}^{k}T_i \ge 0\). Furthermore, since \(m_k\) is a decreasing non-negative sequences, \(\sum _{i=0}^{k}m_iT_i\ge 0\). Therefore,

$$\begin{aligned} {E[v(Y,r,u)]-E[v(X,r,u)]}&\ge -\gamma R \varepsilon _n \int _{-\infty }^{\infty }\left( F(x)-G(x)\right) _{+}dx. \end{aligned}$$

Letting \(n\rightarrow \infty\) yields part [1] holds. This completes the proof of the theorem.

1.2 Proof of Proposition 1

Proof

Let \(F_1(x)\) and \(F_2(x)\) be two cdfs of Z and W, respectively. The proof is main to construct cdfs \(F_1(x)\) and \(F_2(x)\). When \(x_1\le r\le x_2\), define \(F_1(x)\) and \(F_2(x)\) as

$$\begin{aligned} F_{1}(x)=\left\{ \begin{array}{ll} F(x), &{} x\ge x_2~\text{ or }~ x<x_1 \\ G(x)+\zeta _1^{*}, &{} x_1\le x< x_2, \end{array} \right. \end{aligned}$$
(19)

and

$$\begin{aligned} F_{2}(x)=\left\{ \begin{array}{ll} G(x), &{} x\ge x_2~\text{ or }~ x<x_1 \\ F(x)-\zeta _1^{*}, &{} x_1\le x< x_2, \end{array} \right. \end{aligned}$$
(20)

where \(\zeta _1^{*}\) satisfies that

$$\begin{aligned} \gamma \zeta _1^{*}\left[ \frac{1+\eta ^{*}\lambda ^{*}}{1+\eta ^{*}}(r-x_1)+(x_2-r)\right] =\zeta _2(x_4-x_3) \end{aligned}$$
(21)

Obviously, \(0\le \zeta _1^{*}\le \zeta _1\). Hence, \(F_1(x)\le F(x)\) and \(F_2(x)\ge G(x)\) for all \(x\in \mathfrak {R}\), that is, \(X\le _{1-SD}Z\) and \(W\le _{1-SD} Y\). From (13), (19), we have that Z is simple spread of Z. Combing with (21), it holds that \(Z \le _{(1+\gamma )-SD}^{r, \eta ^{*},\lambda ^{*}}Y\) but not \(Z\le _{(1+\gamma )-SD}Y\). And since X is also a simple spread of W based on (13), (20), \(X\le _{(1+\gamma )-SD}^{r, \eta ^{*},\lambda ^{*}}W\) but not \(X\le _{(1+\gamma )-SD}W\) by (21).

When \(x_2<r\le x_3\), we also define cdfs \(F_1(x)\) as (19) and \(F_2(x)\) as (20). But \(\zeta _1^{*}\) satisfies that

$$\begin{aligned} \gamma \zeta _1^{*}(x_2-x_1)\frac{1+\eta ^{*}\lambda ^{*}}{1+\eta ^{*}}=\zeta _2(x_4-x_3) \end{aligned}$$
(22)

Similarly, we can obtain \(X\le _{1-SD}Z\), \(W\le _{1-SD}Y\), \(Z\le _{(1+\gamma )-SD}^{r, \eta ^{*},\lambda ^{*}}Y\) but not \(Z\le _{(1+\gamma )-SD}Y\) and \(X\le _{(1+\gamma )-SD}^{r, \eta ^{*},\lambda ^{*}}W\) but not \(X\le _{(1+\gamma )-SD}W\).

When \(x_3<r\le x_4\), we define

$$\begin{aligned} F_1(x)=\left\{ \begin{array}{ll} F(x), &{} x<x_3~\text{ or }~x\ge x_4 \\ G(x)-\zeta _2^{*}, &{} x_3\le x< x_4 \end{array} \right. \end{aligned}$$
(23)

and

$$\begin{aligned} F_2(x)=\left\{ \begin{array}{ll} G(x), &{} x<x_3~\text{ or }~x\ge x_4 \\ F(x)-\zeta _2^{*}, &{} x_3\le x< x_4, \end{array} \right. \end{aligned}$$
(24)

where \(\zeta _2^{*}\) satisfies that

$$\begin{aligned} \gamma \zeta _1(x_2-x_1)\frac{1+\eta ^{*}\lambda ^{*}}{1+\eta ^{*}}=\zeta _2^{*}\left[ (r-x_3)\frac{1+\eta ^{*}\lambda ^{*}}{1+\eta ^{*}}+(x_4-r)\right] \end{aligned}$$
(25)

Clearly, \(\zeta _2^{*}\ge \zeta _2\). Hence, \(F_1(x)\le F(x)\) and \(F_2(x)\ge G(x)\) for all \(x\in \mathfrak {R}\). That is, \(X\le _{1-SD}Z\) and \(W\le _{1-SD} Y\). Based on (13), (23), (25), we have \(Z \le _{(1+\gamma )-SD}^{r, \eta ^{*},\lambda ^{*}}Y\) but not \(Z\le _{(1+\gamma )-SD}Y\). And from (13), (24), (25), it holds that \(X\le _{(1+\gamma )-SD}^{r, \eta ^{*},\lambda ^{*}}W\) but not \(X\le _{(1+\gamma )-SD}W\). Combing these three cases, we complete the proof.

1.3 Proof of Proposition 2

Proof

Without loss of generality, we assume \(r_1< r_2\). Let X be a random variable with probability mass function \(p_1(x)\). We take \(x_1< x_2 \le x_3 <x_4\) with \(x_2=r_1\) and \(x_4=r_2\). Continue to take \(\eta _1>0\), \(\eta _2>0\), \(0<\gamma \le 1\) such that \(\gamma \eta _1\left( x_2-x_1\right) \frac{1+\eta ^{*}\lambda ^{*}}{1+\eta ^{*}}=\eta _2(x_4-x_3)\). Define random variable Y with probability mass function \(p_2(x)\) as

$$\begin{aligned} p_2(x_1)= & {} p_1(x_1)-\eta _1,\\ p_2(x_2)= & {} p_1(x_2)+\eta _1,\\ p_2(x_3)= & {} p_1(x_3)+\eta _2,\\ p_2(x_4)= & {} p_1(x_4)-\eta _2,\\ p_2(x)= & {} p_1(x) \qquad \text{ for } \text{ all } \text{ other } \text{ values } \text{ x }. \end{aligned}$$

Let F(x) and G(x) be the cdfs of X and Y, respectively. Obviously, we have

$$F(x)-G(x)=\left\{ \begin{array}{cc} \eta _1, &{} x_1\le x<x_2 \\ -\eta _2, &{} x_3\le x<x_4 \\ 0, &{} \text{ otherwise } \end{array} \right.$$

Therefore, X is a simple spread of Y with a single crossing point \(x_2\). And since \(\gamma \eta _1\left( x_2-x_1\right) \frac{1+\eta ^{*}\lambda ^{*}}{1+\eta ^{*}}=\eta _2(x_4-x_3)\), we have

$$\gamma \int _{-\infty }^{x_2}(F(x)-G(x))f_{\eta ^{*}, \lambda ^{*}}(x,r_1)dx=\int _{x_2}^{\infty }(G(x)-F(x))f_{\eta ^{*}, \lambda ^{*}}(x,r_1)dx$$

Thus, \(X\le _{(1+\gamma )-SD}^{r_1, \eta ^{*},\lambda ^{*}} Y\). And since

$$f_{\eta ^{*}, \lambda ^{*}}(x,r_1)< f_{\eta ^{*}, \lambda ^{*}}(x,r_2)~ \text{ on }~[x_3,x_4]$$

and

$$f_{\eta ^{*}, \lambda ^{*}}(x,r_1)= f_{\eta ^{*}, \lambda ^{*}}(x,r_2)~ \text{ on } \text{ otherwise },$$

we have

$$\begin{aligned} \gamma \int _{-\infty }^{x_2}(F(x)-G(x))f_{\eta ^{*}, \lambda ^{*}}(x,r_2)dx= & {} \gamma \int _{-\infty }^{x_2}(F(x)-G(x))f_{\eta ^{*}, \lambda ^{*}}(x,r_1)dx\\= & {} \int _{x_2}^{\infty }(G(x)-F(x))f_{\eta ^{*}, \lambda ^{*}}(x,r_1)dx\\\le & {} \int _{x_2}^{\infty }(G(x)-F(x))f_{\eta ^{*}, \lambda ^{*}}(x,r_2)dx. \end{aligned}$$

Thus, \(X \nleqslant _{(1+\gamma )-SD}^{r_2,\eta ^{*},\lambda ^{*}} Y\). This completes the proof of proposition 2.

1.4 Proof of Theorem 2

To prove Theorem 2, we need the next lemma.

Lemma 1

Given cdfs F and G, let \(F^{-1}(s)=\inf \{x:F(x)\ge s\}\) and \(G^{-1}(s)=\inf \{x:G(x)\ge s\}\). \(F\le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}} G\) if and only if, for all \(p\in [0,1]\),

$$\begin{aligned} \int _{0}^{p}H_{\gamma ,r}(G^{-1}(s))ds\ge \int _{0}^{p}H_{\gamma , r}(F^{-1}(s))ds \end{aligned}$$
(26)

Proof

We can use the similar proof of Müller et al. (2017). Define the function

$$t\rightarrow h(x):=\int _{-\infty }^{x}\left[ F(t)-G(t)\right] dH_{\gamma ,r}(t).$$

Then, \(F\le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}} G\) if and only if for all x, \(h(x)\ge 0\). Since \(H_{\gamma ,r}^{'}(t)\ge 0\), the function h assumes local minima in the points \(b_k\) where the distribution functions F and G cross, going from \(F\le G\) left of \(b_k\) to \(F>G\) right of \(b_k\). If we define \(p_k:=G(b_k)\) and

$$\tilde{h}(p):=\int _{0}^{p}\left[ H_{\gamma ,r}(G^{-1}(s))-H_{\gamma ,r}(F^{-1}(s))\right] ds,$$

then we have \(h(b_k)=\tilde{h}(p_k)\) and the function \(\tilde{h}\) assumes its local minima in the points \(p_k\). Therefore, \(h\ge 0\) if and only if \(\tilde{h}\ge 0\).

Proof

Based on the idea of Müller et al. (2017), define

$$A_{1}(p):=\int _{0}^{p}\left[ H_{\gamma , r}(G^{-1}(s))-H_{\gamma , r}(F^{-1}(s))\right] _{+}ds$$

and

$$A_2(p):=\int _{0}^{p}\left[ H_{\gamma , r}(F^{-1}(s))-H_{\gamma , r}(G^{-1}(s))\right] _{+}ds.$$

Without the loss of generality, we assume that \(A_{2}(1)>0\). Let \(\alpha (a)\) and \(\beta (a)\) be the smallest probabilities that solve

$$A_{1}(\alpha (a))=a ~~~ \text{ and }~~~ A_{2}(\beta (a))=a,~~~0<a< A_{2}(1).$$

It follows from Lemma 1 that \(A_{1}(p)\ge A_{2}(p)\). Hence, \(\alpha (a)\le \beta (a)\) for all \(0<a<A_{2}(1)\). We set

$$x_1(a):= F^{-1}(\alpha (a)),~ x_2(a):=G^{-1}(\alpha (a))$$

and

$$x_3(a):= G^{-1}(\beta (a)),~x_4(a):=F^{-1}(\beta (a)).$$

Since X and Y assume only finitely many values, there is a sequence \(0=a_1<\cdots <a_k\le A_{2}(1)\) such that \(a\mapsto x_1(a), \cdots , x_4(a)\) are constant on \((a_{i-1}, a_i)\). Denote the corresponding values of these functions as

$$x_{l,i}=x_l(a)~~~~~ \text{ for }~ a\in (a_{i-1}, a_i),~l=1,\ldots , 4.$$

Moreover, for \(i=1,\ldots ,k\), at the points \(x_{1,i}\) and \(x_{4,i}\) the function F has jumps of sizes at least \(\zeta _{1i}\) and \(\zeta _{2i}\), and at the corresponding points \(x_{2,i}\) and \(x_{4,i}\) the function G has jumps of sizes at least \(\zeta _{1i}\) and \(\zeta _{2i}\), where \(\zeta _{1i}\) and \(\zeta _{2i}\) are given by the equation

$$\zeta _{1i}\left( H_{\gamma ,r}(x_{2,i})-H_{\gamma ,r}(x_{1,i})\right) =\zeta _{2i}\left( H_{\gamma ,r}(x_{4,i})-H_{\gamma ,r}(x_{3,i})\right) =a_i-a_{i-1}~$$

For \(x>x_{4,k}\), we have \(F(x)>G(x)\). Thus, G is obtained from F by a sequence of k \((\gamma ,r)\)-transfers described by the corresponding x’s and \(\zeta\)’s above, plus a finite number of increasing transfers moving the mass from F to G right of \(x_{4,k}\). This completes the proof.

1.5 Proof of Theorem 3

Proof

Note that for random variables \(X_n\), X with distribution functions \(F_n\), F the convergence \(X_n\Rightarrow X\) mentioned in the theorem holds if and only if

$$\int _{-\infty }^{\infty }|F_{n}(x)-F(x)|dx\rightarrow 0,$$

and since \(0<H_{\gamma ,r}^{'}\le \frac{1+\eta ^{*}\lambda ^{*}}{1+\eta ^{*}}\), it holds that

$$\int _{-\infty }^{\infty }|F_n(x)-F(x)|dH_{\gamma ,r}(x)\rightarrow 0.$$

This implies that, for any \(t\in \mathfrak {R}\),

$$\int _{-\infty }^{t}\left( F_n(x)-F(x)\right) dH_{\gamma ,r}(x)\rightarrow 0.$$

The if-part thus follows from (14).

For the only-if-part, if XY are are bounded, then the proof is similar to Müller et al. (2017). We can define for any \(n\in \mathbb {N}\)

$$X_n=\frac{i}{n},~\text{ if }~ \frac{i}{n}\le X<\frac{i+1}{n}, ~i\in \mathbb {Z}$$

and

$$Y_n=\frac{i+1}{n},~\text{ if }~ \frac{i}{n}\le Y<\frac{i+1}{n},~i \in \mathbb {Z}.$$

Then \(X_n\) and \(Y_n\) have finite support with \(X_n\le _{1-SD} X\) and \(Y\le _{1-SD} Y_n\). Therefore,

$$X_n\le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}}X\le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}}Y \le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}} Y_n$$

and \(X_n\Rightarrow X\) and \(Y_n\Rightarrow Y\).

If X and Y are unbounded, then we define

$$\begin{aligned} X_n:=\left\{ \begin{array}{ll} x_n^{*}, &{} \text{ if }~X<-n, \\ X, &{} \text{ if }~-n\le X\le n, \\ n, &{} \text{ if }~n<X, \end{array} \right. \end{aligned}$$
(27)

and

$$\begin{aligned} Y_n:=\left\{ \begin{array}{ll} -n, &{}~ \text{ if }~Y<-n, \\ Y, &{}~ \text{ if }~-n\le Y\le n, \\ n, &{}~ \text{ if }~n<Y, \end{array} \right. \end{aligned}$$
(28)

where \(x_n^{*}\) satisfies that

$$H_{\gamma ,r}(x_n^{*})=H_{\gamma ,r}(-n)-\frac{\int _{-\infty }^{-n}F(x)dH_{\gamma ,r}(x)}{P(X<-n)}.$$

An easy calculation for the corresponding distribution functions \(F_n\), \(G_n\) shows that

$$\begin{aligned}&\int _{-\infty }^{t}\left( F_n(x)-G_n(x)\right) dH_{\gamma ,r}(x)\\&~~~~=\left\{ \begin{array}{ll} 0, &{} t\le x_n^{*}, \\ P(X<-n)\left( H_{\gamma ,r}(t)-H_{\gamma ,r}(x_n^{*})\right) , &{} x_n^{*}< t\le -n, \\ \int _{-\infty }^{t}\left( F(x)-G(x)\right) dH_{\gamma ,r}(x)+\int _{-\infty }^{-n}G(x)dH_{\gamma ,r}(x), &{} -n<t\le n,\\ 0,&{}t>n. \end{array} \right. \end{aligned}$$

Thus \(X\le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}}Y\), that is,

$$\int _{-\infty }^{t}\left( F(x)-G(x)\right) dH_{\gamma ,r}(x)\ge 0, \text{ for } \text{ all }~~t\in \mathbb {R}$$

implies that

$$\int _{-\infty }^{t}\left( F_n(x)-G_n(x)\right) dH_{\gamma ,r}(x)\ge 0 ~~\text{ for } \text{ all }~~t\in \mathbb {R}.$$

Then, it follows that \(X\le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}}Y\) implies that \(X_n\le _{(1+\gamma )-SD}^{r, \lambda ^{*},\eta ^{*}}Y_n\), and obviously \(X_n\), \(Y_n\) are bounded and \(X_n\Rightarrow X\), and \(Y_n\Rightarrow Y\).

For each fixed n we can approximate \(X_n\) and \(Y_n\) by sequences \(\{X_{nn}\}\) and \(\{Y_{nn}\}\) as in (27) and (28). Then, the sequences \(\{X_{nn}\}\) and \(\{Y_{nn}\}\) fulfill the conditions of the theorem. This completes the proof.

1.6 Proof of Theorem 4

To prove Theorem 4, we need Lemma 2.

Lemma 2

Define

$$\begin{aligned} A(x;H)=\frac{f_{\eta , \lambda }(x; H)}{f_{\eta ^{*},\lambda ^{*}}(x; H)} \end{aligned}$$
(29)

Then, A(xH) is nonegative and nonincreasing.

Proof

 Obviously, A(xH) is greater than zero. Since

$$A(x; H)=\frac{\eta (\lambda -1)(1+\eta ^{*})}{\eta ^{*}(\lambda ^{*}-1)(1+\eta )}+ \frac{\frac{1+\eta \lambda }{1+\eta }-\frac{(1+\eta ^{*}\lambda ^{*})\eta (\lambda -1)}{(1+\eta )\eta ^{*}(\lambda ^{*}-1)}}{f_{\eta ^{*},\lambda ^{*}}(x;H)}$$

and cdf H(x) is nondecreasing in x, to prove A(xH) is nonincreasing is just to prove

$$\frac{1+\eta \lambda }{1+\eta }-\frac{(1+\eta ^{*}\lambda ^{*})\eta (\lambda -1)}{(1+\eta )\eta ^{*}(\lambda ^{*}-1)}$$

is less than zero.

$$\begin{aligned}&\frac{1+\eta \lambda }{1+\eta }-\frac{(1+\eta ^{*}\lambda ^{*})\eta (\lambda -1)}{(1+\eta )\eta ^{*}(\lambda ^{*}-1)}\\&~~~~~~=\frac{1}{\eta ^{*}(1+\eta )(\lambda ^{*}-1)}\left[ \eta ^{*}(\lambda ^{*}-1)-\eta (\lambda -1)-\eta \eta ^{*}(\lambda -\lambda ^{*})\right] . \end{aligned}$$

From \(\eta \ge \eta ^{*}\ge 0\), \(\lambda \ge \lambda ^{*}\ge 1\), we have

$$\frac{1+\eta \lambda }{1+\eta }-\frac{(1+\eta ^{*}\lambda ^{*})\eta (\lambda -1)}{(1+\eta )\eta ^{*}(\lambda ^{*}-1)} \le 0.$$

This completes the proof of Lemma 2.

Proof

 Let \(X\vee R = \max \left\{ X, R\right\}\) and \(X\wedge R = \min \left\{ X,R\right\}\) with cdfs \(F_{X\vee R}(x)\) and \(F_{X\wedge R}(x)\), respectively. Then,

$$\begin{aligned}&E\left[ v(X; R,u)\right] \\&~~~~= E\left[ u(X)+\eta u(X\vee R)+\eta \lambda u(X\wedge R)-\eta (1+\lambda )u(R)\right] \nonumber \\&~~~~=\int _{-\infty }^{\infty }u(x)dF(x)+\eta \int _{-\infty }^{\infty }u(x)dF_{X\vee R}(x)+ \eta \lambda \int _{-\infty }^{\infty }u(x)dF_{X\wedge R}(x)\nonumber \\&~~~~~~~-\eta (1+\lambda )\int _{-\infty }^{\infty }u(x)dH(x)\nonumber \\&~~~~=\int _{-\infty }^{\infty }u(x)dF(x)+\eta \int _{-\infty }^{\infty }u(x)d\left[ F(x)H(x)\right] -\eta (1+\lambda )\int _{-\infty }^{\infty }u(x)dH(x)\nonumber \\&~~~~~~~+\eta \lambda \int _{-\infty }^{\infty }u(x)d\left[ F(x)+H(x)-H(x)F(x)\right] \nonumber \\&~~~~=\int _{-\infty }^{\infty }u(x)d\left[ F(x)\left( 1+\eta \lambda -\eta (\lambda -1)H(x)\right) \right] -\eta \int _{-\infty }^{\infty }u(x)dH(x)\nonumber \\&~~~~ =\int _{-\infty }^{\infty }u(x)d\left[ F(x)(1+\eta )f_{\eta ,\lambda }(x, H)\right] -\eta \int _{-\infty }^{\infty }u(x)dH(x). \end{aligned}$$

Thus,

$$E\left[ v(X; R,u)\right] -E\left[ v(Y; R,u)\right] =(1+\eta )\int _{-\infty }^{\infty }u(x)d\left[ \left( F(x)-G(x)\right) f_{\eta ,\lambda }(x, H)\right]$$

Combing with Lemma 2, the equivalent integral condition of

$$E\left[ v(X; R, H)\right] \le E\left[ v(Y; R, H)\right]$$

for all \(u\in U_{\gamma }\), and \(\eta \ge \eta ^{*}, \lambda \ge \lambda ^{*}\) follows in a similar manner of the proof of Theorem 1.

1.7 Proof of Proposition 3

Proof

Since \(X_1\) is a simple spread of \(Y_1\) with crossing point \(x_0\) and

$$\begin{aligned}&\gamma _0\int _{-\infty }^{x_0}\left[ F_1(x)-G_1(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,G_1)dx\\&~~~~~~~~=\int _{x_0}^{\infty }\left[ G_1(x)-F_1(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,G_1)dx \end{aligned}$$

it holds that \(X_1\le _{(1+\gamma _0)-SD}^{Y_1,\eta ^{*},\lambda ^{*}}Y_1\). But we have

$$\begin{aligned}&\int _{x_0}^{\infty }\left[ G_2(x)-F_2(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,G_2)dx\\&~~~~~~=k\int _{x_0}^{\infty }\left[ G_1(x)-F_1(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,G_1)dx\\&~~~~~=k\gamma _0\int _{-\infty }^{x_0}\left[ F_1(x)-G_1(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,G_1)dx\\&~~~~~~\ge \gamma _0\int _{-\infty }^{x_0}\left[ F_2(x)-G_2(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,G_2)dx. \end{aligned}$$

That is, \(X_2\nleq _{(1+\gamma _0)-SD}^{Y_2,\eta ^{*},\lambda ^{*}}Y_2\). This completes the proof.

1.8 Proof of Proposition 4

Proof

Since X is a simple spread of Y with crossing point \(x_0\) and \(\gamma _1\le 1\), we have \(X\le _{(1+\gamma _1)-SD}^{R_1\eta ^{*},\lambda ^{*}} Y\). For case [1], it holds that \(H_2(x)\ge H_1(x)\) on \([-\infty ,x_0)\) and \(H_2(x)\le H_1(x)\) on \([x_0,\infty )\). Thus, we have

$$\begin{aligned}&\gamma _1\int _{-\infty }^{x_0}\left[ F(x)-G(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,H_2)dx\nonumber \\&~~~~~~\le \gamma _1\int _{-\infty }^{x_0}\left[ F(x)-G(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,H_1)dx \nonumber \\&~~~~~~=\int _{x_0}^{\infty }\left[ G(x)-F(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,H_1)dx \\&~~~~~~\le \int _{x_0}^{\infty }\left[ G(x)-F(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,H_2)dx.\nonumber \end{aligned}$$
(30)

That is, \(X\nleq _{(1+\gamma _1)-SD}^{R_2, \eta ^{*},\lambda ^{*}} Y\). For case [2], it holds that \(H_2(x)\le H_1(x)\) on \([-\infty ,x_0)\) and \(H_2(x)\ge H_1(x)\) on \([x_0,\infty )\). Thus, in contrast to (30), we have

$$\gamma _1\int _{-\infty }^{x_0}\left[ F(x)-G(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,H_2)dx\ge \int _{x_0}^{\infty }\left[ G(x)-F(x)\right] f_{\eta ^{*},\lambda ^{*}}(x,H_2)dx.$$

That is, \(X\le _{(1+\gamma _1)-SD}^{R_2, \eta ^{*},\lambda ^{*}} Y\). This completes the proof.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, J., Zhao, C., Chen, W. et al. Fraction-Degree Reference Dependent Stochastic Dominance. Methodol Comput Appl Probab 24, 1193–1219 (2022). https://doi.org/10.1007/s11009-022-09939-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11009-022-09939-0

Keywords

Mathematics Subject Classification (2000)

Navigation