Risk aversion

Last updated
Risk aversion (red) contrasted to risk neutrality (yellow) and risk loving (orange) in different settings. Left graph: A risk averse utility function is concave (from below), while a risk loving utility function is convex. Middle graph: In standard deviation-expected value space, risk averse indifference curves are upward sloped. Right graph: With fixed probabilities of two alternative states 1 and 2, risk averse indifference curves over pairs of state-contingent outcomes are convex. Risikoeinstellung.svg
Risk aversion (red) contrasted to risk neutrality (yellow) and risk loving (orange) in different settings. Left graph: A risk averse utility function is concave (from below), while a risk loving utility function is convex. Middle graph: In standard deviation-expected value space, risk averse indifference curves are upward sloped. Right graph: With fixed probabilities of two alternative states 1 and 2, risk averse indifference curves over pairs of state-contingent outcomes are convex.

In economics and finance, risk aversion is the tendency of people to prefer outcomes with low uncertainty to those outcomes with high uncertainty, even if the average outcome of the latter is equal to or higher in monetary value than the more certain outcome. [1]

Contents

Risk aversion explains the inclination to agree to a situation with a more predictable, but possibly lower payoff, rather than another situation with a highly unpredictable, but possibly higher payoff. For example, a risk-averse investor might choose to put their money into a bank account with a low but guaranteed interest rate, rather than into a stock that may have high expected returns, but also involves a chance of losing value.

Example

Riskpremium1.png
Utility function of a risk-averse (risk-avoiding) individual
Riskpremium2.png
Utility function of a risk-neutral individual
Riskpremium3.png
Utility function of a risk-loving (risk-seeking) individual
CECertainty equivalent; E(U(W))Expected value of the utility (expected utility) of the uncertain payment W; E(W) – Expected value of the uncertain payment; U(CE)Utility of the certainty equivalent; U(E(W)) – Utility of the expected value of the uncertain payment; U(W0) – Utility of the minimal payment; U(W1) – Utility of the maximal payment; W0 – Minimal payment; W1 – Maximal payment; RPRisk premium

A person is given the choice between two scenarios: one with a guaranteed payoff, and one with a risky payoff with same average value. In the former scenario, the person receives $50. In the uncertain scenario, a coin is flipped to decide whether the person receives $100 or nothing. The expected payoff for both scenarios is $50, meaning that an individual who was insensitive to risk would not care whether they took the guaranteed payment or the gamble. However, individuals may have different risk attitudes. [2] [3] [4]

A person is said to be:

The average payoff of the gamble, known as its expected value, is $50. The smallest guaranteed dollar amount that an individual would be indifferent to compared to an uncertain gain of a specific average predicted value is called the certainty equivalent, which is also used as a measure of risk aversion. An individual that is risk averse has a certainty equivalent that is smaller than the prediction of uncertain gains. The risk premium is the difference between the expected value and the certainty equivalent. For risk-averse individuals, risk premium is positive, for risk-neutral persons it is zero, and for risk-loving individuals their risk premium is negative.

Utility of money

In expected utility theory, an agent has a utility function u(c) where c represents the value that he might receive in money or goods (in the above example c could be $0 or $40 or $100).

The utility function u(c) is defined only up to positive affine transformation – in other words, a constant could be added to the value of u(c) for all c, and/or u(c) could be multiplied by a positive constant factor, without affecting the conclusions.

An agent is risk-averse if and only if the utility function is concave. For instance u(0) could be 0, u(100) might be 10, u(40) might be 5, and for comparison u(50) might be 6.

The expected utility of the above bet (with a 50% chance of receiving 100 and a 50% chance of receiving 0) is

,

and if the person has the utility function with u(0)=0, u(40)=5, and u(100)=10 then the expected utility of the bet equals 5, which is the same as the known utility of the amount 40. Hence the certainty equivalent is 40.

The risk premium is ($50 minus $40)=$10, or in proportional terms

or 25% (where $50 is the expected value of the risky bet: (). This risk premium means that the person would be willing to sacrifice as much as $10 in expected value in order to achieve perfect certainty about how much money will be received. In other words, the person would be indifferent between the bet and a guarantee of $40, and would prefer anything over $40 to the bet.

In the case of a wealthier individual, the risk of losing $100 would be less significant, and for such small amounts his utility function would be likely to be almost linear. For instance, if u(0) = 0 and u(100) = 10, then u(40) might be 4.02 and u(50) might be 5.01.

The utility function for perceived gains has two key properties: an upward slope, and concavity. (i) The upward slope implies that the person feels that more is better: a larger amount received yields greater utility, and for risky bets the person would prefer a bet which is first-order stochastically dominant over an alternative bet (that is, if the probability mass of the second bet is pushed to the right to form the first bet, then the first bet is preferred). (ii) The concavity of the utility function implies that the person is risk averse: a sure amount would always be preferred over a risky bet having the same expected value; moreover, for risky bets the person would prefer a bet which is a mean-preserving contraction of an alternative bet (that is, if some of the probability mass of the first bet is spread out without altering the mean to form the second bet, then the first bet is preferred).

Measures of risk aversion under expected utility theory

There are various measures of the risk aversion expressed by those given utility function. Several functional forms often used for utility functions are represented by these measures.

Absolute risk aversion

The higher the curvature of , the higher the risk aversion. However, since expected utility functions are not uniquely defined (are defined only up to affine transformations), a measure that stays constant with respect to these transformations is needed rather than just the second derivative of . One such measure is the Arrow–Pratt measure of absolute risk aversion (ARA), after the economists Kenneth Arrow and John W. Pratt, [5] [6] also known as the coefficient of absolute risk aversion, defined as

where and denote the first and second derivatives with respect to of . For example, if so and then Note how does not depend on and so affine transformations of do not change it.

The following expressions relate to this term:

The solution to this differential equation (omitting additive and multiplicative constant terms, which do not affect the behavior implied by the utility function) is:

where and . Note that when , this is CARA, as , and when , this is CRRA (see below), as . See [7]

and this can hold only if . Therefore, DARA implies that the utility function is positively skewed; that is, . [8] Analogously, IARA can be derived with the opposite directions of inequalities, which permits but does not require a negatively skewed utility function (). An example of a DARA utility function is , with , while , with would represent a quadratic utility function exhibiting IARA.

Relative risk aversion

The Arrow–Pratt measure of relative risk aversion (RRA) or coefficient of relative risk aversion is defined as [11]

.

Unlike ARA whose units are in $−1, RRA is a dimensionless quantity, which allows it to be applied universally. Like for absolute risk aversion, the corresponding terms constant relative risk aversion (CRRA) and decreasing/increasing relative risk aversion (DRRA/IRRA) are used. This measure has the advantage that it is still a valid measure of risk aversion, even if the utility function changes from risk averse to risk loving as c varies, i.e. utility is not strictly convex/concave over all c. A constant RRA implies a decreasing ARA, but the reverse is not always true. As a specific example of constant relative risk aversion, the utility function implies RRA = 1.

In intertemporal choice problems, the elasticity of intertemporal substitution often cannot be disentangled from the coefficient of relative risk aversion. The isoelastic utility function

exhibits constant relative risk aversion with and the elasticity of intertemporal substitution . When using l'Hôpital's rule shows that this simplifies to the case of log utility, u(c) = log c, and the income effect and substitution effect on saving exactly offset.

A time-varying relative risk aversion can be considered. [12]

Implications of increasing/decreasing absolute and relative risk aversion

The most straightforward implications of increasing or decreasing absolute or relative risk aversion, and the ones that motivate a focus on these concepts, occur in the context of forming a portfolio with one risky asset and one risk-free asset. [5] [6] If the person experiences an increase in wealth, he/she will choose to increase (or keep unchanged, or decrease) the number of dollars of the risky asset held in the portfolio if absolute risk aversion is decreasing (or constant, or increasing). Thus economists avoid using utility functions such as the quadratic, which exhibit increasing absolute risk aversion, because they have an unrealistic behavioral implication.

Similarly, if the person experiences an increase in wealth, he/she will choose to increase (or keep unchanged, or decrease) the fraction of the portfolio held in the risky asset if relative risk aversion is decreasing (or constant, or increasing).

In one model in monetary economics, an increase in relative risk aversion increases the impact of households' money holdings on the overall economy. In other words, the more the relative risk aversion increases, the more money demand shocks will impact the economy. [13]

Portfolio theory

In modern portfolio theory, risk aversion is measured as the additional expected reward an investor requires to accept additional risk. If an investor is risk-averse, they will invest in multiple uncertain assets, but only when the predicted return on a portfolio that is uncertain is greater than the predicted return on one that is not uncertain will the investor will prefer the former. [1] Here, the risk-return spectrum is relevant, as it results largely from this type of risk aversion. Here risk is measured as the standard deviation of the return on investment, i.e. the square root of its variance. In advanced portfolio theory, different kinds of risk are taken into consideration. They are measured as the n-th root of the n-th central moment. The symbol used for risk aversion is A or An.

Von Neumann-Morgenstern utility theorem

The von Neumann-Morgenstern utility theorem is another model used to denote how risk aversion influences an actor’s utility function. An extension of the expected utility function, the von Neumann-Morgenstern model includes risk aversion axiomatically rather than as an additional variable. [14]

John von Neumann and Oskar Morgenstern first developed the model in their book Theory of Games and Economic Behaviour. [14] Essentially, von Neumann and Morgenstern hypothesised that individuals seek to maximise their expected utility rather than the expected monetary value of assets. [15] In defining expected utility in this sense, the pair developed a function based on preference relations. As such, if an individual’s preferences satisfy four key axioms, then a utility function based on how they weigh different outcomes can be deduced. [16]

In applying this model to risk aversion, the function can be used to show how an individual’s preferences of wins and losses will influence their expected utility function. For example, if a risk-averse individual with $20,000 in savings is given the option to gamble it for $100,000 with a 30% chance of winning, they may still not take the gamble in fear of losing their savings. This does not make sense using the traditional expected utility model however;

The von Neumann-Morgenstern model can explain this scenario. Based on preference relations, a specific utility can be assigned to both outcomes. Now the function becomes;

For a risk averse person,  would equal a value that means that the individual would rather keep their $20,000 in savings than gamble it all to potentially increase their wealth to $100,000. Hence a risk averse individuals’ function would show that;

Limitations of expected utility treatment of risk aversion

Using expected utility theory's approach to risk aversion to analyze small stakes decisions has come under criticism. Matthew Rabin has showed that a risk-averse, expected-utility-maximizing individual who,

from any initial wealth level [...] turns down gambles where she loses $100 or gains $110, each with 50% probability [...] will turn down 50–50 bets of losing $1,000 or gaining any sum of money. [17]

Rabin criticizes this implication of expected utility theory on grounds of implausibility—individuals who are risk averse for small gambles due to diminishing marginal utility would exhibit extreme forms of risk aversion in risky decisions under larger stakes. One solution to the problem observed by Rabin is that proposed by prospect theory and cumulative prospect theory, where outcomes are considered relative to a reference point (usually the status quo), rather than considering only the final wealth.

Another limitation is the reflection effect, which demonstrates the reversing of risk aversion. This effect was first presented by Kahneman and Tversky as a part of the prospect theory, in the behavioral economics domain. The reflection effect is an identified pattern of opposite preferences between negative as opposed to positive prospects: people tend to avoid risk when the gamble is between gains, and to seek risks when the gamble is between losses. [18] For example, most people prefer a certain gain of 3,000 to an 80% chance of a gain of 4,000. When posed the same problem, but for losses, most people prefer an 80% chance of a loss of 4,000 to a certain loss of 3,000.

The reflection effect (as well as the certainty effect) is inconsistent with the expected utility hypothesis. It is assumed that the psychological principle which stands behind this kind of behavior is the overweighting of certainty. Options which are perceived as certain are over-weighted relative to uncertain options. This pattern is an indication of risk-seeking behavior in negative prospects and eliminates other explanations for the certainty effect such as aversion for uncertainty or variability. [18]

The initial findings regarding the reflection effect faced criticism regarding its validity, as it was claimed that there are insufficient evidence to support the effect on the individual level. Subsequently, an extensive investigation revealed its possible limitations, suggesting that the effect is most prevalent when either small or large amounts and extreme probabilities are involved. [19] [20]

Bargaining and risk aversion

Numerous studies have shown that in riskless bargaining scenarios, being risk-averse is disadvantageous. Moreover, opponents will always prefer to play against the most risk-averse person. [21] Based on both the von Neumann-Morgenstern and Nash Game Theory model, a risk-averse person will happily receive a smaller commodity share of the bargain. [22] This is because their utility function concaves hence their utility increases at a decreasing rate while their non-risk averse opponents may increase at a constant or increasing rate. [23] Intuitively, a risk-averse person will hence settle for a smaller share of the bargain as opposed to a risk-neutral or risk-seeking individual.

In the brain

Attitudes towards risk have attracted the interest of the field of neuroeconomics and behavioral economics. A 2009 study by Christopoulos et al. suggested that the activity of a specific brain area (right inferior frontal gyrus) correlates with risk aversion, with more risk averse participants (i.e. those having higher risk premia) also having higher responses to safer options. [24] This result coincides with other studies, [24] [25] that show that neuromodulation of the same area results in participants making more or less risk averse choices, depending on whether the modulation increases or decreases the activity of the target area.

Public understanding and risk in social activities

In the real world, many government agencies, e.g. Health and Safety Executive, are fundamentally risk-averse in their mandate. This often means that they demand (with the power of legal enforcement) that risks be minimized, even at the cost of losing the utility of the risky activity. It is important to consider the opportunity cost when mitigating a risk; the cost of not taking the risky action. Writing laws focused on the risk without the balance of the utility may misrepresent society's goals. The public understanding of risk, which influences political decisions, is an area which has recently been recognised as deserving focus. In 2007 Cambridge University initiated the Winton Professorship of the Public Understanding of Risk, a role described as outreach rather than traditional academic research by the holder, David Spiegelhalter. [26]

Children

Children's services such as schools and playgrounds have become the focus of much risk-averse planning, meaning that children are often prevented from benefiting from activities that they would otherwise have had. Many playgrounds have been fitted with impact-absorbing matting surfaces. However, these are only designed to save children from death in the case of direct falls on their heads and do not achieve their main goals. [27] They are expensive, meaning that less resources are available to benefit users in other ways (such as building a playground closer to the child's home, reducing the risk of a road traffic accident on the way to it), and—some argue—children may attempt more dangerous acts, with confidence in the artificial surface. Shiela Sage, an early years school advisor, observes "Children who are only ever kept in very safe places, are not the ones who are able to solve problems for themselves. Children need to have a certain amount of risk taking ... so they'll know how to get out of situations." [28] [ citation needed ]

Game shows and investments

One experimental study with student-subject playing the game of the TV show Deal or No Deal finds that people are more risk averse in the limelight than in the anonymity of a typical behavioral laboratory. In the laboratory treatments, subjects made decisions in a standard, computerized laboratory setting as typically employed in behavioral experiments. In the limelight treatments, subjects made their choices in a simulated game show environment, which included a live audience, a game show host, and video cameras. [29] In line with this, studies on investor behavior find that investors trade more and more speculatively after switching from phone-based to online trading [30] [31] and that investors tend to keep their core investments with traditional brokers and use a small fraction of their wealth to speculate online. [32]

The behavioural approach to employment status

The basis of the theory, on the connection between employment status and risk aversion, is the varying income level of individuals. On average higher income earners are less risk averse than lower income earners. In terms of employment the greater the wealth of an individual the less risk averse they can afford to be, and they are more inclined to make the move from a secure job to an entrepreneurial venture. The literature assumes a small increase in income or wealth initiates the transition from employment to entrepreneurship-based decreasing absolute risk aversion (DARA), constant absolute risk aversion (CARA), and increasing absolute risk aversion (IARA) preferences as properties in their utility function. [33] The apportioning risk perspective can also be used to as a factor in the transition of employment status, only if the strength of downside risk aversion exceeds the strength of risk aversion. [33] If using the behavioural approach to model an individual’s decision on their employment status there must be more variables than risk aversion and any absolute risk aversion preferences.

Incentive effects are a factor in the behavioural approach an individual takes in deciding to move from a secure job to entrepreneurship. Non-financial incentives provided by an employer can change the decision to transition into entrepreneurship as the intangible benefits helps to strengthen how risk averse an individual is relative to the strength of downside risk aversion. Utility functions do not equate for such effects and can often screw the estimated behavioural path that an individual takes towards their employment status. [34]

The design of experiments, to determine at what increase of wealth or income would an individual change their employment status from a position of security to more risky ventures, must include flexible utility specifications with salient incentives integrated with risk preferences. [34] The application of relevant experiments can avoid the generalisation of varying individual preferences through the use of this model and its specified utility functions.

See also

Related Research Articles

In economics, utility is a measure of the satisfaction that a certain person has from a certain state of the world. Over time, the term has been used in at least two different meanings.

<span class="mw-page-title-main">Prospect theory</span> Theory of behavioral economics

Prospect theory is a theory of behavioral economics, judgment and decision making that was developed by Daniel Kahneman and Amos Tversky in 1979. The theory was cited in the decision to award Kahneman the 2002 Nobel Memorial Prize in Economics.

<span class="mw-page-title-main">St. Petersburg paradox</span> Paradox involving a game with repeated coin flipping

The St. Petersburg paradox or St. Petersburg lottery is a paradox involving the game of flipping a coin where the expected payoff of the theoretical lottery game approaches infinity but nevertheless seems to be worth only a very small amount to the participants. The St. Petersburg paradox is a situation where a naïve decision criterion that takes only the expected value into account predicts a course of action that presumably no actual person would be willing to take. Several resolutions to the paradox have been proposed, including the impossible amount of money a casino would need to continue the game indefinitely.

<span class="mw-page-title-main">Risk premium</span> Measure of excess

A risk premium is a measure of excess return that is required by an individual to compensate being subjected to an increased level of risk. It is used widely in finance and economics, the general definition being the expected risky return less the risk-free return, as demonstrated by the formula below.

The expected utility hypothesis is a foundational assumption in mathematical economics concerning decision making under uncertainty. It postulates that rational agents maximize utility, meaning the subjective desirability of their actions. Rational choice theory, a cornerstone of microeconomics, builds this postulate to model aggregate social behaviour.

The equity premium puzzle refers to the inability of an important class of economic models to explain the average equity risk premium (ERP) provided by a diversified portfolio of equities over that of government bonds, which has been observed for more than 100 years. There is a significant disparity between returns produced by stocks compared to returns produced by government treasury bills. The equity premium puzzle addresses the difficulty in understanding and explaining this disparity. This disparity is calculated using the equity risk premium:

The Allais paradox is a choice problem designed by Maurice Allais to show an inconsistency of actual observed choices with the predictions of expected utility theory. Rather than adhering to rationality, the Allais paradox proves that individuals rarely make rational decisions consistently when required to do so immediately. The independence axiom of expected utility theory, which requires that the preferences of an individual should not change when altering two lotteries by equal proportions, was proven to be violated by the paradox.

In accounting, finance, and economics, a risk-seeker or risk-lover is a person who has a preference for risk.

In decision theory and economics, ambiguity aversion is a preference for known risks over unknown risks. An ambiguity-averse individual would rather choose an alternative where the probability distribution of the outcomes is known over one where the probabilities are unknown. This behavior was first introduced through the Ellsberg paradox.

<span class="mw-page-title-main">Cumulative prospect theory</span>

Cumulative prospect theory (CPT) is a model for descriptive decisions under risk and uncertainty which was introduced by Amos Tversky and Daniel Kahneman in 1992. It is a further development and variant of prospect theory. The difference between this version and the original version of prospect theory is that weighting is applied to the cumulative probability distribution function, as in rank-dependent expected utility theory but not applied to the probabilities of individual outcomes. In 2002, Daniel Kahneman received the Bank of Sweden Prize in Economic Sciences in Memory of Alfred Nobel for his contributions to behavioral economics, in particular the development of Cumulative Prospect Theory (CPT).

<span class="mw-page-title-main">Exponential utility</span>

In economics and finance, exponential utility is a specific form of the utility function, used in some contexts because of its convenience when risk is present, in which case expected utility is maximized. Formally, exponential utility is given by:

Consumption smoothing is an economic concept for the practice of optimizing a person's standard of living through an appropriate balance between savings and consumption over time. An optimal consumption rate should be relatively similar at each stage of a person's life rather than fluctuate wildly. Luxurious consumption at an old age does not compensate for an impoverished existence at other stages in one's life.

<span class="mw-page-title-main">Isoelastic utility</span> Concept in economics

In economics, the isoelastic function for utility, also known as the isoelastic utility function, or power utility function, is used to express utility in terms of consumption or some other economic variable that a decision-maker is concerned with. The isoelastic utility function is a special case of hyperbolic absolute risk aversion and at the same time is the only class of utility functions with constant relative risk aversion, which is why it is also called the CRRA utility function. In statistics, the same function is called the Box-Cox transformation.

Goals-Based Investing or Goal-Driven Investing is the use of financial markets to fund goals within a specified period of time. Traditional portfolio construction balances expected portfolio variance with return and uses a risk aversion metric to select the optimal mix of investments. By contrast, GBI optimizes an investment mix to minimize the probability of failing to achieve a minimum wealth level within a set period of time.

In decision theory, the von Neumann–Morgenstern (VNM) utility theorem shows that, under certain axioms of rational behavior, a decision-maker faced with risky (probabilistic) outcomes of different choices will behave as if they are maximizing the expected value of some function defined over the potential outcomes at some specified point in the future. This function is known as the von Neumann–Morgenstern utility function. The theorem is the basis for expected utility theory.

In consumer theory, a consumer's preferences are called homothetic if they can be represented by a utility function which is homogeneous of degree 1. For example, in an economy with two goods , homothetic preferences can be represented by a utility function that has the following property: for every :

In decision theory, economics, and finance, a two-moment decision model is a model that describes or prescribes the process of making decisions in a context in which the decision-maker is faced with random variables whose realizations cannot be known in advance, and in which choices are made based on knowledge of two moments of those random variables. The two moments are almost always the mean—that is, the expected value, which is the first moment about zero—and the variance, which is the second moment about the mean.

In finance, economics, and decision theory, hyperbolic absolute risk aversion (HARA) refers to a type of risk aversion that is particularly convenient to model mathematically and to obtain empirical predictions from. It refers specifically to a property of von Neumann–Morgenstern utility functions, which are typically functions of final wealth, and which describe a decision-maker's degree of satisfaction with the outcome for wealth. The final outcome for wealth is affected both by random variables and by decisions. Decision-makers are assumed to make their decisions so as to maximize the expected value of the utility function.

In mathematical economics, an isoelastic function, sometimes constant elasticity function, is a function that exhibits a constant elasticity, i.e. has a constant elasticity coefficient. The elasticity is the ratio of the percentage change in the dependent variable to the percentage causative change in the independent variable, in the limit as the changes approach zero in magnitude.

Intertemporal portfolio choice is the process of allocating one's investable wealth to various assets, especially financial assets, repeatedly over time, in such a way as to optimize some criterion. The set of asset proportions at any time defines a portfolio. Since the returns on almost all assets are not fully predictable, the criterion has to take financial risk into account. Typically the criterion is the expected value of some concave function of the value of the portfolio after a certain number of time periods—that is, the expected utility of final wealth. Alternatively, it may be a function of the various levels of goods and services consumption that are attained by withdrawing some funds from the portfolio after each time period.

References

  1. 1 2 Werner, Jan (2008). "Risk Aversion". The New Palgrave Dictionary of Economics. pp. 1–6. doi:10.1057/978-1-349-95121-5_2741-1. ISBN   978-1-349-95121-5.
  2. Mr Lev Virine; Mr Michael Trumper (28 October 2013). ProjectThink: Why Good Managers Make Poor Project Choices. Gower Publishing, Ltd. ISBN   978-1-4724-0403-9.
  3. David Hillson; Ruth Murray-Webster (2007). Understanding and Managing Risk Attitude. Gower Publishing, Ltd. ISBN   978-0-566-08798-1.
  4. Adhikari, Binay Kumar; Agrawal, Anup (June 2016). "Does local religiosity matter for bank risk-taking?". Journal of Corporate Finance. 38: 272–293. doi:10.1016/j.jcorpfin.2016.01.009.
  5. 1 2 Arrow, K. J. (1965). "Aspects of the Theory of Risk Bearing". The Theory of Risk Aversion. Helsinki: Yrjo Jahnssonin Saatio. Reprinted in: Essays in the Theory of Risk Bearing, Markham Publ. Co., Chicago, 1971, 90–109.
  6. 1 2 Pratt, John W. (January 1964). "Risk Aversion in the Small and in the Large". Econometrica. 32 (1/2): 122–136. doi:10.2307/1913738. JSTOR   1913738.
  7. "Zender's lecture notes".
  8. Levy, Haim (2006). Stochastic Dominance: Investment Decision Making under Uncertainty (2nd ed.). New York: Springer. ISBN   978-0-387-29302-8.
  9. Friend, Irwin; Blume, Marshall (1975). "The Demand for Risky Assets". American Economic Review . 65 (5): 900–922. JSTOR   1806628.
  10. Bellemare, Marc F.; Brown, Zachary S. (January 2010). "On the (Mis)Use of Wealth as a Proxy for Risk Aversion". American Journal of Agricultural Economics. 92 (1): 273–282. doi:10.1093/ajae/aap006. hdl: 10161/7006 . S2CID   59290774.
  11. Simon, Carl and Lawrence Blume (2006). Mathematics for Economists (Student ed.). Viva Norton. p. 363. ISBN   978-81-309-1600-2.
  12. Benchimol, Jonathan (March 2014). "Risk aversion in the Eurozone". Research in Economics. 68 (1): 39–56. doi:10.1016/j.rie.2013.11.005. S2CID   153856059.
  13. Benchimol, Jonathan; Fourçans, André (March 2012). "Money and risk in a DSGE framework: A Bayesian application to the Eurozone". Journal of Macroeconomics. 34 (1): 95–111. doi:10.1016/j.jmacro.2011.10.003. S2CID   153669907.
  14. 1 2 von Neumann, John; Morgenstern, Oskar; Rubinstein, Ariel (1944). Theory of Games and Economic Behavior (60th Anniversary Commemorative ed.). Princeton University Press. ISBN   978-0-691-13061-3. JSTOR   j.ctt1r2gkx.
  15. Gerber, Anke (2020). "The Nash Solution as a von Neumann–Morgenstern Utility Function on Bargaining Games". Homo Oeconomicus. 37 (1–2): 87–104. doi: 10.1007/s41412-020-00095-9 . ISSN   0943-0180. S2CID   256553112.
  16. Prokop, Darren (2023). "Von Neumann–Morgenstern utility function | Definition & Facts | Britannica". www.britannica.com. Retrieved 2023-04-24.
  17. Rabin, Matthew (2000). "Risk Aversion and Expected-Utility Theory: A Calibration Theorem". Econometrica . 68 (5): 1281–1292. CiteSeerX   10.1.1.295.4269 . doi:10.1111/1468-0262.00158. S2CID   16418792.
  18. 1 2 Kahneman, Daniel; Tversky, Amos (March 1979). "Prospect Theory: An Analysis of Decision under Risk". Econometrica. 47 (2): 263. CiteSeerX   10.1.1.407.1910 . doi:10.2307/1914185. JSTOR   1914185.
  19. Hershey, John C.; Schoemaker, Paul J.H. (June 1980). "Prospect theory's reflection hypothesis: A critical examination". Organizational Behavior and Human Performance. 25 (3): 395–418. doi:10.1016/0030-5073(80)90037-9.
  20. Battalio, RaymondC.; Kagel, JohnH.; Jiranyakul, Komain (March 1990). "Testing between alternative models of choice under uncertainty: Some initial results". Journal of Risk and Uncertainty. 3 (1). doi:10.1007/BF00213259. S2CID   154386816.
  21. Roth, Alvin E.; Rothblum, Uriel G. (1982). "Risk Aversion and Nash's Solution for Bargaining Games with Risky Outcomes". Econometrica. 50 (3): 639–647. doi:10.2307/1912605. ISSN   0012-9682. JSTOR   1912605.
  22. Murnighan, J. Keith; Roth, Alvin E.; Schoumaker, Francoise (1988). "Risk Aversion in Bargaining: An Experimental Study". Journal of Risk and Uncertainty. 1 (1): 101–124. doi:10.1007/BF00055566. ISSN   0895-5646. JSTOR   41760532. S2CID   154784555.
  23. Kannai, Yakar (1977-03-01). "Concavifiability and constructions of concave utility functions". Journal of Mathematical Economics. 4 (1): 1–56. doi:10.1016/0304-4068(77)90015-5. ISSN   0304-4068.
  24. 1 2 Knoch, Daria; Gianotti, Lorena R. R.; Pascual-Leone, Alvaro; Treyer, Valerie; Regard, Marianne; Hohmann, Martin; Brugger, Peter (14 June 2006). "Disruption of Right Prefrontal Cortex by Low-Frequency Repetitive Transcranial Magnetic Stimulation Induces Risk-Taking Behavior". The Journal of Neuroscience. 26 (24): 6469–6472. doi:10.1523/JNEUROSCI.0804-06.2006. PMC   6674035 . PMID   16775134.
  25. Fecteau, Shirley; Pascual-Leone, Alvaro; Zald, David H.; Liguori, Paola; Théoret, Hugo; Boggio, Paulo S.; Fregni, Felipe (6 June 2007). "Activation of Prefrontal Cortex by Transcranial Direct Current Stimulation Reduces Appetite for Risk during Ambiguous Decision Making". The Journal of Neuroscience. 27 (23): 6212–6218. doi:10.1523/JNEUROSCI.0314-07.2007. PMC   6672163 . PMID   17553993.
  26. Spiegelhalter, David (2009). "Don's Diary" (PDF). CAM – the Cambridge Alumni Magazine. 58. The University of Cambridge Development Office: 3. Archived from the original (PDF) on March 9, 2013.
  27. Gill, Tim (2007). No fear: Growing up in a Risk Averse society (PDF). Calouste Gulbenkian Foundation. p. 81. ISBN   9781903080085. Archived from the original (PDF) on 2009-03-06.
  28. Sue Durant, Sheila Sage (10 January 2006). Early Years – The Outdoor Environment. Teachers TV.
  29. Baltussen, Guido; van den Assem, Martijn J.; van Dolder, Dennie (May 2016). "Risky Choice in the Limelight". Review of Economics and Statistics. 98 (2): 318–332. doi:10.1162/REST_a_00505. S2CID   57561510. SSRN   2057134.
  30. Barber, Brad M; Odean, Terrance (1 February 2001). "The Internet and the Investor". Journal of Economic Perspectives. 15 (1): 41–54. doi: 10.1257/jep.15.1.41 .
  31. Barber, Brad; Odean, Terrance (2002). "Online Investors: Do the Slow Die First?". Review of Financial Studies . 15 (2): 455–488. CiteSeerX   10.1.1.46.6569 . doi:10.1093/rfs/15.2.455.
  32. Konana, Prabhudev; Balasubramanian, Sridhar (May 2005). "The Social–Economic–Psychological model of technology adoption and usage: an application to online investing". Decision Support Systems. 39 (3): 505–524. doi:10.1016/j.dss.2003.12.003.
  33. 1 2 Bonilla, Claudio (2021). "Risk aversion, downside risk aversion, and the transition to entrepreneurship". Theory and Decision. 91: 123–133. doi:10.1007/s11238-020-09786-w. S2CID   228879460.
  34. 1 2 Harrison, Glenn (2006). "Risk Aversion and Incentive Effects: Comment". The American Economic Review. 95 (3): 897–901. doi:10.1257/0002828054201378 via The University of Queensland.

U.Sankar (1971), A Utility Function for Wealth for a Risk Averter, Journal of Economic Theory.