Chapter 5

Heterogeneous Agent Models in Finance

Roberto DieciXue-Zhong He,1    University of Bologna, Bologna, Italy
University of Technology Sydney, Sydney, NSW, Australia
1Corresponding author. email address: [email protected]

Abstract

This chapter surveys the state-of-art of heterogeneous agent models (HAMs) in finance using a jointly theoretical and empirical analysis, combined with numerical analysis from the latest development in computational finance. It provides supporting evidence on the explanatory power of HAMs to various stylized facts and market anomalies through model calibration, estimation, and economic mechanisms analysis. It presents HAMs with the mainstream finance a unified framework in continuous time to study the impact of historical price information on price dynamics, profitability and optimality of fundamental and momentum trading. It demonstrates how HAMs can help to understand stock price co-movements and evolutionary CAPM. It also introduces a new HAMs perspective on house price dynamics and an integrate approach to study dynamics of limit order markets. The survey provides further insights into the complexity and efficiency of financial markets and policy implications.

Keywords

Heterogeneity; Bounded rationality; Heterogeneous agent-based models; Stylized facts; Asset pricing; Housing bubbles; Limit order markets; Information efficiency; Comovement

1 Introduction

Economic and finance theory is witnessing a paradigm shift from a representative agent with rational expectations to boundedly rational agents with heterogeneous expectations. This shift reflects a growing evidence on the theoretical limitations and empirical challenges in the traditional view of homogeneity and perfect rationality in finance and economics.

The existence of limitations to fully rational behavior and the roles of psychological phenomena and behavioral factors in individuals' decision making have been emphasized and discussed from a variety of different standpoints in the economics and finance literature (see, e.g. Simon, 1982, Sargent, 1993, Arthur, 1994, Conlisk, 1996, Rubinstein, 1998, and Shefrin, 2005). Due to endogenous uncertainty about the state of the world and limits to information and computational ability, agents are prevented from forming rational forecasts and solving life-time optimization problems. Rather, agents favor simple reasoning and ‘rules of thumb’, such as the well documented technical analysis and active trading from financial market professionals.1 In addition, empirical investigations of financial time series show a number of market phenomena (including bubbles, crashes, short-run momentum and long-run mean reverting in asset prices) and some common features, the so-called stylized facts,2 which are difficult to accommodate and explain within the standard paradigm based on homogeneous agents and rational expectations.

Moreover, agents are heterogeneous in their beliefs and behavioral rules, which may change over time due to social interaction and evolutionary selection (see Lux, 1995, Arthur et al., 1997b, and Brock and Hommes, 1998). Such heterogeneity and diversity in individual behavior in economics, along with social interaction among individuals, can hardly be captured by a ‘representative’ agent at the aggregate level (see Kirman, 1992, 2010 for extensive discussions). For instance, as Heckman (2001), the 2000 Nobel Laureate in Economics, points out (concerning the contribution of microeconometrics to economic theory), “the most important discovery was the evidence on pervasiveness of heterogeneity and diversity in economic life. When a full analysis was made of heterogeneity in response, a variety of candidate averages emerged to describe the average person, and the longstanding edifice of the representative consumer was shown to lack empirical support.” Regarding agents' behavior during crisis periods and the role of policy makers, the former ECB president Jean-Claude Trichet writes “We need to deal better with heterogeneity across agents and the interaction among those heterogeneous agents”, highlighting the potential of alternative approaches such as behavioral economics and agent-based modeling.

Over the last three decades, empirical evidence, unconvincing justification of the assumption of unbounded rationality, and role of investor psychology have led to an incorporation of heterogeneity in beliefs and bounded rationality of agents into financial market modeling and asset pricing theory. This has changed the landscape of finance theory dramatically and led to fruitful development in financial economics, empirical finance, and market practice. In this chapter, we focus on the state-of-the-art of this expanding research field, denoted as Heterogeneous Agent Models (HAMs) in finance.

HAMs start from the contributions of Day and Huang (1990), Chiarella (1992), de Grauwe et al. (1993), Lux (1995), Brock and Hommes (1998), inspired by the pioneering work of Zeeman (1974) and Beja and Goldman (1980). This modeling framework views financial market dynamics as a result of the interaction of heterogeneous investors with different behavioral rules, such as fundamental and technical trading rules. One of the key aspects of these models is the expectation feedback mechanism. Namely, agents' decisions are based upon the predictions of endogenous variables whose actual values are determined by the expectations of agents. This results in the co-evolution of beliefs and asset prices over time. Earlier HAMs develop various nonlinear models to characterize various endogenous mechanisms of market fluctuations and financial crisis resulting from the interaction of heterogeneous agents rather than exogenous shocks or news. Overall, such models demonstrate that asset price fluctuations can be caused endogenously. We refer to Hommes (2006), LeBaron (2006), Chiarella et al. (2009a), Hommes and Wagener (2009), Westerhoff (2009), Chen et al. (2012), Hommes (2013), and He (2014) for surveys of these developments in the literature.

HAMs have strong connections with a broader area of Agent-Based Models (ABMs) and Agent-based Computational Economics (ACE). In fact, HAMs can be regarded as particular types of ABMs. However, generally speaking, ABMs are by nature very computationally oriented and allow for a large number of interacting agents, network structures, many parameters, and thorough descriptions of the underlying market microstructures. As such, they turn out to be extremely flexible and powerful, suitable for simulation, scenario analysis and regulation of real-world dynamic systems (see, e.g. Tesfatsion and Judd, 2006, LeBaron and Tesfatsion, 2008). By contrast, HAMs are typically characterized by substantial simplifications at the modeling level (few belief-types or behavioral rules, simplified interaction structures and reduced number of parameters). This makes HAMs analytically tractable to some extent, mostly within the theoretical framework of nonlinear dynamical systems. However, unlike computationally oriented ABMs, HAMs allow a deeper understanding of the basic dynamic mechanisms and driving forces at work, making it possible to identify different and clear-cut ‘types’ of macro outcomes in connection to specific agents' behavior.

Among the large number of HAMs in finance, this chapter is mostly concerned with analytically tractable models based on the interplay of two broad types of beliefs: extrapolative vs. regressive (or technical vs. fundamental rules, or chartists vs. fundamentalists). Since chartists rely on extrapolative rules to forecast future prices and to take their position in the market, they tend to sustain and reinforce current price trends or to amplify the deviations from the ‘fundamental price’. By contrast, fundamentalists place their orders in view of a mean reversion of asset price to its fundamental in long-run. The interplay between such forces is able to capture, albeit in a simplified manner, a basic mechanism of price fluctuations in financial markets. Support to this kind of behavioral heterogeneity comes from survey evidence (Menkhoff and Taylor, 2007, Menkhoff, 2010), experimental evidence (Hommes et al., 2005, Heemeijer et al., 2009), and empirically grounded discussion on the profitability of momentum and mean reversion strategies in financial markets (e.g. Lakonishok et al., 1994, Jegadeesh and Titman, 2001, Moskowitz et al., 2012).

In this chapter, we focus on the state-of-the-art of HAMs in finance from five main strands of the literature developed approximately over the last ten years since the appearance of the previous contributions in Volume II of this Handbook series. This development can have profound consequences for the interpretation of empirical evidence and the formulation of economic policy.

The first strand of research (Section 2) emphasizes the lasting potential of stylized HAMs in discrete time (in particular, chartist-fundamentalist models) to address key issues in finance. Such models have been largely investigated in the past in a wide range of versions incorporating heterogeneity, adaptation, evolution, and even learning (Hommes, 2001, Chiarella and He, 2002, 2003, and Chiarella et al., 2002, 2006b). They have successfully explained various market behavior, such as the long-term swing of market prices from fundamental price, asset bubbles, and market crashes, showing a potential to characterize and explain the stylized facts (Alfarano et al., 2005, Gaunersdorfer and Hommes, 2007) and various power law behavior (He and Li, 2008 and Lux and Alfarano, 2016) observed in financial markets. In addition, the chartist-fundamentalist framework can still provide insight into various stylized facts and market anomalies, and relate them to the economic mechanisms, parameters and scenarios of the underlying nonlinear deterministic systems. Such promising perspectives have motivated further empirical studies, leading to a growing literature on the calibration and estimation of the HAMs. In particular, in Sections 2.1 and 2.2, we focus on a simple HAM of Dieci et al. (2006) to illustrate its explanatory power to volatility clustering through calibration and empirical estimation, and relate the results to the underlying mechanisms and bifurcations of the nonlinear deterministic ‘skeleton’. Moreover, by considering an integrated approach of HAMs and incomplete information about the fundamental value, we provide a micro-foundation to the endogenous trading heterogeneity and switching behavior wildly characterized in HAMs (Section 2.3). We also survey fund flow effect among competing and evolving investment strategies (Section 2.4).

The second strand (Section 3) is on the development of a general framework in continuous time HAMs to incorporate historical price information in the HAMs. It provides a plausible way to deal with a variety of expectation rules formed from historical prices via moving averages over different time horizons, through a parsimonious system of stochastic delay differential equations. We introduce a time delay parameter to measure the effect of historical price information. Besides being consistent with continuous-time finance, this framework appears promising to understand the impact on market stability of lagged information (incorporated in different moving average rules and in realized profits recorded over different time horizons) and to explain a number of phenomena, particularly the long-range dependence in financial markets. We illustrate this approach and the main results in Section 3.1 by surveying the model in He and Li (2012). We emphasize the similarities to and differences from discrete-time HAMs. Moreover, Sections 3.2 and 3.3 demonstrate how useful the continuous-time HAMs can be in addressing the profitability of momentum and contrarian strategies and the optimal allocation with time series momentum and reversal, two of the most dominating financial market anomalies.

The third strand (Section 4) is on the impact of heterogeneous beliefs, expectations feedback and portfolio diversification on the joint dynamics of prices and returns of multiple risky assets. A related issue concerns the joint dynamics of international asset markets, driven by heterogeneous speculators who switch across markets depending on relative profit opportunities. In such models, often described by dynamical systems of large dimension, the typical nonlinear features of baseline HAMs interact with additional nonlinearities that arise naturally within a multi-asset setting, such as the beliefs about second moments and correlations. Section 4 surveys such models, starting from the basic setup developed by Westerhoff (2004), in which investors can switch not only across strategies but across markets (Section 4.1). Such models are not only able to reproduce various stylized facts, but also to offer some explanations to price comovements and cross-correlations of volatilities reported empirically (Schmitt and Westerhoff, 2014), as well as to address some key regulatory issues (Westerhoff and Dieci, 2006). Further research deals with asset comovements and changes in correlations from a different perspective. Based on models of evolving beliefs and (mean-variance) portfolios of heterogeneous investors, Section 4.2 is devoted to the multi-asset HAM of Chiarella et al. (2013b). This approach appears quite promising to address the issue of ‘time-varying betas’ within an evolutionary CAPM framework. It establishes a link between investors' behavior and changes in risk-return relationships at the aggregate level. Finally, Section 4.3 applies HAMs to illustrate the potentially destabilizing impact of the interlinkages between stock and foreign exchange markets (Dieci and Westerhoff, 2010, 2013b).

The fourth strand (Section 5) investigates the dynamics of house prices from the perspective of HAMs. Similar to financial markets, housing markets have long been characterized by boom-bust cycles and other phenomena apparently unrelated to changes in economic fundamentals, such as short-term positive autocorrelation and long-term mean-reversion, which are at odds with the predictions of the rational representative agent framework. Moreover, peculiar features of the housing market (such as the ‘twofold’ nature of housing, illiquidity, and supply-side elasticity) may interact with investors' demand influenced by behavioral factors. Section 5.1 surveys two recent HAMs of the housing market (Bolt et al., 2014 and Dieci and Westerhoff, 2016) which are based on mean-variance preferences and standard equilibrium conditions, with the fundamental price being regarded as the present value of future expected rental payments. However, within this framework, investors form heterogeneous expectations about future house prices, according to (evolving) regressive and extrapolative beliefs. Estimation of similar models supports the assumption of behavioral heterogeneity changing over time, based on the relative performance of the competing prediction rules. It highlights how such heterogeneity can produce endogenous house price bubbles and crashes (disconnected from the dynamics of the fundamental price). Moreover, the nonlinear dynamic analysis of such models can provide a simple behavioral explanation for the observed role of supply elasticity in ‘shaping’ housing bubbles and crashes, as widely reported and discussed in empirical and theoretical literature (see, e.g. Glaeser et al., 2008). Further ‘disequilibrium’ models, illustrated in Section 5.2, confirm the main findings about the impact of behavioral heterogeneity on housing price dynamics.

The fifth strand (Section 6) is on an integrated approach combining HAMs with traditional market microstructure literature to examine the joint impact of information asymmetry, heterogeneous expectations, and adaptive learning in limit order markets. As shown in Section 6.1, these HAMs are very helpful in examining complexity in market microstructure, providing insight into the impact of heterogeneous trading rules on limit order book and order flows (Chiarella and Iori, 2002, Chiarella et al., 2009b, 2012b, Kovaleva and Iori, 2014), and replicating the stylized facts in limit order markets (Chiarella et al., 2017). Earlier HAMs mainly examine the endogenous mechanism of interaction of heterogeneous agents, less so about information asymmetry, which is the focus of traditional market microstructure literature under rational expectations. Moreover, while the current microstructure literature focuses on informed traders by simplifying the behavior of uninformed traders substantially, a thorough modeling of the learning behavior of uninformed traders appears crucial for trading and market liquidity (O'Hara, 2001). Section 6.2 surveys a contribution in this direction by Chiarella et al. (2015a). By integrating HAMs with asymmetric information and Genetic Algorithm (GA) learning into microstructure literature, they examine the impact of learning on order submission, market liquidity, and price discovery. Finally, very recent contributions (in Sections 6.3 and 6.4) further examine the impact of high frequency trading (Arifovic et al., 2016) and different regulations (Lensberg et al., 2015) on market in a GA learning environment.

Most of the development surveyed in this chapter is based on a jointly theoretical and empirical analysis, combined with numerical simulations and Monte Carlo analysis from the latest development in computational finance. It provides very rich approaches to deal with various issues in equity market, housing market, and market microstructure. The results provide some insights into our understanding of the complexity and efficiency of financial market and policy implications.

2 HAMs of Single Asset Market in Discrete-Time

Empirical evidence of various stylized facts and anomalies in financial markets, such as fat tails in return distribution, long-range dependence in volatility, and time series momentum and reversal, has stimulated increasing research interest in financial market modeling. By focusing on endogenous heterogeneity of investor behavior, HAMs play a very important role in providing insights into the importance of investor heterogeneity and explaining stylized facts and marker anomalies observed in financial time series. Early HAMs consider two types of traders, typically fundamentalists and chartists. Beja and Goldman (1980), Day and Huang (1990), Chiarella (1992), Lux (1995), and Brock and Hommes (1997, 1998) are amongst the first to have shown that interaction of agents with heterogeneous expectations can lead to market instability. These HAMs have successfully explained market booms, crashes, and the deviations of market price from fundamental price and replicated some of the stylized facts, which are nicely surveyed in Hommes (2006), LeBaron (2006), and Chiarella et al. (2009a). The promising perspectives of HAMs have stimulated further studies on empirical testing in different markets, including commodity markets (Baak, 1999, Chavas, 2000), stock markets (Boswijk et al., 2007; Franke, 2009; Franke and Westerhoff, 2011, 2012; Chiarella et al., 2012a, 2014; He and Li, 2015a, 2015b), foreign exchange markets (Westerhoff and Reitz, 2003; de Jong et al., 2010; ter Ellen et al., 2013), mutual funds (Gomes and Michaelides, 2008), option markets (Frijns et al., 2010), oil markets (ter Ellen and Zwinkels, 2010), and CDS markets (Chiarella et al., 2015b). HAMs have also been estimated with contagious interpersonal communication by Gilli and Winker (2003), Alfarano et al. (2005), Lux (2009a, 2012), and other works reviewed in Chen et al. (2012).

This development has spurred recent attempts at theoretical explanations and the underlying economic mechanism analysis, which is nicely summarized in a recent survey of Lux and Alfarano (2016). Several behavioral mechanisms on volatility clustering have been proposed based on the underlying deterministic dynamics (He and Li, 2007, 2015b, 2017, Gaunersdorfer et al., 2008, He et al., 2016b), stochastic herding (Alfarano et al., 2005), and stochastic demand (Franke and Westerhoff, 2011, 2012).

In this section, we use the simple HAM of Dieci et al. (2006) to illustrate the explanatory power of the model to investor behavior and provide some of the underlying mathematical and economic mechanisms to volatility clustering and long-range dependence in volatility. We first introduce the model of boundedly rational and adaptive switching behavior of investors in financial markets in Section 2.1. We then provide two particular mechanisms to explain volatility clustering and long memory in return volatility based on the underlying deterministic dynamics in Section 2.2. Mathematically, the first is based on the local stability and Hopf bifurcation, explored in He and Li (2007), while the second is characterized by the coexistence of two locally stable attractors with different size, proposed initially in Gaunersdorfer et al. (2008) and further developed theoretically in He et al. (2016b). Economically, it demonstrates that the dominance of trend chasing behavior when investors cannot change their strategies or the intensive switching behavior of investors to switch to more profitable strategy can explain volatility clustering and long memory in return volatility, while the noise traders also play a very important role.

In Section 2.3, we briefly discuss He and Zheng (2016) about the emergence of trading heterogeneity due to information uncertainty and strategic trading of agents. Through an integrated approach of HAMs and incomplete information about the fundamental value, He and Zheng (2016) provide an endogenous self-correction mechanism of the market. This mechanism is very different from the HAMs with complete information, in which mean-reverting is channeled through some kind of nonlinear assumptions on the demand or order flow of risky asset and market stability depends exogenously on balanced activities from fundamental and momentum trading. The approach provides a micro-foundation to endogenous trading heterogeneity and switching behavior wildly characterized in HAMs. We complete the section with a discussion about an evolutionary finance framework in Section 2.4 to examine the effect of the flows of funds among competing and evolving investment styles on investment performance.

2.1 Market Mood and Adaptive Behavior

Empirical evidence in foreign exchange markets (Allen and Taylor, 1990, Taylor and Allen, 1992, Menkhoff, 1998, and Cheung et al., 2004) and managing fund industrial (Menkhoff, 2010) suggests that agents have different information and/or beliefs about market processes. They use not only fundamental but also technical analyses, which are consistent with short-run momentum and long-run reversal behavior in financial markets. In addition, although some agents do not change their particular trading strategies, there are agents who may switch to more profitable strategies over time. Recent laboratory experiments in Hommes et al. (2005), Anufriev and Hommes (2012), and Hommes and in't Veld (2015) also show that agents using simple “rule of thumb” trading strategies are able to coordinate on a common prediction rule. Therefore heterogeneity in expectations and adaptive behavior are crucial to describe individual forecasting and aggregate price behavior.

Motivated by the empirical and experiment evidence, Dieci et al. (2006) introduce a simple financial market of fundamentalists and trend followers. Some agents switch between different strategies over time according to their performance, characterizing the adaptively rational behavior of agents. Others are confident and stay with their strategies over time, representing market mood. It turns out that this simple model is rich enough to illustrating the complicated price dynamics and to exploring different mechanisms in generating volatility clustering and long memory in volatility. In the following, we first outline the model, discuss calibration and empirical estimation of the model, and then provide an analysis on the two underlying mechanisms (see Dieci et al., 2006 and He and Li, 2008, 2017 for the detail).

Consider a financial market with one risky asset and one risk free asset. Let r be the constant risk free rate, ptImage the price, and dtImage the dividend of the risky asset at time t. Assume that there are four types of investors, fundamental traders (or fundamentalists), trend followers (or chartists) and noise traders, and one market maker. Let n3Image be the population fraction of the noise traders. Among 1n3Image, the fractions of the fundamentalists and trend followers have fixed, n1Image and n2Image, and switching, n1,tImage and n2,t=1n1,tImage, components respectively. Denote n0=n1+n2,m0=(n1n2)/n0Image and mt=n1,tn2,tImage. Then the market fractions Qh,tImage (h=1,2,3Image) of the fundamentalists, trend followers, and noise traders at time t can be rewritten as, respectively,

(1) {Q1,t=12(1n3)[n0(1+m0)+(1n0)(1+mt)],Q2,t=12(1n3)[n0(1m0)+(1n0)(1mt)],Q3,t=n3.

Image (1)

Let Rt+1=pt+1+dt+1RptImage be the excess return and R=1+rImage. We model the order flow3 zh,tImage of type-h investors from t to t+1Image by zh,t=Eh,t(Rt+1)/(ahVh,t(Rt+1))Image, where Eh,tImage and Vh,tImage are the conditional expectation and variance at time t and ahImage is the risk aversion coefficient of type h traders. The order flow of the noise traders ξtN(0,σ2ξ)Image is an i.i.d. random variable. Then the population weighted average order flow is given by Ze,t=Q1,tz1,t+Q2,tz2,t+n3ξt.Image To determine the market price, we follow Chiarella and He (2003) and assume that the market price is determined by the market maker as follows,

(2) pt+1=pt+λZe,t=pt+μze,t+δt,

Image (2)

where ze,t=q1,tz1,t+q2,tz2,tImage, qh,t=Qh,t/(1n3)Image for h=1,2Image, λ denotes the speed of price adjustment of the market maker, μ=(1n3)λImage and δtN(0,σ2δ)Image with σδ=λn3σξImage.

We now describe briefly the heterogeneous beliefs of the fundamentalists and trend followers and the adaptive switching mechanism. The conditional mean and variance for the fundamental traders are assumed to follow

(3) E1,t(pt+1)=pt+(1α)[Et(pt+1)pt],V1,t(pt+1)=σ21,

Image (3)

where ptImage is the fundamental value of the risky asset following a random walk,

(4) pt+1=ptexp(σ2ε2+σεεt+1),εtN(0,1),σε0,p0=p>0,

Image (4)

εtImage is independent of the noise process δtImage, σ21Image is constant, and hence Et(pt+1)=ptImage. Here (1α)Image measures the speed of price adjustment towards the fundamental price with 0<α<1Image. A high α indicates less confidence on the convergence to the fundamental price, leading to a slower adjustment of the market price to the fundamental. For the trend followers, we assume

(5) E2,t(pt+1)=pt+γ(ptut),V2,t(pt+1)=σ21+b2vt,

Image (5)

where γ0Image measures the extrapolation of the trend, utImage and vtImage are sample mean and variance, respectively. We assume that ut=δut1+(1δ)ptImage and vt=δvt1+δ(1δ)(ptut1)2Image, representing limiting mean and variance of the geometric decay processes when the memory lag tends to infinity. Here δ(0,1)Image measures the geometric decay rate and b20Image measures the sensitivity to the sample variance. For simplicity we assume that investors share a homogeneous belief about the dividend process dtImage, which is i.i.d. and normally distributed with mean ˉdImage and variance σ2dImage. Denote by p=po=ˉd/rImage the long-run fundamental price.

Let πh,t+1Image be the realized profit between t and t+1Image of type-h investors, πh,t+1=zh,t(pt+1+dt+1Rpt)Image for h=1,2Image. Following Brock and Hommes (1997, 1998), the market fraction of investors choosing strategy h at time t+1Image is determined by

nh,t+1=exp[β(πh,t+1Ch)]iexp[β(πi,t+1Ci)],h=1,2,

Image

where β measures the intensity of the choice and Ch0Image the cost. Together with (1) the market fractions and asset price dynamics are determined by the following random dynamic system in discrete-time,

(6) {pt+1=pt+μ(q1,tz1,t+q2,tz2,t)+δt,δtN(0,σ2δ),ut=δut1+(1δ)pt,vt=δvt1+δ(1δ)(ptut1)2,mt=tanh[β2(z1,t1z2,t1(C1C2))(pt+dtRpt1)].

Image (6)

2.2 Volatility Clustering: Calibration and Mechanisms

By conducting econometric analysis via Monte Carlo simulations, He and Li (2015b, 2017) show that the autocorrelations of returns, absolute returns and squared returns of the model developed above share the same pattern as those of the DAX 30. They further characterize the power-law behavior of the DAX 30 and find that the estimates of the power-law decay indices, the (FI)GARCH parameters, and the tail index of the model closely match those of the DAX 30. In the following we first report the calibrated results of the model developed in the previous subsection and then provide some insights into investor behavior and two underlying mechanisms of the volatility clustering.

When there is no switching between the two strategies, the above model reduces to the no-switching model in He and Li (2007), showing that the no-switching model is able to replicate the power-law behavior in return volatility. Based on the daily price index data of the DAX 30 from 11 August, 1975 to 29 June, 2007, He and Li (2015b, 2017) calibrate three scenarios of the above model: the no-switching (N) model with β=0Image, pure-switching (S) model with n0=0Image, and full (F) model of (6). The results are collected in Table 1 (with fixed r=5%p.aImage. and C1=C2=0Image). By conducting econometric analysis via Monte Carlo simulations based on the calibrated models, He and Li (2015b, 2017) find that, for all three scenarios, the estimates of the power-law decay indices d, the (FI)GARCH parameters, and the tail index of the calibrated model closely match those of the DAX 30. By conducting a Wald test Ho:dDAX=dImage at 5% and 1% significant levels (with the critical values of 3.842 and 6.635, respectively), He and Li (2017) show that switching model fits the data better than the no-switching and pure-switching models.

Table 1

Calibrated parameters of the no-switching (N), pure-switching (S), and full (F) models

α γ a 1 a 2 μ n 0 m 0 δ b σ σ δ β Wald
N 0.858 8.464 6.024 0.383 0.946 1 −0.200 0.292 6.763 0.24 3.473 0 112
S 0.513 0.764 7.972 0.231 2.004 0 0.983 3.692 0.231 3.268 0.745 108
F 0.488 1.978 7.298 0.320 1.866 0.313 −0.024 0.983 3.537 0.231 3.205 0.954 106

Image

Comparing the estimates of the three scenarios leads to different investor behavior. The estimated annual return volatility σ is close to the annual return volatility of the DAX 30. Higher a1Image than a2Image implies that the fundamentalists are more risk averse compared to the trend followers. For the no-switching scenario, a higher value of α indicates a slow price adjustment of the fundamentalists toward the fundamental value, while a higher value of γ indicates that the trend followers extrapolate the price trend actively. Without switching, mo=0.2Image indicates that both the fundamentalists and trend followers are active in the market, which is however dominated by the trend followers (about 60%). On the full model, the market is dominated by investors (about 70%) who constantly switch between the fundamental and trend following strategies, although some investors (about 30%) never change their strategies over the time. This is consistent with the empirical findings discussed at the beginning of this section.

We now provide two mechanisms based on the underlying deterministic dynamics. The first one is on the local stability and periodic oscillation due to Hopf bifurcation, explored in He and Li (2007). Essentially, on the parameter space of the deterministic model, near the Hopf bifurcation boundary, the fundamental steady state can be locally stable but globally unstable. Due to the nature of Hopf bifurcation, such global instability leads to switching between the locally stable fundamental price and the periodic oscillations around the fundamental price. Then triggered by the fundamental and market noises, He and Li (2007) show that the interaction of the fundamental, risk-adjusted trend chasing from the trend followers, and the interplay of the noises and the underlying deterministic dynamics can be the source of power-law behavior in return volatility. Mathematically, the calibrated no-switching and switching models share the same underlying deterministic mechanism. Economically, however, they provide different behavioral mechanisms. With no-switching, it is the dominance of the trend followers (about 60%) that drives the power-law behavior. However, with both switching and no-switching investors, dominated by these traders (about 70%) who constantly switch between the two strategies. It is therefore the adaptive behavior of investors that generates the power-law behavior. This is also in line with Franke and Westerhoff (2012, 2016) who estimate various HAMs and show that herding behavior plays a key role in matching the stylized facts. More importantly, the noise traders play an important role in generating insignificant ACs on the returns, while the significantly decayed AC patterns of the absolute returns and squared returns are more influenced by the fundamental noise. As pointed out in Lux and Alfarano (2016), noise traders is probably a central ingredient of these models.

The second mechanism proposed initially in Gaunersdorfer et al. (2008) is characterized by the coexistence of two locally stable attractors with different size, while such coexistence is not required in the previous mechanism. Dieci et al. (2006) show that the above model can display such co-existence of locally stable fundamental steady state and periodic cycle. The interaction of the coexistence of the deterministic dynamics and the noise processes then triggers the switching among the two attractors and endogenously generates volatility clustering. More recently, by applying normal form analysis and center manifold theory, He et al. (2016b) provide the following theoretical result on the coexistence of the locally stable steady state and invariant circle of the underlying deterministic model (we refer to He et al., 2016b for the details).

Proposition 2.1

The underlying deterministic system of (6) has a unique fundamental steady state (p,u,v,m)=(ˉp,ˉp,0,ˉm)Image with ˉm=tanhβ(C2C1)2Image. The fundamental steady state is locally asymptotically stable for γ(0,γ)Image, and it undergoes a Neimark–Sacker bifurcation at γ=γImage, that is, there is an invariant curve near the fundamental steady state. Moreover, the bifurcated closed invariant curve is forward and stable when a(0)<0Image and backward and unstable when a(0)>0Image, and a Chenciner (generalized Neimark–Sacker) bifurcation takes place when a(0)=0Image. Here a(0)Image is the first Lyapunov coefficient.

Note that the market fractions of the fundamentalists and trend followers at the fundamental steady state are given by q1=(1+mq)/2Image and q2=(1mq)/2Image with mq=n0m0+(1n0)ˉmImage, respectively. When the cost of the fundamental strategy C1Image is higher than the cost of the trend following strategy C2Image, an increase in the switching intensity β leads to a decrease in γImage, meaning that the fundamental price becomes less stable when traders switch their strategies more often. This is essentially the rational routes to randomness of Brock and Hommes (1997, 1998).

Fig. 1 illustrates two different types of Neimark–Sacker bifurcation. It is the sign of the first Lyapunov coefficient a(0)Image that determines the bifurcation direction, either forward or backward, and the stability of the bifurcated invariant circles, leading to different bifurcation dynamics. When a(0)<0Image, the bifurcation is forward and stable, meaning that the bifurcated invariant circle occurring for γ>γImage is locally stable. In this case, as γ increases and passes γImage, the fundamental steady state becomes unstable and the trajectory converges to an invariant circle bifurcating from the fundamental steady state. As γ increases further, the trajectory converges to invariant circles with different sizes. This is illustrated in Fig. 1A with γ0.93Image where the two bifurcating curves for γ>γImage indicate the minimum and maximum value boundaries of the bifurcating invariant circles as γ increases.

Image
Figure 1 Bifurcation diagrams of the market price with respect to γ. Here a = a1 = a2 = 0.5, μ = 1, α = 0.3, δ = 0.85, b2 = 0.05, C = C1 − C2 = 0.5, β = 0.5, and m0 = 0. (A) n0 = 0.8; (B) n0 = 0.5.

However, when a(0)>0Image, the bifurcation is backward and unstable, meaning that the bifurcated invariant circle occurring at γ=γImage is unstable, illustrated in Fig. 1B (with γ0.88Image). There is a continuation of the unstable bifurcated circles as γ decreases initially until it reaches a critical value ˆγImage, which is indicated by the two (red) dotted curves of the bifurcating circles for ˆγ<γ<γImage. Then as γ increases from the critical value ˆγImage, the bifurcated circles become forward and stable. This is illustrated by the two (blue) solid curves, which are the boundaries of the bifurcating circles, for γ>ˆγImage in Fig. 1B. Therefore, the locally stable steady state coexists with the locally stable ‘forward extended’ circles for ˆγ<γ<γImage, in between there are backward extended unstable circles. For ˆγ<γ<γImage, even when the fundamental steady state is locally stable, prices need not converge to the fundamental value, while may settle down to a stable limit circle. We call ˆγ<γ<γImage the ‘volatility clustering region’. In addition, a Chenciner (generalized Neimark–Sacker) bifurcation takes place when a(0)=0Image. Based on the above analysis, a necessary condition on the coexistence is that a(0)>0Image. The coexistence of the locally stable steady state and invariant circle illustrated in Fig. 2 shows that the price dynamics depends on the initial values.

Image
Figure 2 The deterministic trajectories of time series of price for (p0,u0,v0,m0)=(ˉp+1,ˉp,0,ˉm)Image in (A) and (p0,u0,v0,m0)=(ˉp+1,ˉp1,0,ˉm)Image in (B) and the phase plot of (p,u) in (C). Here the parameter values are the same as in Fig. 1 and n0 = 0.5. (A) Price time series; (B) price time series; (C) phase plot.

When buffeted with noises, the stochastic model can endogenously generate volatility clustering and long range dependence in volatility, illustrated in Fig. 3. Economically, with strong trading activities of either the fundamental investors or the trend followers, market price fluctuates around either the fundamental value with low volatility or a cyclical price movement with high volatility, depending on market conditions. When the activities of the fundamentalists and trend followers are balanced (to be in the volatility clustering region), the interaction of the fundamental noise and noise traders and the underlying co-existence dynamics then triggers an irregular switching between the two volatility regimes, leading to volatility clustering. In particular, volatility clustering becomes more significant when neither the fundamental nor the trend following traders dominate the market and traders switch their strategies more often. The results verify the endogenous mechanism on volatility clustering proposed by Gaunersdorfer et al. (2008) and provide a behavioral explanation on the volatility clustering.

Image
Figure 3 Time series of (A) the market price (red solid line) and the fundamental price (blue dotted line), (B) the market returns (r); the ACs of (C) the returns and (D) the absolute returns. Here the parameter values are the same as in Fig. 2 and σδ = 2, σε = 0.025.

2.3 Information Uncertainty and Trading Heterogeneity

Traditional finance literature mainly explore the role of asymmetric information and information uncertainty. Most HAMs however mainly focus on endogenous market mechanism through the interaction among heterogeneous agents by assuming a complete information about the fundamental value of risky assets. An integration of HAMs and asymmetric and/or uncertain information would provide a micro-foundation on behavioral heterogeneity and a more broad framework to better explaining various puzzles and anomalies in financial markets. Instead of heuristical heterogeneity assumption of agents' behavior, He and Zheng (2016) model the trading heterogeneity by introducing information uncertainty about the fundamental value to a HAM. Agents are homogeneous ex ante. Conditional on their private information about the fundamental value, agents choose optimally among different trading strategies when optimizing their expected utilities. This approach provides a micro-foundation to trading and behavioral heterogeneity among agents. It also offers a different switching behavior of agents from the current HAMs. In the following, we brief this approach.

Consider a continuum [0,1]Image of agents trading one risky asset and one risk-free asset in discrete-time. For simplicity, the risk-free rate is normalized to be zero. The fundamental value of the risky asset μN(ˉμ,σ2μ)Image is not known publicly. Denote αμ=1/σ2μImage the precision of the fundamental value μ. In each time period, agent i receives a private signal about the fundamental value μ, given by xi,t=μ+εi,tImage, where εi,tN(0,σ2x)Image is i.i.d. normal across agents and over time. Let αx=1/σ2xImage be the precision of the signal. Agents maximize CARA utility function U(Wi,t)=exp(aWi,t),Image with the same risk aversion coefficient a, in which Wi,tImage is the wealth of agent i at time t. Let ptImage be the (cum-)market price of the risky asset and denote It={pt,pt1,}Image the public information of historical price. Conditional on the public information It1Image and her private signal xi,tImage, agent i seeks to maximize her expected utility, leading to the optimal demand qi,t=[E(pt|xi,t,It1)pt1]/[aVar(pt|xi,t,It1)]Image, conditional on the public information It1Image and her signal xi,tImage.

Facing the information uncertainty on the fundamental value, the agent considers both fundamental and momentum trading strategies based on the public information of the history price and her private signal about the fundamental value. More explicitly, the fundamental trading strategy is based on

(7) Ef(pt|xi,t,It1)=(1γ)pt1+γαμˉμ+αxxi,tαμ+αx,

Image (7)

(8) Varf(pt|xi,t,It1)=γ2Var(μ|xi,t,It1)=γ2αμ+αx,

Image (8)

where γ(0,1]Image is a constant, measuring the convergence speed of the market price to the expected fundamental value. Note that αμˉμ+αxxi,tαμ+αxImage and 1αμ+αxImage are agent i's posterior updating of the mean and variance, respectively, of the fundamental value of the risky asset conditional on her signal xi,tImage. Condition (7) means that the predicted price is a weighted average of the latest market price and the posterior updating of the fundamental value conditional on her private signal xi,tImage; while (8) means that the conditional variance is proportional to the posterior variance conditional on the private signal xi,tImage. In particular, when γ=1Image, the conditional mean and variance (7)(8) are reduced to the posterior mean and variance, respectively. Therefore the fundamental trading strategy reflects agent's belief that the future price is expected to converge to the expected fundamental value. Though the private signals xi,tImage are i.i.d. across agents and over time, they are partially incorporated through the current market price ptImage and hence reflected in the prediction of the future prices. Consequently, the optimal demand based on the fundamental analysis becomes qfi,t=[αμˉμ+αxxi,t(αμ+αx)pt1]/(aγ)Image, which is called the fundamental trading strategy f.

The momentum trading is however independent of the private signal xi,tImage, but depends on a price trend,

(9) Ec(pt|xi,t,It1)=pt1+β(pt1vt),Varc(pt|xi,t,It1)=σ2t1,

Image (9)

where vtImage is a reference price or price trend (can be a moving average, a supporting/resistance price level, or any index derived from technical analysis), β measures the extrapolation of the price deviation from the trend, and σ2t1Image is a heuristic prediction on the variance of the asset price. Then the optimal demand becomes qci,t=β(pt1vt)/(aσ2t1)Image, which is called momentum strategy c. In particular, when vtImage is a moving average of the historical prices and β>(<)0Image, strategy c is essentially a time-series momentum (contrarian) strategy (Moskowitz et al., 2012).

Given the information uncertainty, the agent compares the expected value functions based on the two optimal trading strategies and chooses the one with relative higher value. More explicitly, the agent firstly calculates the respective value functions based on strategy f and c,

Efi,t(U)=exp{A[Wi,t1+[αμˉμ+αxxi,t(αμ+αx)pt1]22a(αμ+αx)]},Eci,t(U)=exp{A[Wi,t1+β2(pt1vt)22aσ2t1]}.

Image

The agent then compares the value functions and selects the one that yields a higher value. Note that Efi,tImage is an increasing function of the absolute value of the signal |xi,t|Image, while EciImage is independent of xi,tImage. Therefore there exists threshold signal values ˉxtImage for the private signal such that Efi,t=Eci,tImage, that is,

Efi,t(U)Eci,t(U)=exp{[[αμˉμ+αxˉxt(αμ+αx)pt1]22(αμ+αx)β2(pt1vt)22σ2t1]}=1.

Image

Solving for ˉxtImage yields

(10) x±t=1αx[(αμ+αx)pt1αμˉμ±βαμ+αxσt1(pt1vt)].

Image (10)

Therefore, when vt=pt1Image, the agent chooses strategy f. When vtpt1Image, the agent chooses strategy c if her signal is less informative, falling into the interval (xmt,xMt)Image, and strategy f otherwise, where xmt=min(x±t)Image and xMt=max(x±t)Image. Therefore, the optimal demand of agent i is given by qi,t=qfi,tImage for xi,txmtImage or xi,txMtImage; otherwise qi,t=qci,tImage when xi,t(xmt,xMt)Image. Intuitively, when agent's private signal is near the mean fundamental value, the private information becomes less valuable. However, when agent's private signal is far away from the mean fundamental value, the private information becomes more valuable and hence the agent favors the fundamental trading strategy.

The choice between the two strategies due to the informativeness of the private information about the fundamental value leads to endogenous heterogeneity and switching behavior of agents' choices. More explicitly, by aggregating the demand DtImage in a closed form and considering noisy supply StImage, the market price is determined through a market maker scenario via pt=pt1+λ(Dt+St)Image with λ>0Image. He and Zheng (2016) first conduct an analysis on the underlying deterministic model when σ2t1=σ2Image is a constant and vt=pt2Image (corresponding to a simple momentum trading based on the change in the last price). They show that the fundamental price is locally stable with small precisions of the fundamental information noise. That is, the fundamental price becomes unstable when the level of the fundamental information noise is small, leading to high price volatility. Intuitively, in this case, the fundamental information become more accurate and hence less valuable. Therefore the fundamental strategy becomes less profitable, while the momentum trading strategy becomes more popular. This is consistent with the literature on coordination game with imperfect information (see Angeletos and Werning, 2006).

When the fundamental price becomes unstable, the price dynamics can become very complicated. On the stochastic model, they have shown that the market fraction of the agents choosing the momentum (fundamental) strategy decreases (increases) as the mis-pricing increases. This underlies mean-reverting of market price to its fundamental price when mis-pricing becomes significant, burst of a bubble, and recover of a recession. This mechanism, together with the destabilizing role of the momentum trading and the stabilizing role of the fundamental trading, provides an endogenous self-correction mechanism of the market. This mechanism is very different from the HAMs with complete information, in which the mean-reverting is channeled through some nonlinear assumptions on the demand or order flow of risky asset. The market stability depends exogenously on balanced activities from fundamental and momentum trading. This integrated approach of HAMs and incomplete information about the fundamental value therefore provides a micro-foundation to endogenous trading heterogeneity and switching behavior wildly characterized in HAMs. Furthermore, He and Zheng (2016) conduct a time series analysis on the stylized facts and demonstrate that the model is able to match the S&P 500 in terms of power-law distribution in returns, volatility clustering, long memory in volatility, and leverage effect.

2.4 Switching of Agents, Fund Flows, and Leverage

Similar to Dieci et al. (2006), most HAMs employ the discrete-choice framework4 to capture the way investors switch across different competing strategies/behavioral rules. However, since this approach models the changes of investors' proportions, not directly the flows of funds, it is not very suitable to capture the long-run performance of investment strategies (or ‘styles’) in terms of accumulated wealth, nor the impact of fund flows on the price dynamics. For this reason, LeBaron (2011) defines such forms of switching between strategies as active learning, capturing investors' tendency to adopt the best-performing rule, in contrast to passive learning, by which investors' wealth naturally accumulates on strategies that have been relatively successful. This second form of learning is closely related to the issue of survival and long-run dominance of strategies and to the evolutionary finance approach (see Blume and Easley, 1992, 2006, Sandroni, 2000, Hens and Schenk-Hoppé, 2005, as well as Evstigneev et al., 2009 for a comprehensive survey of early results and recent research in this field).5

LeBaron (2011) argues that the dynamics of real-world markets are likely to be affected by some combinations of active and passive learning, and that exploring their interaction may improve our understanding of the dynamics of asset prices. Moreover, LeBaron (2012) proposes a simple framework that can simultaneously account for wealth dynamics and active search for new strategies, based on performance comparison. Besides reproducing the basic stylized facts of asset returns and trading volume, the model yields some insight into the dynamics of agents' strategies and their impact on market stability.

A further recent contribution on the interplay of active and passive learning is provided by Palczewski et al. (2016). They build an evolutionary finance framework in discrete time with fundamental, trend-following and noise trading strategies. Such strategies are interpreted as portfolio managers with different investment ‘styles’. Individual investors can move (part of) their funds between portfolio managers. The total amount of freely flowing capital is a model parameter, capturing the clients' degree of impatience (similar to the proportion of switching investors in Dieci et al., 2006). Funds are reallocated based on the relative performance of competing fund managers, according to the discrete choice principle. Therefore, portfolio managers may experience an exogenous growth of their wealth, in addition to the endogenous growth due to returns on the employed capital. The model framework appears promising to investigate the market impact of the fund flows and to incorporate different types of ‘behavioral biases’ into HAMs. In particular, Palczewski et al. (2016) show that even a small amount of freely flowing capital can have a large impact on price movements if investors exhibit ‘recency bias’ in evaluating fund performance.

In a somewhat related framework with heterogeneous investment funds using ‘value investing’, Thurner et al. (2012) explore the joint impact of wealth dynamics and the flows of capital among competing investment funds. Evolutionary pressure generated by short-run competition forces fund managers to make leveraged asset purchases with margin calls. Simulation results highlight a new mechanism to fat tails and clustered volatility, which is linked to wealth dynamics and leverage-induced crashes. Moreover, this framework appears promising to test different credit regulation policies (Poledna et al., 2014) and to investigate the impact of bank leverage management on the stability properties of the financial system (Aymanns and Farmer, 2015).

3 HAMs of Single Asset Market in Continuous-Time

Historical information plays a very important role in testing efficient market hypothesis in financial markets. In particular, it is crucial to understand how quickly market prices reflect fundamental shocks and how much information is contained in the historical prices. Empirical evidence shows that stock markets react with a delay to information on fundamentals and that information diffuses gradually across markets (Hou and Moskowitz, 2005, Hong et al., 2007). Based on market underreaction and overreaction hypotheses, momentum and contrarian strategies are widely used by financial market practitioners and their profitability has been extensively investigated by academics. De Bondt and Thaler (1985) and Lakonishok et al. (1994) find supporting evidence on the profitability of contrarian strategies for a holding period of 3–5 years based on the past 3–5 years' returns. In contrast, Jegadeesh and Titman (1993, 2001) among many others, find supporting evidence on the profitability of momentum strategies for holding periods of 3–12 months based on the returns over the past 3–12 months. Time series momentum investigated recently in Moskowitz et al. (2012) characterizes a strong positive predictability of a security's own past returns. It becomes clear that the time horizons of historical prices play crucial roles in the performance of contrarian and momentum strategies. Many theoretical studies have tried to explain the momentum,6 however, as argued in Griffin et al. (2003), “the comparison is in some sense unfair since no time horizon is specified in most behavioral models”.

In the literature of HAMs, the heterogeneous expectations of agents, in particular of chartists, are formed based on price trends such as moving average of historical prices. In discrete-time models, with different time horizon, the dimension of the model is different. To examine the effect of time horizon analytically, we need to study the model with different dimension separately. Also, as the time horizon increases, it becomes more difficult analytically in dealing with high dimensional nonlinear dynamic system. This challenge is illustrated in Chiarella et al. (2006b) when examining the effect of different moving averages on market stability. Therefore, how different time horizons of historical prices affect price dynamics becomes a challenging issue in the current HAMs.

This section introduces some of the recent developments of HAMs of a single risky asset (and a riskless asset) in continuous time to deal with the price delay problems in behavioral finance and HAMs literature. In continuous-time HAMs, the time horizon of historical price information is simply captured by a time delay parameter. Such models are characterized mathematically by a system of stochastic delay differential equations, which provide a more broad framework to investigate the joint effect of adaptive behavior of heterogeneous agents and the impact of historical prices.

Development of deterministic delay differential equation models to characterize fluctuation of commodity prices and cyclic economic behavior has a long history,7 however the application to asset pricing and financial markets is relatively new. This section bridges HAMs with traditional approaches in continuous-time finance to investigate the impact of moving average rules over different time horizon (He and Li, 2012) in Section 3.1, the profitability of fundamental and momentum strategies (He and Li, 2015a) in Section 3.2, and optimal asset allocation with time series momentum and reversal (He et al., 2018) in Section 3.3.

3.1 A Continuous-Time HAM with Time Delay

We now introduce the continuous-time model of He and Li (2012) and demonstrate first that the result of Brock and Hommes on rational routes to market instability in discrete-time also holds in continuous time. That is, adaptive switching behavior of agents can lead to market instability as the switching intensity increases. We then show a double edged effect of an increase in the time horizon of historical price information on market stability. An initial increase in time delay can destabilize the market, leading to price fluctuations. However, as the time delay increases further, the market is stabilized. This double edged effect is a very different feature of continuous-time HAMs from discrete-time HAMs. With noisy fundamental value and liquidity traders, the continuous-time model is able to generate long deviations of market price from the fundamental price, bubbles, crashes, and volatility clustering.

Consider a financial market with a risky asset and let P(t)Image denote the (cum) price per share of the risky asset at time t. The market consists of fundamentalists, chartists, liquidity traders, and a market maker. The fundamentalists believe that the market price P(t)Image is mean-reverting to the fundamental price F(t)Image, and their demand is given by Zf(t)=βf[F(t)P(t)]Image, with βf>0Image measuring the mean-reverting speed of the market price to the fundamental price. The chartists are modeled as trend followers, believing that the future market price follows a price trend u(t)Image, and their demand is given by8 Zc(t)=tanh(βc[P(t)u(t)])Image with βc>0Image measuring the extrapolation of the trend followers to the price trend. Among various price trends used in practice, we consider u(t)Image as a normalized exponentially decaying weighted average of historical prices over a time interval [tτ,t]Image,

(11) u(t)=k1ekτttτek(ts)P(s)ds,

Image (11)

where time delay τ(0,)Image represents time horizon of historical prices, k>0Image measures the decay rate of the weights on the historical prices. In particular, when k0Image, the weights are equal and the price trend u(t)Image in (11) is simply given by the standard moving average (MA) with equal weights, u(t)=1τttτP(s)dsImage. When kImage, all the weights go to the current price so that u(t)P(t)Image. For the time delay, when τ0Image, the trend followers regard the current price as the price trend. When τImage, the trend followers use all the historical prices to form the price trend, u(t)=ktek(ts)P(s)dsImage. In general, for 0<k<Image, Eq. (11) can be expressed as a delay differential equation with time delay τ

du(t)=k1ekτ[P(t)ekτP(tτ)(1ekτ)u(t)]dt.

Image

The demand of liquidity traders is i.i.d. normally distributed with mean of zero and standard deviation of σM(>0)Image.

Let nf(t)Image and nc(t)Image represent the market fractions of agents who use the fundamental and trend following strategies, respectively. Their net profits over a short time interval [t,t+dt]Image can be measured, respectively, by πf(t)dt=Zf(t)dP(t)CfdtImage and πc(t)dt=Zc(t)dP(t)CcdtImage, where Cf,Cc0Image are constant costs of the strategies. To measure strategy performance, we introduce a cumulated profit over the time interval [tτ,t]Image by Ui(t)=η1eητttτeη(ts)πi(s)ds,i=f,cImage, where η>0Image measures the decay of the historical profits. Consequently, dUi(t)=η[πi(t)eητπi(tτ)1eητUi(t)]dtImage for i=f,cImage. Following Hofbauer and Sigmund (1998, Chapter 7), the evolution dynamics of the market populations are governed by

dni(t)=βni(t)[dUi(t)dˉU(t)], for i=f,c,

Image

where dˉU(t)=nf(t)dUf(t)+nc(t)dUc(t)Image is the average performance of the two strategies and β>0Image measures the intensity of choice. The switching mechanism in the continuous-time setup is consistent with the one used in discrete-time HAMs. In fact, it can be verified that the dynamics of the market fraction nf(t)Image satisfy dnf(t)=βnf(t)(1nf(t))[dUf(t)dUc(t)]Image, leading to nf(t)=eβUf(t)/(eβUf(t)+eβUc(t))Image, which is the discrete choice model used in Brock and Hommes (1998).

Finally, the price P(t)Image is adjusted by the market maker according to dP(t)=μ[nf(t)Zf(t)+nc(t)Zc(t)]dt+σMdWM(t)Image, where μ>0Image represents the speed of the price adjustment of the market maker, WM(t)Image is a standard Wiener process capturing the random excess demand process either driven by unexpected market news or liquidity traders, and σM>0Image is a constant. To sum up, the market price of the risky asset is determined according to the stochastic delay differential system

(12) {dP(t)=μ[nf(t)Zf(t)+(1nf(t))Zc(t)]dt+σMdWM(t),du(t)=k1ekτ[P(t)ekτP(tτ)(1ekτ)u(t)]dt,dU(t)=η1eητ[π(t)eητπ(tτ)(1eητ)U(t)]dt,

Image (12)

where U(t)=Uf(t)Uc(t)Image, nf(t)=1/(1+eβU(t))Image, Zf(t)=βf(F(t)P(t))Image, Zc(t)=tanh[βc(P(t)u(t))]Image, C=CfCcImage, and

π(t)=πf(t)πc(t)=μ[nf(t)Zf(t)+(1nf(t))Zc(t)][Zf(t)Zc(t)]C.

Image

By assuming that the fundamental price is a constant F(t)ˉFImage and there is no market noise σM=0Image, system (12) becomes a deterministic delay differential system with (P,u,U)=(ˉF,ˉF,C)Image as the unique fundamental steady state. He and Li (2012) show that the steady state is locally stable for either small or large time delay τ when the market is dominated by the fundamentalists. Otherwise, the steady state becomes unstable through Hopf bifurcations as the time delay increases. This result is in line with the results obtained in discrete-time HAMs. However, different from discrete-time HAMs, the continuous-time model shows that the fundamental steady state becomes locally stable again when the time delay is large enough. This is illustrated by the bifurcation diagram of the market price with respect to τ in Fig. 4A.9 It shows that there are two Hopf bifurcation values 0<τo<τ1Image occurring at τ=τ0(8)Image and τ=τ1(28)Image. The fundamental steady state is locally stable when the time delay is small, τ[0,τ0)Image, then becomes unstable for τ(τo,τ1)Image, and then regains the local stability for τ>τ1Image. Due to the problem of high dimensionality, such analysis on the effect of historical price on market price in discrete-time HAMs can become very complicated, see Chiarella et al. (2006b) that examining the effect of different moving averages on market stability. It is the continuous-time model that facilitates such analysis on the stability effect of time horizon of historical prices and stability switching. The bifurcation diagram of the market price with respect to the switching intensity β is given in Fig. 4B. It shows that the fundamental steady state is locally stable when the switching intensity β is low, becoming unstable as the switching intensity increases, bifurcating to periodic price with increasing fluctuations. This is consistent with the discrete-time HAMs.

Image
Figure 4 Bifurcation diagram of the market price with respect to τ in (A) and β in (B).

For the deterministic model, when the steady state becomes unstable, it bifurcates to stable periodic solutions through a Hopf bifurcation. The periodic fluctuations of the market prices are associated with periodic fluctuations of the market fractions, illustrated in Fig. 5A. Based on the bifurcation diagram in Fig. 4A, the steady state is unstable for τ=16Image. Fig. 5A shows that both price and market fraction fluctuate periodically. It shows that, when the fundamental steady state becomes unstable, the market fractions tend to stay away from the steady state market fraction level most of the time and a mean of nfImage below 0.5 clearly indicates the dominance of the trend following strategy. To examine the effect of population evolution, we compare the case without switching β=0Image to the case with switching β0Image. Fig. 5A clearly shows that the evolution of population increases the fluctuations in both price and market fraction.

Image
Figure 5 Time series of (A) deterministic market price P (solid line) and market fraction nf(t) of fundamentalists (dotted line) for τ = 16 and stochastic fundamental price (the dotted line) and market price (the solid line) for two delays (B) τ = 3 and (C) τ = 16 with and without switching.

For the stochastic model with a random walk fundamental price process, Fig. 5B demonstrates that the market price follows the fundamental price closely when τ=3Image, while Fig. 5C illustrates that the market price fluctuates around the fundamental price in cyclical fashion for τ=16Image. To examine the effect of population evolution, we compare the case without switching β=0Image to the case with switching β0Image. Fig. 5B shows that the evolution of population has insignificant impact on the price dynamics when the fundamental steady state of the underlying deterministic model is locally stable for τ=3Image. However, when the fundamental steady state becomes unstable for τ=16Image, the fluctuations in both price and market fraction become more significant. Therefore the stochastic price behavior is underlined by the dynamics of the corresponding deterministic model. He and Li (2012) further explore the potential of the stochastic model in generating volatility clustering and long range dependence in volatility. The underlying mechanism and the interplay between the nonlinear deterministic dynamics and noises are very similar to the discrete-time HAM by He and Li (2007). The framework can be used to study the joint impact of many heterogeneous strategies based on different time horizons of historical prices on market stability.

3.2 Profitability of Momentum and Contrarian Strategies

Momentum and contrarian strategies are widely used by market practitioners to profit from momentum in the short-run and mean-reversion in the long-run in financial markets. Empirical profitability of these strategies based on moving averages with different time horizon of historical prices and different holding period has been extensively investigated in the literature (Lakonishok et al., 1994, Jegadeesh and Titman, 1993, 2001, and Moskowitz et al., 2012).

To explain the profitability and the underlying mechanism of time series momentum and contrarian strategies, He and Li (2015a) propose a continuous-time HAM consisting of fundamental, momentum, and contrarian traders. They develop an intuitive and parsimonious financial market model of heterogeneous agents to study the impact of different time horizons on market price and profitability of fundamental, momentum and contrarian trading strategies. They show that the performance of momentum strategy is determined by the historical time horizon, investment holding period, and market dominance of momentum trading. More specifically, due to price continuity, the price trend based on the moving average of historical prices becomes very significant (apart from over very short time horizon). Therefore, when momentum traders are more active in the market, the price trend becomes very sensitive to the shocks, which is characterized by the destabilizing role of the momentum trading to the market. This provides a profit opportunity for momentum trading with short, not long, holding time horizons. When momentum traders are less active in the market, they always loose. The results provide some insights into the profitability of time series momentum over short, not long, holding periods. We now brief the main results of He and Li (2015a).

Consider a continuous-time model with fundamentalists who trade according to fundamental analysis and momentum and contrarian traders who trade differently based on price trend calculated from moving averages of historical prices over different time horizons. Let P(t)Image and F(t)Image denote the log (cum dividend) price and (log) fundamental value10 of a risky asset at time t, respectively. The fundamental traders buy (sell) the stock when the current price P(t)Image is below (above) the fundamental value F(t)Image. For simplicity, we assume that the fundamental return follows a pure white noise process dF(t)=σFdWF(t)Image with F(0)=ˉFImage, σF>0Image, and WF(t)Image is a standard Wiener process.

Regarding the momentum and contrarian trading, as in the previous section, we assume that both momentum and contrarian traders trade based on their estimated market price trends, although they behave differently. Momentum traders believe that future market price follows a price trend um(t)Image, while contrarians believe that future market price goes opposite to a price trend uc(t)Image. The price trend used for the momentum traders and contrarians can be different in general. Among various price trends used in practice, the standard moving average (MA) rules with different time horizons are the most popular ones, ui(t)=1τittτiP(s)dsImage for i=m,cImage, where the time delay τi0Image represents the time horizon of the MA. Assume the excess demand of the momentum traders and contrarians are given, respectively, by Zm(t)=gm(P(t)um(t))Image and Dc(t)=gc(uc(t)P(t))Image, where gi(x)Image satisfies gi(0)=0,gi(x)>0,gi(0)=βi>0,xgi(x)<0Image for x0Image and i=m,cImage, and parameter βiImage represents the extrapolation rate when the market price deviation from the trend is small.

Assume a zero net supply in the risky asset and let αfImage, αmImage, and αcImage be the fixed market population fractions of the fundamental, momentum, and contrarian traders,11 respectively, with αf+αm+αc=1Image. Following Beja and Goldman (1980) and Farmer and Joshi (2002), the price P(t)Image at time t is determined by

(13) dP(t)=μ[αfZf(t)+αmZm(t)+αcZc(t)]dt+σMdWM(t),

Image (13)

where μ>0Image represents the speed of the price adjustment of the market maker, WM(t)Image is a standard Wiener process, independent of WF(t)Image, capturing the random demand of either noise or liquidity traders, and σM0Image is constant.

By assuming a constant fundamental price F(t)ˉFImage and no market noise σM=0Image, system (13) becomes a deterministic delay integro-differential equation,

(14) dP(t)dt=μ[αfβf(ˉFP(t))+αmtanh(βm(P(t)1τmttτmP(s)ds))+αctanh(βc(P(t)1τcttτcP(s)ds))].

Image (14)

It is easy to see that P(t)=ˉFImage, the fundamental steady state, is the unique steady state price of system (14). He and Li (2015a) examine different role of the time horizon used in the MA by either the contrarians or momentum traders. When both strategies are employed in the market, the market stability of system (14) can be characterized by the following proposition.

Proposition 3.1

If τm=τc=τImage, then the fundamental steady state price P=ˉFImage of system (14) is

  1. 1.  locally stable for all τ0Image when γm<γc+γf/(1+a)Image;
  2. 2.  locally stable for either 0τ<τlImage or τ>τhImage and unstable for τl<τ<τhImage when γc+γf/(1+a)γmγc+γfImage; and
  3. 3.  locally stable for τ<τlImage and unstable for τ>τlImage when γm>γc+γfImage.

Here τ1=2(γmγc)/(γfγm+γc)2Image, and τl(<τ1)Image and τh((τl,τ1))Image are the minimum and maximum positive roots, respectively, of the equation

τγmγc(γfγm+γc)2cos[2(γmγc)τ(γfγm+γc)2τ2]1=0.

Image

The three conditions (1) γm<γc+γf1+aImage, (2) γc+γf1+aγmγc+γfImage, and (3) γm>γc+γfImage in Proposition 3.1 characterize three different states of market stability, having different implications to the profitability of momentum trading. For convenience, market state k is referred to condition (k) for k=1,2,3Image in the following discussion. Numerical analysis shows that for market state 1, the fundamental price is locally stable, independent of the time horizon; for market state 2, the fundamental price is locally stable when τ[0,τl)(τh,)Image and becomes unstable when τ(τl,τh)Image (the stability switches twice); while for market state 3, the first (Hopf bifurcation) value τl(0.22)Image leads to stable limit cycles for τ>τlImage (the stability switches only once at τlImage).

The profitability of different strategies based on the stochastic model is closely related to the market states and holding period. In market state 1, the market is dominated jointly by the fundamental and contrarian traders (so that γm<γc+γf/(1+a)Image). In this case, the stability of the fundamental price of the underlying deterministic model is independent of the time horizon. Monte Carlo simulations show that the contrarian and fundamental strategies are profitable, but not the momentum strategy and the market maker, underlined by significant and negative ACs for small lags and insignificant ACs for large lags. This corresponds to market overreaction in short-run and hence the fundamental and contrarian trading can generate significant profits. Without under-reaction in this case, the momentum trading is not profitable.

In market state 2, the momentum traders are active, but their activities are balanced by the fundamental and contrarian traders. In this case, the fundamental and contrarian trading strategies are still profitable, but not the momentum traders and the market maker. This is illustrated by the average accumulated profits based on a typical simulation with time horizon τ=0.5Image and holding period h=2Image in Fig. 6A. The return ACs based on Monte Carlo simulation show some significantly negative ACs over short lags, indicating the profitability of the fundamental and contrarian trading due to market overreaction, but not for the momentum trading.

Image
Figure 6 The average accumulated profits based on a typical simulation with different time horizon τ and holding period h in different market state: (A) τ = h = 0.5 in market state 2; (B) τ = h = 0.5 in market state 3; (C) τ = h = 3 in market state 3; and (D) τ = 3, h = 0.5 in market state 3. Here γf = 20, γm = 22.6, and γc = 5 in (A) and γf = 2, γm = 20, and γc = 10 in (B)–(D).

In market state 3, the market is dominated by the momentum traders and their destabilizing role. Over short time horizon, the market price fluctuates due to the unstable fundamental price of the underlying deterministic system. When the market price increases, the price trend follows the market price closely and increases too. The momentum trading with short holding period hence becomes profitable by taking long positions. Similarly, when the market price declines, the price trend follows and hence the momentum trading with short holding period is profitable by taking short positions. Therefore, the momentum trading strategies are profitable, but not the contrarians, illustrated by Fig. 6B for τ=h=0.5Image and Fig. 6C for τ=h=3Image respectively. Over long time horizon, the market price fluctuates widely due to the unstable fundamental value of the underlying deterministic system. A longer time horizon makes the price trend less sensitive to the changes in price and the shocks. The dominance of the momentum trading and market price continuation make the momentum trading with short holding period more profitable, illustrated by Fig. 6D for τ=3Image and h=0.5Image. With long holding period, the momentum trading mis-matches the profitability opportunity and hence becomes less profitable. With long time horizon and long holding period, Fig. 6C also illustrates that the fundamental and contrarian strategies are profitable, but not the momentum strategy. For time horizon and holding period from 1 to 60 months, the model is able to replicate the time series momentum profit explored for the S&P 500. The results are consistent with Moskowitz et al. (2012) who find that the time series momentum strategy with 12 months horizon and one month holding is the most profitable among others.

In summary, the stochastic delay integro-differential system of the model provides a unified approach to deal with different time horizons of momentum and contrarian strategies. The profitability is closely related to the market states defined by the stability of the underlying deterministic model. In particular, in market state 3 where the momentum traders dominate the market, the momentum strategy is profitable with short, but not long, holding periods. Some explanations to the mechanism of the profitability through autocorrelation patterns and the under-reaction and overreaction hypotheses are also provided in He and Li (2015a).

3.3 Optimal Trading with Time Series Momentum and Reversal

Short-run momentum and long-run reversal are two of the most prominent financial market anomalies. Though market timing opportunities under mean reversion in equity return are well documented (Campbell and Viceira, 1999 and Wachter, 2002), time series momentum (TSM) has been explored recently in Moskowitz et al. (2012). Intuitively, if we incorporate both return momentum and reversal into a trading strategy optimally, we would expect to outperform the strategies based only on return momentum or reversal, and even the market index. To capture this intuition, He et al. (2018) develop a continuous-time asset price model, derive an optimal investment strategy theoretically, and test the strategy empirically. They show that, by combining market fundamentals and timing opportunity with respect to market trend and volatility, the optimal strategy based on the time series momentum and reversal significantly outperforms, both in-sample and out-of-sample, the S&P 500 and pure strategies based on either time series momentum or reversal only. We now outline the main results and refer the details to He et al. (2018).

Consider a financial market with two tradable securities. A riskless asset B satisfies dBt/Bt=rdtImage with a constant risk-free rate r. The risky asset StImage satisfies

dSt/St=[ϕmt+(1ϕ)μt]dt+σSdZt,dμt=α(ˉμμt)dt+σμdZt,

Image

where α>0,ˉμ>0Image, and mt=(1/τ)ttτdSuSuImage. Here ϕ is a constant, ˉμImage is the constant long-run expected return, α measures the speed of the convergence of μtImage to ˉμImage, σSImage and σμImage are two-dimensional volatility vectors, and ZtImage is a two-dimensional vector of independent Brownian motions. Therefore, the expected return is given by a combination of a momentum component mtImage based on a moving average of the past returns and a long-run mean-reversion component μtImage based on market fundamentals such as dividend yield.

Consider a typical long-term investor who maximizes the expected log utility of terminal wealth at time T(>t)Image. Let WtImage be the wealth of the investor at time t and πtImage be the fraction of the wealth invested in the stock. Then

(15) dWtWt=(πt[ϕmt+(1ϕ)μtr]+r)dt+πtσSdZt.

Image (15)

By applying the maximum principle for optimal control of stochastic delay differential equations, He et al. (2018) derive the optimal investment strategy

(16) πt=ϕmt+(1ϕ)μtrσSσS.

Image (16)

That is, by taking into account the short-run momentum and long-run reversal, as well as the timing opportunity with respect to market trend and volatility, a weighted average of the momentum and mean-reverting strategies is optimal.

This result has a number of implications. (i) When the asset price follows a geometric Brownian motion process with mean-reversion drift μtImage, namely ϕ=0Image, the optimal investment strategy (16) becomes πt=μtrσSσSImage. This is the optimal investment strategy with mean-reverting returns obtained in the literature (Campbell and Viceira, 1999 and Wachter, 2002). In particular, when μt=ˉμImage is a constant, the optimal portfolio collapses to the optimal portfolio of Merton (1971). (ii) When the asset return depends only on the momentum, namely ϕ=1Image, the optimal portfolio (16) reduces to πt=mtrσSσSImage. If we consider a trading strategy based on the trading signal indicated by the excess moving average return mtrImage only, with τ=12Image months, the strategy of long/short when the trading signal is positive/negative is consistent with the TSM strategy used in Moskowitz et al. (2012). Therefore, if we only take fixed long/short positions and construct simple buy-and-hold momentum strategies over a large range of look-back and holding periods, the TSM strategy of Moskowitz et al. (2012) can be optimal when the mean reversion is not significant in financial markets.

He et al. (2018) then examine the performance of the optimal portfolio in terms of the utility of the portfolio wealth empirically. As a benchmark, the log utility of $1 investment in the S&P 500 index from January 1876 grows to 5.765 at December 2012. With a time horizon of τ=12Image and one month holding period, the optimal portfolio wealth fractions and the evolution of the utility of the optimal portfolio wealth (lnWtImage) based on the estimated model from January 1876 to December 2012 are plotted in Fig. 7A and B, showing that the optimal portfolios outperform the market index measured by the utility of wealth (lnWtImage).

Image
Figure 7 Time series of the optimal portfolio (A) and the utility (B) of the optimal portfolio wealth (lnWtImage) from January 1876 until December 2012 for τ = 12.

4 HAMs of Multi-Asset Markets and Financial Market Interlinkages

A recent literature has been developed to understand the joint dynamics of multiple asset markets from the viewpoint of HAMs. In particular, research in this area investigates how investors' heterogeneity and changing behavior (including dynamic strategy and market selection) affect the comovement of prices, returns and volatilities in a multiple-asset framework. Modeling such interlinkages naturally introduces additional nonlinearities into HAMs and has the potential to address key issues in financial markets.

4.1 Stock Market Comovement and Policy Implications

A number of models extend the single-risky asset frameworks of Brock and Hommes (1998), Chiarella and He (2002), and Westerhoff (2003) to allow agents to switch not only across strategies but also across different asset markets. Westerhoff (2004) provides one of the first HAMs of interconnected financial markets in which both fundamentalists and chartists are simultaneously active. In each market, chartist demand is positively related to the observed price trends but negatively related to the risk of being caught in a bursting bubble. Asset prices react to the excess demand according to a log-linear price impact function. Chartists may switch between markets depending on short-run profit opportunities. The basic model of interacting agents and markets can naturally produce complex dynamics. A simple stochastic extension of the model can mimic the behavior of actual asset markets closely, offering an explanation for the high degree of stock price comovements observed empirically.

Westerhoff and Dieci (2006) extend the basic framework of Westerhoff (2004) to investigate the effect of transaction taxes when speculators can trade in two markets, and the related issue of regulatory coordination. The market fractions of fundamentalists and chartists active in each market evolve depending on the realized profitability of each ‘rule-market’ combination, which is affected by the adoption of transaction taxes. Log-price adjustments depend on excess demand and are subject to i.i.d. random noise (uncorrelated across markets). The joint dynamics of the two markets is investigated with and without transaction taxes. Moreover, the effectiveness12 of transaction taxes is assessed when tax is imposed in one market only and when uniform transaction taxes are imposed in both markets. It turns out that, while the market subject to a transaction tax becomes less distorted and less volatile, the other market may be destabilized. On the contrary, a uniform transaction tax tends to stabilize, by forcing agents to focus more strongly on fundamentals.

Building on the above frameworks, Schmitt and Westerhoff (2014) focus on coevolving stock prices in international stock markets. In their model, the demand of heterogeneous speculators is subject to different types of exogenous shocks (global shocks and shocks specific to markets or to trading rules). Investors switch between strategies and between markets depending on a number of behavioral factors and market circumstances. Besides reproducing a large number of statistical properties of stock markets (‘stylized facts’), the model shows how traders' behavior can amplify financial market interlinkages and generate stock price comovements and cross-correlations of volatilities.

Other recent papers are closely related to the above topics. For instance, Huang and Chen (2014) develop a nonlinear model with chartists and fundamentalists that generalizes the framework of Day and Huang (1990) to the case of two regional stock markets with a common currency, in order to investigate the global effects of financial market integration and of possible stabilization policies. In an agent-based model where portfolio managers allocate their funds between two asset markets, Feldman (2010) shows how fund managers' aggregate behavior can undermine global financial stability, whenever they enter the markets in large numbers, their leverage increases and their investment strategies are affected by behavioral factors (such as loss aversion). Overall, such models demonstrate the potential of HAMs for understanding the global effects of financial market interlinkages.

4.2 Heterogeneous Beliefs and Evolutionary CAPM

A further strand of research investigates the impact of behavioral heterogeneity in an evolutionary CAPM framework. More precisely, this literature adopts standard mean-variance portfolio selection across multiple assets (or asset classes/markets) and develops a dynamic CAPM framework with fundamental and technical traders. Investors update their beliefs about the means, variances and covariances of the prices or returns of the risky assets, based on fundamental information and historical prices. They may either use fixed rules (Chiarella et al., 2007, 2013a) or switch between different strategies based on their performance (Chiarella et al., 2013b). This framework is helpful to understand how investors' behavior can produce changes of the market portfolio and spillovers of volatility and correlation across markets. In particular, through the construction of a consensus belief, Chiarella et al. (2013b) develop a dynamic CAPM relationship between the market-average expected returns of the risky assets and their ex-ante betas in temporary equilibrium. Results show that systematic changes in the market portfolio and risk-return relationships may occur due to changes of investor sentiment (such as chartists acting more strongly as momentum traders). Besides providing behavioral explanations for the debated on time-varying betas, such models allow to compare theoretical ex-ante betas to commonly used ex-post beta estimates based on rolling-windows. The remainder of this section presents the model setup and key findings of Chiarella et al. (2013b).

4.2.1 A Dynamic Multi-Asset Model

Consider an economy with H agent-types, indexed by h=1,,HImage, where the agents within the same group are homogeneous in their beliefs and risk aversion. Agents invest in portfolios of a riskless asset (with a risk-free gross return Rf=1+rfImage) and N risky assets, indexed by j=1,,NImage (with N1Image). Vectors pt=(p1,t,,pN,t)Image, dt=(d1,t,,dN,t)Image, and xt:=pt+dtImage denote prices, dividends, and payoffs of the risky assets at time t. Assume that an agent of type h maximizes expected CARA utility, uh(w)=eθhwImage, of one-period-ahead wealth, where θhImage is the agent's absolute risk aversion coefficient. Then the optimal demand for the risky assets (in terms of number of shares) is determined as the N-dimensional vector zh,t=θ1hΩ1h,t[Eh,t(xt+1)Rfpt]Image, where Eh,t(xt+1)Image and Ωh,t=[Covh,t(xj,t+1,xk,t+1)]N×NImage are the subjective conditional expectation and variance-covariance matrix of the risky payoffs. Moreover, denote by nh,tImage the market fraction of agents of type h at time t. Market clearing requires:

(17) Hh=1nh,tzh,t=Hh=1nh,tθ1hΩ1h,t[Eh,t(xt+1)Rfpt]=zst,

Image (17)

where zst=s+ξtImage is a N-dimensional supply vector of the risky assets, subject to random supply shocks satisfying ξt=ξt1+σκκtImage, where κtImage is standard normal i.i.d. with E(κt)=0Image, Cov(κt)=IImage. Likewise, dividends dtImage are assumed to follow a N-dimensional martingale process, dt=dt1+σζζtImage, where ζtImage is standard normal i.i.d. with E(ζt)=0Image, Cov(ζt)=IImage, independent of κtImage.13 In spite of heterogeneous beliefs about asset prices, conditional beliefs about dividends are assumed to be homogeneous across agents and correct.

4.2.2 Price Dynamics Under Consensus Belief

Solving Eq. (17) one obtains the temporary equilibrium asset prices, ptImage, as functions of the beliefs, risk attitudes, and current market proportions of the H agent-types. The solution can be rewritten as if prices were determined by a homogeneous agent endowed with average risk aversion θa,t:=(Hh=1nh,tθ1h)1Image and a ‘consensus’ belief about the conditional first and second moments of the payoff process, {Ea,t,Ωa,t}Image, where

Ωa,t=θ1a,t(Hh=1nh,tθ1hΩ1h,t)1,Ea,t(xt+1)=θa,tΩa,tHh=1nh,tθ1hΩ1h,tEh,t(xt+1).

Image

From (17) and the assumption of homogeneous and correct beliefs about dividends, one obtains

(18) pt=1Rf[Ea,t(pt+1)+dtθa,tΩa,tzst].

Image (18)

Eq. (18) represents ptImage in a standard way as the discounted value of the expected end-of-period payoffs. The adjustment for the risk takes place through a negative correction to the dividends. The equilibrium prices decrease with the discount rate and increase with the expectations of future prices and dividends (other things being equal), whereas they tend to be negatively affected by risk aversion, risk perceptions, and the supply of assets.

4.2.3 Fitness and Strategy Switching

Based on the discrete choice model adopted in HAMs, the fraction nh,tImage of agents of type h depends on their strategy's fitness vh,t1Image, namely, nh,t=eηvh,t1/ZtImage, where Zt=heηvh,t1Image and η>0Image is the intensity of choice. The fitness is specified as vh,t=πh,tπBh,tChImage, where Ch0Image measures the cost of the strategy, and

(19) πh,t:=zh,t1(pt+dtRfpt1)θh2zh,t1Ωh,t1zh,t1,

Image (19)

(20) πBh,t:=(θa,t1θhs)(pt+dtRfpt1)θh2(θa,t1θhs)Ωh,t1(θa,t1θhs).

Image (20)

This performance measure generalizes the risk-adjusted profit introduced by Hommes (2001) represented by (19).14 It views strategy h as a successful strategy only to the extent that portfolio zh,t1Image outperforms (in terms of risk-adjusted profitability) portfolio zBh,t1:=θa,t1θhsImage. The latter can be naturally interpreted as a ‘benchmark’ portfolio for type-h agents, based on their risk aversion θhImage.15 Moreover, as shown in Chiarella et al. (2013b), the fitness measure vh,tImage is not affected by the differences in risk aversion across agents.

4.2.4 Fundamentalists and Trend Followers

In particular, the model focuses on the interplay of fundamentalists and trend followers, indexed by h{f,c}Image, respectively. Based on their beliefs in mean reversion, the price expectations of the fundamentalists are specified as Ef,t(pt+1)=pt1+α(Ef,t(pt+1)pt1)Image, where pt=(p1,t,,pN,t)Image is the vector of fundamental values at time t, α:=diag[α1,,αN]Image, and αj[0,1]Image reflects their confidence in the fundamental price for asset j. The beliefs of the fundamentalists about the covariance matrix of the payoffs are assumed constant, Ωf,t=Ω0:=(σjk)N×NImage. Fundamental prices ptImage are assumed to evolve exogenously as a martingale process, consistent with the assumed dividend and supply processes. Moreover, ptImage is also consistent with Eq. (18) under the special case of homogeneous and correct first-moment beliefs, constant risk aversion θImage, and constant second moment beliefs Ω0Image. This results in

(21) pt=1rf(dtθΩ0(s+ξt)),

Image (21)

which implies pt+1=pt+ϵt+1Image, where ϵt+1:=1rf(σζζt+1θΩ0σκκt+1)Image i.i.d. normal. The fundamental price process can be treated as ‘steady state’ of the dynamic heterogeneous-belief model.

Unlike the fundamentalists, trend followers form their beliefs about price trends based on the observed prices and (exponential) moving averages. Their conditional mean and covariance matrices are assumed to satisfy Ec,t(pt+1)=pt1+γ(pt1ut1)Image, Ωc,t=Ω0+λVt1Image, where ut1Image and Vt1Image are sample means and covariance matrices of historical prices pt1,pt2,Image. Moreover, γ=diag[γ1,,γN]>0Image, γjImage measures the ‘strength’ of extrapolation for asset j, and λ measures the sensitivity of the second-moment estimate to the sample variance. Quantities utImage and VtImage are updated recursively according to ut=δut1+(1δ)ptImage and Vt=δVt1+δ(1δ)(ptut1)(ptut1)Image, where parameter δ[0,1]Image is related to the weight of past information.

The optimal portfolios of fundamentalists and chartists are then given by, respectively,

(22) zf,t=θ1fΩ10[pt1+dt+α(ptpt1)Rfpt],

Image (22)

(23) zc,t=θ1c[Ω0+λVt1]1[pt1+dt+γ(pt1ut1)Rfpt].

Image (23)

4.2.5 Dynamic Model and Stability Properties

The stochastic nonlinear multi-asset HAM with two belief-types results in the following recursive equation for asset prices

(24) pt=θa,tRfΩa,t[nf,tθfΩ10(pt1+α(ptpt1))+nc,tθc(Ω0+λVt1)1(pt1+γ(pt1ut1))sξt]+1Rfdt,

Image (24)

where the average risk aversion and second-moment beliefs satisfy θa,t=(nf,tθf+nc,tθc)1Image and Ωa,t=1θa,t(nf,tθfΩ10+nc,tθc(Ω0+λVt1)1)1Image. In (24), market fractions evolve based on performances vf,t1Image and vc,t1Image, as follows:

nf,t=11+eη(vf,t1vc,t1),  nc,t=1nf,t,

Image

where

vf,t=(zf,t1θa,t1sθf)[pt+dtRfpt1θf2Ω0(zf,t1+θa,t1sθf)]Cf,vc,t=(zc,t1θa,t1sθc)×[pt+dtRfpt1θc2(Ω0+λVt2)(zc,t1+θa,t1sθc)]Cc,

Image

and CfCc0Image.

Despite the large dimension of the dynamical system, insightful analytical results about the steady state and its stability properties are possible for the ‘deterministic skeleton’, obtained by setting the supply and dividends at their unconditional mean levels ξt=0Image, dt=ˉdImage. The model admits a unique steady state16 (pt,ut,Vt,nf,t)=(p,p,0,nf):=FImage, where p=1rf(ˉdθaΩ0s)Image is the fundamental price vector of the deterministic system, θa=1/(nf/θf+nc/θc)Image is the average risk aversion and nf=1/(1+eη(CfCc))Image, nc=1nfImage are the market fractions of the fundamentalist and chartist, respectively, at the steady state. It turns out that the local stability of FImage is based on clear-cut and intuitive analytical relationships between chartist extrapolation and memory, fundamentalist confidence, and switching intensity. We set θ0:=θf/θcImage, CΔ:=CfCcImage, and denote by Jo{1,,N}Image the subset of assets characterized by ‘sufficiently’ strong extrapolation from the chartists, namely, by γj>Rf/δ1Image. In the typical case CΔ>0Image, the local stability results can be summarized as follows:

  1. (i)  If the chartist extrapolation is not very strong in general (namely, γjRf/δ1Image for all j{1,,N}Image), the steady state FImage is locally stable for any level of the switching intensity η;
  2. (ii)  If chartist extrapolation is sufficiently strong for some (possibly for all) assets (JoImage), then FImage is locally stable when the switching intensity is not too strong, namely η<ˆηm:=minjJoˆηjImage, where ˆηjImage for asset j is defined by

(25) ˆηj:=1CΔlnRfδ(1αj)θ0[δ(1+γj)Rf].

Image (25)

  1. Moreover, for increasing switching intensity FImage undergoes a Neimark–Sacker bifurcation at η=ˆηmImage.

Roughly speaking, investors' switching intensity η is not sufficient, per se, to destabilize the steady state FImage (case (i)), but the possibility that investors' behavior destabilizes the system depends on the joint effect of the switching intensity η and the chartists' strengths of extrapolation γjImage, j=1,2,,NImage. In case (ii), the threshold ˆηjImage is determined for each asset according to (25), depending, amongst others, negatively on γjImage and positively on αjImage. Hence, even when chartist extrapolation is strong enough for some asset j (so that γj>Rf/δ1Image), the system can still be stable when the fundamentalists dominate the market at the steady state and the switching intensity is not too large. Conversely, since the stability depends on the lowest threshold amongst assets (ˆηmImage), a large extrapolation on one or few assets is sufficient for the whole system to be eventually destabilized for large enough η. Numerical investigations confirm that, by increasing η in case (ii), fluctuations are initially ‘confined’ to the asset with the lowest ˆηjImage and then spill over to the whole system of interconnected assets. As for the ‘non asset-specific’ parameters, the above results show that increases in δ, CfImage, and θfImage (respectively RfImage, CcImage, and θcImage) tend to reduce (respectively to increase) all thresholds ˆηjImage, j=1,2,...,NImage. In particular, larger values of the ratio θ0=θf/θcImage of the fundamentalist and chartist risk aversion and of the strategy cost differential CΔ=CfCcImage reduce the stability domain, whereas a larger risk-free return RfImage or a faster decay in chartist moving averages (i.e. a smaller δ) widens the stability domain.

4.2.6 Nonlinear Risk-Return Patterns

Further results concern the impact of the dynamic correlation structure on the global properties of the stochastic model. Although the levels of the fundamental prices do depend on the ‘exogenous’ subjective beliefs about variances and covariances, Ω0Image, such beliefs have no influence on the local stability properties.17 However, second-moment beliefs and their evolution turn out to be very important for the dynamics of the nonlinear system buffeted by exogenous noise. The nonlinear stochastic model is characterized by emerging patterns and systematic changes in risk-return relationships that can by no means be explained by the linearized model. One important example concerns the nonlinear stochastic nature of the time-varying ex-ante beta coefficients implied by the model (based on the consensus beliefs), and of the realized betas, estimated using rolling windows.18 The value at time t and the payoff at time t+1Image of the market portfolio are given by Wm,t=ptsImage and Wm,t+1=xt+1sImage, respectively, while rj,t+1=xj,t+1/pj,t1Image, rm,t+1=Wm,t+1/Wm,t1Image represent the returns of risky asset j and of the market portfolio, respectively. Hence, under the consensus belief, Ea,t(Wm,t+1)=Ea,t(xt+1)sImage, Vara,t(Wm,t+1)=sΩa,tsImage, Ea,t(rj,t+1)=Ea,t(xj,t+1)pj,t1Image, Ea,t(rm,t+1)=Ea,t(Wm,t+1)Wm,t1Image. Following Chiarella et al. (2011), one obtains the CAPM-like return relation19

(26) Ea,t(rt+1)rf1=βa,t[Ea,t(rm,t+1)rf],

Image (26)

where rt+1Image is the vector collecting the risky returns and βa,t=(β1,t,,βN,t)Image, βj,t=Cova,t(rm,t+1,rj,t+1)Vara,t(rm,t+1)Image are the ex-ante beta coefficients, in the sense that they reflect the temporary market equilibrium condition under the consensus beliefs Ea,tImage and Ωa,tImage. In the case of two risky assets, Fig. 8 (from the top-left to bottom-right) shows the time series of asset prices (ptImage), asset returns (rtImage), the aggregate wealth shares invested in the risky assets (i.e. the market portfolio weights, denoted as ωt:=(ω1,t,ω2,t)Image), the ex-ante betas of the risky assets under the consensus belief (βa,tImage), and the estimates of the betas using rolling windows of 100 and of 300 periods.20 In particular, the variation of the ex-ante beta coefficients is significant and seems to indicate substantially different levels over different subperiods. Although the rolling estimates of the betas do not necessarily reflect the nature of the ex-ante betas implied by the CAPM (see also Chiarella et al., 2013a), the 100-period and the (smoother) 300-period rolling betas also reveal systematic changes in risk-return relationships, with patterns similar to the ex-ante betas.

Image
Figure 8 Dynamics of the evolutionary CAPM (monthly time step).

Finally, further numerical results on the relationship between trading volume and volatility indicate that the ACs for both volatility and trading volume are highly significant and decaying over long lags, which is close to what we have observed in financial markets. Moreover, the correlation between price volatility and trading volume of the risky assets is remarkably influenced by the assets' correlation structure.

From a broader perspective, the results described in this section are part of a growing stream of research. They show that asset diversification in a dynamic setting where investors rebalance their portfolios based on heterogeneous strategies and behavioral rules may produce aggregate effects that different substantially from risk reduction and equilibrium risk-return relationships predicted by standard mean-variance analysis and finance theory. Amongst recent work in this area, Brock et al. (2009) show that the introduction of additional hedging instruments in the baseline asset pricing setup of Brock and Hommes (1998) may have destabilizing effects in the presence of heterogeneity and adaptive behavior according to performance-based reinforcement learning. In an evolutionary finance setting that allows for the coexistence of different trading strategies, the stochastic multi-asset model of Anufriev et al. (2012) shows the existence of strong trading-induced excess covariance in equilibrium, which is a key ingredient of systemic risk. Corsi et al. (2016) investigate the dynamic effect of financial innovation and increasing diversification in a model of heterogeneous financial institutions subject to Value-at-Risk constraints. They show that this may lead to systemic instabilities, through increased leverage and overlapping portfolios. Similar channels of contagion and systemic risk in financial networks are investigated by Caccioli et al. (2015).

4.3 Interacting Stock Market and Foreign Exchange Market

The recent work of Dieci and Westerhoff (2010) and Dieci and Westerhoff (2013b) investigates how the trading activity of foreign-based stock market speculators – who care both about stock returns and exchange rate movements – can affect otherwise independent stock markets denominated in different currencies and the related foreign exchange market. We brief the main findings in the following.

Let us abstract from the impact of international trade on exchange rates, and focus on the sole effect of financial market speculators. For simplicity, let us define cross-market traders the investors from one country who are active in the stock market of the other country, in contrast to home-market traders.21 Quantities PtImage, QtImage, and StImage denote the price of the domestic asset (in domestic currency), the price of the foreign asset (in foreign currency) and the exchange rate,22 while PImage, QImage, and SImage denote their fundamental values, respectively. We use lowercase letters for log-prices ptImage, qtImage, stImage, pImage, qImage, sImage, respectively.

Exchange rate movements are driven by the excess demand for domestic currency. As such, they are directly affected by foreign exchange speculators, but they also depend, indirectly, on stock transactions of cross-market traders. This is captured by:

(27) st+1st=αS(Ut+Xt+Yt),αS>0,

Image (27)

where (positive or negative) quantities UtImage, XtImage, and YtImage are different components of the excess demand for domestic currency, expressed in currency units. More precisely, UtImage is the excess demand for domestic currency due to direct speculation in the foreign exchange market (to be specified later), Xt:=Pt˜Dt=˜Dtexp(pt)Image is the currency excess demand from foreign traders active in the domestic stock market (and demanding/supplying ˜DtImage units of domestic asset), and Yt:=Qt˜Zt/St=˜Ztexp(qtst)Image is the excess demand generated by domestic traders active in the foreign stock market (since ˜ZtImage units of foreign stock correspond to Qt˜ZtImage units of foreign currency and thus result in a counter transaction of Qt˜Zt/StImage units of domestic currency).

Similar price adjustment mechanisms are assumed for the two stock markets:

(28) pt+1pt=αPDEt,qt+1qt=αQZEt,αP,αQ>0

Image (28)

where DEtImage and ZEtImage denote the excess demand for the domestic and foreign stock, respectively, including the components ˜DtImage and ˜ZtImage from cross-market traders, as explained below. In a framework with two agent-types, both DEtImage and ZEtImage can be modeled as the sum of four components, representing the demand of domestic and foreign chartists and fundamentalists. At time t, the excess demand DEtImage for the domestic asset is given by:

(29) DEt=β(ptpt1)+θ(ppt)+˜Dt,

Image (29)

where ˜Dt=˜β(st+ptst1pt1)+˜θ(sst+ppt)Image and β,θ,˜β,˜θ0Image. Both β(ptpt1)Image and θ(ppt)Image represent the demand from domestic chartists and fundamentalists, based on the observed price trend and the observed mispricing, respectively. Similar comments hold for demands ˜β(st+ptst1pt1)Image and ˜θ(sst+ppt)Image from foreign chartists and fundamentalists, respectively, which depend also on the observed trend and misalignment of the exchange rate. Symmetrically, demand ZEtImage for the foreign asset is given by:

(30) ZEt=γ(qtqt1)+ψ(qqt)+˜Zt,

Image (30)

where ˜Zt=˜γ(st+qt+st1qt1)+˜ψ(s+st+qqt)Image and γ,ψ,˜γ,˜ψ0Image. The four terms γ(qtqt1)Image, ψ(qqt)Image, ˜γ(st+qt+st1qt1)Image and ˜ψ(s+st+qqt)Image represent the demands from foreign chartists, foreign fundamentalists, domestic chartists and domestic fundamentalists, respectively.

Dieci and Westerhoff (2013b) investigate the case without foreign exchange speculators (Ut0Image in Eq. (27)). Even if demand in the stock markets is linear in (log-)prices, the joint dynamics (27)(30) of the three markets results in a nonlinear dynamical system, by construction, due to the products, ‘price×quantity’, which govern the exchange rate dynamics (27).23 Moreover, although system (27)(30) is 6-dimensional, analytical stability conditions of the unique ‘fundamental’ steady state (FSS henceforth)24 can be derived in the case of symmetric markets, namely, β=γImage, ˜β=˜γImage, θ=ψImage, ˜θ=˜ψImage, q=p+sImage, thanks to a factorization of the characteristic polynomial of the Jacobian matrix at the FSS. This allows an exhaustive comparison of the stability condition for the integrated system with that of otherwise independent stock markets.

Fig. 9 illustrates the impact of parameters ˜βImage and ˜θImage of the cross-market traders on the stability of the steady state of otherwise independent symmetric stock markets. The stability region is represented in the plane of parameters β and θ of home-market traders. In both panels, the area bounded by the axes and by the two (thick) lines of equations θ=2(1+β)Image and β=1Image is the stability region for isolated symmetric markets, which we denote by SImage. Therefore, the markets in isolation may become unstable in the presence of sufficiently large chartist extrapolation (β) or fundamentalist reaction (θ) from the home-market traders. If the two markets interact (˜β,˜θ0Image), the FSS of the resulting integrated system is unstable for at least all the parameter combinations (β,θImage) originally in area SImage and now falling within the dark grey region, say area RSImage. The shape and extension of area RImage depend on the behavioral parameters of the cross-market chartists and fundamentalists, ˜βImage and ˜θImage. In particular, a larger chartist impact ˜βImage tends to enlarge area RImage. The left panel depicts the case ˜β<˜θ/2Image, in which the integration is always destabilizing (the new stability area is strictly a subset of the original one). A destabilizing effect prevails also in the opposite case, as shown in the right panel, for ˜β˜θ/2Image. However, in this case there exists a parameter region (light grey area) in which the otherwise unstable isolated markets (due to overreaction of fundamentalists) may be stabilized by strong extrapolation of the cross-market traders.

Image
Figure 9 Destabilization of two symmetric markets, due to the entry of new cross-market speculators. For parameters (β,θ) in the dark grey region, the markets are stable when considered in isolation, but the system of interacting market has an unstable FSS. Left panel: case ˜β<˜θ/2Image. Right panel: case ˜β˜θ/2Image.

We may interpret parameters β and θ as proportional to the total number of chartists and fundamentalists trading in their home markets, while ˜βImage and ˜θImage represent the number of additional cross-market traders of the two types. From this standpoint, the above results indicate a destabilizing effect of the market entry of additional cross-market speculators, once the two stock markets become interconnected. In addition, an even stronger result holds in the case of simple relocation of the existing mass of speculators across the markets, namely, the case when the total population of chartists (β+˜βImage) and fundamentalists (θ+˜θImage) remains unchanged, while parameters ˜βImage, ˜θImage are increased (and β, θ are decreased accordingly). In this case the stability conditions for the integrated system are definitely more restrictive than for the markets in isolation, as proven in Dieci and Westerhoff (2013b). Further numerical investigations show the robustness of such results to the introduction of asymmetries between the two stock markets.

In a related paper, Dieci and Westerhoff (2010) investigate the case in which instability originates in the foreign exchange market due to speculative currency trading, and then it propagates to the stock markets. Different from Dieci and Westerhoff (2013b), only the fundamental traders are active in the two stock markets, while the foreign exchange market is populated by the speculators who switch between two behavioral rules, based on extrapolative and regressive beliefs, depending on the exchange rate misalignment. Therefore, the general setup (27)(30) is reduced to a special case where β=˜β=γ=˜γ=0Image, whereas currency excess demand UtImage is specified as:

(31) Ut=nc,tDFXc,t+(1nc,t)DFXf,t,

Image (31)

(32) DFXc,t=κ(sts),DFXf,t=φ(sst),nc,t=[1+ν(sst)2]1,

Image (32)

where κ,φ,ν>0Image and nc,tImage is the weight of extrapolative beliefs in period t. By Eqs. (31) and (32), chartist and fundamentalist demand are then proportional to the current exchange rate deviation. That is, the chartists believe that the observed misalignment will increase further, whereas the fundamentalists believe that the exchange rate will revert to the fundamental. However, the more the exchange rate deviates from its fundamental value, the more regressive beliefs gain in popularity at the expense of extrapolative beliefs, as speculators perceive the risk that the bull or bear market might collapse. Moreover, the higher parameter ν is in (32), the more sensitive the mass of speculators becomes with regard to a given misalignment.25 Intuitively, when considered in isolation (˜θ=˜ψ=0Image), the foreign exchange market is unstable (since the extrapolative beliefs prevail and tend to increase the misalignment if stImage is sufficiently close to sImage), whereas the two stock markets converge to their fundamental prices, thanks to the stabilizing activity of fundamental traders.

Dieci and Westerhoff (2010) investigate the dynamics under market integration, which results in a 3-dimensional nonlinear dynamical system, having two additional non-fundamental steady states (NFSS), beside the FSS. Analytical conditions for the FSS to be locally stable can be derived in terms of the model parameters and compared with the stability conditions of each market, considered in isolation. Bifurcation diagrams are particularly useful to understand how the ‘strength’ of the interaction between the stock markets (captured by parameters ˜θImage and ˜ψImage) and chartist extrapolation in the foreign exchange market (parameter κ) jointly affect the stability properties. In the left panels of Fig. 10, the asymptotic behavior of the domestic (log-)stock price p (top) and (log-)exchange rate s (bottom) is plotted against extrapolation parameter κ. In the no-interaction case (illustrated by the superimposed dashed lines), the fundamental (log-)exchange rate sImage is unstable and the exchange rate misalignment in the NFSS increases with κ, whereas the fundamental (log-)prices in the stock markets, pImage and qImage, are stable. The plots show that the connection with stable stock markets can be beneficial, to some extent, by bringing the exchange rate back to its fundamental value (for κ<ˆκ0.6015Image), or by reducing such misalignments. However, if κ is large enough, the integration can destabilize the stock markets, too, and introduce cyclical and chaotic behavior in the whole system of the interacting markets, with fluctuations of increasing amplitude. In particular, for κ>κ4.856Image, the fluctuations range across a much wider area than for κ<κImage. While for κ<κImage, two different attractors coexist, implying that the asymptotic dynamics of prices and exchange rate are confined to different regions depending on the initial condition (‘bull’ or ‘bear’ markets), at κ=κImage they merge into a unique attractor (through a homoclinic bifurcation).26 The right panels of Fig. 10 represent the fluctuations of p (top) and s (bottom) for very large κ, characterized by sudden switching between bull and bear markets. The dynamic analysis thus reveals a double-edged effect of market interlinkages, where behavioral factors appear to play a substantial role.

Image
Figure 10 Bifurcation diagrams of log-price p and log-exchange rate s against extrapolation parameter κ (left panels) and their time paths under strong extrapolation (right panels). The superimposed dashed lines in the left panels depict the case of isolated markets. Parameters are: p = q = s = 0, αP = 1, αQ = 0.8, αS = 1, θ = 1, ψ = 1.5, ˜θ=˜ψ=0.4Image, φ = 0.8, ν = 10000 and (in the right panels) κ = 5.3.

The interaction of foreign and domestic investors using heterogeneous trading rules, and its effect on the dynamics of the foreign exchange market, has been the subject of further research in recent years. Amongst others, Kirman et al. (2007) show that the mere interplay of speculative traders with wealth measured in two different currencies and buying or selling assets of both countries can produce bubbles in foreign exchange market and realistic features of the exchange rate series. Corona et al. (2008) develop and investigate a computationally oriented agent-based model of two stock markets and a related foreign exchange market. They focus, in particular, on the resulting volatility, covariance and correlation of the stock markets, both during quiet periods and during a monetary crisis. Overall, such models highlight a number of dynamic features that are intrinsic to a system of asset markets linked via and with foreign exchange market and that simply arise from the structural properties of such interlinkages combined with the behavior of heterogeneous traders.

5 HAMs and House Price Dynamics

This section surveys recent research on the impact of investors' behavioral heterogeneity on the dynamics of house prices and markets. Similarly to financial market dynamics, the main body of literature on house price dynamics relies on the theoretical framework of fully rational and forward looking investors (see, e.g. Poterba, 1984, Poterba et al., 1991, Clayton, 1996, Glaeser and Gyourko, 2007, Brunnermeier and Julliard, 2008). Broadly speaking, in this framework house price movements are due to sequences of exogenous shocks affecting the fundamentals of the housing market (rents, population growth, the user cost of capital, etc.), and to the resulting ‘well-behaved’ adjustments to new long-run equilibrium levels. Real estate market efficiency is an implication of such rationality assumptions.

Despite the remarkable achievements in this literature, a number of housing market phenomena are far from being fully understood. This includes the existence of boom-bust housing cycles unrelated to changes in underlying fundamentals (Wheaton, 1999, Shiller, 2007) – as the house price bubble and crash of the 2000s. Further empirical evidence challenges real estate market efficiency, in particular the short-term positive autocorrelation and long-term mean-reversion of house price returns (Capozza and Israelsen, 2007, Case and Shiller, 1989, 1990). For this reason, research on housing market dynamics has gradually accepted the view that investors' bounded rationality (optimism and pessimism, herd behavior, adaptive expectations, etc.) may play a role in house price fluctuations, for instance Cutler et al. (1991), Wheaton (1999), Malpezzi and Wachter (2005), Shiller (2005, 2008), Glaeser et al. (2008), Piazzesi and Schneider (2009), Sommervoll et al. (2010), and Burnside et al. (2012).

Recently, a number of HAMs of housing markets have been developed and estimated, inspired by the well-established heterogeneous-agent approach to financial markets. A stylized two-belief (chartist-fundamentalist) framework has been developed to incorporate in a tractable way the behavioral heterogeneity of agents. It proves to be a useful tool to understand housing bubbles and crashes and the way they interact with the ‘real side’ of housing markets, as well as other phenomena that are at odds with the standard approach. The framework of housing models is very close to HAMs of financial markets. It is based on housing demand consistent with mean-variance optimization and on a benchmark ‘fundamental’ price linked to the expected rental earnings (Bolt et al., 2014). However, unlike other asset markets, housing markets have specific features that need to be taken into account (such as the dual nature of housing, endogenous housing supply). Such features generate important interactions between the real and financial side of housing markets, which may be amplified by the interplay of heterogeneous speculators (Dieci and Westerhoff, 2012, 2016).

5.1 An Equilibrium Framework with Heterogeneous Investors

The housing market models developed by Bolt et al. (2014) and Dieci and Westerhoff (2016) are based on a common temporary equilibrium framework for house prices. This framework generalizes standard asset pricing relationships to the case of heterogeneous expectations. Denote by PtImage the price of a housing unit at the beginning of the time interval (t,t+1)Image, Pt+1Image the end-of-period price, and Qt+1Image the (real or imputed) rent in that period. The sum Pt+1+Qt+1Image represents the one-period payoff on the investment in one housing unit. Despite the time subscript, quantity Qt+1Image is assumed to be known with certainty at time t (since rental prices are typically agreed in advance). At time t, housing market investors form expectations about price Pt+1Image by choosing among a number of available rules. Denote by Eh,t()Image and nh,tImage the subjective expectation and the market proportion of investors of type h, respectively, and Pet,t+1:=hnh,tEh,t(Pt+1)Image the average market expectation. Note that price PtImage is not known yet to investors when they form expectations about Pt+1Image. In a single-period setting, the current price is determined by the expectation as follows:

(33) Pt=Pet,t+1+Qt+11+kt+ξt,

Image (33)

where ktImage represents the so-called user cost of housing and ξtImage can be interpreted as the risk premium for buying over renting a house. In particular, the user cost ktImage includes the risk-free interest rate (or mortgage rate), denoted as rtImage, as well as other costs, such as depreciation and maintenance costs, property tax, etc. (see, e.g. Himmelberg et al., 2005). As shown in Bolt et al. (2014) and Dieci and Westerhoff (2016), Eq. (33) is consistent with the assumptions of mean-variance demand and market clearing in the housing market.

5.1.1 Heterogeneous Expectations, Fundamentals, and Temporary Bubbles

In this section we discuss the model of Bolt et al. (2014). They address the issue of house price bubbles and crashes, disconnected from the dynamics of the rent and fundamental price, in a model of the housing market with behavioral heterogeneity and evolutionary selection of beliefs. Following Boswijk et al. (2007), the rent QtImage in (33) follows an exogenous process, namely, a geometric Brownian motion with drift, Qt+1=(1+g)ϵtQtImage, where {ϵt}Image are i.i.d. log-normal, with unit conditional mean. The user cost ktImage (here reduced to the interest rate for simplicity) and the risk premium ξtImage in (33) are assumed constant kt=rt=rImage, ξt=ξImage, with r+ξ>gImage.27 In the reference case of homogeneous and correct expectations, a benchmark ‘fundamental’ solution PtImage can be obtained from Eq. (33), namely, Pt=Et[s=1Qt+s(1+r+ξ)s]=Qt+1[s=1(1+g)s1(1+r+ξ)s]=Qt+1/(r+ξg)Image.

Heterogeneity in expectations is captured by the interplay of regressive (fundamentalist) and extrapolative (chartist) beliefs (indexed by h{f,c}Image, respectively), with time-varying proportions nc,tImage and nf,t=1nc,tImage. More precisely, investors form their beliefs about the relative deviation between the price and the fundamental in the next period, Xt+1:=Image (Pt+1Pt+1)/Pt+1Image, according to the linear rules Eh,t(Xt+1)=ϕhXt1Image, h{f,c}Image, where ϕf<1Image and ϕc>1Image characterize regressive and extrapolative beliefs, respectively. As a consequence,28 asset pricing Eq. (33) takes the following recursive form in relative deviations from the fundamental price, given proportions nc,tImage and nf,tImage:

(34) Xt=(1+g)1+r+ξ(nf,tϕf+nc,tϕc)Xt1,

Image (34)

where (nf,tϕf+nc,tϕc)Xt1Image is the average market expectation of Xt+1Image. It is also clear from (34) that the direction of the price change is remarkably affected by the current belief distribution. Strategies' proportions are determined by a logistic switching model with a-synchronous updating (see, e.g. Diks and van der Weide, 2005), according to nc,t=δnc,t1+(1δ){1+exp[β(Uc,t1Uf,t1)]}1Image, where Uc,t1Image and Uf,t1Image are fitness measures for chartists and fundamentalists, based on the realized excess profits in the previous period.29 The model is described by a high-dimensional nonlinear dynamical system.

Based on earlier literature and on quarterly data on house price and rent indices from OECD databases, Bolt et al. (2014) calibrate the fundamental model parameters and obtain the price-fundamental deviations XtImage for each of eight different countries (US, UK, NL, JP, CH, ES, SE, and BE).30 In a second step, the behavioral parameters of the agent-based model are estimated based on the time series XtImage (with the fundamental parameters fixed during the estimation). Since the model is governed by a nonlinear time-varying AR(1) process, once white noise is added to Eq. (34), it can be estimated by nonlinear least squares. In particular, among the estimated behavioral parameters, ϕcImage is significantly larger than 1 (chartists expect that the bubble will continue in the near future) and the difference Δϕ:=ϕfϕcImage is significant for all countries. This confirms the destabilizing impact of extrapolators and the presence of time-varying heterogeneity in the way agents form expectations. For all countries, long-lasting temporary house price bubbles are identified, driven or amplified by extrapolation (in particular, US, UK, NL SE, and ES display strong housing bubbles over the period 2004–2007). When these bubbles burst, the correction of housing prices is reinforced by investors' switching to a mean-reverting fundamental strategy. Remarkably, for all countries, the estimated parameters are close to regimes of multiple equilibria and/or global instability of the underlying nonlinear switching model. This fact has important policy implications, as the control of certain parameters may prevent the system from getting too close to bifurcation. For instance, the (mortgage) interest rate turns out to be one of the parameters that may shift the nonlinear system closer to multiple equilibria and global instability, whenever it becomes too low. The paper also shows that the qualitative in-sample and out-of-sample predictions of the non-linear switching model differ considerably from those of standard linear benchmark models with a rational representative agent, which is also important from a policy viewpoint.

5.1.2 Heterogeneous Beliefs, Boom-Bust Cycles, and Supply Conditions

In a similar two-beliefs asset pricing framework for housing markets, Dieci and Westerhoff, 2016 investigate how expectations-driven house price fluctuations interact with supply conditions (namely, housing supply elasticity and the existing stock of housing). For this purpose, an evolving mix of extrapolative and regressive beliefs is nested into a traditional stock-flow housing market framework (DiPasquale and Wheaton, 1992, Poterba, 1984) that connects the house price to the rent level and housing stock. Although the house price is still determined by a temporary equilibrium condition formally similar to (33), the model has a number of peculiar features. First, the (constant) user cost kt=kImage now includes also the depreciation rate d, namely, k=r+dImage. Second, the rent paid in period (t,t+1)Image, Qt+1Image, is determined endogenously and, ceteris paribus, negatively related to the current stock of housing HtImage, namely, Qt+1=q(Ht)Image, with q<0Image. This is due to market clearing for rental housing, where supply of housing services is assumed to be proportional to the stock of housing while demand is a downward-sloping function of the rent. Third, the stock of housing evolves due to depreciation and new constructions, where the latter depends positively on the observed price level:

(35) Ht+1=(1d)Ht+h(Pt)h>0.

Image (35)

In each period, investment demand for housing based on standard mean-variance optimization (see, e.g. Brock and Hommes, 1998) results in the following market clearing condition31:

(36) 1α[Pt,t+1e+q(Ht)dPt(1+r)Pt]=Ht,

Image (36)

where Pt,t+1eImage is the average market expectation (across investors) and parameter α>0Image is directly related to investors' risk aversion and second moment beliefs, assumed to be constant and identical across investors. The left-hand side of (36) represents the average individual demand (desired holdings of housing stock) and is proportional to the expected excess profit on one housing unit, taking both rental earnings and depreciation into account. Note that a larger stock HtImage and/or a larger risk perception α require a larger expected excess profit in order for the market to clear, which results in a lower market clearing price, ceteris paribus. By defining the ‘risk-adjusted’ rent q˜(Ht):=q(Ht)αHtImage, one obtains the following house pricing equation32:

(37) Pt=Pt,t+1e+q˜(Ht)1+r+d.

Image (37)

Dynamical system (37) and (35) admits a unique steady state, implicitly defined by P=q˜(H)r+dImage and H=h(P)dImage, which can be regarded as the fundamental steady state (FSS), where the fundamental price PImage obeys to a standard ‘discounted dividend’ representation. Consistently, the price-rent ratio at the FSS can be expressed as the reciprocal of the user cost (including the required housing risk premium):

(38) π=PQ=1r+d+ξ,ξ:=αH/P.

Image (38)

Although the model admits the same FSS under a wide spectrum of expectations schemes, investors' beliefs may remarkably affect the nature of the dynamical system, the way it reacts to shocks, and how it behaves sufficiently far from the FSS. In the reference case of perfect foresight, with homogeneous price expectation satisfying Pt,t+1e=Image Pt+1Image, the FSS is saddle-path stable. In the presence of a ‘fundamental’ shock (e.g., an unanticipated and permanent interest rate reduction) shifting the FSS in the plane (P,H)Image, the adjustment process towards the new FSS implies an initial price overshooting followed by a monotonic decline toward the new equilibrium price PImage, whereas the stock adjusts to level HImage gradually, without overbuilding, as shown in Fig. 11. This dynamic pattern is due to the assumed full rationality of housing market investors, by which the system can jump to the new saddle path immediately after the shock. Remarkably, the qualitative pattern illustrated Fig. 11 is extremely robust to changes of the parameters (in particular, it is unaffected by the response of housing supply).

Image
Figure 11 The case of perfect foresight: ‘well-behaved’ price and stock adjustments in response to an unanticipated shock.

In contrast, by assuming backward-looking and heterogeneous expectations, the stability properties of the FSS and the nature of price and stock fluctuations depend on the way investors' beliefs coevolve with the housing market itself. The average price expectation is specified as

(39) Pt,t+1e=φ(Pt1)=nc,tφc(Pt1)+nf,tφf(Pt1),nf,t=1nc,t,

Image (39)

where φc(P)=P+γ(PP)Image and ϕf(P)=P+θ(PP)Image, γ,θ>0Image, represent the extrapolative and regressive components, respectively. Similar to (32), the market weight of extrapolative and regressive beliefs evolves endogenously, depending on market circumstances. The market proportion of extrapolators is specified as nc,t=w(Pt1)Image, where w(P)=[1+ν(PP)2]1Image, is a ‘bell-shaped’ function of the observed mispricing, governed by a (possibly state-dependent) sensitivity coefficient ν>0Image.33

The rent and the supply of new constructions are modeled as isoelastic functions, namely, q(H)=λ0HλImage, h(P)=μ0PμImage, λ0,μ0,λ,μ>0Image. Dynamical system (37) and (35) has a locally stable FSS34 only for sufficiently weak extrapolation (low parameter γ). For large enough γ, the model predicts that an initial positive deviation from the fundamental price tends to be amplified by investors' behavior. However, the stability loss generated by strong extrapolation may result in different scenarios, depending on the elasticity of housing supply, μ. Under a relatively inelastic housing supply, the extrapolation generates two additional (locally stable) non-fundamental steady states (NFSS), via a so-called pitchfork bifurcation. Such ‘bubble equilibria’ are characterized by higher (respectively lower) levels of the price-rent ratio than the fundamental price-rent ratio πImage in Eq. (38). Therefore, under a weak supply response, a positive mispricing at time t=0Image results in a long-lasting price bubble and overbuilding, in the absence of exogenous shocks (the top left panel of Fig. 12). Things are quite different under a more elastic housing supply. Although the initial price path is very similar, a prompt supply response results in a larger growth of the housing stock, which causes a price decline and, ultimately, the endogenous bursting of the bubble (the top right panel). This second scenario is associated with a stable closed orbit, generated via a Neimark–Sacker bifurcation. The larger the supply response, the larger and faster the growth of the stock, the shorter the bubble period (the bottom panels of Fig. 12).

Image
Figure 12 Impact of different degrees of supply elasticity (from top-left to bottom right: μ = 1, μ = 2.5, μ = 4, μ = 5), in the presence of strong extrapolative behavior. House price (black) and stock (grey) are expressed in relative deviations from their fundamental levels. Other parameters are: P = H = 100, r = d = ξ = 0.5%, γ = 0.15, θ = 0.125, α = 0.005, λ = 4. State-dependent switching coefficient is modeled as ν = 1/100 for P ⩾ 100, whereas ν = ν(P)=(101 − P)/100 for P < 100.

Fig. 13 illustrates a further scenario in which supply elasticity may affect bubbles in a similar manner. The top-left panel is a phase-space representation in the plane of house price and stock (in relative deviations from PImage and HImage, respectively) of the dynamics depicted in the top-left panel in Fig. 12. The underlying regime has three equilibria, two of which are visible in Fig. 13, namely, the FSS and the ‘upper’ NFSS. The light and dark gray regions represent the basins of attraction of the coexisting NFSS, whereas the (saddle) FSS lies on the boundary of the basins. The top panels and the bottom-left panel indicate that, the larger the supply elasticity, the closer the NFSS gets to the boundary of its basin. The bubble equilibrium thus becomes less and less robust to exogenous noise, although it continues to be locally stable. In particular, its basin of attraction may become very small (white area in the bottom-right panel of Fig. 13).35 From the viewpoint of nonlinear dynamics, the phenomena illustrated in Fig. 13 are global, in the sense that they are independent of the local stability properties of the coexisting steady states.

Image
Figure 13 Changes of the basin (boundary) of the bubble steady state, for increasing supply elasticity (from top-left to bottom right: μ = 1, μ = 1.2, μ = 1.4, μ = 1.57). Other parameters are as in Fig. 12.

The qualitative results produced by this model are in agreement with recent research on housing market bubbles and urban economics, reporting that a more elastic housing supply is associated to shorter bubbles, smaller price increases and larger stock adjustments (see, e.g. Glaeser et al., 2008).36 This model thus provides a ‘nonlinear economic dynamics’ interpretation on the observed role of supply elasticity in shaping housing bubbles and crashes, based on bifurcation analysis and on a simple HAM framework.

5.2 Disequilibrium Price Adjustments

Further HAMs of the housing market depart from equilibrium asset pricing equation (33) and rest on the view that prices adjust to excess demand in each period in disequilibrium. This may lead to different dynamics from that observed under market clearing. However, the phenomena reported in the previous section appear to be quite robust to such alternative specifications. In particular, Dieci and Westerhoff (2012) consider the following linear price adjustment equation

(40) Pt+1Pt=ψ(DtR+DtSHt).

Image (40)

Housing stock HtImage evolves similarly to (35), namely, Ht=(1d)Ht1+mPtImage, m>0Image. The housing demand DtR+DtS:=DtImage (interpreted as the desired stock of housing) is made up of ‘real demand’ DtRImage (from consumers of housing services) and speculative demand DtSImage (from investors motivated by short-term capital gains). The two demand components are modeled, respectively, as follows:

(41) DtR=abPt,a,b>0,

Image (41)

(42) DtS=nc,tDc,t+nf,tDf,t=nc,tγˆ(PtP)+nf,tθˆ(PPt),γˆ,θˆ>0,

Image (42)

where Dc,tImage and Df,tImage are chartist and fundamentalist demand, respectively. Again, the proportion of extrapolators nc,t=w(Pt)Image evolves according to weighting function w(P)Image introduced in Section 5.1.2. In particular, while real demand DtRImage depends linearly and negatively on the current price level, speculative demand DtSImage results in a nonlinear, cubic-like function of PtImage.37 In (42), PImage is the FSS, corresponding to the unique steady state of the baseline case without speculative demand, namely, P:=adm+bdImage, H=mdP=amm+bdImage. Using the change of variables πt:=PtPImage, ζt:=Ht1HImage, one obtains the following two-dimensional nonlinear system in deviations from the FSS:

π t + 1 = π t ψ [ ( b + m ) π t γ ˆ π t θ ˆ ν π t 3 1 + ν π t 2 + ( 1 d ) ζ t ] , ζ t + 1 = m π t + ( 1 d ) ζ t .

Image

The analytical and numerical study of the dynamical system delivers clear-cut results about the emergence of housing bubbles and crashes and the joint role played by chartist demand parameter, γˆImage, and the slopes of ‘real’ demand and supply schedules,  b and m. In particular, similar to Dieci and Westerhoff (2016), parameter γˆImage may destabilize the steady state via a pitchfork bifurcation, if the housing supply curve is sufficiently flat (low m), or via a Neimark–Sacker bifurcation, if the supply schedule is sufficiently sloped (large m). Moreover, in both scenarios, large γˆImage results in a ‘route’ to complexity and endogenous irregular bubbles and crashes. In particular, in the pitchfork scenario, two locally attracting NFSS may evolve into more complex (disjoint) attractors and, ultimately, merge into a unique attractor (through a so-called homoclinic bifurcation). The motion of the system on this attractor is characterized by irregular dynamics in the bull or bear market regions, and by sudden, seemingly unpredictable switching between the bull and bear markets (the top-left panel of Fig. 14) and slow change of the stock level (the top-right panel).38 In the Neimark–Sacker scenario, irregular bubbles of different size and duration, followed by sudden crashes, can be observed (the bottom-left panel), with larger and more frequent stock fluctuations (the bottom-right panel). This kind of motion is also due to a complex attractor, originally born as a regular closed curve via a Neimark–Sacker bifurcation.

Image
Figure 14 Irregular bubbles and crashes in the presence of strong extrapolation. Top panels: house price and stock (in deviations from the steady state) in the ‘pitchfork scenario’ (b = 0.6, m = 0.0003, γˆ=7.28Image). Bottom panels: house price and stock in the ‘Neimark–Sacker’ scenario (b = 0.05, m = 0.5, γˆ=6Image). Other parameters are d = 0.02, θˆ=1Image, ν = 1 for all panels.

Kouwenberg and Zwinkels (2015) develop and estimate a housing market model with a structure similar to Dieci and Westerhoff (2012). For estimation purposes, their model is expressed in log price, pt:=lnPtImage, and the log-fundamental pt:=lnPtImage is modeled as a time-varying reference value. The demand functions from consumers and investors are interpreted as flows (desired transactions) and so is supply (identified with the flow of new constructions). While fundamentalist demand is based on current mispricing, chartist demand is based on the extrapolation of a time average of past returns. The proportions of chartists and fundamentalists evolve endogenously based on past performances (related to past observed forecast errors), according to a standard logit switching model. The model is expressed as:

(43) ρt+1:=pt+1pt=ψ(dtht)+ϵt+1,

Image (43)

where ρt+1Image is the log-return on housing investment, ϵt+1Image is a random noise term. The demand and supply are defined as follows:

(44) dt=(abpt)+nc,tγˆl=1Lρtl+1+nf,tθˆ(ptpt),ht=c+mpt.

Image (44)

Chartist proportion is given by nc,t=[1+exp(βAt)]1Image, where At=(Πf,tΠc,t)/(Πf,t+Πc,t)Image, and Πh,t=j=1J|Eh,tj(ρtj+1)ρtj+1|Image is a sum of past absolute forecast errors of agents of type h, h{f,c}Image. Similar to Eqs. (41) and (42), the housing demand dtImage in (44) includes the consumer demand component and the speculative demand terms due to chartists and fundamentalists, respectively.39

The model can be estimated by rewriting it as single non-linear equation and applying maximum likelihood estimation. Estimation results (based on U.S. quarterly time-series data on prices and rents)40 reveal that the coefficients for the fundamentalist and chartist rules are significant and have the expected signs. The positive and significant sign of the estimated intensity of choice parameter (β) implies that agents tend to switch following recent prediction performance. Interestingly, simulation of the deterministic skeleton of the model (with the parameters set equal to the estimated values) shows that the price does not converge to a stable steady state value, but to a stable limit cycle. Hence, an endogenous nonlinear motion appears to be an important part of U.S. housing market dynamics.

A widely reported empirical fact about real estate returns is the presence of short-term positive autocorrelation and long-term mean-reversion (see, e.g. Capozza and Israelsen, 2007, Case and Shiller, 1989, 1990). This fact is, more or less explicitly, part of the motivation for the chartist-fundamentalist framework adopted in the models reviewed in this section. Kouwenberg and Zwinkels (2014) build an econometric model that includes explicitly these two competing components of housing returns. The model is based on a VECM equation, modified to allow for smoothly changing weights of the autoregressive and error correction components, conditional on the value of a transition variable that depends on past relative forecast errors (a so-called smooth transition model). In fact, the econometric model is a particular case of the behavioral model described above (Kouwenberg and Zwinkels, 2015). The analysis shows that house prices are cointegrated with a rent-based estimate of the fundamental value. Estimation results (using quasi-maximum likelihood estimation, based on quarterly US national house price index data) indicate that the strength of the autocorrelation and the long-term mean reversion in housing returns vary significantly over time, depending on recent forecasting performances. The time variation captured by the smooth transition model can produce better out-of-sample forecasts of house price index returns than alternative models.

6 HAMs and Market Microstructure

Limit order markets (LOM) are the most active and dominating financial markets (O'Hara, 2001; Easley et al., 2013; O'Hara, 2015). A core and challenging issue in dynamical LOM models is the endogenous order choice of investors to submit either market or limit orders. It is important to understand how investors trade based on their asymmetric information and what they can learn from order book information. The current literature of limit order market models faces two main challenges. First, they mainly focus on perfectly rational information-based trade and order choice of informed traders. However, within rational expectation equilibrium framework, “a model that incorporates the relevant frictions of limit-order markets (such as discrete prices, staggered trader arrivals, and asymmetric information) does not readily admit a closed-form solution (Goettler et al., 2009)”. This limits the explanatory power of this framework. Second, rational expectation framework simplifies the order choice behavior of uninformed traders by introducing either private value or time preferences exogenously. However, as pointed out by O'Hara (2001), “It is the uninformed traders who provide the liquidity to the informed, and so understanding their behavior can provide substantial insight and intuition into the trading process”. Therefore what uninformed traders can learn from order book information and how learning affects their order choice and the behavior of informed traders are not clear.

Recent development of HAMs and computationally oriented agent-based simulations provide a framework to deal with these challenges in LOM models. With great flexibility in modeling complexity and learning, this framework offers a very promising and integrated approach to the research in market microstructure. Within this framework, many features including asymmetric information, learning, and order choice can be articulated. It can provide an insight into the impact of heterogeneous trading rules on limit order book and order flows (Chiarella and Iori, 2002, Chiarella et al., 2009b, 2012b, Kovaleva and Iori, 2014), interplay of different market architectures and different types of regulatory measures, such as price limits (Yeh and Yang, 2010), transaction taxes (Pellizzari and Westerhoff, 2009), short-sales constraints (Anufriev and Tuinstra, 2013). It also sheds light on the costs and benefits of financial regulations (Lensberg et al., 2015).

This section discusses briefly the recent developments along this line, in particular the contributions of Chiarella et al. (2015a, 2017) and Arifovic et al. (2016). We first focus on how computationally oriented HAMs can be used to replicate the stylized facts in LOM and provide possible mechanism explanation to these stylized facts in Section 6.1. We then discuss how genetic algorithm (GA) learning with a classifier system can help to understand the joint impact of market information, market microstructure mechanisms, and behavioral factors on the dynamics of LOM characterized by information asymmetry and complexity in order flows and trading in Section 6.2. We also examine the impact of high frequency trading (HFT) and learning on information aggregation, market liquidity, and price discovery in Section 6.3, demonstrating that the incentive for high frequency traders not to trade too fast can be consistent with price information efficiency. We also discuss some implications on market design and regulation in Section 6.4.

6.1 Stylized Facts in Limit Order Markets

Agent-based computational finance has made a significant contribution to characterize the stylized facts in financial markets, as discussed in Section 2. As pointed out in Chen et al. (2012) and Gould et al. (2013), after several prototypes have successfully replicated a number of financial stylized facts of the low frequency data, the next milestone is to see whether HAMs can also be used to replicate the features in high frequency domain.

Various stylized facts in limit order markets have been documented in market microstructure literature. According to surveys by Chen et al. (2012) and Gould et al. (2013), apart from the stylized facts in the time series of returns, including fat tails, the absence of autocorrelation in returns, volatility clustering, and long memory in the absolute returns, the limit order market has its own stylized facts. They include long memory in the bid-ask spread and trading volume, hump shapes in order depth profiles of order books, non-linear relationship between trade imbalance and mid-price return, and diagonal effect or event clustering in order submission types, among the most common and important statistical regularities in LOM. They have become the most important criteria to justify the explanatory power of agent-based LOM.

A number of HAMs of market microstructure have been able to replicate some of the stylized facts. They include zero-intelligence models and HAMs (see Chen et al., 2012 and Gould et al., 2013 for surveys). The zero-intelligence models show that some of the stylized facts, such as fat tail and possible volatility and event clusterings, are generated by trading mechanism, instead of agents' strategic behavior. Different from the zero-intelligence models, HAMs consider agents' strategic behaviors as potential explanations to the stylized facts. Chiarella and Iori (2002) argue that substantial heterogeneity must exist between market participants in order for the highly non-trivial properties of volatility to emerge in real limit order markets. By assuming that agents use strategies that blend three components (fundamentalist, chartist, and noisy), Chiarella et al. (2009b) provide a computational HAM of an order-driven market to study order book and flow dynamics. Inspired by the theoretically oriented dynamic analysis of moving average rules in Chiarella et al. (2006b), Chiarella et al. (2012b) conduct a dynamic analysis of a more realistic microstructure model of continuous double auctions. When agents switch between either fundamentalists or chartists based on their relative performance, they show that the model is able to characterize volatility clustering, insignificant autocorrelations (ACs) of returns and significantly slow-decaying ACs of the absolute returns. The result suggests that both behavioral traits and realistic microstructure have a role in explaining several statistical properties of returns.

In a modified version of Chiarella et al. (2009b), Kovaleva and Iori (2014) investigate the interrelation between pre-trade quote transparency and stylized properties of order-driven markets populated by traders with heterogeneous beliefs. The model is able to capture negative skewness of stock returns and volatility clustering once book depth is visible to traders. Their simulation analysis reveals that full quote transparency contributes to convergence in traders' actions, while exogenously partial transparency restriction may exacerbate long-range dependencies. However, replicating most of these stylized facts in LOM simultaneously remains very challenging.

When modeling agents' expectation, behavioral sentiment plays an important role. Barberis et al. (1998) and Daniel et al. (1998) point out that certain well-known psychological biases, including conservatism, representativeness heuristic, overconfidence and biased self-attribution, not only characterize how people actually behave, but can also explain a range of empirical findings, such as underreaction and overreaction of stock prices to news, excess volatility and post-earnings announcement drift. By incorporating behavioral sentiment to a LOM model, Chiarella et al. (2017) show that the behavioral sentiment not only helps to replicate most of the stylized facts simultaneously in LOM, but also plays a unique role in explaining these stylized facts that cannot be explained by noise trading. They include fat tails in the return distribution, long memory in the trading volume, an increasing and non-linear relationship between trading imbalance and mid-price returns, as well as the diagonal effect or event clustering in order submission types.

6.2 Information and Learning in Limit Order Market

Because of information asymmetry and growing complexity in order flows and trading in LOM, the endogenous order choice based on the order book conditions is a core and challenging issue, as highlighted by Rosu (2012). How traders' learning, in particular uninformed traders, from order book information affect their order choice and limit order market becomes important. Recently, Chiarella et al. (2015a) provide a LOM model with adaptive learning through genetic algorithm (GA) with classifier system, trying to explore the joint impact of adaptive learning and information asymmetric on trading behavior, market liquidity, and price discovery.

Since introduced firstly by Holland (1975), GA and classifier system have been used in agent-based models to examine learning and evolution in Santa Fe Institute artificial stock market (SFI-ASM) (Arthur et al., 1997a, LeBaron et al., 1999) and economic models (Marimon et al., 1990, Lettau and Uhlig, 1999, Allen and Carroll, 2001). In LOM, LeBaron and Yamamoto (2008) employ GA to capture the imitation behavior among heterogeneous beliefs. Darley and Outkin (2007) use adaptive learning to evolve trading rules of market makers and apply their simulations to the Nasdaq market in 1998. The adaptive learning has been widely used in financial markets. However most HAMs with adaptive learning and trading are largely driven by the market price instead of asymmetric information, which is the focus of microstructure literature in LOM. This brings a significant difference in the dynamics of LOM.

Unlike informed traders, uninformed traders do not have the information about the current, but lagged fundamental value. By combining information processing of market conditions and order choice into GA with a classifier system, Chiarella et al. (2015a) show that behavior heterogeneity of traders is endogenously emerged from their learning and trading. This approach fills the gap between agent-based computational finance and the mainstream market microstructure since Kyle (1985). They show that, measured by the average usage of different market information, trading rules under the GA learning become stationary and hence effective in the long-run. In particular, the learning of uninformed traders improves market information efficiency, which is not necessarily the case when informed traders learn. The learning also makes uninformed traders submit less aggressive limit orders but more market orders, while it makes informed traders submit less market orders and more aggressive limit orders. In general, both informed and uninformed traders provide liquidity to market at approximately the same rate. The results provide some insight into the effect of learning on order submission behavior, market liquidity and efficiency.

6.3 High Frequency Trading

With technology advance, high frequency trading (HFT) becomes very popular. It also brings a hot debate on the benefit of and market regulation on HFT (O'Hara, 2015). In particular, do financial market participants benefit from HFT and how does HFT affect market efficiency? To examine the effect of HFT and learning in limit order markets, Arifovic et al. (2016) extend the LOM model of Chiarella et al. (2015a) and introduce fast and slow traders with GA learning. Consistent with Grossman and Stiglitz (1980), they show a trade-off between information advantage and profit opportunity for informed HFT. This trade-off leads to a hump-shaped relation between HFT profit, market efficiency, and trading speed. When informed investors trade fast, their information advantage makes HFT more profitable. However, the learning, in particular from uninformed traders, improves information aggregation and efficiency. This then reduces the information advantage of HFT and hence the profit opportunity. HFT in general improves market information efficiency and hence price discovery. However, the trade-off between the information advantage and trading speed of HFT also leads to a hump-shape relation between liquidity consumption and trading speed. HFT improves liquidity consumption and price discovery in general due to information aggregation through the learning. When HFT trade too fast, they submit more market order, which enlarges the spread and reduces market liquidity. This implies that there is an incentive for not trading too fast, which in turn improves price efficiency. The results provide an insight into the profitability of HFT and the current debates and puzzles about the impact of HFT on market liquidity and efficiency.

6.4 HAMs and Microstructure Regulation

Lensberg et al. (2015) build an agent-based framework with market microstructure and delegated portfolio management in order to forecast and compare the equilibrium effects of different regulatory measures: financial transaction tax, short-selling ban and leverage ban. The financial market is characterized by fund managers who trade stocks and bonds in an order-driven market. The process of competition and innovation among different investment styles is modeled through a genetic programming algorithm with tournament selection. However, the heterogeneous trading strategies that emerge from the evolutionary process can be classified by a relatively small number of ‘styles’ (interpreted as value trading, news trading/arbitrage and market making/liquidity supply). The model contributes to understand the pros and cons of different regulations, by providing detailed information on the equilibrium properties of portfolio holdings, order flow, liquidity, cost of capital, price discovery, short-term volatility and long-term price dynamics. By including an exogenous business cycle process, the model also allows to quantify the effects of different regulations during periods of market distress. In particular, it turns out that a financial transaction tax may have a negative impact on liquidity and price discovery, and limited effect on long swings in asset prices.

7 Conclusion and Future Research

This chapter has discussed the latest development of heterogeneous agent models (HAMs) in finance over the last ten years since the publications of the Handbook of Computational Economics in 2006 and, in particular, the Handbook of Financial Markets: Dynamics and Evolution in 2009. It demonstrates a significant contribution of HAMs to finance theory and practice from five broad aspects of financial markets. First of all, inspired by the rich and promising perspectives of the earlier HAMs, we have witnessed a growing supporting evidence on the explanatory power of HAMs to various market anomalies and, in particular, the stylized facts through calibrations and estimations of HAMs to real data in various financial markets over the last decade. More importantly, different from traditional empirical finance and financial econometrics, HAMs provide some insights into economic mechanisms and driving forces of these stylized facts. They therefore lead to some helpful implications in policy and market design. Moreover, the basic framework of earlier HAMs has been naturally developed and extended in two different directions. The first extension to a continuous-time setup provides a unified framework to deal with the effect of historical price information. The framework can be used to examine profitability of fundamental and non-fundamental, such as momentum and contrarian, trading strategies that have been widely used and discussed in financial market practice and finance theory. It also enables to develop optimal asset allocation to incorporate time series momentum and reversal, two of the most important anomalies in financial markets. The second extension to a multiple-risky-asset framework helps to examine the impact of heterogeneous expectations on asset comovements within a financial market, as well as the spill-over effects across markets, and to characterize risk-return relations through an evolutionary CAPM. Moreover, inspired by HAMs of financial markets, a new heterogeneous-agent framework for housing market dynamics has been developed recently. It can well explain house bubbles and crashes, by combining behavioral facts and the real side of housing markets. Finally, the advantage of HAMs in dealing with market complexity plays a unique role in the development of market microstructure modeling. This provides a very promising approach to understand the impact of information, learning, and trading on trading behavior, market liquidity, and price discovery.

The research streams reviewed in this chapter can be developed further in several directions. First, instead of heuristic assumptions on agents' behavioral heterogeneity currently assumed in HAMs, there is a need to provide micro-foundation to endogenize such heuristic heterogeneity among agents. Most of the HAMs investigate the endogenous market mechanism by focusing on the interaction of heterogeneous agents with different expectations (typically fundamentalists and trend followers). Their explanatory power is mainly demonstrated by combining the insights from the nonlinear dynamics of the underlying deterministic model with various noise processes, such as fundamental shocks and noise trading. To a large extent, the HAM literature has not explored the impact of asymmetric information or information uncertainty on agents' behavioral heterogeneity. By considering asymmetric information, which is the focus of traditional finance literature and plays a very important role in financial markets, agents' heterogeneity can be endogenized and micro-founded. This has been illustrated in Section 2.3, based on He and Zheng (2016), by showing how trading heterogeneity can arise endogenously among traders due to uncertainty about the fundamental value information of the risky asset. The development along this line would help to provide economic foundation to the assumed behavioral heterogeneity of agent-based models, which is often critiqued by traditional finance.

Second, as a different aspect of information uncertainty, ambiguity has been introduced in the literature to address various market anomalies and asset pricing (Epstein and Schneider, 2006). More recently, Aliyev and He (2016) discuss the possibilities of capturing efficient market hypothesis and behavioral finance under a general framework based on a broad definition of rationality. They argue that the root of behavioral anomalies comes from the imprecision and reliability of information. A natural combination of heterogeneity and ambiguity would provide a broader framework to financial market modeling and to rationalize market anomalies.

Third, when asset prices are affected by historical price information, we need to develop a portfolio and asset pricing theory in continuous-time to characterize cross-section returns driven by time series momentum in short-run and reversal in long-run. The continuous-time HAMs discussed in Section 3 illustrate the challenging but promising perspectives of this development. Recently, Li and Liu (2016) study the optimal momentum trading strategy when asset prices are affected by historical price information. They provide an optimal way to hedge the momentum crash risk, a newly-found empirical feature, and to significantly improve momentum profits. The techniques developed can potentially be applied to a range of problems in economics and finance, such as momentum, long memory in volatility, post-earnings announcement drift, indexation lags in the inflation-linked bonds.

Fourth, incorporating social interactions and social networks to the current HAMs would be helpful for examining their impact on financial markets and asset pricing. Social interactions are well documented in financial markets, in particular when facing information uncertainty. He et al. (2016a) recently develop a simple evolutionary model of asset pricing and population dynamics to incorporate social interactions among investors with heterogeneous beliefs on information uncertainty. They show that social interactions can lead to mis-pricing and existence of multiple steady state equilibria, generating two different volatility regimes, bi-modal distribution in population dynamics, and stochastic volatility. As pointed out by Hirshleifer (2015), [T]he time has come to move beyond behavioral finance to ‘social finance’. This would provide a fruitful area of research in the near future.

Fifth, HAMS of multiasset markets and financial market interlinkages could be developed further. An interesting research issue is understanding the effect of an increase in the number of risky assets in a setup similar to Chiarella et al. (2013b) and the extent to which standard results on the role of diversification continue to hold in the presence of momentum trading. A related issue concerns the profitability of different trading strategies in a multi-asset framework, their ability to exploit the emerging correlation patterns, and their joint impact on financial market stability. Furthermore, the ability of the evolutionary, heterogeneous-agent CAPM discussed in Section 4.2 to produce a time variation of ex-ante betas has been illustrated through simulation results only. There is a need to have formal statistical tests on the observations based on the numerical simulations. Finally, it would be interesting to see if the time variation of either beta coefficients or risk premia plays a role in explaining the cross section of asset returns.

Sixth, housing market dynamics has only very recently been investigated from the perspective of HAMs. The existing models are mainly aimed at qualitative or quantitative investigations of the role of price extrapolation in generating house price fluctuations. Among the possible interesting developments of this baseline setup is the joint impact of interest rate changes, credit conditions, and investor sentiment on house price fluctuations. In particular, the role of interest rates and credit in triggering house price booms and busts is crucial for policy makers and highly debated in academic literature (see, e.g. Himmelberg et al., 2005 and Jordà et al., 2015). A related issue concerns the dynamic interplay among housing, stock and bond markets, driven by both fundamental shocks, such as interest rate movements, and behavioral factors, such as investors switching to better investment opportunities. Dieci et al. (2017) provide a first attempt in this direction.

Finally, an integrated approach of agent-based models and market microstructure literature would provide a very promising approach, if not the only one, to understand information aggregation, learning, trading, market liquidity and efficiency when facing information asymmetry and growing complexity in market microstructure. This has been illustrated by the discussion in Section 6, but remains largely unexplored.

References

S. Alfarano, T. Lux, F. Wagner, Estimation of agent-based models: the case of an asymmetric herding model, Computational Economics 2005;26:19–49.

N. Aliyev, X. He, Toward a General Model of Financial Markets. [Technical report] Sydney: Quantitative Finance Research Centre, University of Technology; 2016.

H. Allen, M. Taylor, Charts, noise and fundamentals in the London foreign exchange market, Economic Journal 1990;100:49–59.

T. Allen, C. Carroll, Individual learning about consumption, Macroeconomic Dynamics 2001;5:255–271.

G.-M. Angeletos, I. Werning, Crises and prices: information aggregation, multiplicity, and volatility, The American Economic Review 2006;96:1720–1736.

M. Anufriev, G. Bottazzi, M. Marsili, P. Pin, Excess covariance and dynamic instability in a multi-asset model, Journal of Economic Dynamics and Control 2012;36:1142–1161.

M. Anufriev, P. Dindo, Wealth-driven selection in a financial market with heterogeneous agents, Journal of Economic Behavior and Organization 2010;73:327–358.

M. Anufriev, C. Hommes, Evolutionary selection of individual expectations and aggregate outcomes, American Economic Journal: Microeconomics 2012;4:35–64.

M. Anufriev, J. Tuinstra, The impact of short-selling constraints on financial market stability in heterogeneous agents model, Journal of Economic Dynamics and Control 2013;37(8):1523–1543.

J. Arifovic, C. Chiarella, X. He, L. Wei, High Frequency Trading and Learning in a Dynamic Limit Order Market. [Working Paper] Sydney: University of Technology; 2016.

W. Arthur, Inductive reasoning and bounded rationality, American Economic Review 1994;84(2):406–411.

W. Arthur, J. Holland, B. LeBaron, R. Palmer, P. Tayler, Asset pricing under endogeneous expectations in an artificial stock market, Economic Notes 1997;26(2):297–330.

W. Arthur, J. Holland, B. LeBaron, R. Palmer, P. Tayler, Asset pricing under endogenous expectations in an artificial stock market, W.B. Arthur, S.N. Durlauf, D.A. Lane, eds. The Economy as an Evolving Complex System II. Reading, MA: Addison–Wesley; 1997:15–44.

C. Aymanns, J. Farmer, The dynamics of the leverage cycle, Journal of Economic Dynamics and Control 2015;50:155–179.

S. Baak, Test for bounded rationality with a linear dynamics model distorted by heterogeneous expectations, Journal of Economic Dynamics and Control 1999;23:1517–1543.

T. Bao, C. Hommes, When Speculators Meet Constructors: Positive and Negative Feedback in Experimental Housing Markets. [Working Paper 15-10] CeNDEF, University of Amsterdam; 2015.

N. Barberis, A. Shleifer, R. Vishny, A model of investor sentiment, Journal of Financial Economics 1998;49:307–343.

C. Bauer, P. de Grauwe, S. Reitz, Exchange rate dynamics in a target zone—a heterogeneous expectations approach, Journal of Economic Dynamics and Control 2009;33(2):329–344.

A. Beja, M. Goldman, On the dynamic behavior of prices in disequilibrium, Journal of Finance 1980;35:235–247.

L. Blume, D. Easley, Evolution and market behavior, Journal of Economic Theory 1992;58:9–40.

L. Blume, D. Easley, If you're so smart, why aren't you rich? Belief selection in complete and incomplete markets, Econometrica 2006;74:929–966.

W. Bolt, M. Demertzis, C. Diks, C. Hommes, M. van der Leij, Identifying Booms and Busts in House Prices Under Heterogeneous Expectations. [Technical Report 14-13, CeNDEF Working Paper] University of Amsterdam; 2014.

H. Boswijk, C. Hommes, S. Manzan, Behavioral heterogeneity in stock prices, Journal of Economic Dynamics and Control 2007;31:1938–1970.

G. Bottazzi, P. Dindo, Evolution and market behavior with endogeneous investment rules, Journal of Economic Dynamics and Control 2014;48:121–146.

W. Brock, C. Hommes, A rational route to randomness, Econometrica 1997;65:1059–1095.

W. Brock, C. Hommes, Heterogeneous beliefs and routes to chaos in a simple asset pricing model, Journal of Economic Dynamics and Control 1998;22:1235–1274.

W. Brock, C. Hommes, F. Wagener, More hedging instruments may destabilize markets, Journal of Economic Dynamics and Control 2009;33:1912–1928.

M. Brunnermeier, C. Julliard, Money illusion and housing frenzies, Review of Financial Studies 2008;21(1):135–180.

C. Burnside, M. Eichenbaum, S. Rebelo, Understanding Booms and Busts in Housing Markets. [CQER Working Paper 12-02] Federal Reserve Bank of Atlanta; 2012.

F. Caccioli, J. Farmer, N. Foti, D. Rockmore, Overlapping portfolios, contagion, and financial stability, Journal of Economic Dynamics and Control 2015;51:50–63.

J. Campbell, L. Viceira, Consumption and portfolio decisions when expected returns are time varying, The Quarterly Journal of Economics 1999;114:433–495.

D. Capozza, R. Israelsen, Predictability in equilibrium: the price dynamics of real estate investment trusts, Real Estate Economics 2007;35(4):541–567.

K. Case, R. Shiller, The efficiency of the market for single-family homes, The American Economic Review 1989;79(1):125–137.

K. Case, R. Shiller, Forecasting prices and excess returns in the housing market, Real Estate Economics 1990;18(3):253–273.

J. Chavas, On the information and market dynamics: the case of the U.S. beef market, Journal of Economic Dynamics and Control 2000;24:833–853.

S.-H. Chen, C. Chang, Y.R. Du, Agent-based economic models and econometrics, Knowledge Engineering Review 2012;27(2):187–219.

Y.-W. Cheung, M. Chinn, I. Marsh, How do UK-based foreign exchange dealers think their market operates? International Journal of Finance and Economics 2004;9:289–306.

C. Chiarella, The dynamics of speculative behaviour, Annals of Operations Research 1992;37:101–123.

C. Chiarella, R. Dieci, L. Gardini, Speculative behaviour and complex asset price dynamics, Journal of Economic Behavior and Organization 2002;49:173–197.

C. Chiarella, R. Dieci, L. Gardini, Asset price and wealth dynamics in a financial market with Heterogeneous agents, Journal of Economic Dynamics and Control 2006;30:1775–1786.

C. Chiarella, R. Dieci, X. He, Heterogeneous expectations and speculative behaviour in a dynamic multi-asset framework, Journal of Economic Behavior and Organization 2007;62:402–427.

C. Chiarella, R. Dieci, X. He, Heterogeneity, market mechanisms and asset price dynamics, T. Hens, K.R. Schenk-Hoppe, eds. Handbook of Financial Markets: Dynamics and Evolution. Elsevier; 2009:277–344.

C. Chiarella, R. Dieci, X. He, Do heterogeneous beliefs diversify market risk? European Journal of Finance 2011;17(3):241–258.

C. Chiarella, R. Dieci, X. He, Time-varying beta: a boundedly rational equilibrium approach, Journal of Evolutionary Economics 2013:609–639.

C. Chiarella, R. Dieci, X. He, K. Li, An evolutionary CAPM under heterogeneous beliefs, Annals of Finance 2013;9:189–215.

C. Chiarella, X. He, Asset price and wealth dynamics under heterogeneous expectations, Quantitative Finance 2001;1:509–526.

C. Chiarella, X. He, Heterogeneous beliefs, risk and learning in a simple asset pricing model, Computational Economics 2002;19:95–132.

C. Chiarella, X. He, Heterogeneous beliefs, risk and learning in a simple asset pricing model with a market maker, Macroeconomic Dynamics 2003;7:503–536.

C. Chiarella, X. He, C. Hommes, A dynamic analysis of moving average rules, Journal of Economic Dynamics and Control 2006;30:1729–1753.

C. Chiarella, X. He, W. Huang, H. Zheng, Estimating behavioural heterogeneity under regime switching, Journal of Economic Behavior and Organization 2012;83(3):446–460 10.1016/j.jebo.2012.02.014.

C. Chiarella, X. He, P. Pellizzari, A dynamic analysis of the microstructure of moving average rules in a double auction market, Macroeconomic Dynamics 2012;16:556–575.

C. Chiarella, X. He, L. Shi, L. Wei, A behavioral model of investor sentiment in limit order markets, Quantitative Finance 2017;17:71–86.

C. Chiarella, X. He, L. Wei, Learning, information processing and order submission in limit order markets, Journal of Economic Dynamics and Control 2015;61:245–268.

C. Chiarella, X. He, R. Zwinkels, Heterogeneous expectations in asset pricing: empirical evidence from the s&p500, Journal of Economic Behavior and Organization 2014;105:1–16.

C. Chiarella, G. Iori, A simulation analysis of the microstructure of double auction markets, Quantitative Finance 2002;2(5):346–353.

C. Chiarella, G. Iori, J. Perello, The impact of heterogeneous trading rules on the limit order book and order flows, Journal of Economic Dynamics and Control 2009;33(3):525–537.

C. Chiarella, S. ter Ellen, X. He, E. Wu, Fear or fundamentals? Heterogeneous beliefs in the European sovereign CDS market, Journal of Empirical Finance 2015;32:19–34.

J. Clayton, Rational expectations, market fundamentals and housing price volatility, Real Estate Economics 1996;24(4):441–470.

J. Conlisk, Why bounded rationality? Journal of Economic Literature 1996;34:669–700.

E. Corona, S. Ecca, M. Marchesi, A. Setzu, The interplay between two stock markets and a related foreign exchange market: a simulation approach, Computational Economics 2008;32:99–119.

F. Corsi, S. Marmi, F. Lillo, When micro prudence increases macro risk: the destabilizing effects of financial innovation, leverage, and diversification, Operations Research 2016;64(5):1073–1088.

D. Cutler, J. Poterba, L. Summers, Speculative dynamics, Review of Economic Studies 1991;58:529–546.

K. Daniel, D. Hirshleifer, A. Subrahmanyam, A theory of overconfidence, self-attribution, and security market under- and over-reactions, Journal of Finance 1998;53:1839–1885.

V. Darley, A.V. Outkin, A NASDAQ Market Simulation, Vol. 1 of Complex Systems and Interdisciplinary Science. World Scientific; 2007.

R. Day, W. Huang, Bulls, bears and market sheep, Journal of Economic Behavior and Organization 1990;14:299–329.

P. de Grauwe, H. Dewachter, M. Embrechts, Exchange Rate Theory: Chaotic Models of Foreign Exchange Markets. Blackwell; 1993.

E. de Jong, W. Verschoor, R. Zwinkels, Heterogeneity of agents and exchange rate dynamics: evidence from the EMS, Journal of International Money and Finance 2010;29(8):1652–1669.

W. DeBondt, R. Thaler, Does the stock market overreact? Journal of Finance 1985;40:793–808.

R. Dieci, I. Foroni, L. Gardini, X. He, Market mood, adaptive beliefs and asset price dynamics, Chaos, Solitons and Fractals 2006;29:520–534.

R. Dieci, N. Schmitt, F. Westerhoff, Interactions Between Stock, Bond and Housing Markets. [Working Paper] University of Bamberg; 2017.

R. Dieci, F. Westerhoff, Heterogeneous speculators, endogenous fluctuations and interacting markets: a model of stock prices and exchange rates, Journal of Economic Dynamics and Control 2010;34:743–764.

R. Dieci, F. Westerhoff, A simple model of a speculative housing market, Journal of Evolutionary Economics 2012;22(2):303–329.

R. Dieci, F. Westerhoff, Modeling house price dynamics with heterogeneous speculators, G.I. Bischi, C. Chiarella, I. Sushko, eds. Global Analysis of Dynamic Models in Economics and Finance. Berlin/Heidelberg: Springer; 2013:35–61.

R. Dieci, F. Westerhoff, On the inherent instability of international financial markets: natural nonlinear interactions between stock and foreign exchange markets, Applied Mathematics and Computation 2013;221:306–328.

R. Dieci, F. Westerhoff, Heterogeneous expectations, boom-bust housing cycles, and supply conditions: a nonlinear economic dynamics approach, Journal of Economic Dynamics and Control 2016;71:21–44.

C. Diks, R. van der Weide, Herding, a-synchronous updating and heterogeneity in memory in a CBS, Journal of Economic Dynamics and Control 2005;29:741–763.

C. Diks, J. Wang, Can a stochastic cusp catastrophe model explain housing market crashes? Journal of Economic Dynamics and Control 2016;69:68–88.

D. DiPasquale, W. Wheaton, The markets for real estate assets and space: a conceptual framework, Real Estate Economics 1992;20(2):181–198.

D. Easley, M.L. de Prado, M. O'Hara, High Frequency Trading—New Realities for Traders, Markets and Regulators. Risk Books; 2013 Chapter The Volume Clock: Insights into the High-Frequency Paradigm.

P. Eichholtz, R. Huisman, R. Zwinkels, Fundamentals or trends? A long-term perspective on house prices, Applied Economics 2015;47(10):1050–1059.

L. Epstein, M. Schneider, Ambiguity, information quality, and asset pricing, Journal of Finance 2006;63:197–228.

I. Evstigneev, T. Hens, K. Schenk-Hoppé, Evolutionary finance, T. Hens, K.R. Schenk-Hoppé, eds. Handbook of Financial Markets: Dynamics and Evolution. Elsevier; 2009:509–564.

E. Fama, K. French, Multifactor explanations of asset pricing anomalies, Journal of Finance 1996;51:55–84.

J. Farmer, S. Joshi, The price dynamics of common trading strategies, Journal of Economic Behavior and Organization 2002;49:149–171.

T. Feldman, Portfolio manager behavior and global financial crises, Journal of Economic Behavior and Organization 2010;75:192–202.

R. Franke, Applying the method of simulated moments to estimate a small agent-based asset pricing model, Journal of Empirical Finance 2009;16:804–815.

R. Franke, F. Westerhoff, Estimation of a structural stochastic volatility model of asset pricing, Computational Economics 2011;38:53–83.

R. Franke, F. Westerhoff, Structural stochastic volatility in asset pricing dynamics: estimation and model contest, Journal of Economic Dynamics and Control 2012;36:1193–1211.

R. Franke, F. Westerhoff, Why a simple herding model may generate the stylized facts of daily returns: explanation and estimation, Journal of Economic Interaction and Coordination 2016;11:1–34.

B. Frijns, T. Lehnert, R. Zwinkels, Behavioral heterogeneity in the option market, Journal of Economic Dynamics and Control 2010;34:2273–2287.

A. Gaunersdorfer, C. Hommes, A nonlinear structural model for volatility clustering, G. Teyssiere, A. Kirman, eds. Long Memory in Economics. Berlin/Heidelberg: Springer; 2007:265–288.

A. Gaunersdorfer, C. Hommes, F. Wagener, Bifurcation routes to volatility clustering under evolutionary learning, Journal of Economic Behavior and Organization 2008;67:27–47.

M. Gilli, P. Winker, A global optimization heuristic for estimating agent-based model, Computational Statistics and Data Analysis 2003;42:299–312.

E. Glaeser, J. Gyourko, Housing Dynamics. [Discussion Paper 2137] HIER; 2007.

E. Glaeser, J. Gyourko, A. Saiz, Housing supply and housing bubbles, Journal of Urban Economics 2008;64(2):198–217.

R.L. Goettler, C.A. Parlour, U. Rajan, Informed traders and limit order markets, Journal of Financial Economics 2009;93:67–87.

F. Gomes, A. Michaelides, Asset pricing with limited risk sharing and heterogeneous agents, Review of Financial Economics 2008;21:415–448.

R. Goodwin, The nonlinear accelerator and the persistence of business cycles, Econometrica 1951;19:1–17.

M. Gould, M. Porter, S. Williams, M. McDonald, D. Fenn, S. Howison, Limit order books, Quantitative Finance 2013;13:1709–1742.

J. Griffin, X. Ji, J. Martin, Momentum investing and business cycle risk: evidence from pole to pole, Journal of Finance 2003;58:2515–2547.

S. Grossman, J. Stiglitz, On the impossibility of informationally efficient markets, American Economic Review 1980;70:393–408.

X. He, Recent developments in asset pricing with heterogeneous beliefs and adaptive behavior of financial markets, G.I. Bischi, C. Chiarella, I. Sushko, eds. Global Analysis of Dynamical Models in Economics and Finance. Springer; 2014:3–34.

X. He, K. Li, Heterogeneous beliefs and adaptive behaviour in a continuous-time asset price model, Journal of Economic Dynamics and Control 2012;36:973–987.

X. He, K. Li, Profitability of time series momentum, Journal of Banking and Finance 2015;53:140–157.

X. He, K. Li, Y. Li, Asset allocation with time series momentum and reversal, Journal of Economic Dynamics & Control 2018 Forthcoming.

X. He, K. Li, L. Shi, Social Interaction, Stochastic Volatility, and Momentum. [Technical Report] Sydney: Quantitative Finance Research Centre, University of Technology; 2016.

X. He, K. Li, C. Wang, Volatility clustering: a nonlinear theoretical approach, Journal of Economic Behavior and Organization 2016;130:274–297.

X. He, Y. Li, Power law behaviour, heterogeneity, and trend chasing, Journal of Economic Dynamics and Control 2007;31:3396–3426.

X. He, Y. Li, Heterogeneity, convergence and autocorrelations, Quantitative Finance 2008;8:58–79.

X. He, Y. Li, Testing of a market fraction model and power-law behaviour in the DAX 30, Journal of Empirical Finance 2015;30:1–17.

X. He, Y. Li, The adaptiveness in stock markets: testing the stylized facts in the DAX 30, Journal of Evolutionary Economics 2017;27(5):1071–1094.

X. He, H. Zheng, Trading heterogeneity under information uncertainty, Journal of Economic Behavior and Organization 2016;130:64–80.

J. Heckman, Micro data, heterogeneity, and evaluation of public policy: Nobel lecture, Journal of Political Economy 2001;109(4):673–748.

P. Heemeijer, C. Hommes, J. Sonnemans, J. Tuinstra, Price stability and volatility in markets with positive and negative expectations feedback: an experimental investigation, Journal of Economic Dynamics and Control 2009;33:1052–1072.

T. Hens, K. Schenk-Hoppé, Evolutionary stability of portfolio rules, Journal of Mathematical Economics 2005;41:43–66.

C. Himmelberg, C. Mayer, T. Sinai, Assessing high house prices: bubbles, fundamentals and misperceptions, Journal of Economic Perspectives 2005;19(4):67–92.

D. Hirshleifer, Behavioral finance, Annual Review of Financial Economics 2015;7:133–159.

J. Hofbauer, K. Sigmund, Evolutionary Games and Population Dynamics. Cambridge University Press; 1998.

J. Holland, Adaptation in Natural and Artificial Systems: An Introductory Analysis with Applications to Biology, Control, and Artificial Intelligence. University of Michigan Press; 1975.

C. Hommes, Financial markets as nonlinear adaptive evolutionary systems, Quantitative Finance 2001;1:149–167.

C. Hommes, Heterogeneous agent models in economics and finance, L. Tesfatsion, K.L. Judd, eds. Handbook of Computational Economics, Vol. 2, Agent-based Computational Economics. North-Holland; 2006:1109–1186.

C. Hommes, Behavioral Rationality and Heterogeneous Expectations in Complex Economic Systems. New York: Cambridge University Press; 2013.

C. Hommes, D. in't Veld, Booms, Busts and Behavioural Heterogeneity in Stock Prices. [Technical Report 15-088/II] Amsterdam, The Netherlands: Tinbergen Institute Discussion Paper; 2015.

C. Hommes, J. Sonnemans, J. Tuinstra, H.V.D. Velden, Coordination of expectations in asset pricing experiments, Review of Financial Studies 2005;18:955–980.

C. Hommes, F. Wagener, Complex evolutionary systems in behavioral finance, (Chapter 4) T. Hens, K. Schenk-Hoppe, eds. Handbook of Financial Markets: Dynamics and Evolution. Handbooks in Finance. North-Holland; 2009:217–276.

H. Hong, J. Stein, A unified theory of underreaction, momentum trading, and overreaction in asset markets, Journal of Finance 1999;54:2143–2184.

H. Hong, W. Torous, R. Valkanov, Do industries lead stock markets? Journal of Financial Economics 2007;83:367–396.

K. Hou, T. Moskowitz, Market frictions, price delay, and the cross-section of expected returns, Review of Financial Studies 2005;18:981–1020.

W. Huang, Z. Chen, Modeling regional linkage of financial markets, Journal of Economic Behavior and Organization 2014;99:18–31.

R. Jagannathan, Z. Wang, The conditional CAPM and cross-section of expected returns, Journal of Finance 1996;51:3–53.

N. Jegadeesh, S. Titman, Returns to buying winners and selling losers: implications for stock market efficiency, Journal of Finance 1993;48:65–91.

N. Jegadeesh, S. Titman, Profitability of momentum strategies: an evaluation of alternative explanations, Journal of Finance 2001;56:699–720.

O. Jordà, M. Schularick, A. Taylor, Betting the house, Journal of International Economics 2015;96:S2–S18.

M. Kalecki, A macroeconomic theory of the business cycle, Econometrica 1935;3:327–344.

A. Kirman, Whom or what does the representative agent represent? Journal of Economic Perspectives 1992;6:117–136.

A. Kirman, The economic crisis is a crisis for economic theory, CESifo Economic Studies 2010;56(4):498–535.

A. Kirman, R. Ricciotti, R. Topol, Bubbles in foreign exchange markets: it takes two to tango, Macroeconomic Dynamics 2007;11:102–123.

R. Kouwenberg, R. Zwinkels, Forecasting the US housing market, International Journal of Forecasting 2014;30:415–425.

R. Kouwenberg, R. Zwinkels, Endogenous price bubbles in a multi-agent system of the housing market, PLoS ONE 2015;10(6).

P. Kovaleva, G. Iori, Heterogenous beliefs and quote transparency in an order-driven market, R. Dieci, X. He, C. Hommes, eds. Nonlinear Economic Dynamics and Financial Modelling: Essays in Honour of Carl Chiarella. Springer; 2014:163–181.

A. Kyle, Continuous auctions and insider trading, Econometrica 1985;53:1315–1335.

J. Lakonishok, A. Shleifer, R. Vishny, Contrarian investment, extrapolation and risk, Journal of Finance 1994;49:1541–1578.

A. Larson, The hog cycle as harmonic motion, Journal of Farm Economics 1964;46:375–386.

B. LeBaron, Agent-based computational finance, L. Tesfatsion, K.L. Judd, eds. Handbook of Computational Economics, Vol. 2, Agent-based Computational Economics. North-Holland; 2006:1187–1233.

B. LeBaron, Active and passive learning in agent-based financial markets, Eastern Economic Journal 2011;37:35–43.

B. LeBaron, Heterogeneous gain learning and the dynamics of asset prices, Journal of Economic Behavior & Organization 2012;83:424–445.

B. LeBaron, W. Arthur, R. Palmer, Time series properties of an artificial stock market, Journal of Economic Dynamics and Control 1999;23:1487–1516.

B. LeBaron, L. Tesfatsion, Modeling macroeconomies as open-ended dynamic systems of interacting agents, American Economic Review 2008;98(2):246–250.

B. LeBaron, R. Yamamoto, The impact of imitation on long-memory in an order-driven market, Eastern Economic Journal 2008;34:504–517.

T. Lensberg, K. Schenk-Hoppé, D. Ladley, Costs and benefits of financial regulation: short-selling bans and transaction taxes, Journal of Banking and Finance 2015;51:103–118.

M. Lettau, H. Uhlig, Rules of thumb versus dynamic programming, American Economic Review 1999;89:148–174.

M. Levy, H. Levy, S. Solomon, A microscopic model of the stock market, Economics Letters 1994;45:103–111.

K. Li, J. Liu, Optimal Dynamic Momentum Strategies. 2016. SSRN Working Paper http://ssrn.com/abstract=2746561.

T. Lux, Herd behaviour, bubbles and crashes, Economic Journal 1995;105:881–896.

T. Lux, Rational forecasts or social opinion dynamics? Identification of interaction effects in a business climate survey, Journal of Economic Behavior and Organization 2009;32:638–655.

T. Lux, Stochastic behavioural asset pricing and stylized facts, T. Hens, K.R. Schenk-Hoppe, eds. Handbook of Financial Markets: Dynamics and Evolution. Elsevier; 2009:161–215.

T. Lux, Estimation of an agent-based model of investor sentiment formation in financial markets, Journal of Economic Behavior and Organization 2012;36:1284–1302.

T. Lux, S. Alfarano, Financial power laws: empirical evidence, models, and mechanisms, Chaos, Solitons and Fractals 2016;88:3–18.

M. Mackey, Commodity price fluctuations: price dependent delays and nonlinearities as explanatory factors, Journal of Economic Theory 1989;48:495–509.

S. Malpezzi, S. Wachter, The role of speculation in real estate cycles, Journal of Real Estate Literature 2005;13(2):141–164.

R. Marimon, E. McGrattan, T. Sargent, Money as a medium of exchange in an economy with artificially intelligent agents, Journal of Economic Dynamics and Control 1990;14:329–373.

A. Matsumoto, F. Szidarovszky, Delay differential neoclassical growth model, Journal of Economic Behavior and Organization 2011;78:272–289.

L. Menkhoff, The noise trading approach – questionnaire evidence from foreign exchange, Journal of International Money and Finance 1998;17:547–564.

L. Menkhoff, The use of technical analysis by fund managers: international evidence, Journal of Banking and Finance 2010;34:2573–2586.

L. Menkhoff, P. Taylor, The obstinate passion of foreign exchange professionals: technical analysis, Journal of Economic Literature 2007;45:936–972.

R. Merton, Optimum consumption and portfolio rules in a continuous time model, Journal of Economic Theory 1971;3:373–413.

T. Moskowitz, Y.H. Ooi, L.H. Pedersen, Time series momentum, Journal of Financial Economics 2012;104:228–250.

M. O'Hara, Overview: market structure issues in market liquidity, Market Liquidity: Proceedings of a Workshop Held at the BIS. 2001 Bank for International Settlements (BIS) Papers.

M. O'Hara, High frequency market microstructure, Journal of Financial Economics 2015;116:257–270.

A. Pagan, The econometrics of financial markets, Journal of Empirical Finance 1996;3:15–102.

J. Palczewski, K. Schenk-Hoppé, T. Wang, Itchy feet vs cool heads: flow of funds in an agent-based financial market, Journal of Economic Dynamics and Control 2016;63:53–68.

P. Pellizzari, F. Westerhoff, Some effects of transaction taxes under different microstructures, Journal of Economic Behavior and Organization 2009;72(3):850–863.

A. Phillips, Stabilization policy and time forms of lagged responses in a closed economy, Economic Journal 1957;67:265–277.

M. Piazzesi, M. Schneider, Momentum traders in the housing market: survey evidence and a search model, American Economic Review 2009;99(2):406–411.

S. Poledna, S. Thurner, J. Farmer, J. Geanakoplos, Leverage-induced systemic risk under Basle II and other credit risk policies, Journal of Banking and Finance 2014;42:199–212.

J. Poterba, Tax subsidies to owner-occupied housing: an asset-market approach, The Quarterly Journal of Economics 1984;99(4):729–752.

J. Poterba, D. Weil, R. Shiller, House price dynamics: the role of tax policy and demography, Brookings Papers on Economic Activity 1991;1991(2):143–203.

I. Rosu, Order choice and information in limit order markets, Market Microstructure: Confronting Many Viewpoints. Wiley; 2012:41–60.

A. Rubinstein, Modeling Bounded Rationality. Cambridge, MA: MIT Press; 1998.

A. Sandroni, Do markets favor agents able to make accurate predictions? Econometrica 2000;68:1303–1342.

T. Sargent, Bounded Rationality in Macroeconomics. Oxford: Clarendon Press; 1993.

N. Schmitt, F. Westerhoff, Speculative behavior and the dynamics of interacting stock markets, Journal of Economic Dynamics and Control 2014;45:262–288.

H. Shefrin, A Behavioral Approach to Asset Pricing. London: Academic Press Inc.; 2005.

R. Shiller, Irrational Exuberance. 2nd edn Princeton: Princeton University Press; 2005.

R. Shiller, Understanding Recent Trends in House Prices and Home Ownership. [Technical Report] National Bureau of Economic Research; 2007.

R. Shiller, Historic turning points in real estate, Eastern Economic Journal 2008;34(1):1–13.

H. Simon, Models of Bounded Rationality. Cambridge, MA: MIT Press; 1982.

D. Sommervoll, T. Borgersen, T. Wennemo, Endogenous housing market cycles, Journal of Banking and Finance 2010;34(3):557–567.

M. Taylor, H. Allen, The use of technical analysis in the foreign exchange market, Journal of International Money and Finance 1992;11:304–314.

S. ter Ellen, W. Verschoor, R. Zwinkels, Dynamic expectation formation in the foreign exchange market, Journal of International Money and Finance 2013;37:75–97.

S. ter Ellen, R. Zwinkels, Oil price dynamics: a behavioral finance approach with heterogeneous agents, Energy Economics 2010;32:1427–1434.

L. Tesfatsion, K. Judd, Agent-Based Computational Economics. Handbook of Computational Economics. Elsevier; 2006;vol. 2.

S. Thurner, J. Farmer, J. Geanakoplos, Leverage causes fat tails and clustered volatility, Quantitative Finance 2012;12(5):695–707.

F. Tramontana, L. Gardini, R. Dieci, F. Westerhoff, The emergence of bull and bear dynamics in a nonlinear model of interacting markets, Discrete Dynamics in Nature and Society 2009;2009:1–30.

F. Tramontana, L. Gardini, R. Dieci, F. Westerhoff, Global bifurcations in a three-dimensional financial model of bull and bear interactions, G.-I. Bischi, C. Chiarella, L. Gardini, eds. Nonlinear Dynamics in Economics, Finance and Social Sciences: Essays in Honour of John Barkley Rosser Jr. Springer; 2010:333–352.

J. Wachter, Portfolio and consumption decisions under mean-reverting returns: an exact solution for complete markets, Journal of Financial and Quantitative Analysis 2002;37:63–91.

F. Westerhoff, Speculative markets and the effectiveness of price limits, Journal of Economic Dynamics and Control 2003;28:439–508.

F. Westerhoff, Multiasset market dynamics, Macroeconomic Dynamics 2004;8:591–616.

F. Westerhoff, Exchange rate dynamics: a nonlinear survey, Handbook of Research on Complexity. Cheltenham: Edward Elgar; 2009:287–325.

F. Westerhoff, R. Dieci, The effectiveness of Keynes–Tobin transaction taxes when heterogeneous agents can trade in different markets: a behavioral finance approach, Journal of Economic Dynamics and Control 2006;30:293–322.

F. Westerhoff, S. Reitz, Nonlinearities and cyclical behavior: the role of chartists and fundamentalists, Studies in Nonlinear Dynamics and Econometrics 2003;7(4) Article no. 3.

W. Wheaton, Real estate “cycles”: some fundamentals, Real Estate Economics 1999;27(2):209–230.

C. Yeh, C. Yang, Examining the effectiveness of price limits in an artificial stock market, Journal of Economic Dynamics and Control 2010;34:2089–2108.

H. Yoshida, T. Asada, Dynamic analysis of policy lag in a Keynes–Goodwin model: stability, instability, cycles and chaos, Journal of Economic Behavior and Organization 2007;62:441–469.

E. Zeeman, The unstable behavior of stock exchange, Journal of Mathematical Economics 1974;1:39–49.


 Acknowledgments: We are grateful to the editors, Cars Hommes and Blake LeBaron, and three reviewers for their very helpful comments. We also thank the participants to the Workshop “Handbook of Computational Economics, Volume 4, Heterogeneous Agent Models”, hosted by the Amsterdam School of Economics, University of Amsterdam, for insightful discussions and suggestions. We would like to dedicate this survey to the memory of Carl Chiarella who inspired and collaborated with us on a large body of research covered partially in this chapter. Financial support from the Australian Research Council (ARC) under Discovery Grant (DP130103210) is gratefully acknowledged.”

1  “See Allen and Taylor (1990) for foreign exchange rate markets and Menkhoff (2010) for fund managers.”

2  “They include excess volatility, excess skewness, fat tails, volatility clustering, long range dependence in volatility, and various power-law behavior, as detailed in Pagan (1996) and Lux (2009b).”

3  “This order flow can be motivated by assuming that investors maximize their expected CARA utility under their beliefs. This is particular the case when prices or payoffs of the risky asset are assumed to be normally distributed, agents make a myopic mean-variance decision, and linear price adjustment rule is used by market maker. When prices are assumed to be log-normal, the order flow and price adjustment in log-linear price would be more appropriate (see Franke and Westerhoff, 2011, 2012 for the related discussion), though their micro-economic foundation becomes less clear with heterogeneous expectations.”

4  “A further example of switching based on the discrete-choice approach is contained in the multi-asset model discussed in Section 4.2, whereas in the models described in Sections 5.1.2 and 5.2 investors' shares evolve through a simplified mechanism based on current market conditions.”

5  “Note that, while most HAMs with strategy switching are based on CARA utility maximization, the evolutionary finance approach is consistent with CRRA utility. Other models where endogenous dynamics emerge due to the evolution of the wealth shares of heterogeneous investors are Levy et al. (1994), Chiarella and He (2001), Chiarella et al. (2006a), Anufriev and Dindo (2010), Bottazzi and Dindo (2014).”

6  “See, for example, Fama and French (1996), Daniel et al. (1998), and Hong and Stein (1999).”

7  “See, for example, Kalecki (1935), Goodwin (1951), Larson (1964), Mackey (1989), Phillips (1957), Yoshida and Asada (2007), and Matsumoto and Szidarovszky (2011).”

8  “The fact that the S-shaped demand function captures the trend following behavior is well documented in the HAM literature (see, for example, Chiarella et al., 2009a).”

9  “Unless specified otherwise, the parameter values for Figs. 4 and 5 are: k=0.05Image, μ=1Image, βf=1.4Image, βc=1.4Image, η=0.5Image, β=0.5Image, C=0.02Image, F¯=1Image, σF=0.12Image, and σM=0.05Image.”

10  “For convenience of return calculations, we use log-price instead of price.”

11  “To track the profitability of the trading strategies easily, we do not consider the adaptive evolution of the market fractions.”

12  “Effectiveness refers to the ability of transaction taxes to reduce volatility, distortion (i.e. misalignment from the fundamental price), and weight of chartism.”

13  “Matrix σζImage is not necessarily diagonal; that is, the exogenous dividend processes may be correlated across assets. The same holds for matrix σκImage, characterizing the supply process.”

14  Hommes (2001) considers a simplified case where the stock of the risky asset is endogenous (zts0Image), in which case market clearing leads to Ea,t(xt+1)=RfptImage and the performance measure reduces to the risk-adjusted profit (corrected for the strategy cost), πh,tChImage.”

15  “Benchmark portfolio zh,t1BImage, proportional to the ‘market portfolio’ s, is more (less) aggressive than the market portfolio iff θhImage is smaller (larger) than the average risk aversion θa,t1Image.”

16  “For consistency between the model's unique steady state and the fundamental price, we set θ¯=θaImage in Eq. (21).”

17  “Note that the threshold (25) for asset j is independent of the parameters specific to any other asset, since the fitness measure and the variance-covariance matrices are in higher order terms. They can affect the nonlinear dynamics, but not the dynamics of the linearized system.”

18  “A large literature on time-varying betas has been developed within the conditional CAPM, which proves successful in explaining the cross-section of returns and a number of empirical ‘anomalies’ (see, e.g. Jagannathan and Wang, 1996). However, most models of the time-varying betas are motivated by econometric estimation and generally lack economic intuition.”

19  “The CAPM relation (26) is evolutionary, since asset and market returns, as well as the corresponding consensus beliefs, co-evolve endogenously, based on the dynamic HAM with expectations feedback.”

20  “The parameter used in Fig. 8 are θf=θc=1Image, Cf=4Image, Cc=1Image, γ=diag[0.3,0.3]Image, α=diag[0.4,0.5]Image, λ=1.5Image, δ=0.98Image, η=1.5Image, s=(0.1,0.1)TImage, rf:=Rf1=0.025Image, d=(0.08,0.05)TImage, Ω0=[σ12,ρσ1σ2;ρσ1σ2,σ22]Image, where σ1=0.6Image, σ2=0.4Image, ρ=0.5Image. Parameters rfImage, Ω0Image, α, γ, δ, dImage, CfImage, and CcImage are expressed in annual terms and converted to monthly via the factor 1/12 (δ is converted to a monthly value of 0.9983, in such a way to preserve the average memory length). Supply and dividend noise parameters are σκ=diag[0.001,0.001]Image and σζ=diag[0.002,0.002]Image. The parameter setting is one where the underlying deterministic model has a stable fundamental steady state, namely, η<ηˆm:=minjJoηˆjImage. When the system is no longer stable due to larger switching intensity η, even stronger effects can be observed.”

21  “In general, we use a ‘tilde’ to denote demand components and behavioral parameters characterizing cross-market traders, whereas analogous quantities without the tilde be related to home-market traders.”

22  “For convenience, we define the exchange rate S as the price of one unit of domestic currency in terms of the foreign currency.”

23  “Further nonlinearities may result from speculative demand UtImage, as shown below.”

24  “At the FSS, stock prices and the exchange rate are at their fundamental values.”

25  “Similar weighting mechanisms have also been used in de Grauwe et al. (1993), Bauer et al. (2009), and Gaunersdorfer and Hommes (2007).”

26  Tramontana et al. (2009, 2010) investigate how bull and bear market phases may arise in a HAM of stock and foreign exchange markets similar to Dieci and Westerhoff (2010), using techniques from nonlinear dynamics and the theory of global bifurcations.”

27  “Quantity ξtImage is positively related to investors' second-moment beliefs and risk aversion, and to the stock of housing at time t. This quantity is kept constant both for analytical tractability and for estimation purposes.”

28  “Under the assumed belief types, Eq. (33) simplifies to (34) provided that Xt+1Image and Qt+2Image are regarded as conditionally and mutually independent in agents' beliefs at time t.”

29  “Performance measures Uc,t1Image and Uf,t1Image are related to investors' demand and realized returns in the previous period. Under simplifying assumptions, they can be rewritten as nonlinear functions of past relative deviations XtiImage (i=1,2,3Image), as well.”

30  “Calibration of the fundamental model parameters R¯:=(1+r)/(1+g)Image and ξ¯:=ξ/(1+g)Image is based on estimates of average housing risk premia from earlier literature (in particular Himmelberg et al., 2005) and on average quarterly rental yields (average of Qt/PtImage) obtained from OECD housing datasets. Based on the datasets of prices and rents and the calibrated fundamental parameters, the time series Xt=lnPtlnPtImage is obtained. See Section 3 in Bolt et al. (2014) for detailed data description and parameter calibration.”

31  “Note that HtImage is interpreted as the current housing stock per investor.”

32  “In Eq. (37), the adjustment for risk affects the expected payoff instead of the discount rate in the denominator (similar to Eq. (18) in Section 4.2.2). This equation can be reduced to the standard form (33) by simple algebraic manipulations.”

33  “See Section 4.3 for a behavioral interpretation of this endogenous rule. In Figs. 12 and 13, ν=ν(P)Image is specified in such a way that the bell-shaped function w(P)Image is asymmetric, featuring stronger reaction to negative mispricing.”

34  “Note, however, that the local stability of the FSS in this model is conceptually different from the saddle-path stability in the model with perfect foresight.”

35  “In the bottom-right panel of Fig. 13, the dark gray region represents the basin of a coexisting attracting closed orbit.”

36  “Further experimental evidence on the negative feedback and the stabilizing role of elastic housing supply is provided by Bao and Hommes (2015) in a related heterogeneous-agent setting.”

37  “In an interesting recent paper, Diks and Wang (2016) find a similar cubic-type nonlinearity, by applying stochastic catastrophe theory to housing market dynamics.”

38  “See also Dieci and Westerhoff (2013a) for similar dynamics in a housing market model with different specifications of housing supply and demand.”

39  “Chartist and fundamentalist speculative demand is assumed to be proportional to their expected log-returns.”

40  Eichholtz et al. (2015) develop and estimate a similar HAM based on a long-term time series of house prices in Amsterdam.”

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.119.132.223