Chapter 19 ASSET PRICES, CONSUMPTION, AND THE BUSINESS CYCLE * JOHN Y. CAMPBELL Harvard University and NBER. Department of Economics, Littauer Center, Harvard University, Cambridge, MA 02138, USA Contents Abstract 1232 Keywords 1232 1. Introduction 1233 2. International asset market data 1238 3. The equity premium puzzle 1245 3.1. The stochastic discount factor 1245 3.2. Consumption-based asset pricing with power utility 1249 3.3. The riskfree rate puzzle 1252 3.4. Bond returns and the equity premium and riskfrce rate puzzles 1255 3.5. Separating risk aversion and intertemporal substitution 1256 4. The dynamics of asset returns and consumption 1260 4.1. Time-variation in conditional expectations 1260 4.2. A loglinear asset pricing framework 1264 4.3. The stock market volatility puzzle 1268 4.4. Implications for the equity premium puzzle 1272 4.5. What does the stock market forecast? 1275 4.6. Changing volatility in stock returns 1277 4.7. What does the bond market forecast? 1280 5. Cyclical variation in the price of risk 1284 5.1. Habit formation 1284 5.2. Models with heterogeneous agents 1290 * This chapter draws heavily on John Y. Campbell, "Consumption and the Stock Market: Interpreting International Experience", Swedish Economic Policy Review 3:251-299, Autumn 1996. I am grateful to the National Science Foundation for financial support, to Tim Chue, Vassil Konstantinov, and Luis Viceira for able research assistance, to Andrew Abel, Olivier Blanchard, Ricardo Caballero, Robert Shiller, Andrei Shleifer, John Taylor, and Michael Woodford for helpful comments, and to Barclays de Zoete Wedd Securities Limited, Morgan Stanley Capital International, David Barr, Bjorn Hansson, and Paul S6derlind for providing data. Handbook of Mactveconomics, Volume 1, Edited by J.B. lhylor and M. WoodJbrd © 1999 Elsevier Science B.V. All tqghts reserved 1231 1232 J..Y Campbell 5.3. Irrationalexpectations 6. Some implications for macroeconomics References 1293 1296 1298 Abstract This chapter reviews the behavior of financial asset prices in relation to consumption. The chapter lists some important stylized facts that characterize US data, and relates them to recent developments in equilibrium asset pricing theory. Data from other countries are examined to see which features of the US experience apply more generally. The chapter argues that to make sense of asset market behavior one needs a model in which the market price of risk is high, time-varying, and correlated with the state of the economy. Models that have this feature, including models with habitformation in utility, heterogeneous investors, and irrational expectations, are discussed. The main focus is on stock returns and short-term real interest rates, but bond returns are also considered. Keywords JEL classification: G12 Ch. 19." Asset Prices, Consumption, and the Business Cycle 1233 1. Introduction The behavior of aggregate stock prices is a subject of enduring fascination to investors, policymakers, and economists. In recent years stock markets have continued to show some familiar patterns, including high average returns and volatile and procyclical price movements. Economists have struggled to understand these patterns. If stock prices are determined by fundamentals, then what exactly are these fundamentals and what is the mechanism by which they move prices? Researchers, working primarily with US data, have documented a host of interesting stylized facts about the stock market and its relation to short-term interest rates and aggregate consumption: (1) The average real return on stock is high. In quarterly US data over the period 1947.2 to 1996.4, a standard data set that is used throughout this chapter, the average real stock return has been 7.6% at an annual rate. (Here and throughout the chapter, the word return is used to mean a log or continuously compounded return unless otherwise stated.) (2) The average riskless real interest rate is low. 3-month Treasury bills deliver a return that is riskless in nominal terms and close to riskless in real terms because there is only modest uncertainty about inflation at a 3-month horizon. In the postwar quarterly US data, the average real return on 3-month Treasury bills has been 0.8% per year. (3) Real stock returns are volatile, with an annualized standard deviation of 15.5% in the US data. (4) The real interest rate is much less volatile. The annualized standard deviation of the ex post real return on US Treasury bills is 1.8%, and much of this is due to short-run inflation risk. Less than half the variance of the real bill return is forecastable, so the standard deviation of the ex ante real interest rate is considerably smaller than 1.8%. (5) Real consumption growth is very smooth. The annualized standard deviation of the growth rate of seasonally adjusted real consumption of nondurables and services is 1.1% in the US data. (6) Real dividend growth is extremely volatile at short horizons because dividend data are not adjusted to remove seasonality in dividend payments. The annualized quarterly standard deviation of real dividend growth is 28.8% in the US data. At longer horizons, however, the volatility of dividend growth is intermediate between the volatility of stock returns and the volatility of consumption growth. At an annual frequency, for example, the volatility of real dividend growth is only 6% in the US data. (7) Quarterly real consumption growth and real dividend growth have a very weak correlation of 0.06 in the US data, but the correlation increases at lower frequencies to just over 0.25 at a 4-year horizon. (8) Real consumption growth and real stock returns have a quarterly correlation of 0.22 in the US data. The correlation increases to 0.33 at a 1-year horizon, and declines at longer horizons. 1234 J.Y. Campbell (9) Quarterly real dividend growth and real stock returns have a very weak correlation of 0.04 in the US data, but the correlation increases dramatically at lower frequencies to reach 0.51 at a 4-year horizon. (10) Real US consumption growth is not well forecast by its own history or by the stock market. The first-order autocorrelation of the quarterly growth rate of real nondurables and services consumption is a modest 0.2, and the log pricedividend ratio forecasts less than 5% of the variation of real consumption growth at horizons of 1 to 4 years. (11) Real US dividend growth has some short-run forecastability arising from the seasonality of dividend payments. But it is not well forecast by the stock market. The log price-dividend ratio forecasts no more than about 8% of the variation of real dividend growth at horizons of 1 to 4 years. (12) The real interest rate has some positive serial correlation; its first-order autocorrelation in postwar quarterly US data is 0.5. However the real interest rate is not well forecast by the stock market, since the log price-dividend ratio forecasts less than 1% of the variation of the real interest rate at horizons of 1 to 4 years. (13) Excess returns on US stock over Treasury bills are highly forecastable. The log price-dividend ratio forecasts 18% of the variance of the excess return at a 1-year horizon, 34% at a 2-year horizon, and 51% at a 4-year horizon. These facts raise two important questions for students of macroeconomics and finance: • Why is the average real stock return so high in relation to the average short-term real interest rate? • Why is the volatility of real stock returns so high in relation to the volatility of the short-term real interest rate? Mehra and Prescott (1985) call the first question the "equity premium puzzle". 1 Finance theory explains the expected excess return on any risky asset over the riskless interest rate as the quantity of risk times the price of risk. In a standard consumptionbased asset pricing model of the type studied by Hansen and Singleton (1983), the quantity of stock market risk is measured by the covariance of the excess stock return with consumption growth, while the price of risk is the coefficient of relative risk aversion of a representative investor. The high average stock return and low riskless interest rate (stylized facts 1 and 2) imply that the expected excess return on stock, the equity premium, is high. But the smoothness of consumption (stylized fact 5) makes the covariance of stock returns with consumption low; hence the equity premium can only be explained by a very high coefficient of risk aversion. Shiller (1982), Hansen and Jagannathan (1991), and Cochrane and Hansen (1992) have related the equity premium puzzle to the volatility of the stochastic discount factor, or equivalently the volatility of the intertemporal marginal rate of substitution of a representative investor. Expressed in these terms, the equity premium puzzle is I For excellentrecent surveys,see Cochraneand l-lansen(1992) or Kocherlakota(1996). Ch. 19: Asset Prices, Consumption, and the Business Cycle 1235 that an extremely volatile stochastic discount factor is required to match the ratio of the equity premium to the standard deviation of stock returns (the Sharpe ratio of the stock market). Some authors, such as Kandel and Stambaugh (1991), have responded to the equity premium puzzle by arguing that risk aversion is indeed much higher than traditionally thought. However this can lead to the "riskfree rate puzzle" of Weil (1989). If investors are very risk averse, then they have a strong desire to transfer wealth from periods with high consumption to periods with low consumption. Since consumption has tended to grow steadily over time, high risk aversion makes investors want to borrow to reduce the discrepancy between future consumption and present consumption. To reconcile this with the low real interest rate we observe, we must postulate that investors are extremely patient; their preferences give future consumption almost as much weight as current consumption, or even greater weight than current consumption. In other words they have a low or even negative rate of time preference. I will call the second question the "stock market volatility puzzle". To understand the puzzle, it is helpful to classify the possible sources of stock market volatility. Recall first that prices, dividends, and returns are not independent but are linked by an accounting identity. If an asset's price is high today, then either its dividend must be high tomorrow, or its return must be low between today and tomorrow, or its price must be even higher tomorrow. If one excludes the possibility that an asset price can grow explosively forever in a "rational bubble", then it follows that an asset with a high price today must have some combination of high dividends over tile indefinite future and low returns over the indefinite future. Investors must recognize this fact in forming their expectations, so when an asset price is high investors expect some combination of high future dividends and low future returns. Movements in prices must then be associated with some combination of changing expectations ("news") about future dividends and changing expectations about future returns; the latter can in turn be broken into news about future riskless real interest rates and news about future excess returns on stocks over short-term debt. Until the early 1980s, most financial economists believed that there was very little predictable variation in stock returns and that dividend news was by far the most important factor driving stock market fluctuations. LeRoy and Porter (1981) and Shiller (1981) challenged this orthodoxy by pointing out that plausible measures of expected future dividends are far less volatile than real stock prices. Their work is related to stylized facts 6, 9, and 11. Later in the 1980s Campbell and Shiller (1988), Fama and French (1988a,b, 1989), Poterba and Summers (1988) and others showed that real stock returns are highly forecastable at long horizons. The variables that predict returns are ratios of stock prices to scale factors such as dividends, earnings, moving averages of earnings, or the book value of equity. When stock prices are high relative to these scale factors, subsequent long-horizon real stock returns tend to be low. This predictable variation in stock returns is not matched by any equivalent variation in long-term real interest rates, which are comparatively stable and do not seem to move with the stock market. 1236 J.Y. Campbell in the late 1970s, for example, real interest rates were unusually low yet stock prices were depressed, implying high forecast stock returns; the 1980s saw much higher real interest rates along with buoyant stock prices, implying low forecast stock returns. Thus excess returns on stock over Treasury bills are just as forecastable as real returns on stock. This work is related to stylized facts 12 and 13. Campbell (1991) uses this evidence to show that the great bulk of stock market volatility is associated with changing forecasts of excess stock returns. Changing forecasts of dividend growth and real interest rates are much less important empirically. The stock market volatility puzzle is closely related to the equity premium puzzle. A complete model of stock market behavior must explain both the average level of stock prices and their movements over time. One strand of work on the equity premium puzzle makes this explicit by studying not the consumption covariance of measured stock returns, but the consumption covariance of returns on hypothetical assets whose dividends are determined by consumption. The same model is used to generate both the volatility of stock prices and the implied equity premium. This was the approach of Mehra and Prescott (1985), and many subsequent authors have followed their lead. Unfortunately, it is not easy to construct a general equilibrium model that fits all the stylized facts given above. The standard model of Mehra and Prescott (1985) gets variation in stock price-dividend ratios only from predictable variation in consumption growth which moves the expected dividend growth rate and the riskless real interest rate. The model is not consistent with the empirical evidence for predictable variation in excess stock returns. Bond market data pose a further challenge to this standard model of stock returns. In the model, stocks behave very much like long-term real bonds; both assets are driven by long-term movements in the riskless real interest rate. Thus parameter values that produce a large equity premium tend also to produce a large term premium on real bonds. While there is no direct evidence on real bond premia, nominal bond premia have historically been much smaller than equity premia. Since the data suggest that predictable variation in excess returns is an important source of stock market volatility, researchers have begun to develop models in which the quantity of stock market risk or the price of risk change through time. ARCH models and other econometric methods show that the conditional variance of stock returns is highly variable. If this conditional variance is an adequate proxy for the quantity of stock market risk, then perhaps it can explain the predictability of excess stock returns. There are several problems with this approach. First, changes in conditional variance are most dramatic in daily or monthly data and are much weaker at lower frequencies. There is some business-cycle variation in volatility, but it does not seem strong enough to explain large movements in aggregate stock prices [Bollerslev, Chou and Kroner (1992), Schwert (1989)]. Second, forecasts of excess stock returns do not move proportionally with estimates of conditional variance [Harvey (1989, 1991), Chou, Engle and Kane (1992)]. Finally, one would like to derive stock market volatility endogenously within a model rather than treating it as an exogenous variable. There is little evidence of cyclical variation in consumption or dividend volatility that could explain the variation in stock market volatility. Ch. 19: Asset Prices, Consumption, and the Business Cycle 1237 A more promising possibility is that the price of risk varies over time. Time-variation in the price of risk arises naturally in a model with a representative agent whose utility displays habit-formation. Campbell and Cochrane (1999), building on the work of Abel (1990), Constantinides (1990), and others, have proposed a simple asset pricing model of this sort. Campbell and Cochrane suggest that assets are priced as if there were a representative agent whose utility is a power function of the difference between consumption and "habit", where habit is a slow-moving nonlinear average of past aggregate consumption. This utility fi.mction makes the agent more risk-averse in bad times, when consumption is low relative to its past history, than in good times, when consumption is high relative to its past history. Stock market volatility is explained by a small amount of underlying consumption (dividend) risk, amplified by variable risk aversion; the equity premium is explained by high stock market volatility, together with a high average level of risk aversion. Time-variation in the price of risk can also arise from the interaction of heterogeneous agents. Constantinides and Duffle (l 996) develop a simple framework with many agents who have identical utility functions but heterogeneous streams of labor income; they show how changes in the cross-sectional distribution of income can generate any desired behavior of the market price of risk. Grossman and Zhou (1996) and Wang (1996) move in a somewhat different direction by exploring the interactions of agents who have different levels of risk aversion. Some aspects of asset market behavior could also be explained by irrational expectations of investors. If investors are excessively pessimistic about economic growth, for example, they will overprice short-term bills and underprice stocks; this would help to explain the equity premium and riskfree rate puzzles. If investors overestimate the persistence of variations in economic growth, they will overprice stocks when growth has been high and underprice them when growth has been low, producing time-variation in the price of risk [Barsky and DeLong (1993)]. This chapter has three objectives. First, it tries to summarize recent work on stock price behavior, much of which is highly technical, in a way that is accessible to a broader professional audience. Second, the chapter summarizes stock market data from other countries and asks which of the US stylized facts hold true more generally. The recent theoretical literature is used to guide the exploration of the international data. Third, the chapter systematically compares stock market data with bond market data. This is an important discipline because some popular models of stock prices are difficult to reconcile with the behavior of bond prices. The organization of the chapter is as follows. Section 2 introduces the international data and reviews stylized facts 1-9 to see which of them apply outside the USA. (Additional details are given in a Data Appendix available on the author's web page or by request from the author.) Section 3 discusses the equity premium puzzle, taking the volatility of stock returns as given. Section 4 discusses the stock market volatility puzzle; this section also reviews stylized facts 10-13 in the international data. Sections 3 and 4 drive one towards the conclusion that the price of risk is both high and time-varying. It must be high to explain the equity premium puzzle, and it 1238 J. Y Campbell must be time-varying to explain the predictable variation in stock returns that seems to be responsible for the volatility of stock returns. Section 5 discusses models which produce this result, including models with habit-formation in utility, heterogeneous investors, and irrational expectations. Section 6 draws some implications for other topics in macroeconomics, including the modelling of investment, labor supply, and the welfare costs of economic fluctuations. 2. International asset market data The stylized facts described in the previous section apply to postwar quarterly US data. Most empirical work on stock prices uses this data set, or a longer annual US time series originally put together by Shiller (1981). But data on stock prices, interest rates, and consumption are also available for many other countries. In this chapter I use an updated version of the international developed-country data set in Campbell (1996a). The data set includes Morgan Stanley Capital International (MSCI) stock market data covering the period since 1970. ! combine the MSCI data with macroeconomic data on consumption, short- and long-term interest rates, and the price level from the International Financial Statistics (IFS) of the International Monetary Fund. For some countries the IFS data are only available quarterly over a shorter sample period, so I use the longest available sample for each country. Sample start dates range from 1970.1 to 1982.2, and sample end dates range from 1995.1 to 1996.4. I work with data from 11 countries: Australia, Canada, France, Germany, Italy, Japan, the Netherlands, Sweden, Switzerland, the United Kingdom, and the United States 2. For some purposes it is useful to have data over a much longer span of calendar time. I have been able to obtain annual data for Sweden over the period 1920-1994 and the UK over the period 1919-1994 to complement the US annual data for the period 1891-1995. The Swedish data come from Frennberg and Hansson (1992) and Hassler, Lundvik, Persson and S6derlind (1994), while the UK data come from Barclays de Zoete Wedd Securities (1995) and The Economist (1987) 3. In working with international stock market data, it is important to keep in mind that different national stock markets are of very different sizes, both absolutely and in 2 The first version of this paper, following Campbell (1996a), also presented data for Spain. However Spain, unlike the other countries in the sample, underwent a major political change to democratic governmentduringthe sampleperiod, and both asset returns and inflationshow dramatic shifts fi'omthe 1970s to the 1980s. It seems more conservativeto consider Spain as an emerging market and exclude it from the developed-countrydata set. 3 1acknowledgethe invaluableassistance of Bjorn Hansson and Paul S6derlindwith the Swedishdata, and David Barr with the UK data. Full details about the construction of the quarterly and annual data are givenin a Data Appendixavailable on the attthor's web page or by request fi'omthe author. Ch. 19." Asset Prices, Consumption, and the Business Cycle Table 1 MSCI market capitalization, 1993a 1239 vl v, vi Country V/ -- (%) - - (%) (Bill. of US$) GDPi VusMxcl ~f';4Vi - - (%) AUL 117.9 41.55 4.65 1.85 CAN 167.3 30.62 6.60 2.63 FR 272.5 22.49 10.75 4.29 GER 280.7 16.83 11.07 4.41 ITA 86.8 9.45 3.42 1.37 JAP 1651.9 39.74 65.16 25.98 NTH 136.7 45.91 5.39 2.15 SWD 62.9 36.22 2.48 0.99 SWT 205.6 87.46 8.12 3.23 UK 758.4 79.52 29.91 11.93 USA MSCI 2535.3 37.25 100.00 39.88 USA CRSP 4875.6 71.64 192.30 a vi is the stock indexmarket capitalizationin billions of 1993 US dollars. All stock index data are from Morgan StanleyCapital International(MSCI),exceptfor USA-CRSPwhich is from the Center for Research in SecurityPrices. Vi/GDPi is the indexmarket capitalization as a percentageof 1993 GDP, Vi/VtjsMscI is the indexmarket capitalizationas a percentage of the market capitalizationof the US MSCI index, and Vi/(~ i Vi) is the percentage share of the index market capitalizationin the total market capitalizationof all the MSCI indexes. Abbreviations:AUL, Australia; CAN, Canada; FR, France; GER, Germany; 1TA,Italy; JPN, Japan; NTH, Netherlands; SWD, Sweden; SWT, Switzerland;UK, United Kingdom; USA, United States of America. proportion to national GDP's. Table 1 illustrates this by reporting several measures of stock market capitalization for the quarterly MSCI data. Column 1 gives the market capitalization for each country's MSCI index at the end of 1993, in billions of $US. Column 2 gives the market capitalization for each country as a fraction of its GDP. Column 3 gives the market capitalization for each country as a fraction of the US MSCI index capitalization. Column 4 gives the market capitalization for each country as a fraction of the value-weighted world MSCI index capitalization. Since the MSCI index for the United States is only a subset of the US market, the last row of the table gives the same statistics for the value-weighted index of New York Stock Exchange and American Stock Exchange stocks reported by the Center for Research in Security Prices (CRSP) at the University of Chicago. Table 1 shows that most countries' stock markets are dwarfed by the US market. Column 3, for example, shows that the Japanese MSCI index is worth only 65% of the US MSCI index, the UK MSCI index is worth only 30% of the US index, the French and German MSCI indexes are worth only 11% of the US index, and all 1240 J.Y. Campbell other countries' indexes are worth less than 10% of the US index. Column 4 shows that the USA and Japan together account for 66% of the world market capitalization, while the USA, Japan, the UK, France, and Germany together account for 86%. In interpreting these numbers one must keep in mind that the MSCI indexes do not cover the whole market in each country (the US MSCI index, for example, is worth about half the US CRSP index), but they do give a guide to relative magnitudes across countries. Table 1 also shows that different countries' stock market values are very different as a fraction of GDP. If one thinks that total wealth-output ratios are likely to be fairly constant across countries, then this indicates that national stock markets are very different fractions of total wealth in different countries. In highly capitalized countries such as the UK and Switzerland, the MSCI index accounts for about 80% of GDP, whereas in Germany and Italy it accounts for less than 20% of GDR The theoretical convention of treating the stock market as a claim to total consumption, or as a proxy for the aggregate wealth of an economy, makes much more sense in the highly capitalized countries 4. Table 2 reports summary statistics for international asset returns. For each country the table reports the mean, standard deviation, and first-order autocorrelation of the real stock return and the real return on a short-term debt instrument 5. The first line of Table 2 gives numbers for the standard postwar quarterly US data set summarized in the introduction. The next panel gives numbers for the 11-country quarterly MSCI data, and the bottom panel gives numbers for the long-term annual data sets. The table shows that the first four stylized facts given in the introduction are fairly robust across countries. (1) Stock markets have delivered average real returns of 5% or better in almost every country and time period. The exceptions to this occur in short-term quarterly data, and are concentrated in markets that are particularly small relative to GDP (Italy), or that predominantly represent claims on natural resources (Australia and Canada). (2) Short-term debt has rarely delivered an average real return above 3%. The exceptions to this occur in two countries, Germany and the Netherlands, whose sample periods begin in the late 1970s and thus exclude much of the surprise inflation of the oil-shock period. 4 Stock ownership also tends to be much more concentrated in the countries with low capitalization. La Porta, Lopez-de-Silanes, Shleifer and Vishny (1997) have related these international patterns to differences in the protections afforded outside investors by different legal systems. 5 As explained in the Data Appendix, the best available short-term interest rate is sometimes a Treasury bill rate and sometimes another money market interest rate. Both means and standard deviations are given in almualizedpercentage points. To annualize the raw quarterly numbers, means are multiplied by 400 while standard deviations are multiplied by 200 (since standard deviations increase with the square root of the time interval in serially uncorrelated data). Ch. 19." Asset Prices, Consumption, and the Business Cycle 1241 Table 2 International stock and bill returns a Country Sample period re o(r~) p(rc) rj-~ a(rf ) p(rj ) USA 1947.2-1996.4 7.569 15.453 0.104 0.794 1.761 0.501 AUL 1970.1 1996.3 2.633 23.459 0.008 1.820 2,604 0.636 CAN 1970.1-1996.3 4,518 16,721 0.119 2.738 1.932 0.674 FR 1973.2-1996.3 7.207 22.877 0,088 2.736 1.917 0.714 GER 1978.4-1996.3 8.135 20.326 0,066 3.338 1.161 0.322 ITA 1971.2-1995.3 0.514 27.244 0.071 2.064 2.957 0.681 JPN 1970.2--1996.3 5.831 21,881 0,017 1.538 2.347 0.493 NTH 1977.2-1996.2 12.721 15.719 0.027 3.705 1.542 -0.099 SWD 1970.1-1995.1 7.948 23.867 0.053 1.520 2.966 0.218 SWT 1982.~1996,3 11.548 20.431 0.112 1.466 1.603 0.255 UK 1970.1-1996,3 7,236 21,555 0,103 1,081 3.067 0,474 USA 1970.1-1996.4 5.893 17.355 0.076 1.350 1,722 0.568 SWD 1920-1994 6.219 18.654 0.064 2.073 5.918 0,708 UK 1919-1994 7.314 22.675 -0.024 1.198 5.446 0.591 USA 1891 1995 6.697 18.634 0.025 1.955 8.919 0.338 a ~ is the mean log real return on the stock market index, multiplied by 400 in quarterly data or 100 in annual data to express in annualized percentage points, tY(re) is the standard deviation of the log real return on the market index, multiplied by 200 in quarterly data or 100 in annual data to express in annualized percentage points, p(re) is the first-order autocorrelation of the log real return on the market index, rT, o(rf), and p(t).) are defined in the same way for the real return on a 3-month money market instrument. The money market instruments vary across countries and are described in detail in the Data Appendix. Abbreviations: AUL, Australia; CAN, Canada; FR, France; GER, Germany; ITA, Italy; JPN, Japan; NTH, Netherlands; SWD, Sweden; SWT, Switzerland; UK, United Kingdom; USA, United States of America. (3) The annualized standard deviation of stock returns ranges from 15% to 27%. It is striking that the market with the highest volatility, Italy, is the smallest market relative to GDP and the one with the lowest average return, (4) In quarterly data the annualized volatility of real returns on short debt is around 3% for the UK, Italy, and Sweden, around 2.5% for Australia and Japan, and below 2% for all other countries. Volatility is higher in long-term annual data because of large swings in inflation in the interwar period, particularly in t919-21. Much of the volatility in these real returns is probably due to unanticipated inflation and does not reflect volatility in the ex ante real interest rate. 1242 J Y. Campbell These numbers show that high average stock returns, relative to the returns on shortterm debt, are not unique to the United States but characterize many other countries as well. Recently a number of authors have suggested that average excess returns in the USA may be overstated by sample selection or survivorship bias. if economists study the USA because it has had an unusually successful economy, then sample average US stock returns may overstate the true mean US stock return. Brown, Goetzmann and Ross (1995) present a formal model of this effect. While survivorship bias may affect data from all the countries included in Table 2, it is reassuring that the stylized facts are so consistent across these countries 6. Table 3 turns to data on aggregate consumption and stock market dividends. The table is organized in the same way as Table 2. It illustrates the robustness of two more of the stylized facts given in the introduction. (5) In the postwar period the annualized standard deviation of real consumption growth is never above 3%. This is true even though data are used on total consumption, rather than nondurables and services consumption, for all countries other than the USA. Even in the longer annual data, which include the turbulent interwar period, consumption volatility slightly exceeds 3% only in the USA. (6) The volatility of dividend growth is much greater than the volatility of consumption growth, but generally less than the volatility of stock returns. The exceptions to this occur in countries with highly seasonal dividend payments; these countries have large negative autocorrelations for quarterly dividend growth and much smaller volatility when dividend growth is measured over a full year rather than over a quarter. Table 4 reports the contemporaneous correlations among real consumption growth, real dividend growth, and stock returns. It turns out that these correlations are somewhat sensitive to the timing convention used for consumption. A timing convention is needed because the level of consumption is a flow during a quarter rather than a point-in-time observation; that is, the consumption data are timeaveraged 7. If we think of a given quarter's consumption data as measuring consumption at the beginning of the quarter, then consumption growth for the quarter is next quarter's consumption divided by this quarter's consumption. If on the other hand 6 Goetzmalm and Jorion (1997) consider imernational stock-price data from earlier in the 20th Century and argue that the long-term average real growth rate of stock prices has been higher in the US than elsewhere. However they do not have data on dividend yields, which are an important component of total return and are likely to have been particularly important in Europe dtmng the troubled intei~ar period. 7 Tilne-averaging is one of a number of interrelated issues that arise in relating measured consumption data to the theoretical concept of consumption. Other issues include measurement error, seasonal adjustment, and the possibilitythat some goods classified as nondurable in the national income accounts are in fact durable. Grossman, Melino and Shiller (1987), Wheatley (1988), Miron (1986), and Heaton (1995) handle time-averaging, measurement error, seasonality, and dinability, respectively, in a much more careful way than is possible here, while Wilcox (1992) provides a detailed account of the sampling procedures used to constxuct US consumption data. Ch. 19." Asset Prices, Consumption, and the Business Cycle 1243 Table 3 International consumption and dividends a Country Sample period Ac o(Ac) p(Ac) Ad ~r(Ad) p(Ad) USA 1947.2-1996.4 1.921 1.085 0.221 2.225 28.794 -0.544 AUL 1970.1-1996.3 1.886 2.138 0.351 0.883 36.134 -0.451 CAN 1970.1 1996.3 1.853 2.083 0.113 -0.741 5.783 0.540 FR 1973.2 1996.3 1.600 2.121 --0.093 1.214 13.383 -0.159 GER 1978.4 1996.3 1.592 2.478 -0.328 1.079 8.528 0.018 ITA 1971.2-1995.3 2.341 1.724 0.253 -4.919 19.635 0.294 JPN 1970.2-1996.3 3.384 2.347 -0.225 2.489 4.504 0.363 NTH 1977.2 1996.2 1.661 2.772 -0.265 4.007 4.958 0.277 SWD 1970.1-1995.1 0.705 1.920 0.305 1.861 13.595 0.335 SWT 1982.2-1996.3 0.376 2.246 -0.4t9 4.143 6.156 0.165 UK 1970.1-1996.3 1.991 2.583 -0.017 0.681 7.125 0.335 USA 1970.1 1996.4 1.722 0.917 0.390 0.619 17.229 0.581 SWD 1920 1994 1.790 2.866 0.159 0.423 12.215 0.214 UK 1919-1994 1.443 2.898 0.281 1.844 7.966 0.225 USA 1891-1995 1.773 3.256 -0.117 1.485 14.207 -0.087 a Ac is the mean log real consumption growth rate, multiplied by 400 in quarterly data or 100 in annual data to express in annualized percentage points, cr(Ac) is the standard deviation of the log real consumption growth rate, multiplied by 200 in quarterly data or 100 in annual data to express in annualized percentage points, p(Ac) is the first-order autocorrelation of the log real consumption growth rate. Ad, a(Ad), and p(Ad) are defined in the same way for the real dividend growth rate. Consumption is nondurables and selwices consumption in the USA, and total consumption elsewhere. Abbreviations: AUL, Australia; CAN, Canada; FR, France; GER, Germany; ITA, Italy; JPN, Japan; NTH, Netherlands; SWD, Sweden; SWT, Switzerland; UK, United Kingdom; USA, United States of America. we think of the consumption data as measuring consumption at the end of the quarter, then consumption growth is this quarter's consumption divided by last quarter's consumption. Table 4 uses the former, "beginning-of-quarter" timing convention because this produces a higher contemporaneous correlation between consumption growth and stock returns. The timing convention has less effect on correlations when the data are measured at longer horizons. Table 4 also shows how the correlations among real consumption growth, real dividend growth, and real stock returns vary with the horizon. Each pairwise correlation among these series is calculated for horizons of 1, 4, 8, and 16 quarters in the quarterly data and for horizons of 1, 2, 4, and 8 years in the long-term annual data. The table illustrates three more stylized facts from the introduction. 1244 J.Y. Campbell % o m 0 0 I l l II II [I I I I I II I I I I I I I I I I I ;a4zaaza;a a ~ d e m ~ : d ~ d ~ d o NN~ ', ~ '~ o "~, . ~ b ~ fa, >~ N ~ ca Ch. 19: AssetPrices',Consumption,and the BusinessCycle 1245 (7) Real consumption growth and dividend growth are generally weakly positively correlated in the quarterly data. In many countries the correlation increases strongly with the measurement horizon. However long-horizon correlations remain close to zero for Australia and Switzerland, and are substantially negative for Italy (with a very small stock market) and Japan (with anomalous dividend behavior). The correlations of consumption and dividend growth are positive and often quite large in the longer-term annual data sets. (8) The correlations between real consumption growth rates and stock returns are quite variable across countries. They tend to be somewhat higher in high-capitalization countries (with the notable exception of Switzerland), which is consistent with the view that stock returns proxy more accurately for wealth returns in these countries. Correlations typically increase with the measurement horizon out to 1 or 2 years, and are moderately positive in the longer-term annual data sets. (9) The correlations between real dividend growth rates and stock returns are small at a quarterly horizon but increase dramatically with the horizon. This pattern holds in every country. The correlations also increase strongly with the horizon in the longer-term annual data. After this preliminary look at the data, I now use some simple finance theory to interpret the stylized facts. 3. The equity premium puzzle 3.1. The stochastic discount factor To understand the equity premium puzzle, consider the intertemporal choice problem of an investor, indexed by k, who can trade freely in some asset i and can obtain a gross simple rate of return (1 +Ri,t+l) on the asset held from time t to time t + 1. If the investor consumes Ck~ at time t and has time-separable utility with discount factor 6 and period utility U(Ckt), then her first-order condition is g'((~t) = 6E~ [(1 +Ri, l+l)Ut(Ck,t+l)]. (1) The left-hand side of Eqnation (1) is the marginal utility cost of consuming one real dollar less at time t; the right-hand side is the expected marginal utility benefit from investing the dollar in asset i at time t, selling it at time t + 1, and consuming the proceeds. The investor equates marginal cost and marginal benefit, so Equation (1) must describe the optimum. Dividing Equation (1) by U'(C~) yields v'(G,,+,)]1 = Et (1 +Ri,,~,)6 ~ j --E, [(1 +Ri,,+,)Ma,,+lJ, (2) where Mk,t.~l = 6U'(Ck, t+l)/U/(Ct) is the intertemporal marginal rate of substitution of the investor, also known as the stochastic discountjdctor. This way of writing the 1246 J..Y Campbell model in discrete time is due originally to Grossman and Shiller (1981), while the continuous-time version of the model is due to Breeden (1979). Cochrane and Hansen (1992) and Hansen and Jagannathan (1991) have developed the implications of the discrete-time model in detail. The derivation just given for Equation (2) assumes the existence of an investor maximizing a time-separable utility function, but in fact the equation holds more generally. The existence of a positive stochastic discount factor is guaranteed by the absence of arbitrage in markets in which non-satiated investors can trade freely without transactions costs. In general there can be many such stochastic discount factors for example, different investors k whose marginal utilities follow different stochastic processes will have different M~,t+l - but each stochastic discount factor must satisfy Equation (2). It is common practice to drop the subscript k from this equation and simply write 1 = E, [(1 +&t+,lMi+~]. (3) In complete markets the stochastic discount factor Mt+l is unique because investors can trade with one another to eliminate any idiosyncratic variation in their marginal utilities. To understand the implications of Equation (3) it is helpful to write the expectation of the product as the product of expectations plus the covariance, E1[(1 +Ri,,~l)m~+l] = Et[(l + Ri,,+I)]E~[M,+I] + Covt[R~,l.~,M,+~]. (4) Substituting into Equation (3) and rearranging gives 1 - Covt[Ri, ¢+1,Mr+l] 1 + E,[R~,,+I] - (5) Et[Mt+l] An asset with a high expected simple return must have a low covariance with the stochastic discount factor. Such an asset tends to have low retunas when investors have high marginal utility. It is risky in that it fails to deliver wealth precisely when wealth is most valuable to investors. Investors therefore demand a large risk premium to hold it. Equation (5) must hold for any asset, including a riskless asset whose gross simple return is 1 + Ry;t~1. Since the simple riskless return has zero covariance with the stochastic discount factor (or any other random variable), it is just the reciprocal of the expectation of the stochastic discount factor: 1 1 +R/;,+t - Et[M,+I]" (6) This can be used to rewrite Equation (5) as 1 +Et[Ri,,+t] = (1 +RLt+I)(1 -Cov,[Ri,~+L,Mi+l]). (7) For simplicity I now follow Hansen and Singleton (1983) and assume that the joint conditional distribution of asset returns and the stochastic discount factor is lognormal Ch. 19." Asset Prices, Consumption, and the Business Cycle 1247 and homoskedastic. While these assumptions are not literally realistic - stock returns in particular have fat-tailed distributions with variances that change over time - they do make it easier to discuss the main forces that should determine the equity premium. When a random variable X is conditionally lognormally distributed, it has the convenient property that log EtX E~ logX + ½Vart logX, (8) where VartlogX = G[(logX -EtlogX)2]. If in addition X is conditionally homoskedastic, then Var, logX = E[(logX - E, logX) 2] = Var(logX - Et logX). Thus with joint conditional lognormality and homoskedasticity of asset returns and consumption, I can take logs of Equation (3) and obtain 2 0 = Etri, t+l + Etmt+l + (1) [02 + O£,+ 20,m]. (9) Here rnt - log(Mr) and r, -- log(1 + Ri,), while 02 denotes the unconditional variance of log return innovations Var(r/, t+t -Gri, t+l), a2 denotes the unconditional variance of innovations to the stochastic discount factor Var(mt~l Etmt+l), and G,, denotes the unconditional covariance of innovations Cov(ri, t+l - Elri, tvl, rntel - Etmt+l). Equation (9) has both time-series and cross-sectional implications. Consider first an asset with a riskless real return rt; ¢)1. For this asset the return innovation variance c~2 and the covariance aim are both zero, so the riskless real interest rate obeys rj;l+l =-Etmt+l 02 2 (10) This equation is the log counterpart of Equation (6). Subtracting Equation (10) from Equation (9) yields an expression for the expected excess return on risky assets over the riskless rate: o.2 Et[ri,/+1 - 9; tM] + ~- - -oi .... 2 (ll) The variance term on the left-hand side of Equation (11) is a Jensen's Inequality adjustment arising from the fact that we are describing expectations of log returns. This term would disappear if we rewrote the equation in terms of the log expectation of the ratio of gross simple returns: log G [(1 + Ri, t +I)/(1 + Rf, ~+1)] = -aim. The righthand side of Equation (11) says that the log risk premium is determined by the negative of the covariance of the asset with the stochastic discount factor. This equation is the log counterpart of Equation (7). The covariance O,m can be written as the product of the standard deviation of the asset return a,., the standard deviation of the stochastic discount factor (7,,,, and the 1248 J Y. Campbell ~P f~ L) q, 3 o o o V~=7 >,~° ~ •~ ,.= 03 r.~ '~ :~ Z ~ ~'" •~ o=° ~o ~ ,.~ II rm ~. o,~o 1 (12) This inequality was first derived by Shiller (1982); a multi-asset version was derived by Hansen and Jagannathan (1991) and developed further by Cochrane and Hansen (1992). The right-hand side of Equation (12) is the excess return on an asset, adjusted for Jensen's Inequality, divided by the standard deviation of the asset's return - a logarithmic Sharpe ratio for the asset. Equation (12) says that the standard deviation of the log stochastic discount factor must be greater than this Sharpe ratio for all assets i, that is, it must be greater than the maximum possible Sharpe ratio obtainable in asset markets. Table 5 uses Equation (12) to illustrate the equity premium puzzle. For each data set the first column of the table reports the average excess return on stock over shortterm debt, adjusted for Jensen's inequality by adding one-half the sample variance of the excess log return to get a sample estimate of the numerator in Equation (12). This adjusted average excess return is multiplied by 400 to express it in annualized percentage points. The second column of the table gives the aunualized standard deviation of the excess log stock return, a sample estimate of the denominator in Equation (12). This standard deviation was reported earlier in Table 2. The third column gives the ratio of the first two columns, multiplied by 100; this is a sample estimate of the lower bound on the standard deviation of the log stochastic discount factor, expressed in annualized percentage points. In the postwar US data the estimated lower bound is a standard deviation greater than 50% a year; in the other quarterly data sets it is below 10% for Italy, between 15% and 20% for Australia and Canada, and above 30% for all the other countries, in the long-run annual data sets the lower bound on the standard deviation exceeds 30% for all three countries. 3.2. Consumption-based asset pricing with power utility To understand why these numbers are disturbing, I now follow Mehra and Prescott (1985) and other classic papers on the equity premium puzzle and assume that there is a representative agent who maximizes a time-separable power utility function defined over aggregate consumption G: C¢-Y-1 u(G) - - - , (13) 1-7 where y is the coefficient of relative risk aversion. This utility function has several important properties. First, it is scale-invariant; with constant return distributions, risk premia do not change over time as aggregate wealth and the scale of the 1250 Jg CampbeH economy increase. Related to this, if different investors in the econonW have different wealth levels but the same power utility function, then they can be aggregated into a single representative investor with the same utility function as the individual investors. A possibly less desirable property of power utility is that the elasticity of intertemporal substitution, which I write as ~p, is the reciprocal of the coefficient of relative risk aversion y. Epstein and Zin (1989, 1991) and Weil (1989) have proposed a more general utility specification that preserves the scale-invariance of power utility but breaks the tight link between the coefficient of relative risk aversion and the elasticity of intertemporal substitution. I discuss this form of utility in section 3.4 below. Power utility implies that marginal utility UI(C,) = Ct r, and the stochastic discount factor Mt+l = CS(Ct+t/CI)-r. The assumption made previously that the stochastic discount factor is conditionally lognormal will be implied by the assumption that aggregate consumption is conditionally lognormal [Hansen and Singleton (1983)]. Making this assumption for expositional convenience, the log stochastic discount factor is mt+l = log(cS)- 7Ac,+1, where c, --- log(Q), and Equation (9) becomes 0 = Etri,,+l +log (5- yE, Act~l + (½) [a,.2+ y20~2- 2ro,c]. (14) Here a2 denotes Var(ct+l -Etct+l), the unconditional variance of log consumption innovations, and eric denotes Cov(ri, t+l- Etri, t Fl, ct+l - EtCt+l), the unconditional covariance of innovations. Equation (10) now becomes ~,2at2 tj/;,+t = log C5+ ]/EtAct+l 2 (15) This equation says that the riskless real rate is linear in expected consumption growth, with slope coefficient equal to the coefficient of relative risk aversion. The conditional variance of consumption growth has a negative effect on the riskless rate which can be interpreted as a precautionary savings effect. Equation (11) becomes Et[ri, t+~ --rj;t+l] + ~- = 7oic. (16) The log risk premium on any asset is the coefficient of relative risk aversion times the covariance of the asset return with consumption growth. Intuitively, an asset with a high consumption covariance tends to have low returns when consumption is low, that is, when the marginal utility of consumption is high. Such an asset is risky and commands a large risk premium. Table 5 uses Equation (16) to illustrate the equity premium puzzle. As already discussed, the first column of the table reports a sample estimate of the left-hand Ch. 19." Asset Prices, Consumption, and the Business Cycle 1251 side of Equation (16), multiplied by 400 to express it in annualized percentage points. The second column reports the annualized standard deviation of the excess log stock return (given earlier in Table 2), the fourth column reports the annualized standard deviation of consumption growth (given earlier in Table 3), the fifth column reports the correlation between the excess log stock return and consumption growth, and the sixth column gives the product of these three variables which is the annualized covariance a,~.between the log stock return and consumption growth. Finally, the table gives two columns with implied risk aversion coefficients. The column headed RRA(1) uses Equation (16) directly, dividing the adjusted average excess return by the estimated covariance to get estimated risk aversion 8. The column headed RRA(2) sets the correlation of stock returns and consumption growth equal to one before calculating risk aversion. While this is of course a counterfactual exercise, it is a valuable diagnostic because it indicates the extent to which the equity premium puzzle arises from the smoothness of consumption rather than the low correlation between consumption and stock returns. The correlation is hard to measure accurately because it is easily distorted by short-term measurement errors in consumption, and Table 4 indicates that the sample correlation is quite sensitive to the measurement horizon. By setting the correlation to one, the RRA(2) column indicates the extent to which the equity premium puzzle is robust to such issues. A correlation of one is also implicitly assumed in the volatility bound for the stochastic discount factor, Equation (12), and in many calibration exercises such as Mehra and Prescott (1985), Campbell and Cochrane (1999), or Abel (1999). Table 5 shows that the equity premium puzzle is a robust phenomenon in international data. The coefficients of relative risk aversion in the RRA(1) column are generally extremely large. They are usually many times greater than 10, the maximum level considered plausible by Mehra and Prescott (1985). In a few cases the risk aversion coefficients are negative because the estimated covariance of stock returns with consumption growth is negative, but in these cases the covariance is extremely close to zero. Even when one ignores the low correlation between stock returns and consumption growth and gives the model its best chance by setting the correlation to one, the RRA(2) column still has risk aversion coefficients above 10 in most cases. Thus the fact shown in Table 4, that for some countries the correlation of stock returns and consumption increases with the horizon, is unable by itself to resolve the equity premium puzzle. The risk aversion estimates in Table 5 are of course point estimates and are subject to sampling error. No standard errors are reported for these estimates. However authors such as Cecchetti, Lam and Mark (1993) and Kocherlakota (1996), studying the long- 8 The calculationis donecorrectly,in naturalunits,eventhoughthe tablereportsaverageexcessreturns and covafiancesin percentagepoint units. Equivalently,the ratio of the quantities givenin the table is multipliedby 100. 1252 J E Campbell run annual US data, have found small enough standard errors that they can reject risk aversion coefficients below about 8 at conventional significance levels. Of course, the validity of these tests depends on the characteristics of the data set in which they are used. Rietz (1988) has argued that there may be a peso problem in these data. A peso problem arises when there is a small positive probability of an important event, and investors take this probability into account when setting market prices. If the event does not occur in a particular sample period, investors will appear irrational in the sample and economists will mis-estimate their preferences. While it may seem unlikely that this could be an important problem in 100 years of annual data, Rietz (1988) argues that an economic catastrophe that destroys almost all stock-market value can be extremely unlikely and yet have a major depressing effect on stock prices. One difficulty with this argument is that it requires not only a potential catastrophe, but one which affects stock market investors more seriously than investors in short-term debt instruments. Many countries that have experienced catastrophes, such as Russia or Germany, have seen very low returns on short-term government debt as well as on equity. A peso problem that affects both asset returns equally will affect estimates of the average levels of returns but not estimates of the equity premium 9. The major example of a disaster for stockholders that did not negatively affect bondholders is the Great Depression of the early 1930s, but of course this is included in the long-run annual data for Sweden, the UK, and the USA, all of which display an equity premium puzzle. Also, the consistency of the results across countries requires investors in all countries to be concerned about catastrophes. If the potential catastrophes are uncorrelated across countries, then it becomes less likely that the data set includes no catastrophes; thus the argument seems to require a potential international catastrophe that affects all countries simultaneously. 3.3. The riskf?ee rate puzzle One response to the equity premium puzzle is to consider larger values for the coefficient of relative risk aversion ~/. Kandel and Stambaugh (1991) have advocated 9 Thispoint is relevant for the studyof Goetzmann and Jorion (1997). These authors measure average growth rates &real stock prices, as a proxy for real stock returns, bnt they do not look at real returns on short-term debt. Theyfind low real stock-pricegrowth rates in many countriesin the early 20th Century; in somecases these may havebeen accompaniedby low returns to holders of short-term debt. Note also that stock-price growth rates are a poor proxy for total stock returns in periods where investors expect low growth rates, since dividend yields will tend to be higher in such periods. Ch. 19: Asset Prices, Consumption, and the Business Cycle 1253 this l°. However this leads to a second puzzle. Equation (15) implies that the unconditional mean riskless interest rate is y2~ (17)Er];t+l =-log6+gg- 2 ' where g is the mean growth rate of consumption. Since g is positive, as shown in Table 3, high values of 7 imply high values of 7g. Ignoring the term -y2o~2/2 for the moment, this can be reconciled with low average short-term real interest rates, shown in Table 2, only if the discount factor 6 is close to or even greater than one, corresponding to a low or even negative rate of time preference. This is the riskfree rate puzzle emphasized by Weil (1989). Intuitively, the riskfree rate puzzle is that if investors are risk-averse then with power utility they must also be extremely unwilling to substitute intertemporally. Given positive average consumption growth, a low riskless interest rate and a high rate of time preference, such investors would have a strong desire to borrow from the future to reduce their average consumption growth rate. A low riskless interest rate is possible in equilibrium only if investors have a low or negative rate of time preference that reduces their desire to borrow 11 Of course, if the risk aversion coefficient g is high enough then the negative quadratic term -V2a~/2 in Equation (17) dominates the linear term and pushes the riskless interest rate down again. The quadratic term reflects precautionary savings; risk-averse agents with uncertain consumption streams have a precautionary desire to save, which can work against their desire to borrow. But a reasonable rate of time preference is obtained only as a knife-edge case. Table 6 illustrates the riskfree rate puzzle in international data. The table first shows the average riskfree rate from Table 2 and the mean consumption growth rate and standard deviation of consumption growth from Table 3. These moments and the risk aversion coefficients calculated in Table 5 are substituted into Equation (17), and the equation is solved for an implied time preference rate. The time preference rate is reported in percentage points per year; it can be interpreted as the riskless real interest rate that would prevail if consumption were known to be constant forever at its current level, with no growth and no volatility. Risk aversion coefficients in the RRA(2) range imply negative time preference rates in every country except Switzerland, whereas larger risk aversion coefficients in the RRA(I) range imply time preference rates that are often positive but always implausible and vary wildly across countries. J0 One might think that introspectionwould be sufficientto rule out very largevalues of V,but Kandel and Stambaugh (1991) point out that introspection can deliver very different estimatesof risk aversion depending on the size of the gamble considered. This suggests that introspection can be misleading or that some more general model of utility is needed. I~ As Abel (1999) and Kocherlakota(1996) point out,negative time preference is consistentwith finite utility in a time-separable model provided that consumption is growing, and marginal utility shrinking, sufficientlyrapidly.The question is whether negative thne preference is plausible. 1254 J.Y. Campbell Table 6 The riskfree rate puzzlea Country Sample period r~ Ac o(Ae) RRA(1) TPR(1) RRA(2) TPR(2) USA 1947.2-1996.3 0.794 1.908 1.084 246.556 112.474 47.600 -76.710 AUL 1970.1 1996.2 1.820 1.854 2.142 45.704 -34.995 7.107 -10.196 CAN 1970.1-1996.2 2.738 1.948 2.034 56.434 41.346 8.965 -13.066 FR 1973.~1996.2 2.736 1.581 2.130 < 0 N/A 14.634 -15.536 GER 1978.4-1996.2 3.338 1.576 2.495 343.133 >1000 13.327 12.142 ITA 1971.2 1995.2 2.064 2.424 1.684 >1000 >1000 4.703 -9.021 JPN 1970.2-1996.2 1.538 3.416 2.353 134.118 41.222 13.440 -39.375 NTH 1977.2-1996.1 3.705 1.466 2.654 >1000 >1000 23.970 -11.201 SWD 1970.1 1994.4 1.520 0.750 1.917 >1000 >1000 20.705 -6.126 SWT 1982.2-1996.2 1.466 0.414 2.261 < 0 N/A 26.785 8.698 UK 1970.1 1996.2 1.08i 2.025 2.589 156.308 503.692 14.858 -21.600 USA 1970.1-1996.3 1.350 1.710 0.919 150.136 -160.275 37.255 -56.505 SWD 1920 1993 2.073 1.748 2.862 65.642 63.778 11.091 -12.274 UK 1919-1993 1.198 1.358 2.820 39.914 10.364 14.174 -10.057 USA 1891-1994 1.955 1.742 3.257 20.861 11.305 10.366 10.406 a ~: is the mean money market return from Table 2, in annualized percentage points. Ae and cr(Ae) are the mean and standard deviation of consmnption growth from Table 3, in annualized percentage points. RRA(1) and RRA(2) are the risk aversion coefficients from Table 5. TPR(1)=7- RRA(1)Ac + RRA(1)2g2(Ac)/200, and TPR(2)= ~- RRA(2)Ac + RRA(2)2oZ(Ac)/200.From Equation (17), these time preference rates give the real interest rate, in annualized percentage points, that would prevail if consumption growth had zero mean and zero standard deviation and risk aversion were RRA(1) or RRA(2), respectively. Abbreviations: AUL, Australia; CAN, Canada; FR, France; GER, Germany; ITA, Italy; JPN, Japan; NTH, Netherlands; SWD, Sweden; SWT, Switzerland; UK, United Kingdom; USA, United States of America. An interesting issue is how mismeasurement of average inflation might affect these calculations. There is a growing consensus that in recent years conventional price indices have overstated true inflation by failing to fully capture the effects of quality improvements, consumer substitution to cheaper retail outlets, and price declines in newly introduced goods. If inflation is overstated by, say, 1%, the real interest rate is understated by 1%, which by itself might help to explain the riskfree rate puzzle. Unfortunately the real growth rate of consumption is also understated by 1%, which worsens the riskfree rate puzzle. When y > 1, this second effect dominates and understated inflation makes the riskfree rate puzzle even harder to explain. Ch. 19." Asset Prices, Consumption, and the Business Cycle 1255 Table 7 International yield spreads and bond excess returns a Country Sample period ~ a(s) p(s) er~ a(erb) p(erb) USA 1947.2-1996.4 1.199 0.999 0.783 0.011 8.923 0.070 AUL 1970.1-1996.3 0.938 1.669 0.750 0.156 8.602 0.162 CAN 1970.1 1996.3 1.057 1.651 0.819 0.950 9.334 -0.009 FR 1973.2 1996.3 0.917 1.547 0.733 1.440 8.158 0.298 GER 1978.4-1996.3 0.99l 1.502 0.869 0.899 7.434 0.117 ITA 197t.~1995.3 0.200 2.025 0.759 1.386 9.493 0.335 JPN 1970.2-1996.3 0.593 1.488 0.843 1.687 9.165 -0.058 NTH 1977.2-1996.2 1.212 1.789 0.574 1549 7.996 0.032 SWD 1970.1-1995.1 0.930 2.046 0.724 0.212 7.575 0.244 SWT 1982.2 1996.3 0.471 1.655 0.755 1.071 6.572 0.268 UK 1970.1 1996.3 1.202 2.106 0.893 0.959 11.611 0.057 USA 1970.1-1996.4 1.562 1.190 0.737 1.504 10.703 0.033 SWD 1920-1994 0.284 1.140 0.280 -0.075 6.974 0.185 UK 1919-1994 1.272 1.505 0.694 0.318 8.812 -0.098 USA 1891 1995 0.720 1.550 0.592 0.172 6.499 0.153 a Sis the mean of the log yield spread, the difference between the log yield on long-term bonds and the log 3-month money market return, expressed in annualized percentage points. ~7(s)is the standard deviation of the log yield spread and p(s) is its first-order autocorrelation, erh, a(ert,), and p(erb) are defined in the same way for the excess 3-month return on long-term bonds over money market instruments, where the bond return is calculated from the bond yield using the par-bond approximation given in Campbell, Lo and MacKinlay (1997), Chapter 10, equation (10.1.I9). Full details of this calculation are given in thc Data Appendix. Abbreviations: AUL, Australia; CAN, Canada; FR, France; GER, Germany; ITA, Italy; JPN, Japan; NTH, Netherlands; SWD, Sweden; SWT, Switzerland; UK, United Kingdom; USA, United States of America. 3.4. Bond returns and the equity premium and riskfree rate puzzles Some authors have argued that the riskfree interest rate is low because short-term government debt is more liquid than long-term financial assets. Short-term debt is "moneylike" in that it facilitates transactions and can be traded at minimal cost. The liquidity advantage of debt reduces its equilibrium return and increases the equity premium [Bansal and Coleman (1996), Heaton and Lucas (1996)]. The difficulty with this argument is that it implies that all long-term assets should have large excess returns over short-term debt. Long-term government bonds, for example, are not moneylike and so the liquidity argument implies that they should offer a large term premium. But historically, the term premium has been many times smaller than the equity premium. This point is illustrated in Table 7, which reports two 1256 Jg Campbell alternative measures of the term premium. The first measure is the average log yield spread on long-term bonds over the short-term interest rate, while the second is the average quarterly excess log return on long bonds. In a long enough sample these two averages should coincide if there is no upward or downward drift in interest rates. The average yield spread is typically between 0.5% and 1.5%. A notable outlier is Italy, which has a negative average yield spread in this period. Average long bond returns are quite variable across countries, reflecting differences in inflationary experiences; however in no country does the average excess bond return exceed 1.7% per year. Thus both measures suggest that term premia are far smaller than equity premia. Table 8 develops this point further by repeating the calculations of Table 5, using bond returns rather than equity returns. The average excess log return on bonds over short debt, adjusted for Jensen's Inequality, is divided by the standard deviation of the excess bond return to calculate a bond Sharpe ratio which is a lower bound on the standard deviation of the stochastic discount factor. The Sharpe ratio for bonds is several times smaller than the Sharpe ratio for equities, indicating that term premia are small even after taking account of the lower volatility of bond returns. This finding is not consistent with a strong liquidity effect at the short end of the term structure, but it is consistent with a consumption-based asset pricing model if bond returns have a low correlation with consumption growth. Table 8 shows that sample consumption correlations often are lower for bonds, so that RRA(1) risk aversion estimates for bonds, which use these correlations, are often comparable to those for equities. A direct test of the liquidity story is to measure excess returns on stocks over long bonds, rather than over short debt. If the equity premium is due to a liquidity effect on short-term interest rates, then there should be no "equity-bond premium" puzzle. Table 9 carries out this exercise and finds that the equity-bond premium puzzle is just as severe as the standard equity premium puzzle 12. 3.5. Separating risk aversion and intertemporal substitution Epstein and Zin (1989, 1991) and Weil (1989) use the theoretical framework of Kreps and Porteus (1978) to develop a more flexible version of the basic power utility model. That model is restrictive in that it makes the elasticity of intertemporal substitution, % the reciprocal of the coefficient of relative risk aversion, 7. Yet it is not clear that these two concepts should be linked so tightly. Risk aversion describes the consumer's reluctance to substitute consumption across states of the world and is meaningful even 12 The excess return of equities over bonds must be measured with the appropriatecorrection for Jensen's Inequality.From Equation (16), the appropriatemeasure is the log excessreturn on equities overshort-termdebt,lessthe log excessreturn on bondsovershort-termdebt,plus one-halfthe variance of the log equityreturn, less one-halfthe varianceof the log bond return. Ch, 19.• Asset Prices, Consumption, and the Business Cycle % 0 V ~,, ~ . , ~ ,.n ~ ~ ~. ~.. I • " ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ 2 ~, r¢3 o ~ t'4 V V t-'q ~ t"~ ~ ~ rz t'-q ,,6 ~6 ,,6 1257 •~ z ,.~ Z t r~ 1258 JK Campbell % ~o <~ ~2 o o ~ V V ~ oo ~ . . ~ ~ . ~ . . ¢S"~ ~ o O9 ~ , ~ ~~ ~ L~ .~q i~~ _ ~ °<% ~ ~ J ~.~ "~'" I!~ ~ "~ < -~ ~.~ ~ o ~ ,~ ~.~ Ch. 19: Asset Prices, Consumption, and the Business Cycle 1259 in an atemporal setting, whereas the elasticity of intertemporal substitution describes the consumer's willingness to substitute consumption over time and is meaningful even in a deterministic setting. The Epstein-Zin--Weil model retains many of the attractive features of power utility but breaks the link between the parameters y and ~p. The Epstein-Zin-Weil objective function is defined recursively by 0 U,= (1-6)C 7+6 G_.f\ , (18) where 0 = (1 -y)/(l - l/W). When y - I/W, 0 = 1 and Equation (18) becomes linear; it can then be solved forward to yield the familiar time-separable power utility model. The integemporal budget constraint for a represemative agent can be written as V/t+1 - (1 + Rw,t+l) (J4zt- - CI), (19) where Wt+l is the representative agent's wealth, and (1 + Rw,t+l) is the gross simple return on the portfolio of all invested wealth 13. This form of the budget constraint is appropriate for a complete-markets model in which wealth includes human capital as well as financial assets. Epstein and Zin use dynamic programming arguments to show that Equations (18) and (19) together imply an Euler equation of the form 1I=G \CT-t / (1 +Rw, t<) (1 +R,,,+I) . (20) If I assume that asset returns and consumption are homoskedastic and jointly lognormal, then this implies that the riskless real interest rate is 0-1 2 0 2 rj;t+l = -log6+ E,[Act+l] + ~2-- cG - ~2 o~. (21) The riskless interest rate is a constant, plus 1/~p times expected consumption growth. In the power utility model, 1/ip = y and 0 = 1, so Equation (21) reduces to Equation (15). The premium on risky assets, including the wealth portfolio itself, is 62 o,, E,[ri,,+l] - rj;,+l + " - 0 +(1 - O)oiw. (22) 2 ~0 The risk premium on asset i is a weighted combination of asset i's covariance with consumption growth (divided by the elasticity of intertemporal substitution W) and 13 This is often called the "market" return and written Rm,t~i, but l have already used m to denote the stochastic discount factor so I write R,,,t~l to avoid confusion. 1260 J.Y. Campbell asset i's covariance with the return on wealth. The weights are 0 and 1 - 0 respectively. The Epstein-Zin-Weil model thus nests the consumption CAPM with power utility (0 = 1) and the traditional static CAPM (0 = 0). Equations (21) and (22) seem to indicate that Epstein-Zin-Weil utility might be helpful in resolving the equity premium and riskfree rate puzzles. First, in Equation (21) a high risk aversion coefficient does not necessarily imply a low average riskfree rate, because 0--1 2 0 Erj;t+l =-log6+ g + ~rw- or2. (23) The average consumption growth rate is divided by ~p here, and in the Epstein-ZinWell framework ~pneed not be small even if ~ is large. Second, Equation (22) suggests that it might not even be necessary to have a high risk aversion coefficient to explain the equity premium. If 0 ~ 1, then the risk premium on an asset is determined in part by its covariance with the wealth portfolio, a/w. If the return on wealth is more volatile than consumption growth, as implied by the common use of a stock index return as a proxy for the return on wealth, then Oiw may be much larger than oic, and this may help to explain the equity premium. Unfortunately, there are serious difficulties with both these potential escape routes from the equity premium and riskfree rate puzzles. The difficulty with the first is that there is direct empirical evidence for a low elasticity of intertemporal substitution in consumption. The difficulty with the second is that consumption and wealth are linked through the intertemporal budget constraint; if consumption is smooth and wealth is volatile, this itself is a puzzle that must be explained, not an exogenous fact that can be used to resolve other puzzles. I now develop these points in detail by analyzing the dynamic behavior of stock returns and short-term interest rates in relation to consumption. 4. The dynamics of asset returns and consumption 4.1. Time-variation in conditional expectations Equations (21) and (22) imply a tight link between rational expectations of asset returns and of consumption growth. Expected asset returns are perfectly correlated with expected consumption growth, with a standard deviation 1/~p times as large. Equivalently, the standard deviation of expected consumption growth is ~p times as large as the standard deviation of expected asset returns. Ch. 19: Asset Prices', Consumption, and the Business Cycle 1261 This suggests a way to estimate % Hansen and Singleton (1983), followed by Campbell and Mankiw (1989), Hall (1988), and others, have proposed an instrumental variables (IV) regression approach. If we define an error term t/i,t+l ~ ri, t+l - - Et[ri, t+l] - 7(Act+t - Et[ACt+l]), then we can rewrite Equations (21) and (22) as a regression equation, (24) In general the error term t/i,t+l will be correlated with realized consumption growth so OLS is not an appropriate estimation method. However t/i,t+l is uncorrelated with any variables in the information set at time t. Hence any lagged variables correlated with asset returns can be used as instruments in an IV regression to estimate 1/% Table 10 illustrates two-stage least squares estimation of Equation (24). In each panel the first set of results uses the short-term real interest rate, while the second set uses the real stock return. The instruments are the asset return, the consumption growth rate, and the log price-dividend ratio. The instruments are lagged twice to avoid difficulties caused by time-aggregation of the consumption data ]Campbell and Mankiw (1989, 1991), Wheatley (1988)]. For each asset and set of instruments, the table first reports the R2 statistics and significance levels for first-stage regressions of the asset return and consumption growth rate onto the instruments. The table then shows the IV estimate of 1/~p with its standard error, and (in the column headed "Test (1)") the R2 statistic for a regression of the residual on the instruments together with the associated significance level of a test of the over-identifying restrictions of the model. The quarterly results in Table 10 show that the short-term real interest rate is highly forecastable in every country except Germany. The real stock return is also forecastable in many countries, but there is weaker evidence for forecastability in consumption growth. In fact the R2 statistic for forecasting consumption growth is lower than the R2 statistic for stock returns in all but four of the quarterly data sets. The IV estimates of 1/~p are very imprecise; they are sometimes large and positive, often negative, but they are almost never significantly different from zero. The overidentifying restrictions of the model are often strongly rejected, particularly when the short-term interest rate is used in the model. Results are similar for the annual data sets in Table 10, except that twice-lagged instruments have almost no ability to forecast real interest rates or stock returns in the annual US data 14 14 Campbell, Lo and MacKinlay (1997), Table 8.2, showsmuch greater fbrecastabilityof returns using once-laggedinstruments in a similarannual US data set. Even with twice-laggedhlstruments, US annual returns become forecastableonce one increasesthe return horizon beyond one year,as shownin Table12 below. 1262 Table 10 Predictable variation in returns and consumption growth a J Y. Campbell Count~2¢ Sample period Asset First-stage (1/~--~) regressions (s.e.) (s.e.) ri Ac Testb 1 2 USA AUL CAN FR GER ITA JPN NTH SWD 1947.~1996.3 rf 0.160 0.000 re 0.065 0.003 1970.2 1996.2 rf 0.404 0.000 r,, 0.060 0.034 1970.2 1996.2 rf 0.292 0.000 r~ 0.040 0.269 1973.2-1996.2 r! 0.519 0.000 r~ 0.111 0.006 1978.4-1996.2 1)- 0.062 0.328 r~ 0.046 0.050 1971.2-1995.2 rj 0.405 0.000 re 0.048 0.278 1970.~1996.2 rf 0.203 0.002 r~ 0.115 0,001 1977.2-1996.1 rj 0.248 0.000 re 0.021 0.756 1970.2 1994.4 rj 0.262 0.000 r~, 0.110 0.039 0.037 0.260 0.025 0.165 0.037 0.077 0.740 0.114 0.000 0.027 0.037 -8.187 0.021 0.035 0.025 0.077 7.069 0.028 0.033 0.090 0.013 4.450 0.099 0.017 0,008 0.432 2.973 0.107 0.419 0,676 0.013 20.250 0.038 0.004 0.003 0.432 13.145 0.026 0.828 0.856 0.048 -0.970 -0.174 0.142 0.041 0.042 0.677 0.177 0.001 0.123 0.048 6.635 0.130 0.004 0.004 0,042 4.536 0.092 0.822 0.819 0.010 -2.189 -0.051 0.073 0.009 0.751 2.170 0.133 0.037 0,667 0.010 -27.662 -0.021 0.006 0.004 0.751 29.994 0.026 0,750 0.833 0.057 0.481 1.773 0.005 0.005 0.085 0.354 1.141 0.840 0.841 0.057 -6.117 0.079 0.017 0.018 0.085 4.992 0.066 0.569 0.547 0.010 -2.432 -0.019 0.171 0.010 0.877 3.353 0.113 0.000 0,624 0.010 19.919 0.016 0.013 0.007 0.877 26.244 0.034 0.540 0.734 0.044 -0.446 -0.093 0.162 0.04 l 0.081 0.464 0.266 0.000 0.121 0.044 11.028 0.047 0.026 0.019 0.081 5.458 0.027 0.260 0.376 0.024 0,167 0.052 0.218 0.023 0.373 0.385 0.428 0,000 0.428 0.024 -4.532 -0.138 0.005 0,005 0.373 6.571 0.162 0,835 0.832 0.005 -1.056 -0.007 0.197 0.005 0.806 2,949 0.085 0000 0,779 0.005 15.210 0.004 0.047 0.005 0.806 21.187 0.017 0.107 0.790 continued on next page Ch. 19." Asset Prices, Consumption, and the Business Cycle 1263 Table 10, continued Country Sample period Asset First-stage (1/~-~) ~ Test b regressions (s.e.) (s.e.) 1 2 ri Ac SWT 1982.2 1996.2 rf 0.194 0.007 0.731 0.065 0.074 0.006 0.000 0.887 1.273 0.397 0.136 0.844 re 0.033 0.007 20.084 0.048 0.000 0.000 0.270 0.887 31.100 0.070 0.996 0.996 UK 1970.~1996.2 rf 0.306 0.057 1.992 0.260 0.047 0.028 0.000 0.042 0.988 0.136 0.090 0.238 re 0.097 0.057 -4.493 0.038 0.056 0.040 0.094 0.042 3.793 0.034 0.058 0.132 USA 1970.2-1996.3 J) 0.307 0.071 1.573 0.t02 0.188 0.062 0.000 0.015 0.704 0.111 0.000 0.041 rC 0.069 0.071 4.977 0.016 0.069 0.071 0.095 0.015 7.677 0.023 0.029 0.025 SWD 1920-1993 rf 0.302 0.052 2.740 0.194 0.037 0.023 0.000 0.202 1.466 0.t61 0.266 0.437 r~ 0.041 0.052 -1.537 0.043 0.034 0.041 0.342 0.202 3.349 0.082 0.304 0.236 UK 1920-1993 r~/ 0.265 0.061 2.499 0.197 0.056 0.033 0.000 0.140 1.509 0.123 0.139 0.314 r~, 0.147 0.061 5.861 0.037 0.115 0.055 0.096 0.140 4.569 0.021 0.017 0.144 USA 1891-1994 ~/ 0.013 0.065 -0.293 -0.202 0.012 0.049 0.783 0.004 0.892 0.341 0.552 0.085 r,, 0.037 0.065 0.723 0.038 0.040 0.074 0.184 0.004 2.003 0.070 0.132 0.024 a This table reports two-stage least squares eshination results for Equations (24) and (25). The first set of results for each country uses the short-term real interest rate, while the second set uses the real stock return. The instruments are the asset return, the consumption growth rate, and the log price-dividend ratio, lagged twice. For each asset and set of instruments, the first two colunms show the R2 statistics, with significance levels below, lbr first-stage regressions of the asset return and consumption growth rate onto the instruments. The third column shows the IV estimate of 1/~p from Equation (24) with its standard error below, and the fourth column shows the IV estimate of ~p from Equation (25) with its standard error below. The fifth column, headed "Test (1)", shows the R2 statistic for a regression of the residual from Equation (24) on the instruments, with the associated significance level below of a test of the over-identifying restrictions of the model. The sixth column, headed "Test (2)" is the equivalent of the fifth column for Equation (25). Abbreviations: AUL, Australia; CAN, Canada; FR, France; GER, Germany; ITA, Italy; JPN, Japan; NTH, Netherlands; SWD, Sweden; SWT, Switzerland; UK, United Kingdom; USA, United States of America. b Tests: (1)ri,t~ j =t,ti-t-(1/~O)Act+t+rli,t+l; (2) Act+t - Ti + ~/)1~,1~1 +~z,t41. 1264 JE Campbell Campbell and Mankiw (1989, 1991) have explored this regression in more detail, using both US and international data, and have found that predictable variation in consumption growth is often associated with predictable variation in income growth. This suggests that some consumers keep their consumption close to their income, either because they follow "rules of thumb", or because they are liquidity-constrained, or because they are "buffer-stock" savers [Deaton (1991), Carroll (1992)]. After controlling for the effect of predictable income growth, there is little remaining predictable variation in consumption growth to be explained by consumers' response to variation in real interest rates. One problem with IV estimation of Equation (24) is that the instruments are only very weakly correlated with the regressor because consumption growth is hard to forecast in this data set. Nelson and Startz (1990) have shown that in this situation asymptotic theory can be a poor guide to inference in finite samples; the asymptotic standard error of the coefficient tends to be too small and the overidentifying restrictions of the model may be rejected even when it is true. To circumvent this problem, one can reverse the regression (24) and estimate Act+ 1 = ~ -t- ~gri, t+ 1 + ~i,t+l. (25) If the orthogonality conditions hold, then the estimate of 'qJ in Equation (25) will asymptotically be the reciprocal of the estimate of 1/~p in Equation (24). In a finite sample, however, if ~p is small then IV estimates of Equation (25) will be better behaved than IV estimates of Equation (24). In Table 7 ~p is almost always estimated to be close to zero. The estimates are much more precise than those for 1/% The overidentifying restrictions of the model are sometimes rejected, but less often and less strongly than when Equation (24) is estimated. These results suggest that the elasticity of intertemporal substitution ~p is small, so that the generality of the Epstein-Zin-Weil model, which allows ~p to be large even if ~/is large, does not actually help one fit the data on consumption and asset returns 15. 4.2. A loglinear asset pricing J?amework in order to understand the second momems of stock returns, it is essential to have a framework relating movements in stock prices to movements in expected future dividends and discount rates. The present value model of stock prices is intractably nonlinear when expected stock returns are time-varying, and this has forced researchers to use one of several available simplifying assumptions. The most common approach is to assume a discrete-state Markov process either for dividend growth [Mehra and 15 Attanasioand Weber(1993) and Beaudryand van Wincoop(1996)havearguedthatthis conclusion depends on the use of aggregateconsumptiondata. Theywork with cohort-leveland state-leveldata, respectively,and fred some evidencefor a larger elasticityof intertemporalsubstitution. Ch. 19: Asset Prices', Consumption, and the Business Cycle 1265 Prescott (1985)] or, following Hamilton (1989), for conditionally expected dividend growth [Abel (1994, 1999), Cecchetti, Lam and Mark (1990, 1993), Kandel and Stambaugh (1991)]. The Markov structure makes it possible to solve the present value model, but the derived expressions for returns tend to be extremely complicated and so these papers usually emphasize numerical results derived under specific numerical assumptions about parameter values 16. An alternative framework, which produces simpler closed-form expressions and hence is better suited for an overview of the literature, is the loglinear approximation to the exact present value model suggested by Campbell and Shiller (1988). Campbell and Shiller's loglinear relation between prices, dividends, and returns provides an accounting framework: High prices must eventually be followed by high future dividends or low future returns, and high prices must be associated with high expected future dividends or low expected future returns. Similarly, high returns must be associated with upward revisions in expected future dividends or downward revisions in expected future returns. The loglinear approximation starts with the definition of the log return on some asset i, ri, t+l ~ log(Pi, t+l + Di, t+l) - log(Pit). The timing convention here is that prices are measured at the end of each period so that they represent claims to next period's dividends. The log return is a nonlinear function of log prices Pit and pi, t+l and log dividends di, t+l, but it can be approximated around the mean log dividend-price ratio, (dit -Pa), using a first-order Taylor expansion. The resulting approximation is ri, t+l ~ k +/)Pi, t tl +(1 -/))di, t+l-Pit, (26) where/) and k are parameters of linearization defined by/) = 1/(1 + exp(~)) and k - log(/)) - (1 -/))log(l//) - 1). When the dividend-price ratio is constant, then p = Pi/(Pi + Di), the ratio of the ex-dividend to the cum-dividend stock price. In the postwar quarterly US data shown in Table 3, the average price-dividend ratio has been 26.4 on an annual basis, implying that/) should be about 0.964 in annual data 17 The Taylor approximation (26) replaces the log of the sum of the stock price and the dividend in the exact relation with a weighted average of the log stock price and the log dividend. The log stock price gets a weight/) close to one, while the log dividend gets a weight 1 -p close to zero because the dividend is on average much smaller than the stock price, so a given percentage change in the dividend has a much smaller effect on the return than a given percentage change in the price. Ic, A partial exceptionto thisstatementis that Abel (1994)derivesseveralanalyticalresultsfor the first momentsof retarns in a Markovmodel for expecteddividendgrowth. 17 Strictly speakingbothp and k should have asset subscripts i, but 1 omit these for simplicity.The assetpricingformulaslater in this chapterassumethat all assetshavethe samep, whichsimplifiessome expressionsbut doesnot changeany of the qualitativeconclusions. 1266 JE Campbe# Equation (26) is a linear difference equation for the log stock price. Solving forward, imposing the terminal condition that limj~o~ PJPi, t+j = 0, taking expectations, and subtracting the current dividend, one gets 0<3 pit-d#- k kE~Zp.i[Adi,~+l+j-ri,,+l+j]. (27) 1 -p .i=o This equation says that the log price-dividend ratio is high when dividends are expected to grow rapidly, or when stock returns are expected to be low. The equation should be thought of as an accounting identity rather than a behavioral model; it has been obtained merely by approximating an identity, solving forward subject to a terminal condition, and taking expectations. Intuitively, if the stock price is high today, then from the definition of the return and the terminal condition that the stock price is non-explosive, there must either be high dividends or low stock returns in the future. hwestors must then expect some combination of high dividends and low stock returns if their expectations are to be consistent with the observed price. The terminal condition used to obtain Equation (27) is perhaps controversial. Models of "rational bubbles" do not impose this condition. Blanchard and Watson (1982) and Froot and Obstfeld (1991) have proposed simple, explicit models of explosive bubbles in asset prices. There are however several reasons to rule out such bubbles. The theoretical circumstances under which bubbles can exist are quite restrictive; Tirole (1985), for example, uses an overlapping generations framework and finds that bubbles can only exist if the economy is dynamically inefficient, a condition which seems unlikely on prior grounds and which is hard to reconcile with the empirical evidence of Abel, Mankiw, Summers and Zeckhauser (1989). Santos and Woodford (1997) also conclude that the conditions under which bubbles can exist are fragile. Empirically, bubbles imply explosive behavior of prices in relation to dividends and other measures of fundamentals; there is no evidence of this, although nonlinear bubble models are hard to reject using standard linear econometric methods is Equation (27) describes the log price-dividend ratio rather than the log price itself. This is a useful way to write the model because in many data sets dividends appear to follow a loglinear unit root process, so that log dividends and log prices are nonstationary. In this case changes in log dividends are stationary, so from Equation (27) the log price-dividend ratio is stationary provided that the expected stock return is stationary. Thus log stock prices and dividends are cointegrated, and the stationary linear combination of these variables involves no unknown parameters since it is just the log ratio. Table 11 reports some summary statistics for international stock prices in relation to dividends. The table gives the average price-dividend ratio, the standard deviation t8 Campbell,Lo and MacKinlay(1997),Chapter7, givesa somewhatmoredetailedtextbookdiscussion of the literatureon rationalbubbles. Ch. 19." Asset Prices, Consumption, and the Business Cycle 1267 Table 11 International stock prices and dividends a Country Sample period P/D a(p-d) p(p d) ADF(1) Ap Ad Ap-d USA 1947.~1996.4 27.121 0.265 0.941 -1.752 3.547 2.225 1.688 AUL 1970.1 1996.3 25.919 0.267 0.856 3.273 -1.410 0.883 -2.477 CAN 1970.1-1996.3 30.108 0.221 0.902 -1.900 0.754 -0.741 1.200 FR 1973.2-1996.3 22.718 0.541 0.971 -1.3t0 1.358 -1.214 2.538 GER 1978.4-1996.3 27.787 0.300 0.922 -1.660 4.186 1.079 3.853 ITA 1971.2-1995.3 41.345 0.318 0.882 -3.743 2.172 4.919 3.531 JPN 1970.2-1996.3 91.251 0.642 0.964 -1.574 4.192 2.489 6.974 NTH 1977.2 1996.2 21.139 0.272 0.932 -0.727 7.540 4.007 3.637 SP 1984.2 1996.2 22.509 0.319 0.823 -3.075 6.843 -3.086 10.078 SWD 1970.1-1995.1 35.021 0.439 0.941 -1.632 4.922 1.861 3.499 SWT 1982.2-1996.3 47.320 0.217 0.814 -1.588 9.291 4.143 6.074 UK 1970.1 1996.3 18.434 0.280 0.913 -1.657 1.464 0.681 0.579 USA 1970.1 1996.4 27.882 0.235 0.904 -1.372 2.034 0.619 1.582 SWD 1920-1994 26.706 0.333 0.746 0.768 2.129 0.423 2.054 UK 1919 1994 20.806 0.238 0.514 4.093 2.064 1.844 0.220 USA 1891-1995 22.733 0.279 0.778 -1.868 2.064 1.485 0.477 a P/D is the mean price-dividendratio, c~(p-d) is the standard deviation of the log price-dividendratio in natural units (not annualizedpercentage points), p(p - d) is the first-order autocorrelation of the log price-dividend ratio. ADF(1) is the augmented Dickey-Fuller t-ratio for the lagged log price--dividend ratio when the change in the log price-dividend ratio is regressed on a constant, four lagged changes, and the lagged log price dividend ratio. Ap, Ad, and Ap- d are the mean changes in log prices, log dividends, and the log price-dividend ratio respectively, in annualizedpercentage points. Abbreviations: AUL, Australia; CAN, Canada; FR, France; GER, Germany; ITA, Italy; JPN, Japan; NTH, Netherlands; SWD, Sweden; SWT, Switzerland; UK, United Kingdom; USA, United States of America. of the log price-dividend ratio in natural units, the first-order autocorrelation of the log price-dividend ratio, average growth rates of prices, dividends, and the log pricedividend ratio in percentage points per year, and a test statistic for the null hypothesis that the log price-dividend ratio has a unit root. Following standard practice, the pricedividend ratio is measured as the ratio of the current stock price to the total of dividends paid during the past year. Average price-dividend ratios vary considerably across countries but generally lie between 20 and 30. The extreme outlier is Japan, which has an average price-dividend ratio of 91. The volatility and first-order autocorrelation of the log price-dividend ratio are also unusually high for Japan, reflecting an upward trend in the Japanese log price- 1268 J.Y. Campbell dividend ratio for much of the sample period which is also visible in the average growth rates of prices and dividends at the right of the table. Other countries in the quarterly data set, with the exception of France, have firstorder autocorrelation coefficients for the log price-dividend ratio of between 0.85 and 0.95. Unit root tests do not reject the unit root null hypothesis for most of these countries, but this may reflect low power of the tests in short data samples. Equation (27) implies that the log price-dividend ratio must be stationary if real dividend growth and stock returns are stationary, so this gives some reason to assume stationarity for the series. So far I have written asset prices as linear combinations of expected future dividends and returns. Following Campbell (1991), I can also write asset returns as linear combinations of revisions in expected future dividends and returns. Substituting Equation (27) into Equation (26), I obtain OO OO ri, l+l -Et ri, t+l = (Et+l -Et) ~PJA~,t,1 i:j- (Et ~l -Et) ~ pJri, t+l+j. .i- o j - 1 (28) This equation says that unexpected stock returns must be associated with changes in expectations of future dividends or real returns. An increase in expected future dividends is associated with a capital gain today, while an increase in expected future returns is associated with a capital loss today. The reason is that with a given dividend stream, higher future returns can only be generated by future price appreciation from a lower current price. 4.3. The stock market oolatility puzzle I now use this accounting framework to illustrate the stock market volatility puzzle. The intertemporal budget constraint for a representative agent, Equation (19), implies that aggregate consumption is the dividend on the portfolio of all invested wealth, denoted by subscript w: dwt = ct. (29) Many authors, including Grossman and Shiller (1981), Lucas (1978), and Mehra and Prescott (1985), have assumed that the aggregate stock market, denoted by subscript e for equity, is equivalent to the wealth portfolio and thus pays consumption as its dividend. Here I follow Campbell (1986) and Abel (1999) and make the slightly more general assumption that the dividend on equity equals aggregate consumption raised to a power )~. In logs, we have det - ,~ct. (30) Abel (1999) shows that the coefficient )~ can be interpreted as a measure of leverage. When )~ > 1, dividends and stock returns are more volatile than the returns on the Ch. 19." AssetPrices, Consumption,and the BusinessCycle 1269 aggregate wealth portfolio. This framework has the additional advantage that a riskless real bond with infinite maturity - an inflation-indexed consol, denoted by subscript b can be priced merely by setting )~= 0. The representative-agent asset pricing model with Epstein-Zin-Weil utility, conditional lognormality, and homoskedasticity [Equations (21) and (22)] implies that Etre,t+l=~e+(@)EtAct+l, (31) where g~ is an asset-specific constant term. The expected log return on equity, like the expected log return on any other asset, is just a constant plus expected consumption growth divided by the elasticity of intertemporal substitution % Power utility is the special case where the coefficient of relative risk aversion y is the reciprocal of ~p so the effect of expected consumption growth on expected asset returns is proportional to y; but this is not true in general. Substituting Equations (30) and (31) into Equations (27) and (28), I find that ~-t- )~--~)) Et ZpJmct+l±f, j=0 (32) and re,t+l-Et re,t+l = Z(Act+l -E/AG+I)+ Z - (Et+l -Et) Zp/Act+I+j. j=[ (33) Expected future consumption growth has offsetting effects on the log price-dividend ratio. It has a direct positive effect by increasing expected future dividends X-forone, but it has an indirect negative effect by increasing expected future real interest rates (1/~0)-for-one. The unexpected log return on the stock market is X times contemporaneous unexpected consumption growth (since contemporaneous consumption growth increases the contemporaneous dividend X-for-one), plus (3,- 1/~p) times the discounted sum of revisions in expected future consumption growth. For future reference I note that Equation (33) can be inverted to express consumption growth as a function of tile unexpected return on equity and revisions in expectations about future returns on equity. Rearranging Equation (33) and using Equation (31), ACt+l - Et Act+l = (re,t ~q- Efre,t+l) + -~P (Et~l-Et)ZpJr.,t+l+j. j=l (34) An innovation in the equity return raises wealth by a factor (1/;~), and this raises consumption by the same factor. Increases in expected future equity returns have offsetting income and substitution effects on consumption; the positive income effect is (t/)~), and the negative substitution effect is -% 1270 1 Y. Campbell These equations can be simplified if I assume that expected aggregate consumption growth, which I write as zt, follows an AR(1) process with mean g and positive persistence O: Act+l = zt + Cc, t+l, (35) zt+l = (1 - O)g + ~zt + ez, t+l. (36) This is a linear version of the model used by Cecchetti, Lam and Mark (1990, 1993) and Kandel and Stambaugh (1991), in which expected consumption growth follows a persistent discrete-state Markov process. The contemporaneous shocks to realized consumption growth ~c,t+~ and expected future consumption growth c~,t+~ may be positively or negatively correlated. The correlation between these contemporaneous shocks controls the univariate autocovariances of consumption growth; the first-order autocovariance is ~bVar(zt)+ Cov(ez, t ~1, co,t+l), and higher-order autocovariances die out geometrically at rate ~b. Thus consumption growth inherits the positive serial correlation of the zt process unless the contemporaneous shocks are sufficiently negatively correlated. An important special case of the model sets Cz,t+l = ~ec, t+l to make consumption growth itself an AR(1) process; this is a linear version of the model of Mehra and Prescott (1985) 19 From Equation (21), the riskless interest rate is linear in expected consumption growth zs, so this model implies a homoskedastic AR(1) process for the riskless interest rate, with persistence 0. It is a discrete-time version of the Vasicek (1977) model of the term structure of interest rates. Campbell, Lo and MacKinlay (1997), Chapter 11, gives a detailed textbook exposition of this model following Backus (1993), Singleton (1990), and Sun (1992). Equations (35) and (36) allow me to rewrite Equations (32) and (33) as - - + ,~- + (37) _Pc, det 1-p ~ 1-p~bJ ' and = - ez,~+l. (38) Equation (38) shows why it is difficult to match the volatility of stock returns within this standard framework. The most obvious way to generate volatile stock returns is 19 The empiricalevidenceonunivariateserial correlationin consumptiongrowthis mixed.Table4 shows small negative autocorrelation in 8 out of 12 quarterly data sets, but only 1 out of 3 annual data sets. Measurement problems may bias these autocorrelations in either direction. Durability of consumption tends to bias autocorrelation downwards, but time-averaging and seasonal adjustment tend to bias it upwards. Empirical estimates of discrete-state Markov models by Cecchetti, Lain and Mark (1990, 1993), Kandel and Stambaugh (1991), and Mehra and Prescott (1985) find some evidence for modest but persistent predictable variation in consumption growth. Ch. 19: Asset Prices',Consumption,and the Business Cycle 1271 to assume a large ,t, that is, a volatile dividend, increasing )~, however, has mixed effects; it increases the volatility of the first term in Equation (38) proportionally, but as long as '1 < 1/*p it diminishes the volatility of the second term because the dividend and real interest rate effects of expected consumption growth offset each other more exactly. The overall volatility of stock returns may actually fall, or grow only slowly, with '1 until the point is reached where '1 > 1/~p. The empirical evidence for small ~ppresented in Table 10 suggests that very high ,t will be needed to generate volatile stock returns. A similar point has been made by Abel (1999), who emphasizes that predictable variation in expected consumption growth can dampen stock market volatility and exacerbate the equity premium puzzle. This model also tends to produce highly volatile returns on real (inflation-indexed) bonds. By setting '1 = 0 in Equations (37) and (38), the log yield and unexpected return on a real consol bond, denoted by a subscript b, are kl, ~ + (39) Ybt = db~ Pbt - 1- p 1- p~ ] ' and rb,t+l-Etrb, t+l---(;) (1 P_~)ez, t+l. (40) When ~p is small, even modest variation in zt will tend to produce large variation in the riskffee interest rate and in the yields and returns on long-term real bonds. The correlation of stock and real bond returns is positive if )~ < 1/% but turns negative if '1 is large enough so that ,~ > 1/~p. Of course, all these calculations are dependent on the assumption made at the beginning of this subsection, that the log dividend on stocks is a multiple )~ of log aggregate consumption. More general models, allowing separate variation in dividends and consumption, can in principle generate volatile stock returns without excessive variation in real interest rates. For example, we might modify Equation (30) to allow a second autonomous component of the dividend: det -'1c~ ~ at, (41) where Aat ~qhas a similar structure to consumption growth, being forecast by an AR(1) state variable: Aat~l - Yt + ~a,t+l, Yt+l = (1 - 0)v + Oy, -~ey,~+l. (42) (43) This modification of the basic model would add a term v/(1 ---p) + (Yt v)/(1 -pO) to the formula for the log price-dividend ratio, Equation (37), and would add a term 1272 JY. Campbell ~a,t+l + p~y,t+l/(l -pO) to the formula for the unexpected log stock return, (38). Cecchetti, Lain and Mark (1993) estimate a discrete-state Markov model allowing for this sort of separate variability in consumption and dividends. While such a model provides a more realistic description of dividends, it requires large predictable movements in dividends to explain stock market volatility. Unfortunately, as section 4.5 shows, there is little evidence for this. 4.4. Implications Jbr the equity premium puzzle I now return to the basic model in which the log dividend is a multiple of log aggregate consumption, and use the formulas derived in the previous subsection to gain a deeper understanding of the equity premium puzzle. The discussion of the puzzle in section 3 treated the covariance of stock returns with consumption as exogenous, but given a tight link between stock dividends and consumption the covariance can be derived from the stochastic properties of consumption itself. This is the approach of many papers including Abel (1994, 1999), Kandel and Stambaugh (1991), Mehra and Prescott (1985), and Rietz (1988). An advantage of this approach is that it clarifies the implications of Epstein-ZinWeil utility. The Epstein-Zin-Weil Euler equation is derived by imposing a budget constraint that links consumption and wealth, and it explains risk premia by the covariances of asset returns with both consumption growth and the return on the wealth portfolio. The stochastic properties of consumption, together with the budget constraint, can be used to substitute either consumption or wealth out of the EpsteinZin-Weil model. To understand this point, note that Equation (33) applies to the return on the wealth portfolio when ,~= 1. Setting e = w and )~= 1, Equation (33) becomes rw,t+~-Etrw, t+l Act+l-EtAct~l + 1 - (Et+~-Et)ZpJAc~+l~/, (44) j=l an equation derived by Restoy and Weil (1998) applying the approach of Campbell (1993). It follows that the covariance of any asset return with the wealth portfolio must satisfy aiw- o~c+ (1-~) ai~, (45) where agz denotes the covariance of asset return i with revisions in expectations of future consumption growth: oc aig =~Coy(ri,t+~- E~ri,t+1,(Et+l - Et) Z pJAct+l+j). (46) j-1 The letter g is used here as a mnemonic for consumption growth. Ch. 19." Asset Prices, Consumption, and the Business Cycle 1273 Substituting this expression into the formula for risk premia in the Epstein-Zin-Weil model, Equation (22), that formula simplifies to Et[ri,t+l]-rj;t+l +~- = ]/(Yic-~ Y- Oig. (47) The risk premium on any asset is the coefficient of risk aversion 3/times the covariance of that asset with consumption growth, plus (7- 1/~p) times the covariance of the asset with revisions in expected future consumption growth. The second term is zero if X = 1/% the power utility case, or if there are no revisions in expected future consumption growth 20. I now return to the assumption made in the previous subsection that expected consumption growth is an AR(1) process given by Equation (36). Under this assumption, (E,+I - Et) Z pJAct+L+j= ez,t+1. (48) j=l Equations (38), (47) and (48) imply that E,[r~,,t+l]- rf,l+l+-~ y (49) This expression nests many of the leading cases explored in the literature on the equity premium puzzle. To understand it, it is helpful to break the equity premium into two components, the premium on real consol bonds over the riskless interest rate, and the premium on equities over real consol bonds: 4E,[rt,,,+ll-rt,t+l+T=7[-~ (1 P--@) _}_(~/_~) [_@{1_~)20z21" (50) Et[re,,+l- r<,+I] + 022 ~-Y'~Ia~2+( Xt. Second, the linear model typically implies a highly volatile riskless real interest rate. The process (66) with a non-constant sensitivity function Z(st) allows one to control or even eliminate variation in the riskless interest rate. To derive the real interest rate implied by this model, one first calculates the marginal utility of consumption as d(Ct) = (G-X,) 7 = S rCr. (68) The gross simple risktess rate is then (i +RIll) = (0E, Ut(G'l)~ '= (OEt (~tiJ 7 \-GI.,/(C'4'~-Y)'U'(Q) ) . (69) Taking logs, and using Equations (62) and (66), the log riskless real interest rate is 2 2 ' - 7Gr/) 1 = - log(0) + yg - y(1 - cp)(s, - s) - -~- [Z(s,) + 1]2 . (70) The first two terms on the right-hand side of Equation (70) are familiar from the power utility model (17), while the last two terms are new. The third term (linear in 1288 JY. Campbell (st -~)) reflects intertemporal substitution. If the surplus consumption ratio is low, the marginal utility of consumption is high. However, the surplus consumption ratio is expected to revert to its mean, so marginal utility is expected to fall in the future. Therefore, the consumer would like to borrow and this drives up the equilibrium riskfree interest rate. Note that what determines intertemporal substitution is meanreversion in marginal utility, not mean-reversion in consumption itself. In this model consumption follows a random walk so there is no mean-reversion in consumption; but habit formation causes the consumer to adjust gradually to a new level of consunlption, creating mean-reversion in marginal utility. The fourth term (linear in D~(s~)+ l]2) reflects precautionary savings. As uncertainty increases, consumers become more willing to save and this drives down the equilibrium riskless interest rate. Note that what determines precautionary savings is uncertainty about marginal utility, not uncertainty about consumption itself. In this model the consumption process is homoskedastic so there is no time-variation in uncertainty about consumption; but habit formation makes a given level of consumption uncertainty more serious for marginal utility ,when consumption is low relative to habit. Equation (70) can be made to match the observed stability of real interest rates in two ways. First, it is helpful if the habit persistence parameter q~is close to one, since this limits the strength of the intertemporal substitution effect. Second, the precautionary savings effect offsets the intertemporal substitution effect if A(s~)declines with st. In fact, Campbell and Cochrane parametrize the ,~(st) function so that these two effects exactly offset each other everywhere, implying a constant riskless interest rate. With a constant riskless rate, real bonds of all maturities are also riskless and there are no real term premia. Thus in the Campbell-Cochrane model the equity premium is also an equity-bond premium. The sensitivity function ,~(st) is not fully determined by the requirement of a constant riskless interest rate. Campbell and Cochrane choose the function to satisfy three conditions: (1) The riskless real interest rate is constant. (2) Habit is predetermined at the steady state s~ = 3. (3) Habit is predetermined near the steady state, or, equivalently, positive shocks to consumption may increase habit but never reduce it. To understand conditions (2) and (3), recall that the traditional notion of habit makes it a predetermined variable. On the other hand habit cannot be predetermined everywhere, or a sufficiently low realization of consumption growth would leave consumption below habit. To make habit "as predetermined as possible", Campbell and Cochrane assume that habit is predetermined at and near the steady state. This also eliminates the counterintuitive possibility that positive shocks to consumption cause declines in habit. Using these three conditions, Campbell and Cochrane show that the steady-state surplus consumption ratio must be a function of the other parameters of the model, and that the sensitivity function )~(st) must take a particular form. Campbell and Cochrane pick parameters for the model by calibrating it to fit postwar quarterly US data. They choose the mean consumption growth rate g = 1.89% per year and the standard Ch. 19: Asset Prices, Consumption, and the Business Cycle 1289 deviation of consumption growth oc = 1.50% per year to match the moments of the US consumption data. Campbell and Cochrane follow Mehra and Prescott (1985) by assuming that the stock market pays a dividend equal to consumption. They also consider a more realistic model in which the dividend is a random walk whose innovations are correlated with consumption growth. They show that results in this model are very similar because the implied regression coefficient of dividend growth on consumption growth is close to one, which produces similar asset price behavior. They use numerical methods to find the price-dividend ratio for the stock market as a function of the state variable st. They set the persistence of the state variable, ~, equal to 0.87 per year to match the persistence of the log price-dividend ratio. They choose y = 2.00 to match the ratio of unconditional mean to unconditional standard deviation of return in US stock returns. These parameter values imply that at the steady state, the surplus consumption ratio = 0.057 so habit is about 94% of consumption. Finally, Campbell and Cochrane choose the discount factor 6 = 0.89 to give a riskless real interest rate of just under 1% per year. It is important to understand that with these parameter values the model uses high average risk aversion to fit the high unconditional equity premium. Steady-state risk aversion is y/S = 2.00/0.057 = 35. In this respect the model resembles a power utility model with a very high risk aversion coefficient. There are however two important differences between the model with habit formation and the power utility model with high risk aversion. First, the model with habit formation avoids the riskfree rate puzzle. Evaluating Equation (70) at the steadystate surplus consumption ratio and using the restrictions on the sensitivity function )~(&), the constant riskless interest rate in the Campbell-Cochrane model is r/+j - log(6) + yg- ~-. (71) In the power utility model the same large coefficient y would appear in the consumption growth term and the consumption volatility term [Equation (17)]; in the CampbellCochrane model the curvature parameter ]e appears in the consumption growth term, and this is much lower than the steady-state risk aversion coefficient y/5: which appears in the consumption volatility term. Thus a much lower value of the discount factor 6 is consistent with the average level of the risk free interest rate, and the model implies a less sensitive relationship between mean consumption growth and interest rates. Second, the model with habit formation has risk aversion that varies with the level of consumption, whereas a power utility model has constant risk aversion. The time.variation in risk aversion generates predictable movements in excess stock returns like those documented in Table 12, enabling the Campbell-Cochrane model to match the volatility of stock prices even with a smooth consumption series and a constant riskless interest rate. 1290 J Y.Campbell 5.2. Models' with heterogeneous agents All the models considered so far assume that assets can be priced as if there is a representative agent who consumes aggregate consumption. An alternative view is that aggregate consumption is not an adequate proxy for the consumption of stock market investors. One simple explanation for a discrepancy between these two measures of consumption is that there are two types of agents in the economy: constrained agents who are prevented from trading in asset markets and simply consume their labor income each period, and unconstrained agents. The consumption of the constrained agents is irrelevant to the determination of equilibrium asset prices, but it may be a large fraction of aggregate consumption. Campbell and Mankiw (1989) argue that predictable variation in consumption growth, correlated with predictable variation in income growth, suggests an important role for constrained agents, while Mankiw and Zeldes (1991) and Brav and Geczy (1996) use US panel data to show that the consumption of stockholders is more volatile and more highly correlated with the stock market than the consumption of non-stockholders. Such effects are likely to be even more important in countries with low stock market capitalization and concentrated equity ownership. The constrained agents in the above model do not directly influence asset prices, because they are assumed not to hold or trade financial assets. Another strand of the literature argues that there may be some investors who buy and sell stocks for exogenous, perhaps psychological reasons. These "noise traders" can influence stock prices because other investors, who are rational utility-maximizers, must be induced to accommodate their shifts in demand. If utility-maximizing investors are risk-averse, then they will only buy stocks from noise traders who wish to sell if stock prices fall and expected stock returns rise; conversely they will only sell stocks to noise traders who wish to buy if stock prices rise and expected stock returns fall. Campbell and Kyle (1993), Cutler, Poterba and Summers (1991), DeLong, Shleifer, Summers and Waldmalm (1990), and Shiller (1984) develop this model in some detail. The model implies that rational investors do. not hold the market portfolio - instead they shift in and out of the stock market in response to changing demand from noise traders - and do not consume aggregate consumption since some consumption is accounted for by noise traders. This makes the model hard to test without having detailed information on the investment strategies of different market participants 23. It is also possible that utility-maximizing stock market investors are heterogeneous in important ways. If investors are subject to large idiosyncratic risks in their labor income and can share these risks only indirectly by trading a few assets such as stocks 23 Recentworksurveyedby Shiller(1999)attempts to placethe behaviorof noise traders on a firmer psychologicalfolmdation.BenartziandThaler(1995),fbr example,arguethatpsychologicalbiasesmake noise tradersreluctantto hold stocks,and that this helpsto explainthe equitypremiumpuzzle. Ch. 19: Asset Prices, Consumption,and the Business Cycle 1291 and Treasury bills, their individual consumption paths may be much more volatile than aggregate consumption. Even if individual investors have the same power utility function, so that any individual's consumption growth rate raised to the power -y would be a valid stochastic discount factor, the aggregate consumption growth rate raised to the power -y may not be a valid stochastic discount factor. This problem is an example of Jensen's Inequality. Since marginal utility is nonlinear, the average of investors' marginal utilities of consumption is not generally the same as the marginal utility of average consumption. The problem disappears when investors' individual consumption streams are perfectly correlated with one another as they will be in a complete markets setting. Grossman and Shiller (1982) point out that it also disappears in a continuous-time model when the processes for individual consumption streams and asset prices are diffusions. Recently Constantinides and Duffle (1996) have provided a simple framework within which the effects of heterogeneity can be understood. Constantinides and Duffle postulate an economy in which individual investors k have different consumption levels Ckt. The cross-sectional distribution of individual consumption is lognormal, and the change from time t to time t + 1 in individual log consumption is cross-sectionally uncorrelated with the level of individual log consumption at time t. All investors have the same power utility function with time discount factor 6 and coefficient of relative risk aversion y, In this economy each investor's own intertemporal marginal rate of substitution is a valid stochastic discount factor. Hence the cross-sectional average of investors' intertemporal marginal rates of substitution is a valid stochastic discount factor. I write this as M,+~ = 6E,\~ L\~-k, ) / ' (72) where E[ denotes an expectation taken over the cross-sectional distribution at time t. That is, for any cross-sectionally random variable Xk~, K 1 ~Xkt,E:x,, - lim k-I the limit as the number of cross-sectional units increases of the cross-sectional sample average of Xkt24. Note that E[Xkt will in general vary over time and need not be lognormally distributed conditional on past information. 24 Constantinidesand Duffle(1996)present a morerigorousdiscussion. 1292 J.Y. Campbell The assumption of cross-sectional lognormality means that the log stochastic discount factor, m[+1 = log(M~ 1), can be written as a function of the cross-sectional mean and variance of the change in log consumption: mr+t = -log(b)- yEt+jAck, t+1 + Var~+lAck,t+l, (73) where Var[ is defined analogously to E[ as K Var/Xl, = lim 1 ~(X~ lEt&,)2 ' K--,ooK k=l and like E[ will in general vary over time. An economist who knows the underlying preference parameters of investors but does not understand the heterogeneity in this economy might attempt to construct a representative-agent stochastic discount factor, M/~, using aggregate consumption: /E/+l[G,t+l] ) r -'uF+'l - b k (74) The log of this stochastic discount factor can also be related to the cross-sectional mean and variance of the change in log consumption: me+]=-log(b)-yEt+,Ack, t+l- (~)[Vart+lck, t+1- Var/c?,l] --log(O)-yE2i.,Ack,,+,- (7)[Var,*~_iAc'k,,+,], (7s) where the second equality follows from the relation ck,tH ckf+ Ack,t+l and the fact that Ack,t+l is cross-sectionally uncorrelated with ckt. The diflbrence between these two variables can now be written as m/~1 I~A Y(Y+ 1) .-m"l = 2 Vart+lACk, t kl. (76) The time series of this difference can have a nonzero mean, helping to explain the riskfree rate puzzle, and a nonzero variance, helping to explain the equity premium puzzle. If the cross-sectional variance of log consumption growth is negatively correlated with the level of aggregate consumption, so that idiosyncratic risk increases in economic downturns, then the true stochastic discount factor m[+1 will be more strongly countercyclical than the representative-agent stochastic discount factor constructed using the same preference parameters; this has the potential to explain the high price of risk without assuming that individual investors have high risk aversion. Mankiw (1986) makes a similar point in a two-period model. Ch. 19: Asset Prices, Consumption,and the Business Cycle 1293 An important unresolved question is whether the heterogeneity we can measure has the characteristics that are needed to help resolve the asset pricing puzzles. In the Constantinides-Duffie model the heterogeneity must be large to have important effects on the stochastic discount factor; a cross-sectional standard deviation of log consumption growth of 20%, for example, is a cross-sectional variance of only 0.04, and it is variation in this number over time that is needed to explain the equity premium puzzle. Interestingly, the effect of heterogeneity is strongly increasing in risk aversion since Var~*+lAck,t+l is multiplied by y(g + 1)/2 in Equation (76). This suggests that heterogeneity may supplement high risk aversion but cannot altogether replace it as an explanation for the equity premium puzzle 25. It is also important to note that idiosyncratic shocks have large effects in the Constantinides-Duffie model because they are permanent. Heaton and Lucas (1996) calibrate individual income processes to micro data from the Panel Study of Income Dynamics (PSID). Because the PSID data show that idiosyncratic income variation is largely transitory, Heaton and Lucas find that investors can minimize its effects on their consumption by borrowing and lending. This prevents heterogeneity from having any large effects on aggregate asset prices. To get around this problem, several recent papers have combined heterogeneity with constraints on borrowing. Heaton and Lucas (1996) and Krusell and Smith (1997) find that borrowing constraints or large costs of trading equities are needed to explain the equity premium. Constantinides, Donaldson and Mehra (1998) focus on heterogeneity across generations; in a stylized three-period overlapping generations model they find that they can match the equity premium if they prevent young agents from borrowing to buy equities. All of these models assume that agents have identical preferences. But heterogeneity in preferences may also be important. Several authors have recently argued that trading between investors with different degrees of risk aversion or time preference, possibly in the presence of market frictions, can lead to time-variation in the market price of risk [Aiyagari and Gertler (1998), Grossman and Zhou (1996), Sandroni (1997), Wang (1996)]. This seems likely to be an active research area in the next few years. 5.3. Irrational expectations So far I have maintained the assumption that investors have rational expectations and understand the time-series behavior of dividend and consumption growth. A number of papers have explored the consequences of relaxing this assumption. [See for example 25 Lettau (1997) reaches a similar conclusion by assuming that individuals consume their income, and calculating the risk-aversioncoefficientsneeded to put model-based stochastic discount factors inside the Hansen-Jagannathanvolatility bounds. This procedure is conservativein that individuals trading in financialmarkets are normallyable to achievesome smoothingof consumptionrelativeto income.NeverthelessLettaufindsthathigh individualrisk aversionis stillneededto satisfythe Hansen,~ Jagannathanbounds. 1294 JY CampbeH Barberis, Shleifer and Vishny (1998), Barsky and DeLong (1993), Cecchetti, Lam and Mark (1998), Chow (1989), or Hansen, Sargent and Tallarini (1997)] 26 In the absence of arbitrage, there exist positive state prices that can rationalize the prices of traded financial assets. These state prices equal subjective state probabilities multiplied by ratios of marginal utilities in different states. Thus given any model of utility, there exist subjective probabilities that produce the necessary state prices and in this sense explain the observed prices of traded financial assets. The interesting question is whether these subjective probabilities are sufficiently close to objective probabilities, and sufficiently related to known psychological biases in behavior, to be plausible. Many of the papers in this area work in partial equilibrium and assume that stocks are priced by discounting expected future dividends at a constant rate. This assumption makes it easy to derive any desired behavior of stock prices directly from assumptions on dividend expectations. Barsky and DeLong (1993), for example, assume that investors believe dividends to be generated by a doubly integrated process, so that the dividend growth rate has a unit root. These expectations imply that rapid dividend growth increases stock prices more than proportionally, so that the price-dividend ratio rises when dividends are growing strongly. If dividend growth is in fact stationary, then the high price-dividend ratio is typically followed by dividend disappointments, low stock returns, and reversion to the long-run mean pric~dividend ratio. Thus Barsky and DeLong's model can account for the volatility puzzle and the predictability of stock returns. In general equilibrium, dividends are linked to consumption so investors' irrational expectations about dividend growth should be linked to their irrational expectations about consumption growth, interest rates are not exogenous, but like stock prices, are determined by investors' expectations. Thus it is significantly harder to build a general equilibrium model with irrational expectations. To see how irrationality can affect asset prices, consider first a static model in which log consumption follows a random walk (q} = 0) with drift g. Investors understand that consumption is a random walk, but they expect it to grow at rate ~ instead of g. Equation (37) implies that the log price-dividend ratio is Pet - det -- i - p + ~ -- (77) Equation (21) implies that the riskless imerest rate is 0-1 2 0 2 rL,+1 - - log 6 + ~ + ~-- ow- ~-~- o7, (78) 26 There is also import. Ch. 19: Asset Prices, Consumption, and the Business Cycle 1295 and the rationally expected equity premium is 4Et[r~,,+,l - rf, t-t-1 -t- T = ~I-~0"2-I- ~L(M- D)" (79) The first term on the right-hand side of Equation (79) is the standard formula for the equity premium in a model with serially uncorrelated consumption growth. This is investors' irrational expectation of the equity premium. The second term arises because dividend growth is systematically different from what investors expect. This model illustrates that irrational pessimism among investors @ < g) can lower the average riskfree rate and increase the equity premium. Thus pessimism has the same effects on asset prices as a low rate of time preference and a high coefficient of risk aversion, and it can help to explain both the riskfree rate puzzle and the equity premium puzzle 27. To explain the volatility puzzle, a more complicated model of irrationality is needed. Suppose now that log consumption growth follows an AR(1) process, a special case of Equation (35), but that investors believe the persistence coefficient to be ~}when in fact it is q)28. In this case the riskfree interest rate is given by ^ r/;t+l =/~f+ ~(Act-g), (80) while the rationally expected equity premium is < E,fr~,,~,I-,r~;,,,+T=~-(,}-O) ~ ,~- +,~ (A<-g), (81) where/~f and/J~ are constants. If 0 is larger than ~b, and if the term in square brackets in Equation (81) is positive, then the equity premium falls when consumption growth has been rapid, and rises when consumption growth has been weak. This model, which can be seen as a general equilibrium version of Barsky and DeLong (1993), fits the apparent cyclical variation in the market price of risk. One difficulty with this explanation for stock market behavior is that it has strong implications for bond market behavior. Consumption growth drives up the riskless 27 The effect of pessimism on the averageprice-dividend ratio is ambiguous, for the usual reason that lower riskfree rates and lower expected dividend growth have offsetting effects. Hansen, Sargent and Tallarini (1997) alsoemphasizethat irrational pessimism can be observationallyequivalentto lowertime preference and higher risk aversion. 28 All alternativeformulationwouldbe to assume, followingEquation(35), that log consumptiongrowth is predicted by a state variablex~that investorsobserve,but that investorsmisperceivetile persistence of this process to be ~ rather than ~. In this case investorscorrectly forecast consumptiongrowth overthe next period, but incorrectly forecast subsequent consumption growth. Their irrationality has no effect on the riskfiee interest rate but causes time-variation in equity and bond premia. 1296 JY. Campbell interest rate and the real bond premium even while it drives down the equity premium. Barsky and DeLong (1993) work in partial equilibrium so they do not confront this problem. Cecchetti, Lain and Mark (1998) handle it by allowing the degree of investors' irrationality itself to be stochastic and time-varying 29. 6. Some implications tbr macroeconomics The research summarized in this chapter has important implications for various aspects of macroeconomics. I conclude by briefly discussing some of these. A first set of issues concerns the modelling of production, and hence of investment. This chapter has followed the bulk of the asset pricing literature by concentrating on the relation between asset prices and consumption, without asking how consumption is determined in relation to investment and production. Ultimately this is unsatisfactory, and authors such as Cochrane (1991, 1996) and Rouwenhorst (1995) have argued that asset pricing should place a renewed emphasis on the investment decisions of firms. Standard macroeconomic models with production, such as the canonical real business cycle model of Prescott (1986), imply that asset prices are extremely stable. The real interest rate equals the marginal product of capital, which is perturbed only by technology shocks and changes in the quantity of capital; when the model is calibrated to US data the standard deviation of the real interest rate is only a few basis points. The return on capital is equally stable because capital can costlessly be transformed into consumption goods, so its price is always fixed at one and uncertainty in the return comes only from uncertainty about dividends. If real business cycle models are to generate volatile asset returns, they must be modified to include adjustment costs in investment so that changes in the demand for capital cause changes in the value of installed capital, or Tobin's q, rather than changes in the quantity of capital. Baxter and Crucini (1993), Jermann (t998), and Christiano and Fisher (1995), among others, show how this can be done. The adjustment costs affect not only asset prices, but other aspects of the model; the response of investment to shocks falls, for example, so larger shocks are needed to explain the cyclical behavior of investment. The modelling of labor supply is an equally difficult problem. Any model in which workers choose their labor supply implies a first-order condition of the form OU OU OC~G~- ON,' (82) where Gt is the real wage and Nt is labor supply. A well-known difficulty in business cycle theory is that with a constant real wage, the marginal utility of consumption 29 The workofRietz(1988)canbe understoodin a similarway.Rictzarguesthatinvestorsare concerned about an unlikelybut serious eventthat has not actually occurred. Given the data we have, investors appearto be irrationalbut in fact, with a longenoughdatasample,they will proveto be rational. Ch. 19: Asset Prices, Consumption,and the Business Cycle 1297 OU/OCt will be perfectly correlated with the marginal disutility of work -OU/ON~. Since the marginal utility of consumption is declining in consumption while the marginal disutility of work is increasing in hours, this implies that consumption and hours worked will be negatively correlated. In the data, of course, consumption and hours worked are positively correlated since they are both procyclical. This problem can be resolved if the real wage is procyclical; then when consumption and hours increase in an expansion the decline in marginal utility of consumption is more than offset by an increase in the real wage. In a standard model with log utility of consumption only a 1% increase in the real wage is needed to offset the decline in marginal utility caused by a 1% increase in consumption. But preferences of the sort suggested by the asset pricing literature, with high risk aversion and low intertemporal elasticity of substitution, have rapidly declining marginal utility of consumption. These preferences imply that a much larger increase in the real wage will be needed to offset the effect on labor supply of a given increase in consumption. Boldrin, Christiano and Fisher (1995) and Lettau and Uhlig (1996) confront this problem; Boldrin, Christiano and Fisher try to resolve it by using a two-sector framework with limited mobility of labor between sectors. In their framework the first-order condition (82) does not hold contemporaneously, but only in expectation. Models with production also help one to move away from the common assumption that stock market dividends equal consumption or equivalently, that the aggregate stock market equals total national wealth. This assumption is clearly untrue even for the United States, and is even less appropriate for countries with smaller stock markets. While one can relax the assumption by writing down exogenous correlated timeseries processes for dividends and consumption in the manner of section 4.3, it will ultimately be more satisfactory to derive both dividends and consumption within a general equilibrium model. Another important set of issues concerns the links between different national economies and their financial markets. In this chapter I have treated each national stock market as a separate entity with its own pricing model. That is, I have assumed that national economies are entirely closed so that there is no integrated world capital market. This assumption may be appropriate for examining long-term historical data, but it seems questionable under modern conditions. There is much work to be done on the pricing of national stock markets in a model with a perfectly or partially integrated world capital market. Finally, the asset pricing literature is important in understanding the welfare costs of macroeconomic fluctuations. There has recently been a tendency for economists to downplay the importance of economic fluctuations in favor of an emphasis on long-term economic growth. But models of habit formation imply that consumers take fluctuations extremely seriously. Fluctuations have important negative effects on welfare because they move consumption in the short term, when agents have little time to adjust; reductions in long-term growth, on the other hand, allow agents' habit levels to adjust gradually. 1298 J.E Campbell This conclusion is not an artifact of a particular utility function and habit formation process. As Atkeson and Phelan (1994) emphasize, it must result from any utility function that explains the level of the equity premium. The choice between risky stocks and stable money market instruments offers investors a tradeoff between the mean growth rate of their wealth and the volatility of this growth rate. The fact that so much extra mean growth is available from volatile stock market investments implies that investors find volatility to be a serious threat to their welfare. Economic policymakers should take this into account when they face policy tradeoffs between economic growth and macroeconomic stability. References Abel, A.B. (1990), "Asset prices under habit formation and catching up with the Joneses", American Economic Review Papers and Proceedings 80:38 42. Abel, A.B. (1994), "Exact solutions for expected rates of return under Markov regime switching: implications for the equity premium puzzle", Journal of Money, Credit and Banking 26:345 361. Abel, A.B. (1999), "Risk premia and term premia in general equilibrium", Journal ofMonetaryEconomics 43:3-33. Abel, A.B., N.G. Mankiw, L.H. Sumxners and R.J. Zeckhauser (1989), "Assessing dynamic efficiency: theory and evidence", Review of Economic Studies 56:1-20. Aiyagari, S.R., and M. Gertler (1998), "Overreaction of asset prices in general equilibrium", Working Paper No. 6747 (NBER). Atkeson, A., and C. Phelan (1994), "Reconsidering the costs of business cycles with incomplete markets", in: S. Fischer and J.J. Rotentherg, eds., NBER Macroeconomics Annual 1994 (MIT Press, Cambridge, MA) 187507. Attanasio, O.R, and G. Weber (1993), "Consumptiongrowth, the interest rate, and aggregation", Review of Economic Studies 60:631-649. Backus, D. (1993), "Cox-Ingersoll Ross in discrete time", unpublished paper (New YorkUniversity). Bansal, R., and W.J.Coleman II (1996), "A monetal7 explanationof the equity premium, tern1premium, and risk-free rate puzzles", Journal of Political Economy 104:1135-1171. Barberis, N., A. Shleifer and R.W. Vishny (1998), "A model of investor sentiment", Journal of Financial Economics 49:307 343. Barclays de Zoete Wedd Securities (1995), The BZW Equity-GiltStudy: investmentin the London Stock Market since 1918 (London). Barsky, R.B., and J.B. DeLong (1993), "Why does the stock market fluctuate?", Quarterly Jomnal of Economics 107:291-311. Baxter, M., and M.J. Crucini (t993), "Explaining saving investment correlations", American Economic Review 83:416-436. Beaudry, R, and E. van Wincoop (1996), "The intertemporal elasticity of substitution: an exploration using a US panel of state data", Economica 63:495 512. Bekaert, G., R.J. Hodrick and D.A. Marshall (1997), "'Peso problem' explanations for term structure anomalies", Working Paper No. 6147 (NBER). Benartzi, S., and R.H. Thaler (1995), "Myopic loss aversion and the equity premium puzzle", Quarterly Journal of Economics 110:73-92. Black, E (1976), "Studies of stock price volatility changes", Proceedings of the 1976 Meetings of the Business and Economic Statistics Section (American Statistical Association) 177--181. Ch. 19: Asset Prices, Consumption, and the Business Cycle 1299 Blanchard, O.J., and M.W. Watson (1982), "Bubbles, rational expectations, and financial markets", in: E Wachtel, ed., Crises in the Economic and Financial Structure: Bubbles, Bursts, and Shocks (Lexington Publishers, Lexington, MA). Boldrin, M., L.J. Christiano and J. Fisher (1995), "Asset pricing lessons for modeling business cycles", Working Paper No. 5262 (NBER). Bollerslev, T., R. Engle and J. Wooldridge (1988), "A capital asset pricing model with time varying covariances", Journal of Political Economy 96:116 131. Bollerslev, T., R.Y. Chou and K.E Kroner (1992), "ARCH modeling in finance: a review of the theory and empirical evidence", Journal of Econometrics 52:5-59. Bray, A., and C.C. Geczy (1996), "An empirical resurrection of the simple consumption CAPM with power utility", unpublished paper (University of Chicago). Breeden, D. (1979), "An intertemporal asset pricing model with stochastic consumption and investment opportunities", Journal of Financial Economics 7:265-296. Brown, S., W. Goetzmann and S. Ross (1995), "Survival", Journal of Finance 50:853-873. Campbell, J.Y. (1986), "Bond and stock returns in a simple exchange model", Quarterly Journal of Economics 101:785-804. Campbell, J.Y. (1987), "Stock returns and the term structure", Journal of Financial Economics 18: 373~99. Campbell, J.Y. (1991 ), "A variance decomposition for stock returns", Economic Journal 101:157-179. Campbell, J.Y. (1993), "lntertemporal asset pricing without consumption data", American Economic Review 83:487-512. Campbell, J.Y. (1996a), "Consumption and the stock market: interpreting international experience", Swedish Economic Policy Review 3:251-299. Campbell, J.Y. (1996b), "Understanding risk and return", Journal of Political Economy 104:298-345. Campbell, J.Y., and J.H. Cochrane (1999), "By force of habit: a consumption-based explanation of aggregate stock market behavior", Journal of Political Economy 107:205-251. Campbell, J.Y., and A.S. Kyle (1993), "Smart money, noise trading, and stock price behavior", Review of Economic Studies 60:1-34. Campbell, J.Y., and N.G. Mankiw (1989), "Consumption, income, and interest rates: reinterpreting the time series evidence", in: O.J. Blanchard and S. Fischer, eds., National Bureau of Economic Research Macroeconomics Annual 4:185~ 16. Campbell, J.Y., and N.G. Mankiw (1991), "The response of consumption to income: a cross-cotmtry investigation", European Economic Review 35: 723-767. Campbell, J.Y., and R.J. Shiller (1988), "The dividend-price ratio and expectations of future dividends and discount factors", Review of Financial Studies 1:195-227. Campbell, J.Y., and R.J. Shiller (1991), "Yield spreads and interest rate movements: a bird's eye view", Review of Economic Studies 58:495 514. Campbell, J.Y., A.W. Lo and A.C. MacKinlay (1997), The Econometrics of Financial Markets (Princeton University Press, Princeton, NJ). Carroll, C.D. (1992), "The buffer-stock theory of saving: some macroeconornic evidence", Brookings Papers on Economic Activity t992(2):61-156. Cecchetti, S.G., E-S. Lain and N.C. Mark (1990), "Mean reversion in equilibrium asset prices", American Economic Review 80:3984 18. Cecchetti, S.G., E-S. Lain and N.C. Mark (1993), "The equity premium and the risk-fiee rate: matching the moments", Journal of Monetary Economics 31:2t~45. Cecchetti, S.G., R-S. Lain and N.C. Mark (1998), "Asset pricing with distorted beliefs: are equity returns too good to be true?", Working Paper No. 6354 (NBER). Chen, N. (1991), "Financial investment opportunities and the macroeconomy", Journal of Finance 46:52%554. Chou, R.Y., R.E Engle and A. Kane (1992), "Measuring risk aversion from excess returns on a stock index", Journal of Econometrics 52:201~24. 1300 J E Campbell Chow, G.C. (1989), "Rational versus adaptive expectations in present value models", Review of Economics and Statistics 71:376 384. Christiano, L.J., and J. Fisher (1995), "Tobin's q and asset returns: implications for business cycle analysis", Working Paper No. 5292 (NBER). Cochrane, J.H. (1991), "Production-based asset pricing and the link between stock returns and economic fluctuations", Journal of Finance 46:209-237. Cochrane, J.H. (1996), "A cross-sectional test of an investment-based asset pricing model", Journal of Political Economy 104:572~521. Cochrane, J.H., and L.R Hansen (1992), "Asset pricing lessons for macroeconomics", in: O.J. Blanchard and S. Fischer, eds., NBER Maeroeconomics Annual 1992 (The MIT Press, Cambridge). Constantinides, G.M. (1990), "Habit formation: a resolution of the equity premium puzzle", Journal of Political Economy 98:519-543. Constantinides, G.M., and D. Duffle (1996), "Asset pricing with heterogeneous consumers", Journal of Political Economy 104:219-240. Constantinides, G.M., J.B. Donaldson and R. Mehra (1998), "Junior can't borrow: a new perspective on tile equity premium puzzle", Working Paper No. 6617 (NBER). Cutler, D.M., J.M. Poterba and L.H. Summers (1991), "Speculative dynamics", Review of Economic Studies 58:529-546. Deaton, A.S. (1991), "Saving and liquidity constraints", Econometrica 59:1221-1248. DeLong, J.B., A. Shleifer, L.H. Summers and R.J. Waldmann (1990), "Noise trader risk in financial markets", Journal of Political Economy 98:703-738. Dunn, K.B., and K.J. Singleton (1986), "Modeling the term structure of interest rates under non-separable utility and durability of goods", Journal of Financial Economics 17:27-55. Epstein, L.G., and S.E. Zin (1989), "Substitution, risk aversion, and the temporal behavior of consumption and asset returns: a theoretical fiamework", Econometrica 57:937-968. Epstein, L.G., and S.E. Zin (1991), "Substitution, risk aversion, and the temporal behavior of consmnption and asset returns: an empirical investigation", Journal of Political Economy 99:263-286. Estrella, A., and G.A. Hardouvelis (1991), "The term structure as a predictor of real economic activity", Journal of Finance 46:555-576. Fama, E.E, and R. Bliss (1987), "The information in long-maturity forward rates", American Economic Review 77:68(~692. Fama, E.E, and K.R. French (1988a), "Permanent and temporary components of stock prices", Journal of Political Economy 96:246-273. Fama, E.E, and K.R. French (1988b), "Dividend yields and expected stock returns", Journal of Financial Economics 22:3 27. Fama, E.E, and K.R. French (1989), "Business conditions and expected returns on stocks and bonds", Journal of Financial Economics 25:23~49. Ferson, W.E., and G.M. Constantinides (1991), "Habit persistence and durability in aggregate consumption: empirical tests", Journal of Financial Economics 29:199-240. French, K., G.W. Schwert and R.E Stambaugh (1987), "Expected stock returns and volatility", Journal of Financial Economics 19:3-30. Frennberg, R, and B. Hansson (1992), "Computation of a monthly index for Swedish stock returns 1919-t989", Scandinavian Economic History Review 40:3-27. Froot, K., and M. Obstfeld (1991), "Intrinsic bubbles: the case of stock prices", American Economic Review 81:t189-1217. Glosten, L, R. Jagannathan and D. Runkle (1993), "On the relation between the expected value and the volatility of the nominal excess return on stocks", Journal of Finance 48:177%180I. Goetzmmur, W.N., and R Jorion (1997), "A century of global stock markets", Working Paper No. 5901 (NBER). Grossman, S.J., and R.J. Shiller (198l), "The determinants of the variability of stock market prices", American Economic Review 71:222 227. Ch. 19: Asset Prices', Consumption, and the Business Cycle 1301 Grossman, S.J., and R.J. Shiller (1982), "Consumption correlatedness and risk measurement in economies with non-traded assets and heterogeneous information", Journal of Financial Economics 10:195-210. Grossman, S.J., and Z. Zhou (1996), "Equilibrium analysis of portfolio insurance", Journal of Finance 51:1379 1403. Grossman, S.J., A. Melino and R.J. Shiller (1987), "Estimating the continuous time consumption based asset pricing model", Journal of Business and Economic Statistics 5:315-328. Hall, R.E. (1988), "Intertemporal substitution in consumption", Journal of Political Economy 96:221~273. Hamilton, J.D. (1989), "A new approach to the analysis of nonstationary returns and the business cycle", Econometa%a 57:357-384. Hansen, L.P, and R. Jagannathan (1991), "Restrictions on intertemporal marginal rates of substitution implied by asset returns", Journal of Political Economy 99:225-262. Hansen, L.P., and K.J. Singleton (1983), "Stochastic consumption, risk aversion, and the temporal behavior of asset returns", Journal of Political Economy 91:249-268. Hansen, L.P, T.J. Sargent and T.D. Tallarini Jr (1997), "Robust permanent income and pricing", unpublished paper (University of Chicago and Carnegie Mellon University); Review of Economic Studies, fbrthcoming. Hardouvelis, G.A. (1994), "The term structure spread and future changes in long and short rates in the G7 countries: Is there a puzzle?", Journal of Monetary Economics 33:255283. Harvey, C.R. (1989), "Time-varying conditional covariances in tests of asset pricing models", Journal of Financial Economics 24:289-317. Harvey, C.R. (1991), "The world price of covariance risk", Journal of Finance 46:111-157. Hassler, J., E Lundvik, T. Persson and E S6derlind (1994), "The Swedish business cycle: stylized facts over 130 years", in: V. Bergstr/Sm and A. Vredin, eds., Measuring and Interpreting Business Cycles (Clarendon Press, Oxford). Heaton, J. (1995), "'An empirical investigation of asset pricing with temporally dependent preference specifications", Econometrica 63:681-717. Heaton, J., and D.J. Lucas (1996), "Evaluating the effects of incomplete markets on risk sharing and asset pricing", Journal of Political Economy 104:443-487. Jermann, U.J. (1998), "Asset pricing in production economies", Journal of Monetary Economics 41: 257-275. Kandel, S., and R.E Stambaugh (t991), "Asset returns and intertemporal preferences", Journal of Monetary Economics 27:39-71. Kocherlakota, N. (1996), "The equity premium: it's still a puzzle", Journal of Economic Literature 34:42~ 1. Kreps, D.M., and E.L. Porteus (1978), "Temporal resolution of uncertainty and dynamic choice theory", Econometriea 46:185-200. Kxusell, E, and A.A. Smith Jr (1997), "Income and wealth heterogeneity, portfolio choice, and equilibrium asset returns", Macroeconomic Dynamics 1:387-422. Kugler, E (1988), "An empirical note on the term structure and interest rate stabilization policies", Quarterly Journal of Economics 103:78%792. La Porta, R., E Lopez-de-Silanes, A. Shleifer and R.W. Vishny (1997), "Legal determinants of external finance", Journal of Finance 52:1131-1150. LeRoy, S.E, and R.D. Porter (1981), "The present value relation: tests based on variance bounds", Econometriea 49:555-577. Lettan, M. (1997), "Idiosyncratic risk and volatility bounds", unpublished paper (CentER, Tilburg University). Lettau, M., and H. Uhlig (1996), "Asset prices and business cycles: successes and pitfalls of the general equilibrium approach", unpublished paper (CentER, Tilburg University). Lucas Jr, R.E. (1978), "Asset prices in an exchange economy", Econometrica 46:1429-1446. 1302 J.Y. Campbell Mankiw, N.G. (1986), "The equity premium and the concentration of aggregate shocks", Journal of Financial Economics 17:211-219. Mankiw, N.G., and J.A. Miron (1986), "The changing behavior of the term structure of interest rates", Quarterly Journal of Economics 101:211228. Mankiw, N.G., and S.E Zeldcs (1991), "The consumption of stockholders and non-stockholders", Journal of Financial Economics 29:97-112. Mehra, R., and E.C. Prescott (1985), "The equity premium puzzle", Journal of Monetary Economics 15:145-161. Merton, R. (1973), "An intertemporal capital asset pricing model", Econometa@a 41:867-887. Miron, J.A. (1986), "Seasonal fluctuations and the life cycle-permanent income hypothesis of consumption", Journal of Political Economy 94:1258 1279. Nelson, C.R., and R. Startz (1990), "The distribution of the instrumental variables estimator and its t-ratio when the instrument is a poor one", Journal of Business 63:S125-$140. Poterba, J.M., and L.H. Summers (1988), "Mean reversion in stock returns: evidence and implications", Journal of Financial Economics 22:27-60. Prescott, E.C. (1986), "Theory ahead of business cycle measurement", Carnegie-Rnchester Conference Series on Public Policy 25:11-66. Restoy, E, and E Weil (1998), "Approximate equilibrium asset prices", Working Paper No. 6611 (NBER). Rietz, 32 (1988), "The equity risk premium: a solution?", Journal of Monetary Economics 21:117-132. Rouwenhorst, K.G. (1995), "Asset pricing implications of equilibrium business cycle models", in: T.E Cooley, ed., Frontiers of Business Cycle Research (Princeton University Press, Princeton, NJ). Ryder Jr, H.E., and G.M. Heal (1973), "Optimum growth with intertemporally dependent preferences", Review of Economic Studies 40:1-33. Sandroni, A. (1997), "Assetprices, wealth distribution, and intertemporal preference shocks", unpublished paper (University of Pennsylvania). Santos, M.S., and M. Woodford (1997), "Rational asset pricing bubbles", Econometrica 65:19 57. Schwert, G.W. (1989), "Why does stock market volatility change over time?", Journal of Finance 44:1115 1153. Shiller, R.J. (i 981), "Do stock prices move too much to be justified by subsequent changes in dividends?", American Economic Review 71:421-436. Shiller, R.J. (1982), "Consumption, asset markets, and macroeconomic fluctuations", Carnegie Mellon Conference Series on Public Policy 17:203-238. Shiller, R.J. (1984), "Stock prices and social dynamics", Brookings Papers on Economic Activity 1984(2):457M98. Shiller, R.J. (1999), "Human behavior and the efficiency of the financial system", ch. 20, this Handbook. Singleton, K. (1990), "Specification and estimation ofintertemporal asset pricing models", in B. Friedman and E Hahn, eds., Handbook of Monetary Economics (North-Holland, Amsterdam). Sun, T. (I992), "Real and nominal interest rates: a discrete-time model and its continuous-time limit", Review of Financial Studies 5:581-611. Stmdaresan, S.M. (1989), "Intertemporally dependent preferences and the volatility of consumption and wealth", Review of Financial Studies 2:73 88. Svensson, L.E.O. (1989), "Portfolio choice with non-expected utility in continuous time", Economics Letters 30:313--317. The Economist (1987), One Hundred Years of Economic Statistics (The Economist, London). Tirole, J. (1985), "Asset bubbles and overlapping generations", Econometrica 53:1499-1527. Vasicek, O. (1977), "An equilibrium characterization of the term structure", Journal of Financial Economics 5:177-188. Wang, J. (1996), "The term structure of interest rates in a pure exchange economy with heterogeneous investors", Journal of Financial Economics 41:75 110. Weil, E (i989), "The equity premium puzzle and the risk-t?ec rate puzzle", Journal of Monetary Economics 24:401-421. Ch. 19: Asset Prices, Consumption, and the Business Cycle 1303 Wheatley, S. (1988), "Some tests of the consumption-based asset pricing model", Journal of Monetary Economics 22:193~ 18. Wilcox, D. (1992), "The construction of US consumption data: some facts and their implications for empirical work", American Economic Review 82:922-941.