technological growth, asset pricing, and consumption risk

58
Technological Growth, Asset Pricing, and Consumption Risk Stavros Panageas The Wharton School University of Pennsylvania Jianfeng Yu The Wharton School University of Pennsylvania May, 2006 Abstract In this paper we develop a theoretical model in order to understand comovements between asset returns and consumption over longer horizons. We develop an intertemporal general equilibrium model featuring two types of shocks: "small", frequent and disembodied shocks to productivity and "large" technological innovations, which are embodied into new vintages of the capital stock. The latter type of shocks aect the economy with lags, since rms need to make irreversible investments in the new types of capital and there is an option value to waiting. The model produces endogenous cycles, countercyclical variation in risk premia, and only a very modest degree of predictability in consumption and dividend growth as observed in the data. The delayed reaction of consumption to a major technology shock can also help explain why short run correlations between returns and consumption growth are weaker than their long run counterparts. Because of this eect, the model can reproduce recent empirical ndings that document the success of the CAPM over longer horizons. Contact: Stavros Panageas, 2326 SH-DH, 3620 Locust Walk, Philadelphia PA 19106, USA. email: [email protected]. We would like to thank Andy Abel, Ricardo Caballero, Adlai Fischer, Tano Santos, Motohiro Yogo, and participants of seminars at Wharton, the Swedish School of Economics (Hagen), the Helsinki School of Economics GSF, the University of Cyprus, the Athens University of Economics and Business, the Univer- sity of Piraeus, the NBER 2005 EFG Summer Institute and the NBER 2006 Chicago AP meeting for very helpful discussions and comments. 1

Upload: vuongkhanh

Post on 11-Feb-2017

241 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Technological Growth, Asset Pricing, and Consumption Risk

Technological Growth, Asset Pricing, and Consumption Risk

Stavros Panageas∗

The Wharton SchoolUniversity of Pennsylvania

Jianfeng Yu

The Wharton SchoolUniversity of Pennsylvania

May, 2006

Abstract

In this paper we develop a theoretical model in order to understand comovements between

asset returns and consumption over longer horizons. We develop an intertemporal general

equilibrium model featuring two types of shocks: "small", frequent and disembodied shocks to

productivity and "large" technological innovations, which are embodied into new vintages of

the capital stock. The latter type of shocks affect the economy with lags, since firms need to

make irreversible investments in the new types of capital and there is an option value to waiting.

The model produces endogenous cycles, countercyclical variation in risk premia, and only a very

modest degree of predictability in consumption and dividend growth as observed in the data.

The delayed reaction of consumption to a major technology shock can also help explain why

short run correlations between returns and consumption growth are weaker than their long run

counterparts. Because of this effect, the model can reproduce recent empirical findings that

document the success of the CAPM over longer horizons.

∗Contact: Stavros Panageas, 2326 SH-DH, 3620 Locust Walk, Philadelphia PA 19106, USA. email:

[email protected]. We would like to thank Andy Abel, Ricardo Caballero, Adlai Fischer, Tano Santos,

Motohiro Yogo, and participants of seminars at Wharton, the Swedish School of Economics (Hagen), the Helsinki

School of Economics GSF, the University of Cyprus, the Athens University of Economics and Business, the Univer-

sity of Piraeus, the NBER 2005 EFG Summer Institute and the NBER 2006 Chicago AP meeting for very helpful

discussions and comments.

1

Page 2: Technological Growth, Asset Pricing, and Consumption Risk

1 Introduction

The process of invention, development and diffusion of new technologies has been widely studied

in the economic literature. Hardly anyone would dispute that technological progress is the most

important factor in determining living standards over the long run. It appears equally plausible

that technological advancement is a key determinant of asset price movements during many periods

of economic history.

Our goal in this paper is a) to illustrate how adoption of large technological innovations will lead

to cycles in output and asset prices and b) to use our model in order to understand the economic

forces behind the empirical success of recent literature1 that has re-ignited interest in consumption

based asset pricing, by emphasizing the correlation between returns and consumption growth over

long horizons (“eventual consumption risk” or “long run risk”).

The key idea behind our theoretical framework is that productivity growth comes in the form

of two shocks. The first type are "small", frequent, disembodied shocks, that affect earnings in

the entire economy. One should think of them as daily news that appear in the financial press

(variations in the supply of raw materials, political decisions that affect production, bad weather

etc.). However, these types of shocks do not fundamentally alter the technology used to produce

output. The second type of shocks are Poisson arrivals of major technological or organizational

innovations, like automobiles, the internet, just in time manufacturing etc.. These shocks will not

affect the economy on impact, but only with a lag. The reason is that firms will need to make

investments in order to take advantage of these innovations, since they are embodied in new types

of capital. Given the irreversibility of the investment decisions, and the high relative cost of these

new technologies on arrival, there is going to be an endogenous lag between the impact of the

second type of shock and its eventual effects on output. Importantly, we show that the process of

adoption of new technologies leads to endogenous persistence and cycles, even though all shocks in

the model arrive in an unpredictable pure i.i.d. fashion.

The link between the macroeconomy and asset pricing in our model revolves around the idea

that growth options of firms exhibit a “life cycle” as technologies diffuse. On impact of a major

technological shock, growth options emerge in the prices of all securities. We show that these

1See e.g. Bansal and Yaron [2004], Bansal, Dittmar, and Lundblad [2004] and for the purposes of this paper

Parker and Julliard [2005] in particular.

2

Page 3: Technological Growth, Asset Pricing, and Consumption Risk

growth options are riskier than assets in place. Hence, in the initial phases of the technological

cycle expected returns in the stock market are higher, simply because most growth options have

not been exercised. As time passes, firms start to convert growth options into assets in place, hence

reducing the implicit riskiness of their stock. Eventually, the new technology enters the region of

diminishing marginal returns at the aggregate, most growth options get exercised and expected

returns become particularly low.

Hence, the model produces countercyclical variation in risk premia. On impact of a major

technological shock, few companies invest in the new technologies, and hence the economy is be-

low its stochastic trend. Anticipations of strong growth over the long run are associated with

high expected returns, as most growth options have not been exercised. Once a technology be-

comes widely adopted, both financial returns and anticipations of growth over the long run decline

simultaneously, as growth options become converted into physical assets.

The model also allows for non-trivial cross sectional heterogeneity across firms: The continuum

of firms in the economy are affected differently by the arrival of technological innovations. Therefore,

we can endogenously obtain the cross sectional distribution of returns, as a function of size and

value. We then use our model as a laboratory in order to examine the consumption CAPM over

different horizons. We show that we can replicate certain patterns in the data, namely the success

of long horizon versions of the CAPM and the failure of short horizon versions.

The intuition for these findings is that in our model returns “anticipate” the increases in con-

sumption that will happen over the medium-long run. However consumption over the short run

does not adjust in response to a major technological shock. Hence, only covariances between re-

turns and long run consumption growth will reveal what is driving asset returns. This helps us

explain the wedge between the short-run and long-run covariances and hence reproduce the findings

of recent empirical literature in a calibrated version of our model.

1.1 Relation to the literature

The literature closest to this paper is the growth-option based asset pricing literature (Cochrane

[1996], Berk, Green, and Naik [1999], Berk, Green, and Naik [2004], Kogan [2001], Kogan [2004],

Gomes, Kogan, and Zhang [2003], Carlson, Fisher, and Giammarino [2004], Zhang [2005], Cooper

[2004], Gourio [2004], Gala [2005]). The papers by Gomes, Kogan, and Zhang [2003] and Carlson,

3

Page 4: Technological Growth, Asset Pricing, and Consumption Risk

Fisher, and Giammarino [2004] are the closest to ours.

Gomes, Kogan, and Zhang [2003] examine the cross section and the time series of returns in a

general equilibrium setting, as we do. The two most significant differences between their model and

ours is a) the distinction between “embodied” and “disembodied” aggregate technological shocks,

and b) the presence of a true timing decision as to the exercise of the growth options. Gomes,

Kogan, and Zhang [2003] assume a single persistent and predictable technological shock. Options

arrive in an i.i.d. fashion across firms, and the firms must decide “on the spot” if they want to

proceed with the investment or not. By contrast in our model, all firms are presented with a (firm

specific) opportunity to plant a tree at the same time. However, they have full discretion as to

the timing. This is not a mere technicality. It is the very reason for substantial simultaneity in

the exercise of growth options that leads to our endogenous cycles. Alternatively put, in Gomes,

Kogan, and Zhang [2003] cycles emerge out of the assumption of a trend stationary productivity

process. In our model, all exogenous shocks follow random walks. Cycles emerge fully endogenously

as the result of technological adoption. The most important practical payoff of our assumptions is

that the consumption process in our model has a strong random walk component, unlike Gomes,

Kogan, and Zhang [2003]. In Gomes, Kogan, and Zhang [2003] a decomposition between trend

and cycle in the Beveridge-Nelson tradition would attribute all variation in consumption to cyclical

fluctuations and none to a random walk component. Our model preserves a strong random walk

component in consumption, and hence is better able to match the variability of consumption over

the short and the long run. This appears particularly important when studying consumption based

asset pricing using consumption growth over varying horizons.

The paper by Carlson, Fisher, and Giammarino [2004] develops the intuition that the exercise

of growth options can lead to variation in expected returns in a partial equilibrium setting. In our

paper, a similar mechanism is at operation in general equilibrium. By taking the model to general

equilibrium, we are able to obtain consumption and the stochastic discount factor endogenously.

Moreover, by obtaining the small predictable components of consumption growth endogenously, we

are able to discuss short and long run correlations between consumption and returns.

We also relate to the recent literature on long run risks (Bansal and Yaron [2004], Hansen,

Heaton, and Li [2005], Bansal, Dittmar, and Lundblad [2004] among others), and in particular to

the results reported in Parker and Julliard [2005]. Papers in the long run risk literature use an

4

Page 5: Technological Growth, Asset Pricing, and Consumption Risk

Epstein-Zin utility specification that introduces long run risk into the pricing kernel. From that

point on, the papers take the empirical correlations between long run consumption growth and

cross sectional returns as given, and derive implications for the average excess returns. In this

paper our purpose is different: We try to understand economically why the correlation between

returns is stronger with long horizon consumption growth and why certain types of stocks correlate

with eventual consumption risk stronger than others. Importantly, almost all of our results emerge

as a result of the modelling of technology, while keeping preferences within an expected utility

framework. Additionally, we are able to link the presence of long run risks to recent findings in

the macroeconomics literature on the delayed reaction of the economy to technological shocks (see

Greenwood and Yorukoglu [1997], Basu, Fernald, and Kimball [2002], Vigfusson [2004]). Hence our

work should be viewed as an attempt to deepen our theoretical understanding of consumption risk

over long horizons and it naturally complements the existing asset pricing literature.

Our paper also complements the work of Menzly, Santos, and Veronesi [2004]. In that paper the

behavior of consumption and dividends are assumed exogenously. Consumption is a random walk

and hence by construction there can be no difference between “short run” and “long run” covariances

between returns and consumption growth. Interestingly, our analysis will endogenously produce a

process for the dividends of a firm that will resemble Menzly, Santos, and Veronesi [2004], in the

sense that the total dividends of an individual firm will be cointegrated with aggregate consumption

growth. However, given that our consumption process has some small predictable components, we

can additionally characterize the differences between its short horizon and long horizon covariance

with returns.

A recent paper that is related to the present one is Pastor and Veronesi [2005]. In their model,

Pastor and Veronesi [2005] connect the arrival of technological growth with the “bubble”-type

behavior of asset prices around these events2. Our model produces some patterns that are similar.

However, the focus of the two papers and the mechanisms are different. Our mechanism uses the

endogenous exercise of growth options to produce variations in expected returns, and we focus on

providing a link between technological arrivals and correlations between consumption and returns

in the time series and the cross section.2Other papers that have analyzed the recent upswing in prices include Pastor and Veronesi [2004], Jermann and

Quadrini [2002].

5

Page 6: Technological Growth, Asset Pricing, and Consumption Risk

Finally, there is a vast literature in macroeconomics and growth that analyzes innovation,

dissemination of new technologies or the impact of the arrival of new capital vintages. A partial

listing would include Jovanovic and Rousseau [2004], Jovanovic and MacDonald [1994], Jovanovic

and Rousseau [2003], Greenwood and Jovanovic [1999], Atkeson and Kehoe [1999], Atkeson and

Kehoe [1993], Helpman [1998], Comin and Gertler [2003]. Our paper has a fundamentally different

scope than this literature. In most of these models, uncertainty and the pricing of risk are not

the focus of the analysis. By contrast these papers analyze innovation decisions in much greater

depth than we do. The trade-off is that they cannot allow for sufficiently rich uncertainty, and an

endogenous determination of the stochastic discount factor as is possible in the simpler setup of

our paper. This is why most of this literature cannot be readily used for an in-depth asset pricing

analysis in the time series and the cross section.

An important technical contribution of our work is that it provides a tractable solution to a

general equilibrium model, where the micro-decsions are "lumpy" and exhibit optimal stopping

features. The micro decision of the firm has a similar structure to the recent sequence of papers

by Abel and Eberly [2003], Abel and Eberly [2002], Abel and Eberly [2004]. Just as firms in these

papers adapt to the technological frontier at an optimally chosen time, firms in our framework

decide on the optimal time to plant new trees. Moreover, by having cross sectional heterogeneity

only at the beginning of an epoch, we can aggregate over firms in a much simpler way than the

existing literature (Caballero and Engel [1999], Caballero and Engel [1991], Caballero and Pindyck

[1996], Novy-Marx [2003]).

The structure of the paper is as follows: Section 2 presents the model and Section 3 the resulting

equilibrium allocations. Section 4 presents the qualitative and quantitative implications of the

model. Section 5 concludes. All proofs are given in the appendix.

2 The model

2.1 Trees, Firms and Technological Epochs

There exists a continuum of firms indexed by j ∈ [0, 1]. Each firm owns a collection of trees that

have been planted in different technological epochs3, and its total earnings is just the sum of the

3We shall also use word “round” to refer to an epoch.

6

Page 7: Technological Growth, Asset Pricing, and Consumption Risk

earnings produced by the trees it owns. Each tree in turn produces earnings that are the product

of three components: a) a vintage specific component that is common across all trees of the same

technological epoch, b) a time invariant tree specific component and c) an aggregate productivity

shock. To introduce notation, let YN,i,t denote the earnings stream of tree i at time t, which was

planted in the technological epoch N ∈ (−∞..− 1, 0, 1, ..+∞). In particular, assume the followingfunctional form for YN,i,t:

YN,i,t =¡A¢N

ζ(i)θt (1)¡A¢Ncaptures the vintage effect. A > 1 is a constant. ζ(·) is a positive strictly decreasing function

on [0, 1], so that ζ(i) captures a tree specific effect. θt is the common productivity shock and evolves

as a geometric Brownian Motion:dθtθt= µdt+ σdBt (2)

where µ > 0, σ > 0 are constants, and Bt is a standard Brownian Motion.

Technological epochs arrive at the Poisson rate λ > 0. Once a new epoch arrives, the index N

becomes N + 1, and every firm gains the option to plant a single tree of the new vintage at a time

of its choosing. Since A > 1, and N grows to N +1, equation (1) reveals that trees of a later epoch

are on average "better" than previous trees.

Firm heterogeneity is introduced as follows: Once epoch N arrives, each firm j draws a random

number ij,N from a uniform distribution on [0, 1]. This number informs the firm of the type of tree

that it can plant in the new epoch. In particular a firm that drew the number ij,N can plant a tree

with tree specific productivity ζ(ij,N). These numbers are drawn in an i.i.d fashion across epochs:

It is possible that firm j draws a low ij,N in epoch n, a high ij,N+1 in epoch N + 1 etc.

To simplify the setup, we shall assume that once an epoch changes, the firm loses the option to

plant a tree that corresponds to any previous epoch. It can only plant a tree corresponding to the

technology of the current epoch.

Let:

Xj,t =X

n=−∞..N

Anζ(ij,n)1{χn,j=1} (3)

where N denotes the technological epoch at time t and 1{χn,j=1} is an indicator function that is 1

if firm j decided to plant a tree in technological epoch n and 0 otherwise. A firm’s total earnings

7

Page 8: Technological Growth, Asset Pricing, and Consumption Risk

are then given by:

Yj,t = Xj,tθt

Any given firm determines the time at which it plants a tree in an optimal manner. Planting

a tree at time t requires a fixed cost of qt. This cost is the same for all trees of a given epoch and

represents payments that need to be given to “gardeners” who will plant these trees. To keep with

the usual assumptions of a Lucas tree economy, we shall assume that the company finances these

fixed payments by issuing new equity in the amount qt.

Assuming complete markets, the firm’s objective is to maximize its share price. Given that

options to plant a tree arrive in an i.i.d fashion across epochs, there is no linkage between the

decision to plant a tree in this epoch and any future epochs. Thus, the option to plant a tree can

be studied in isolation in each round.

The optimization problem of firm j in epoch N amounts to choosing the optimal stopping time

τ :

P oN,j,t ≡ sup

τEt

½1{τ<τN+1}

∙µANζ(ij,N )

Z ∞

τ

Hs

Htθsds

¶− Hτ

Htqτ

¸¾(4)

where Hs is the (endogenously determined) stochastic discount factor, τN+1 is the random time

at which the next epoch arrives, while P oN,j,t denotes the (real) option of planting a new tree in

epoch N.

Given the setup, a firm’s price will consist of three components: a) the value of assets in place,

b) the value of the growth option in the current technological epoch and c) the value of the growth

options in all subsequent epochs. To see this, let:

PAj,t ≡ Xj,t

µEt

Z ∞

t

Hs

Htθsds

¶(5)

denote the value of assets in place (with Xj,t as defined in [3]). Then the price of firm j, assuming

it has not planted a tree (yet) in technological round N is

PN,j,t = PAj,t + P o

N,j,t + P fN,t (6)

where:

P fN,t = Et

à Xn=N+1..∞

Hτn

HtP on,j,τn

!and τn denotes the time at which technological round n = N +1..∞ arrives. The first term on the

right hand side of (6) is the value of assets in place, while the second term is the value of the growth

8

Page 9: Technological Growth, Asset Pricing, and Consumption Risk

option in the current epoch. The third term is the value of all future growth options. Naturally, for

a firm that has planted a tree in the current technological epoch there is no longer a “live” option

and hence its value is given by:

PN,j,t = PAj,t +Et

à Xn=N+1...∞

Hτn

HtP on,j,τn

!

2.2 Aggregation

The total output in the economy at time t is given by

Yt =

Z 1

0Yt(j)dj =

µZ 1

0Xj,tdj

¶θt = Xtθt (7)

with Xj,t defined in (3) and Xt defined as

Xt =

Z 1

0Xj,tdj (8)

It will be particularly useful to introduce one extra piece of notation. Let KN,t ∈ [0, 1] denote themass of firms that have updated their technology in technological epoch N up to time t. We show

formally later that KN,t coincides with the index of the most profitable tree that has not been

planted in epoch N .

Since investment in new trees is irreversible, KN,t (when viewed as a function of time) will be

an increasing process. Given the definition of KN,t, the aggregate output is given as

Yt =

" Xn=−∞..N−1

A(n−N)

µZ Kn,τn

0ζ(i)di

¶+

Z KN,t

0ζ(i)di

#ANθt

where τn = τn+1 denotes the time at which epoch n ended (and epoch n + 1 started). To

analyze this decomposition it will be easiest to define

F (x) =

Z x

0ζ(i)di

It can easily be verified that, Fx ≥ 0 (since ζ(·) > 0) and Fxx < 0, (since ζ(·) is declining).Hence F (x) has the two key properties of a production function. Using the definition of F (·), Ytcan accordingly be rewritten as

Yt =

" Xn=−∞..N−1

A(n−N)

F (Kn,τn) + F (KN,t)

#ANθt (9)

9

Page 10: Technological Growth, Asset Pricing, and Consumption Risk

Aggregate output is thus the product of two components: A stationary component (inside the square

brackets) and a stochastic trend³ANθt

´which captures the joint effects of aggregate technological

progress and aggregate productivity growth. The term inside the square brackets is a weighted

average of the contributions of the different vintages of trees towards the aggregate product. The

weight on trees that were planted in previous epochs decays geometrically at the rate A. In this

sense, A is simultaneously the rate of technological progress (in terms of new trees) and technological

obsolescence (in terms of existing ones).

2.3 Markets

As is typically assumed in “Lucas Tree” models, each firm is fully equity financed and the repre-

sentative agent holds all its shares. Moreover, claims to the output stream of these firms are the

only assets in positive supply, and hence the total value of positive supply assets in the economy is:

PN,t =

Z 1

0PN,j,tdj

Next to the stock market for shares of each company there exists a (zero net supply) bond

market, where agents can trade zero-coupon bonds of arbitrary maturity. We shall assume that

markets are complete.4 Accordingly, the search for equilibrium prices can be reduced to the search

for a stochastic discount factor Ht, which will coincide with the marginal utility of consumption

for the representative agent. (See Karatzas and Shreve [1998], Chapter 4)

2.4 Consumers, Gardeners, and Preferences

To keep with Lucas’s analogy of “trees”, we shall assume that trees can only be planted by “gar-

deners”. The economy is populated by a continuum of identical consumers/gardeners that can be

aggregated into a single representative agent. The representative agent owns all the firms in the

economy, and is also the (competitive) provider of gardening services.

We shall allow the agent’s utility to exhibit (external) habit formation w.r.t. to the running

maximum of consumption for both substantive and technical reasons that will become clear in the4 In particular there exist markets where agents can trade securities (in zero net supply) that promise to pay 1 unit

of the numeraire when technological round N arrives. These markets will be redundant in general equilibrium, since

agents will be able to create dynamic portfolios of stocks and bonds that produce the same payoff as these claims.

However, it will be easiest to assume their existence throughout to guarantee ex-ante that markets are complete.

10

Page 11: Technological Growth, Asset Pricing, and Consumption Risk

next subsection. The representative consumer’s preference over consumption streams is character-

ized by a utility function of the form

U(Ct,MCt )

where:

MCt = max

s≤t{Cs} (10)

denotes the running maximum of consumption up to time t, and U¡Ct,M

Ct

¢satisfies UC > 0,

UCC < 0, UMC < 0, UCMC > 0.

Gardeners have a disutility of effort for planting new trees and need to be compensated ac-

cordingly. Planting a tree creates a fixed disutility of UC(s)η(s) per tree planted. Hence, the

representative agent’s utility function is given by:

maxCs,dls

Et

∙Z ∞

te−ρ(s−t)U(Cs,M

Cs )ds−

Z ∞

te−ρ(s−t)UC(s)η(s)dls

¸(11)

where dl(s) denotes the increments in the number of trees that the representative consumer /

gardener has planted.

This utility specification for the representative agent will allow us to keep the supply of gardening

services perfectly elastic at the price ηt. To see this, let VW denote the derivative of the gardener’s

value function with respect to wealth. A gardener will have an incentive to plant a tree if and only

if:

qtVW ≥ ηtUC

Imposing the envelope condition, we obtain VW = UC . Furthermore, assuming perfect competition

among gardeners reveals that the price for planting trees will be given by:

qt = ηt (12)

This observation will allow us substantial technical simplifications, since it will safeguard that

only the demand for gardening services will determine the number of trees planted. (The analogous

statement in neoclassical investment theory is that Tobin’s q=1, and hence there is elastic supply

of capital). Moreover, it will allow us to keep the variation in tree planting costs constant within an

epoch, as we show in the next section. This will safeguard that variations in the price of gardening

services will not drive asset valuations.

11

Page 12: Technological Growth, Asset Pricing, and Consumption Risk

The consumer maximizes (11) over consumption plans in a complete market:

maxCs,dls

Et

∙Z ∞

te−ρ(s−t)U(Cs,M

Cs )ds−

Z ∞

te−ρ(s−t)UC(s)η(s)dls

¸(13)

s.t.

Et

µZ ∞

t

Hs

HtCsds

¶≤

Z 1

0PN,j,tdj +Et

à Xn=−∞..Nt

Z ∞

t

Hs

Htqsdls

!(14)

Note that the representative consumer owns all the trees and receives gardening fees qt every

time a firm plants a tree.

2.5 Functional Forms and Discussion

Before proceeding, we need to make certain assumptions on functional forms, in order to solve the

model explicitly. The assumptions that we make are intended either a) to allow for tractability or

b) to ensure that the solution of the model satisfies certain desirable properties.

The first assumption on functional form concerns the utility U¡Ct,M

Ct

¢.We shall assume that

U¡Ct,M

Ct

¢=¡MC

t

¢γ C1−γt

1− γ=

³CtMCt

´−γCt

1− γ, γ > 1 (15)

It can be easily verified that UC > 0, UCC < 0, UCMC > 0, UMC < 0. This utility is a special

case of the utilities studied in Abel [1990] and exhibits both “envy” (UMC < 0) and catching up

with the Joneses (UCMC > 0) in the terminology of Dupor and Liu [2003]. The main difference is

that the habit index is in terms of the past consumption maximum, not some exponential average

of past consumption as in Campbell and Cochrane [1999] or Chan and Kogan [2002]. Using the

running maximum of consumptionMCt as the habit index is particularly attractive for our purposes,

because of the analytic tractability that it will allow. As most habit level specifications already

proposed in the literature, it has the very attractive property that it is “cointegrated” with aggregate

consumption in the sense that the difference between log(Ct) and log(MCt ) will be stationary.

Moreover, the ratio between Ct and MCt will be bounded between 0 and 1 (as is the surplus in

Campbell and Cochrane [1999]).

At a substantive level, this utility specification will serve four purposes: First, it will allow us

to match first and second moments of the equity premium. Second, it will imply that the growth

cycles that will arise in the model will leave interest rates unaffected. To see this, note that

UC =

µCt

MCt

¶−γ12

Page 13: Technological Growth, Asset Pricing, and Consumption Risk

In equilibrium, it will turn out thatCt

MCt

=θt

maxs<t θs(16)

The right hand side of (16) is unaffected by the investment decisions of firms and this will in turn

be true for the stochastic discount factor and therefore the real interest rate. This is a key property

of this specification. Without habit formation most of the effects of technological innovations will

work through the real interest rate, which is unattractive5.Third, keeping interest rates unaffected

will overcome an additional severe problem of constant relative risk aversion (CRRA) utilities:

Agents with CRRA utilities will try to smooth the future consumption gains that arise at the onset

of a technological epoch by dissaving. In general equilibrium, savings must remain at 0, and so real

interest rates will have to rise in order to bring savings back to zero. With a coefficient of relative

risk aversion above 1, the smoothing motive will be sufficiently strong as to push the interest rate

so high, that the market price to earnings ratio will decline. This appears to be at odds with the

stock market booms that are typically observed in periods of rapid technological innovation as the

nineties and the twenties. By leaving interest rates unaffected by growth cycles, our specification

avoids this problem. Finally, these preferences will imply constant relative risk aversion, so that we

can attribute all our results to the effect of fluctuations in the relative weight of growth options on

expected returns. Clearly, with time varying risk aversion, our results would look even stronger.

Our next choice of functional form concerns the specification of the disutility of effort for

gardening services ηt. By equation (12) this amounts to specifying the properties of the gardening

fees qt directly. In general , we think of “gardening” services as compensation for the “know how”

that is provided by experts who need to invent, create and install the new capital stock. Our choice

for the functional form of these costs is motivated by four main considerations: First, we want the

magnitude of this compensation to share the same trend as aggregate output. Second, we want to

keep the amount of gardening services provided stationary. Third, we want to keep the gardening

fees constant within each epoch, in order to keep the analysis tractable, and provide a link to the

partial equilibrium literature on growth options. Fourth, we want to capture the idea that the

costs of planting a tree are prohibitively high at the beginning of an epoch, so that even the most

productive firm will have an incentive to wait.

5See Campbell and Cochrane [1999].

13

Page 14: Technological Growth, Asset Pricing, and Consumption Risk

To give a specification that satisfies all three objectives simultaneously, define

Mt = maxs≤t

θs (17)

and let

qt = qANMτN (18)

where q > 0 is a constant, ANis the vintage specific productivity of trees in the current epoch and

MτN is the value of the historical maximum of θt at the start of the technological epoch. Note that

these costs will grow between epochs6. However they will stay constant within an epoch. Moreover,

they will share the same trend growth as consumption.

A final assumption that is made purely for technical convenience is that

ζ(i) = ζ0(1− i)ν , i ∈ [0, 1] (19)

where ζ0, ν > 0 are constants.

2.6 Equilibrium

The equilibrium definition is standard. It requires that all markets clear and that all actions be

optimal given prices.

Definition 1 A competitive equilibrium is a set of stochastic processes hCt,Kn,t,Ht, dlt, qti s.t.a) Ct, dlt solve the optimization problem (13) subject to (14)

b) Firms determine the optimal time to plant a tree by solving the optimization problem (4) and

Kn,t =

Z 1

0eχn,j,tdj (20)

where eχn,j,t is an indicator that takes the value 1 if firm j has planted a tree in epoch n by time

t and 0 otherwise.

c) The goods market clears7:

Ct = Yt for all t ≥ 0 (21)

6Since N will grow when an epoch changes, and MτN+1 will be higher than MτN7This condition might seem surprising at first. One would expect that investment in new trees should introduce

a wedge between output and consumption in this economy. The resolution of the puzzle is that new trees in this

economy are created by having gardeners exert effort. Hence investment presents a pure transfer between shareholders

and gardeners. In particular, it does not require a sacrifice of current consumption goods.

14

Page 15: Technological Growth, Asset Pricing, and Consumption Risk

d) The market for gardening services clears:

dlt = dKn,t (22)

e) The markets for all assets clear.

If one could determine the optimal processes Kn,t, assuming that the costs of gardening are

given by (12), then the optimal consumption process could be readily determined by (21), and (9).

This would in turn imply that the equilibrium stochastic discount factor is given by:

Ht = e−ρtUC (23)

This observation suggests that the most natural way to proceed in order to determine an

equilibrium is to make a conjecture about the stochastic discount factor Ht, solve for the optimal

stopping times in equation (4), aggregate in order to obtain the processes Kn,t for n = N, ...∞, andverify that the resulting consumption process satisfies (23). This is done in section 3.

3 Equilibrium Allocations

We first start by making a guess about the stochastic discount factor in general equilibrium. In

particular we assume that the equilibrium stochastic discount factor is:

Ht = e−ρtµθtMt

¶−γ(24)

with Mt defined as in (17). In Proposition 1 in the appendix we present the solution to the firm’s

optimal stopping problem under this stochastic discount factor. We also show that the consumption

and investment process that results at the aggregate will satisfy (16) and hence verify that it forms

a competitive equilibrium.

The solution to the optimal stopping problem of the firm has an intuitive “threshold” form:

A firm should update when the ratio of aggregate productivity θt to its running maximum at the

beginning of the current epoch (MτN ) crosses the threshold θ(j)given by

θ(j)=

Ξ

ζ(iN,j)

15

Page 16: Technological Growth, Asset Pricing, and Consumption Risk

where Ξ > 0 is an appropriate constant given explicitly in the appendix. Formally, the optimal

time for firm j to plant a tree in round N is when:

τ∗j,N = infτN≤t<τN+1

½t :

θtMτN

≥ Ξ

ζ(iN,j)

¾(25)

The optimal policies of the firms possess three desirable and intuitive properties: First, no firm

will find it optimal to plant a tree immediately when the new epoch arrives, as long as8:

Ξ

ζ(0)> 1 (26)

which we assume throughout.

Second, a key implication of (25) is that the firms that have the option to plant a more “pro-

ductive” tree will always go first, since the threshold θ(j)will be lower for them. This is intuitive:

A firm which can profit more from planting a tree has a higher opportunity cost of waiting and

should always plant a tree first.

Third, and most importantly, there are going to be strong correlations between the optimal

investment decisions of the firms. Conditional on θtMτN

reaching the relevant investment threshold

Ξζ(0) for the first firm, a number of other firms will also find it optimal to invest in close proximity.

9

Figure 1 gives a visual impression of these facts by plotting the impulse response function of a

shock to Nt on consumption.

As can be seen, in the short run consumption is unaffected, as all firms are waiting to invest.

Once however the threshold for the first firm is reached, then the growth rate of consumption

peaks and starts to decline thereafter. The intuition for this decline is the following: the most

profitable firms start investing first, and hence the most productive investment opportunities are

depleted. This leaves less attractive investment opportunities unexploited and hence a moderation

in the anticipated growth rate of the economy going forward. This delayed reaction of the economy

to a major technological shock is consistent with recent findings in the macroeconomic literature

(Vigfusson [2004]).

Another interesting implication of the behavior of aggregate consumption can be seen upon

8To see why this condition is sufficient to induce waiting, examine (25) and note that at the beginning of an epochθτN

MτN

≤ 1. Hence all firms (even the most productive one) will be “below” their investment thresholds.9This is simply because ζ(i) is a continuous function of i and θt is a continuous function of time.

16

Page 17: Technological Growth, Asset Pricing, and Consumption Risk

tClog∆

time

tClog∆

time

Figure 1: Impulse response function of a shock to the Poisson process Nt. The shock impacts the

economy at time 0, which is given by the intersection of the x and y axes.

examining formula (9). Taking logs, this equation becomes:

log(Yt) = log(Ct) = log(θt) +N log(A) + xt (27)

where xt is equal to:

xt = log

µXt

AN

¶= log

" Xn=−∞..N−1

A(n−N)

F (Kn,τn) + F (KN,t)

#(28)

xt is a geometrically declining average (at the rate 1A) of the random terms F (Kn,τn). This

means that xt would behave exactly as an AR(1) process (across epochs) if the terms F (Kn,τn)

were perfectly i.i.d. across epochs10.

Hence, the model is able to produce endogenous propagation and cycles, on top of the pure

random walk stochastic trend log(θt) + N log(A) that we assumed exogenously. The fact that

consumption will exhibit a strong random walk component, is desirable from an empirical point of

view: It is well known that consumption in the data behaves “close” to a random walk with only

mild predictable components.

10One can show that there is small but positive persistence in the F (Kn,τn), terms that further amplifies the

persistence in xt.

17

Page 18: Technological Growth, Asset Pricing, and Consumption Risk

Finally, it will be useful for future reference to apply a Beveridge Nelson decomposition to

log(Ct) in order to obtain:

log(θt) +N log(A) +E(x) = limk→∞

{Et log(Ct+k)− kE[log(Ct+1)− log(Ct)]} = (29)

= log(Ct) +

Z ∞

t[Et (d log(Ct+j))−E (d log(Ct+j))]

Combining (27), (28) and (29) gives:

xt −E(x) = −Z ∞

t[Et (d log(Ct+j))−E (d log(Ct+j))] (30)

This equation shows that xt−E(x) can be thought of as a measure of the distance between actualoutput and stochastic trend. Whenever this difference is negative, this means that the economy

has not absorbed the full benefit of existing technology which is captured in the stochastic trend.

Therefore future growth rates will be large. By contrast whenever xt−E(x) is positive, this meansthat the economy is above its trend line, and the future growth rates will be moderate. Figure 2

illustrates these notions graphically.

4 Qualitative and Quantitative Implications

In this section, we present a qualitative discussion of results along with a quantitative assessment

obtained through simulation.

Table 1 presents our choice of parameters for the baseline calibration exercise. These parameters

were chosen so as to match 22 unconditional moments. These unconditional moments include first

and second moments of consumption growth, the one year real interest rate, the yearly equity

premium, the log (P/E) ratio and the aggregate book to market ratio. These 10 time series

moments were complemented by another 12 cross sectional moments, which correspond to the

cross sectional distribution of size quantiles in the model. These are given in the bottom two rows

of Table 4 along with their empirical counterparts.

As can be seen from the Tables 2 and 4 the model fit is satisfactory. Most time series moments

are within 20-50% of their empirical counterparts. The only exception is the volatility of the one

18

Page 19: Technological Growth, Asset Pricing, and Consumption Risk

)log()log( ttNA θ+

)log( A

)log( A

)log( tC

2+tN1+tNtN

{ })log()log()log( tttt NACx θ+−=

)(xE

tx

( ) )log(log

),log(

tt

t

NA

C

θ+

t

t

0

Conditional - Unconditional Expected Growth over the Long Run( ):)(xExt −−

)log()log( ttNA θ+

)log( A

)log( A

)log( tC

2+tN1+tNtN

{ })log()log()log( tttt NACx θ+−=

)(xE

tx

( ) )log(log

),log(

tt

t

NA

C

θ+

t

t

0

Conditional - Unconditional Expected Growth over the Long Run( ):)(xExt −−

Figure 2: This figure depicts the trend log(A)Nt+log (θt) and the actual level of (log) consumption

log(Ct), as well as the difference between the two. To illustrate the behavior of a "typical" path,

we have set the Brownian increments (dBt) to be equal to 0 so that log(θt) =³µ− σ2

2

´t.

µ 0.010 γ 7 ζ(0) 1

σ 0.033 ρ 0.05 v 2

λ 0.050 A 1.55 q 32

Table 1: Parameters used for the calibration

19

Page 20: Technological Growth, Asset Pricing, and Consumption Risk

Data Model

Mean of consumption growth 0.022 0.030

Volatility of consumption growth 0.033 0.033

Mean of 1-year zero coupon yield 0.018 0.029

Volatility of 1-year zero coupon yield 0.030 0.066

Mean of Equity Premium 0.042 0.036

Volatility of Equity Premium 0.177 0.208

Mean (log) Price to Earnings Ratio 3.091 3.496

Volatility of (log) Price to Earnings Ratio 0.280 0.335

Mean of Book to Market 0.668 0.707

Volatility of Book to Market 0.230 0.316

Table 2: Unconditional Moments of the model and the data. Data for consumption growth are from

the Website of Robert Shiller. Consumption growth refers to (yearly) differences in log consumption. Both

means and standard deviations are computed for the entire sample and for post 1918 data and then averaged

to avoid issues with mismeasurement of consumption in pre World War I data. The rest of the data are

from Chan and Kogan [2002] except for the mean and the volatility of the book to market, which is taken

from Pontiff and Schall [1998]. The unconditional moments for the model are computed from a Monte Carlo

Simulation involving 20000 years of data, dropping the initial 8000 to ensure that initial quantitites are

drawn from their stationary distribution. We approximate the volatility of log consumption growth with its

intantaneous quadratic variation σ.

20

Page 21: Technological Growth, Asset Pricing, and Consumption Risk

year interest rate, which is higher in the model than in the data. This is to be expected, since the

utility specification contains (multiplicative) habit formation. The cross sectional distribution of

log size implied by the model is less disperse than in the data, especially so for the outlier portfolios.

4.1 Time Series Properties of Aggregate Consumption

Before discussing any implications of the model for returns, we first need to make sure that the

model is able to match the features of the consumption data. Since the model produces some

predictability in consumption growth, we need to make sure that this predictability is very weak,

as is the case in the data.

A useful visual depiction of the time series properties of (differences in log) consumption is

afforded by the log-periodogram (see Hamilton [1994] for details). A flat log-periodogram is an

indication of white noise, while a downward sloping log periodogram is an indication of time series

dependence.

The top subplot of Figure 3 depicts the smoothed log periodogram for consumption growth in

the data along with 5-95% confidence bands, bootstrapped under the assumption that consumption

growth is i.i.d. The bottom subplot depicts the log periodogram obtained from simulations of the

model and compares it to the data. As can be seen, the model performs remarkably well. The

strong random walk component contained in simulated consumption process allows us to match

the very weak time series dependence of real-world consumption data.

4.2 Countercyclical Variation in Returns

Having obtained the quantitative and qualitative properties of the model for aggregate consumption,

we now turn to a discussion of the model implications for asset returns, which are the main focus

of this paper.

The price of a firm in general equilibrium is given by (6). Equation (6) decomposes the price

of a firm into three components: 1) the value of assets in place, 2) The value of growth options in

the current technological epoch and 3) The value of growth options in all subsequent technological

epochs. In the appendix (Proposition 4) we give closed form expressions for both the value of a

single firm and the value of the aggregate stock market.

21

Page 22: Technological Growth, Asset Pricing, and Consumption Risk

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45−3.5

−3

−2.5

−2

−1.5

−1Log Periodogram of differences in log consumption

Frequency

Log

Per

iodo

gram

Data95% Model5% ModelMedian

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45−3.5

−3

−2.5

−2

−1.5

−1Log Periodogram of differences in log consumption

Frequency

Log

Per

iodo

gram

Data95% Conf. Int5% Conf. Int

Figure 3: Log Periodogram of the consumption process for the data and the model. The top figure

presents the log periodogram for post- world war I yearly differences in log consumption. Confidence bands

under the H0 that consumption is a random walk, are computed by bootstrapping the first differences in

log consumption to produce 1000 artificial series with a length of 87 years. The bottom figure presents the

same exercise for the model: 3000 years of simulated consumption growth were used to produce repeated

series of a length of 87 years. The figure plots the median of these simulations, along with 5% and 95% range

intervals. It also shows the log periodogram for the actual data. An equally weighted “nearest neighbor”

kernel was used to perform the smoothing, equally weighting the 7 nearest frequencies.22

Page 23: Technological Growth, Asset Pricing, and Consumption Risk

In what follows we study the relative importance of these three terms at the aggregate. In

particular, let the ratio of growth options embedded in aggregate stock market valuation be defined

as:

wo+ft =

³R 1KN,t

P oN,j,tdj

´+ P f

N,t³R 10 P

AN,j,tdj

´+³R 1

KN,tP oN,j,tdj

´+ P f

N,t

We shall also denote wot and w

ft as the weight of current epoch growth options and future growth

options respectively.

Lemma 1 The fraction of total growth options wo+ft in the stock market is given by:

wo+ft =

e−xt

θtMt

, MtMτN

´+ e−xt

(31)

where xt was defined in equation (28) and f³

θtMt

, MtMτN

´is an appropriate positive function given

explicitly in the appendix.

A simple differentiation of (31) shows that wx < 0. That is, the relative weight of growth options

is countercyclical. This is intuitive: When the current level of consumption is below its stochastic

trend, this is an indication that there is a large number of unexploited investment opportunities

for firms. Accordingly, the relative weight of growth options will be substantial. By contrast, when

consumption is above its trend level, this is the indication that the most profitable investment

opportunities have been exploited, and the relative importance of growth options is small.

The importance of growth options in the aggregate stock market is central for asset pricing,

since the expected return on the aggregate stock market is equal to the expected returns of assets

in place and growth options weighted by their relative importance:

µ− r = (1− wo+ft )

¡µA − r

¢+ wo+f

t

"wot

wo+ft

(µo − r) +wft

wo+ft

³µf − r

´#(32)

The following Lemma compares the excess returns on assets in place and the two types of growth

options. The result is similar to Gomes, Kogan, and Zhang [2003]:

Lemma 2 The expected (excess) return of current epoch growth options (µo − r) is strictly larger

than the expected (excess) return of future growth options¡µf − r

¢, which in turn is strictly larger

than the expected (excess) return of assets in place¡µA − r

¢.

23

Page 24: Technological Growth, Asset Pricing, and Consumption Risk

This result shows how countercyclical variation in the relative importance of growth options

translates into countercyclical variation in expected returns: When the economy is below its sto-

chastic trend, there are numerous growth options, which are risky in light of Lemma 2. This pushes

aggregate expected excess returns upwards. However, as growth opportunities get exploited, the

relative importance of growth options and hence the expected excess returns in the stock market

decline.

The countercyclical variation in returns helps explain why valuation ratios (such as the Price

to Dividend ratio) do not strongly predict consumption growth or dividend growth11 in this model,

but rather returns. Consider the well known Campbell-Shiller decomposition of the P/D ratio:

log (Pt)− log(Dt) = φ1 +Et

∞Xj=1

φj−12 [∆dt+j −E (∆dt+j)]| {z }Expected dividend growth

− Etφj−12 rt+j| {z }

Expected future returns

(33)

for appropriate constants φ1, φ2. Countercyclical variation in expected returns implies that ex-

pected dividend growth and expected future returns are positively correlated. Since expected future

returns enter with a negative sign in (33), this implies that variations of predictable components in

dividends will be counterbalanced by opposite variation in expected returns. Hence, only a limited

amount of the variation in anticipated future dividend growth will “pass through” to the P/D ratio.

By contrast, a considerable fraction of the variation in the P/D ratio in our model will be captured

by variations in the stochastic discount factor (24). Clearly, the stochastic discount factor is not

affected by the state variable xt, which captures the predictable parts of consumption growth. We

also note that there is empirical evidence that countercyclical variation in expected returns is the

source of the limited ability of the P/D ratio to predict future dividend growth12.

To obtain the fraction of the variability in the model that is due to variation in dividend growth

quantitatively, we performed the following exercise: Using our closed form solutions, we computed

the “fundamental” price P f of a claim that would deliver the same dividends as the aggregate stock

market, but imposing a constant rate of return for discounting dividends. That constant return

11Note that consumption and dividends per share are not the same thing in this model, because of the constant

equity issuance that is taking place. In fact, one can show that they are not even cointegrated. However, their

predictable components are strongly correlated.12For details on this argument, see Lettau and Ludvigson [2005],Larrain and Yogo [2005] and Menzly, Santos, and

Veronesi [2004].

24

Page 25: Technological Growth, Asset Pricing, and Consumption Risk

was chosen so that the average (log) P/E ratio of the “fundamental” price exactly equal the average

(log) P/E ratio implied by our model. Simply put, P f would be the price of the stock market, if

one applied a constant discount rate to the stock market.

Given the price of such a “fundamental” claim, we decomposed the variance of the log P/D

ratio into its covariance with P f and the difference P − P f . I.e.:

var(log(P/D)) = cov(log(P/D), log(P f/D)) + cov(log(P/D), log (P/D)− log(P f/D)) (34)

This variance decomposition is the “ideal” version of the variance decompositions typically

performed in the empirical literature, since the two components on the right hand side of equation

(34) have to add up to 100% of the variance of the log P/D ratio by construction. In simulations

of the model the average fraction of variation explained by the first term on the right hand side of

(34) was about 33%.

This number should be viewed as a theoretical upper bound to the variability of the P/D ratio

that is due to dividend growth: If we were to perform a variance decomposition in the manner that

is typically done in the empirical literature (i.e. truncating the covariances around 7-15 years) the

variation due to dividend growth would be substantially lower.

We conclude the discussion of the time series properties of returns by performing the usual

predictability regressions of aggregate excess returns on the aggregate log P/D ratio. Table 3

tabulates the results of these regressions, and compares them to the data. We simulate 100 years

of data and obtain several independent samples of such 100-year spans of artificial data. We run

predictability regressions for each of these samples and report the average coefficient along with a

95% distribution band. We then compare these simulations to the equivalent point estimates in

the data.

The coefficients in the simulations have the right sign, but are about 1/3 of their empirical coun-

terparts. Moreover, the empirical point estimates are within the 95% distribution band according

to the model, with the exception of the longest horizons.

The results of table 3 may seem surprising at first. One would anticipate that the P/D ratio

will be high when growth options have not been exercised, and low if they have. Hence that should

imply a positive relationship between expected excess returns and the P/D ratio, rather than the

negative relation that we obtain in the simulations.

25

Page 26: Technological Growth, Asset Pricing, and Consumption Risk

PE Predictive Ability

Data Model

Horizon(years) Coefficient R-square Coefficent R-square

1 -0.120 0.040 -0.063 0.015

(-0.282, 0.057) (0.000, 0.071)

2 -0.300 0.100 -0.104 0.025

(-0.413, 0.156) (0.000, 0.099)

3 -0.350 0.110 -0.132 0.033

(-0.553, 0.227) (0.000, 0.141)

5 -0.640 0.230 -0.191 0.047

(-0.650, 0.378) (0.000, 0.181)

7 -0.730 0.250 -0.247 0.059

(-0.919, 0.453) (0.000, 0.225)

Table 3: Results of predictive Regressions. Excess returns in the aggregate stock market between t and

t+T for T = 1, 2, 3, 5, 7 are regressed on the P/E ratio at time t. A constant is included but not reported.

The data column is from Chan and Kogan [2002]. The simulations were performed by drawing 100 time

series of a length equal to the data and performing the same predictive regressions. For each draw out of

the 100, we simulate 5000 years of data and only keep the last 100 years of data to run the regressions. We

report the means of these simulations next to the data. The numbers in parentheses are the 95% confidence

interval of the estimates obtained in the simulations.

26

Page 27: Technological Growth, Asset Pricing, and Consumption Risk

The resolution of the puzzle lies in the difference between long horizon and instantaneous

expected returns. The easiest way to see this is to consider an individual firm and study the

evolution of its P/D ratio over a technological epoch: The top plot of figure 4 depicts the P/D ratio

and the instantaneous expected return of the firm. Clearly, the two are positively correlated: As

long as a firm has not planted a tree, the fraction of growth options in its price is large and so is its

P/D ratio and its expected return in light of Lemma 2. Once the firm plants a tree, its P/D ratio

experiences a discontinuous drop, and so does its instantaneous expected return. This reflects the

transformation of growth options into assets in place.

To compare, the bottom plot depicts the P/D ratio against the average instantaneous expected

return between t and t+ T , for any T that is larger or equal to the average time it takes to plant

a tree13, starting at the beginning of an epoch. Now there is a negative relation between the P/D

and the average expected return, at least before a firm decides to invest. The reason is that a high

P/D ratio anticipates the decline in expected returns that will occur over the long run, when the

firm plants the new tree.

By aggregating over firms we can extend these results to the aggregate stock market, since the

investment decisions of firms are strongly correlated. The main difference between the picture at

the aggregate level and the individual firm level is that the decline in the P/D ratio does not occur

in a discontinuous fashion, but is more gradual and smooth.

In conclusion, as long as we predict returns over long horizons, we should expect a negative

relationship between the P/D return and expected returns, as the one found in the data.

4.3 Cross Sectional Predictability

The last section developed the implications of the model for the aggregate stock market. We now

turn to the cross sectional implications of the model. Our focus in this subsection will be to show

why the model is able to produce a size and a value premium. In the next subsection we discuss

why these cross sectional phenomena can be explained by a consumption CAPM including “long

run” consumption growth instead of quarterly consumption growth.

13This qualitative pattern for the average expected return would hold as long as we averaged over any T1 > T

periods. For intervals shorter than T we would obtain a hump shaped pattern for the average expected return and

hence no clear positive or negative relationship.

27

Page 28: Technological Growth, Asset Pricing, and Consumption Risk

P/E ratio

Instantaneous Expected return

Firm plants a tree at these times

New epoch arrives here

P/E ratio

Long Run Expected return

Ttime

time

P/E, Expected Returns

P/E ratio

Instantaneous Expected return

Firm plants a tree at these times

New epoch arrives here

P/E ratio

Long Run Expected return

Ttime

time

P/E, Expected Returns

Figure 4: The top plot depicts the P/E ratio and the instantaneous expected return. The bottom

plot depicts the P/E ratio against the average expected return over T periods, where T is the

average time it takes to plant a tree. To pick a “typical” path we set the Brownian increments

(dBt) equal to 0.

28

Page 29: Technological Growth, Asset Pricing, and Consumption Risk

To show the first assertion, we start by defining the relative weight of growth options in a

company’s price:

w(j),o+ft =

P oN,j,t + P f

N,t

PAN,j,t + P o

N,j,t + P fN,t

(35)

where P oN,j,t = 0 if company j has already planted a tree in round N by time t.

To show why the model is able to produce a size premium, it will be easiest to consider a firm

j that has a higher market value of equity (size) than firm j0:

PN,j,t > PN,j0,t

To simplify the analysis, assume further that both of these firms have exercised their growth option

in the current epoch, so that

P oN,j,t = P o

N,j0,t = 0.

Under these assumptions, one can easily show that the relative weight of growth options is higher

for firm j0 than for j :

w(j),o+ft =

P fN,t

PAN,j,t + P f

N,t

=P fN,t

PN,j,t<

P fN,t

PN,j0,t= w

(j0),o+ft

Using Lemma 2, and applying formula (32) to compute the excess returns of the two firms reveals

that

µ(j) − r = (1−w(j),o+ft )

¡µA − r

¢+ w

(j),o+ft

³µf − r

´< (1−w

(j0),o+ft )

¡µA − r

¢+w

(j0),o+ft

³µf − r

´Hence, assuming that one could safely ignore current epoch growth options, a sorting of com-

panies based on size will produce a size effect: Companies with higher market value will have lower

expected returns than companies with lower market value.

The presence of current epoch growth options will distort the perfect ranking of expected returns

implied by size. It is easy to construct examples whereby a firm with high market value could have

a higher expected return than a firm with lower market value. To see this most easily, consider two

firms j and j00that have the same value of assets in place:

PAN,j,t = PA

N,j00 ,t.

29

Page 30: Technological Growth, Asset Pricing, and Consumption Risk

Assume however, that firm j has exercised its current epoch growth option, so that P oN,j,t = 0, while

firm j00 hasn’t, and accordingly P oN,j00,t > 0. Therefore, it must be the case that PN,j00,t > PN,j,t.

However, it is also clear by (35) that w(j00),o+f

t > w(j),o+ft and w

(j00)ot > w

(j)ot = 0. Using Lemma 2

it is easy to see that µ(j00) > µ(j).

In simulations, we find however that current epoch growth options are not quantitatively im-

portant enough to affect the size effect. An unconditional univariate sort on size replicates the

empirical finding that stocks with high market value have lower expected returns, simply because

the fraction of their value that is tied up in growth options is smaller.

However, current epoch growth options do affect the strength of the size effect over an epoch:

The size effect is weak in the early stages of technological adoption, when current epoch growth

options are prevalent, and it becomes stronger after firms exercise their growth options. This is

consistent with a prima facie look at data from the nineties: the size and value effects disappeared

for a number of years and re-emerged only after 2001.

The model is also consistent with the value premium. This may seem counterintuitive at first,

since one would expect that firms with a high market to book ratio should have a substantial

fraction of their value tied up in growth options, and hence should be riskier. The resolution of the

puzzle is that trees are heterogenous in this economy, and accordingly the market to book ratio of

a given firm will primarily reflect the average productivity of its existing trees, and not the share

of growth options.

The easiest way to see this, is to consider two firms j and j0, that have planted a tree in every

single epoch, including the current one. Clearly, the two firms will have identical book values and

identical growth options. However, suppose that firm j has always been “luckier” than firm j0 in

terms of the productivity of the trees it has had the opportunity to plant. Then the market value

of firm j will be higher than the market value of firm j0, because the value of its assets in place will

be higher:

PAN,j,t > PA

N,j0,t (36)

The growth options of the two firms are identical, and hence it must be the case that

PN,j,t > PN,j0,t (37)

30

Page 31: Technological Growth, Asset Pricing, and Consumption Risk

But (36) and (37) imply that:

w(j),o+ft =

P fN,t

PAN,j,t + P f

N,t

=P fN,t

PN,j,t<

P fN,t

PN,j0,t= w

(j0),o+ft

and hence firm j has a smaller fraction of its value tied up in growth options. Accordingly firm j

has a lower expected return than firm j0. Also, since the book values of the two firms are identical,

equation (37) implies that firm j has a lower book to market ratio than firm j0. This is consistent

with the well known fact that a firm with a low book to market ratio have a low expected return.

Since high size (and/or high growth) firms are firms that will typically have trees with higher

productivity on average, the model is also consistent with the empirical evidence reported in Fama

and French [1995], who show that sorting on size and value will produce predictability for a firm’s

profitability (earnings to book ratio).

However, the model cannot produce a size and a value effect as independent effects: Sorting on

size will leave little or no room for a value effect and vice versa. For this reason, we focus on the

size effect henceforth, and note that sorting on value produces similar results. We note in passing

that modifications of the model could produce a size and a value effect as two independent effects,

but they are beyond the scope of the present paper.

Quantitatively, the cross sectional distribution of expected returns is smaller than in the data.

Table 4 presents the average returns on size sorted portfolios and compares them to the returns

reported in Fama and French [1992].

Ignoring portfolio 1A of Fama and French [1992], the difference in monthly returns between the

highest and the lowest size portfolio in our model is about a third of the equivalent value in the

data. Hence, the model has a similar performance to Gomes, Kogan, and Zhang [2003] in terms of

obtaining a quantitatively plausible size premium. This is partly driven by the fact that the model

produces a smaller spread in log size than what is observed in the data.

4.4 Eventual Consumption Risk

The conditional consumption CAPM, which holds in this model, asserts that the following rela-

tionship determines the expected return of any firm j :

µ(j)t − r = −covt

ÃdP

(j)t

P(j)t

;dHt

Ht

!= γcovt

ÃdP

(j)t

P(j)t

;dθtθt

!= γσ

∂P(j)t

∂θt

P(j)t

(38)

31

Page 32: Technological Growth, Asset Pricing, and Consumption Risk

Portfolios formed on Size (Stationary Distribution)

Deciles 1A 1B 2 3 4 5 6 7 8 9 10A 10B

Returns -Data 1.64 1.16 1.29 1.24 1.25 1.29 1.17 1.07 1.10 0.95 0.88 0.90

Returns -Simulated 0.70 0.69 0.68 0.67 0.66 0.64 0.63 0.62 0.61 0.60 0.60 0.59

Log Size - Data 1.98 3.18 3.63 4.10 4.50 4.89 5.30 5.73 6.24 6.82 7.39 8.44

Log Size - Simulated 3.34 3.80 4.07 4.31 4.56 4.83 5.10 5.39 5.69 6.01 6.29 6.60

Table 4: Portfolios sorted by size - model and data. The data are from Fama and French [1992], who report

nominal montly returns, which are affected by the high inflation rates between 1963 and 1990. We report

real montlhy returns for the simulations. For details on the number of simulations used, see the caption to

table 2. To compare, note that the average monthly inflation between 1963 and 1990 was about 0.8, and

hence this number should be subtracted from the Fama-French returns in order to make them comparable

to the simulated numbers.

The first equality is the usual CAPM relationship in continuous time (Karatzas and Shreve

[1998], Chapter 4). The second equality follows from (24) and the fact that the running maximum

of θt is an increasing process, and hence has bounded variation. Accordingly, it has no quadratic

covariation with the increments in P(j)t . The final equality exploits the homoskedasticity in the

increments of θt.

An important implication of (38) is that only the covariation between increments to θt and

returns matter for pricing purposes. This is just a restatement of the obvious fact that the stochastic

discount factor (24) is not affected by the arrival of epochs or the endogenous investment decisions

of firms. Moreover, the conditional CAPM implied by the present model, conditions “down” to an

unconditional CAPM:

Eµ(j)t − r = γcov

ÃdP

(j)t

P(j)t

;dθtθt

!(39)

because the price of risk is constant in this model. (See Cochrane [2005], Page 138).

A natural question that emerges is whether an econometrician could substitute increments in

consumption growth dCtCt

in place of the unobserved dθtθt. If the econometrician had access to the

continuous time path of aggregate consumption, the answer would be affirmative. However, for an

econometrician that has access to time aggregated data the answer is no. Let ∆ be a difference

32

Page 33: Technological Growth, Asset Pricing, and Consumption Risk

operator over a short interval of time, say a quarter. By equation (27):

∆ log(Ct) = ∆ log(θt) +∆£xt +N log(A)

¤and hence quarterly (log) consumption differences ∆ log(Ct) measure quarterly changes in

log(θt) with error. Intuitively, consumption growth captures not only increments to the trend

θt, but also to the stationary component ∆£xt +N log(A)

¤that absorbs the effect of new tree

planting.

The covariance between asset returns and ∆£xt +N log(A)

¤turns out to be negative in this

model, as long as ∆ is not very large (say a quarter). The reason is intuitive: At the beginning

of an epoch expected returns are high, while ∆£xt +N log(A)

¤is practically 0 as very few firms

are planting trees. Once firms start planting trees, the term ∆£xt +N log(A)

¤becomes large while

expected returns become low. Hence, there is a negative correlation between expected returns and

∆£xt +N log(A)

¤that makes the correlation between returns and ∆ log(Ct) a downward biased

estimate of the covariance between returns and ∆ log(θt).

Figure 5 gives a visual impression of this effect. It plots the covariance between∆£xt +N log(A)

¤and the excess returns on various portfolios sorted on size. As can be seen, all covariances are neg-

ative. This implies that the presence of the term ∆£xt +N log(A)

¤makes the covariance between

returns and consumption a downward biased estimate between returns and increments to the shock

log(θt) :

cov³R(j)t ,∆ log(Ct)

´< cov

³R(j)t ,∆ log(θt)

´Hence, the standard consumption CAPM that uses the covariance between consumption increments

and returns to measure the riskiness of stocks, will produce upwards biased estimates of the risk

aversion of the representative agent.

Equally importantly, figure 5 shows that the downward bias will be stronger for small size

firms. Hence the regular consumption CAPM, using quarterly consumption growth will particularly

understate the true riskiness of small size portfolios.

The practical implication of this observation is that a consumption CAPM using quarterly

consumption growth ∆ log(Ct) to proxy for increments to ∆ log(θt) fails to capture the cross section

of returns in this model. The top part of figure 6 illustrates this effect by plotting the actual returns

33

Page 34: Technological Growth, Asset Pricing, and Consumption Risk

0 5 10 15 20 25−9.5

−9

−8.5

−8

−7.5

−7

−6.5

−6x 10

−4

<−−small size portfolios large size portfolios −−>

Cov

aria

nce

betw

een

quar

tely

ret

urns

and

∆(x

t+lo

g(A

)Nt)

Figure 5: The covariance between excess returns and ∆£xt +N log(A)

¤. Portfolios are aranged

along the x−axis ranging from small size to large size.

34

Page 35: Technological Growth, Asset Pricing, and Consumption Risk

on size sorted portfolios against the returns that would be predicted by a regular consumption

CAPM using quarterly consumption growth.

It is important to re-iterate the intuition for this effect: The pricing kernel (24) is not affected

by the cycles induced by technological updates. Hence it is only innovations to the trend shock

θt that drive risk compensation. Since consumption growth over discrete intervals will not only

reflect increments to log (θt) but also the effect of tree planting (∆£xt +N log(A)

¤), the covariance

between returns and consumption growth over short intervals mismeasures the covariance between

R(j)t and ∆ log(θt), which captures the true riskiness of a stock.

Noteworthy, the model also predicts that these biases will be reversed if one uses the covariance

between returns and consumption growth over longer intervals as Parker and Julliard [2005] do.

The bottom subplot of 6 illustrates that an econometrician who would estimate the consumption

CAPM using the covariance between quarterly returns and consumption growth over the subsequent

5 years, would uncover a tight link between these covariances and average returns on size sorted

portfolios.

The reason is the following. Let Rt→t+0.25 be the quarterly return of any asset and let ∆ be

the quarterly difference operator. Consider then the following computation for any T > 1:

cov (Rt→t+0.25, log(Ct+0.25T )− log(Ct)) = cov

ÃRt→t+0.25,

TXi

∆ log (θi)

!+ (40)

+cov

ÃRt→t+0.25, log(A)

TXi

∆Ni

!+cov (Rt→t+0.25, xt+0.25T − xt)

Since log (θt) and Nt are both random walks, one obtains:

cov

ÃRt→t+0.25,

TXi

∆ log (θi)

!= cov (Rt→t+0.25,∆ log (θt)) (41)

cov

ÃRt→t+0.25, log(A)

TXi

∆Ni

!= log(A)cov (Rt→t+0.25,∆Nt) ≥ 0 (42)

The sign of the expression in (42) is positive, since the arrival of new epochs creates growth options

for all firms, and hence boosts valuations upward. Moreover, the covariance in (42) is stronger for

small firms because the importance of growth options is larger for them. Accordingly, the arrival

of growth options that coincides with the arrival of a new epoch (∆Nt > 0) will affect the prices

and hence the returns of small firms more than the returns of large firms.

35

Page 36: Technological Growth, Asset Pricing, and Consumption Risk

Finally, and most importantly, for large enough T, we have that:

cov (Rt→t+0.25, xt+T − xt) ' cov (Rt→t+0.25,−xt) > 0 (43)

since expected returns are countercyclical, as was shown in section 4.2. Combining (42), (43), (41)

and (40) yields:

cov (Rt→t+0.25, log(Ct+T )− log(Ct)) > cov (Rt→t+0.25,∆ log (θt))

Hence the covariance between long run consumption and quarterly returns will be higher than the

covariance between quarterly returns and ∆ log θt, which helps eliminate the bias contained in short

run correlations. It is also possible to show analytically14 that the positive bias between the left and

the right hand side is larger for small firms than for large firms. Hence, the spread in covariances

between quarterly returns and long run (say 3-5 year) consumption growth will be larger than

the spread in covariances between quarterly returns and quarterly differences in log (θt) . This will

make the former covariances particularly successful at explaining the cross section of returns as the

bottom part of figure 6 illustrates.

We conclude by noting that we obtained similar results when we used covariances between 5-

year returns and 5-year consumption growth, instead of covariances between quarterly returns and

5-year consumption growth. The reason is intuitive: Covariances between “long run” returns and

“long run” consumption will isolate comovements in trends between the two quantities and will

eliminate the effects of cyclical variations in consumption growth as shown in Bansal, Dittmar, and

Kiku [2005]. The fact that both approaches work in terms of explaining the cross section in returns

is consistent with the empirical evidence presented in Parker and Julliard [2005].

Finally, we point out that the results presented here rely on three features of the model: 1)

the fact that consumption adjusts in a sluggish manner in the short run, while returns reflect

anticipations of future growth immediately, 2) the fact that only increments to d log θt are priced,

because of the form of the stochastic discount factor and 3) the fact that discrete time consumption

data over different horizons will provide imperfect proxies for the unobserved covariances between

returns and the innovations to trend d log θt over various horizons. Importantly, all three properties

would continue to be true if one were to set up the model in discrete time15: it is not just time14We make this derivation available upon request to save space.15The main convenience allowed by continuous time is that the only source of quadratic variation in the stochastic

36

Page 37: Technological Growth, Asset Pricing, and Consumption Risk

0.005 0.01 0.015 0.02 0.0250.005

0.01

0.015

0.02

0.025

Fitted Returns

Ave

rage

Ret

urns

Contemporaneous Consumption Growth Rate

0.005 0.01 0.015 0.02 0.0250.005

0.01

0.015

0.02

0.025

Fitted Returns

Ave

rage

Ret

urns

5 Years of Consumption Growth Rate

Figure 6: In both subplots average returns are plotted against the returns that would be predicted by the

consumption CAPM using short and long consumption differences. The top subplot depicts results for the

consumption CAPM using 1-quarter consumption growth and the bottom subplot using 5-year consumption

growth rates to evaluate the covariation between consumption growth and quarterly excess returns.

37

Page 38: Technological Growth, Asset Pricing, and Consumption Risk

aggregation that produces the results, but the interaction of all of the above factors, in particular

properties 1) and 2).

5 Conclusion

Why does consumption based asset pricing seem to work better over longer horizons, than over short

horizons? In this paper we proposed a general equilibrium framework that served as a laboratory in

order to investigate this question. The key ingredient of our model is the joint presence of “small”

frequent disembodied productivity shocks and “large” infrequent embodied technology shocks. The

first type of shocks affect the economy on impact and behave exactly as a random walk. The latter

arrive also in a random walk fashion. However, there are delays between their impact and their

effects on the economy.

This setup allowed us to obtain a consumption process that has strong random walk components,

along with some small predictable parts. The conditional expected returns exhibit countercyclical

variation. Intuitively, the extent to which the economy has absorbed a major technological shock,

will be informative for pricing purposes, as it will determine the relative weight of growth options

in the prices of companies.

We also showed that the immediate reaction of asset prices to the arrival of new technological

epochs, along with the sluggish adjustment of consumption will drive a wedge between short run and

long run covariances between returns and consumption growth. This helps account for the empirical

findings in Parker and Julliard [2005] and Bansal, Dittmar, and Kiku [2005] among others, who

illustrate that long run correlations are better at explaining the cross section of returns.

Hence, we gave a production based theory for the existence of predictable components in con-

sumption, the nature of stochastic trends, and their implications for asset prices both in the time

series and in the cross section. Importantly, our model appears able to match reasonably well a

number of key facts in the data, even with a constant relative risk aversion and non-Epstein Zin

preferences. We believe that incorporation of Epstein Zin preferences and/or time varying risk

aversion in our framework would most likely strengthen our conclusions even further. We leave this

to future research.

discount factor comes from θt. In discrete time, one would also have to account for the covariance between returns

and the running maximum of θt. That would appear though to strengthen the results further.

38

Page 39: Technological Growth, Asset Pricing, and Consumption Risk

A Appendix

A.1 Propositions and Proofs

A.1.1 Proof of the Main Proposition

Proposition 1 Define the constants Z∗, γ1, γ∗1 and Ξ by

Z∗ =1

ρ− µ(1− γ)− σ2

2 γ(γ − 1)(44)

γ1 =

q¡µ− σ2

2

¢2+ 2σ2(ρ+ λ)−

³µ− σ2

2

´σ2

(45)

γ2 =−q¡

µ− σ2

2

¢2+ 2σ2(ρ+ λ)−

³µ− σ2

2

´σ2

(46)

γ∗1 =

q¡µ− σ2

2

¢2+ 2σ2ρ−

³µ− σ2

2

´σ2

(47)

Ξ =q

Z∗γ1

γ1 − 1γ∗1 − 1

γ∗1 + γ − 1and assume that:

γ∗1 > 1

γ2 < 1− γ

Ξ

ζ(0)> 1

Assume moreover that Ht is given by (24), and qt is given by (18). Then, firm j faced with the optimal

stopping problem (4) will plant a tree the first time that θt crosses the threshold θ

θ =MτN

Ξ

ζ(iN,j)(48)

Formally, the optimal stopping time τ∗ is given by

τ∗ = inf{t : θt = θ}

Finally, if firms follow the above threshold policies, then

Ct

MCt

=θtMt

(49)

with Mt,MCt defined in (17) and (10). Therefore

Ht = e−ρtUC = e−ρtµ

Ct

MCt

¶−γ= e−ρt

µθtMt

¶−γas conjectured in (24).

39

Page 40: Technological Growth, Asset Pricing, and Consumption Risk

This is the key Proposition of the paper. We start by assuming that the state price density is indeed

given by:

Ht = e−ρtµ

θtMt

¶−γ(50)

and prove that (48) provides the solution to the firm’s optimal stopping problem. A useful intermediate

first result is the following:

Lemma 3 The conditional expectation:

Z(θt,Mt) ≡ Et

"Z ∞t

e−ρ(s−t)µ

θsMs

¶−γθsds

#(51)

can be computed explicitly as

Z (θt,Mt) = Z∗µ

θtMt

¶−γθt

"1 +

γ

γ∗1 − 1µ

θtMt

¶γ+γ∗1−1#(52)

with Z∗ and γ∗1 as defined in (44) and (47) respectively.

Proof. (Lemma 3) We give a sketch. One can verify directly that the function Z in equation (52)

satisfies: µθtMt

¶−γθt + µθZθ +

σ2

2θ2Zθθ − ρZ = 0 (53)

whenever θt < Mt and it also satisfies a reflection condition:

ZM (θt,Mt) = 0 (54)

at θt =Mt. By Ito’s Lemma,

e−ρTZ (θT,MT )− e−ρtZ (θt,Mt) =

Z T

t

e−ρs∙µθZθ +

σ2

2θ2Zθθ − ρZ

¸ds+ (55)

+

Z T

t

e−ρsσθsZθdBs +

Z T

t

e−ρsZMdMs

Using (53) and (54), we can rewrite the above as:

e−ρTZ (θT,MT )− e−ρtZ (θt,Mt) = −Z T

t

e−ρsµ

θsMs

¶−γθsds+

+

Z T

t

e−ρsσθsZθdBs

Taking expectations, letting T go to infinity, and multiplying by eρt on both sides gives (51).

Corollary 2 The value of assets in place for firm j is given by

PAj,t = Z∗Xj,tθt

"1 +

γ

γ∗1 − 1µ

θtMt

¶γ+γ∗1−1#

40

Page 41: Technological Growth, Asset Pricing, and Consumption Risk

Proof. (Corollary 2) Combine (52) and (5).

With this Lemma we are now in a position to discuss the solution to the firm’s optimization problem.

The option to plant a tree in epoch N does not affect the option to plant a tree in any subsequent epoch.

Therefore, the firm chooses its optimal strategy to plant a tree “epoch by epoch”.

The individual firm takes the state price density (50) and the costs of planting a tree (18) as given. With

these functional specifications, the optimization problem (4) becomes:

P oN,j,t =

µθtMt

¶γ·

·maxτ

Et

(1{τ≤τN+1}

"Z ∞τ

e−ρ(s−t)ζ (ij,N ) AN

µθsMs

¶−γθsds− qANMτN

µθτMτ

¶−γe−ρ(τ−t)

#)

Using Lemma 3 this optimization problem can be rewritten as

P oN,j,t = AN

µθtMt

¶γMτN · (56)

·maxτ≥t

Et

"e−(ρ+λ)(τ−t)

Ãζ (ij,N )Z

∗"1 +

γ

γ∗1 − 1µ

θτMτ

¶γ+γ∗1−1# θτMτNt

µθτMτ

¶−γ− q

µθτMτ

¶−γ!#

To solve the optimization problem inside the square brackets we proceed as follows: We start by restricting

our attention to trigger strategies, i.e. strategies where the firm invests the first time that the ratio θtMτNt

crosses a threshold θ. Formally, consider strategies of the form:

τ θ = inf{s ≥ t :θs

MτNt

≥ θ} (57)

The proof of the following result is standard and is omitted16:

Et

³e−(ρ+λ)(τ θ−t)

´=

Ãθt

θMτNt

!γ1

(58)

where γ1 is defined in (45). Defining

vτ θ(θt,Mt,MτNt; θ) = Et

"e−(ρ+λ)(τ θ−t)

Ãζ (ij,N )Z

∗"1 +

γ

γ∗1 − 1µ

θτ θMτ θ

¶γ+γ∗1−1# θτ θMτNt

µθτ θMτ θ

¶−γ− q

µθτ θMτ θ

¶−γ!#

and using (58) we obtain:

v(θt,Mt,MτNt; τ θ) = ζ (ij,N )Z

∗"1 +

γ

γ∗1 − 1µ

θ

Mt/MτNt

∧ 1¶γ∗1+γ−1#µ θ

Mt/MτNt

∧ 1¶−γ µ

θtMτNt

¶γ1θ1−γ1

−qµ

θ

Mt/MτNt

∧ 1¶−γ µ

θtMτNt

¶γ1θ−γ1

16See e.g. Karatzas and Shreve [1991] page 197 for details.

41

Page 42: Technological Growth, Asset Pricing, and Consumption Risk

We use the notation x ∧ y = min{x, y}. The termθ

Mt/MτNt

∧ 1

captures the idea that the set of optimal stopping strategies that we consider satisfy either θτMτ

= 1, orθτMτ

= θMt/MτNt

. Given the simple structure of the policies that we consider, finding the optimal stopping

strategy in the constrained set (57) amounts to a simple one-dimensional optimization maximization over θ.

The next Lemma determines the optimal strategies in the constrained set (57). We omit the proof17 both

in order to save space, and because the steps of the proof are fairly straightforward differentiations. Most

importantly, the next Lemma is only useful in that it will allow us to form a conjecture on the value function

and the optimal stopping policies for arbitrary stopping policies. We then show via a formal verification

theorem that the conjectured value function is indeed the appropriate value function.

It will be useful to refer to figure 7 that gives a graphical depiction of the regions described in the proof.

Lemma 4 Let θ1, θ2 and θ2 defined as

θ2 =q

ζ (ij,N )Z∗γ + γ1

(γ1 + γ − 1) + γ(γ1−γ∗1)γ∗1−1

θ2 ≡ q

ζ (ij,N )Z∗γ + γ1

γ1 + γ − 1θ1 = q

1

Z∗γ1

γ1 − 1γ∗1 − 1

γ∗1 + γ − 11

ζ (ij,N )

Then θ1 < θ2 < θ2 and there exists a unique positive solution to the nonlinear equation

h¡θ2;Mt/MτNt

¢ ≡ (1− γ1 − γ) θ2 − γ (γ1 − γ∗1)γ∗1 − 1

µθ2

Mt/MτNt

¶γ∗1+γ−1θ2 + (γ + γ1) q

1

ζ (ij,N )Z∗= 0 (59)

as long as Mt/MτNt> θ2 . Let θ2(Mt/MτNt

) denote this unique positive solution. Then

θ2(Mt/MτNt) < θ2

and the solution to the problem

v(θt,Mt,MτNt) = max

θvτ θ(θt,Mt,MτNt

; θ)

is given as follows:

In region I, i.e. when θt/MτNt> θ2(Mt/MτNt

), then:

v(θt,Mt,MτNt) = ζ (ij,N )Z

∗ θtMτNt

µθtMt

¶−γ+

γ

γ∗1 − 1ζ (ij,N )Z

∗ θtMτNt

µθtMt

¶γ∗1−1− q

µθtMt

¶−γ17Available upon request

42

Page 43: Technological Growth, Asset Pricing, and Consumption Risk

In region II, i.e. when Mt/MτNt≥ θ2 and θ2(Mt/MτNt

) ≥ θt/MτNt, then:

v(θt,Mt,MτNt) =

"ζ (ij,N )Z

∗µ

θ2Mt/MτNt

¶−γθ1−γ12 + ζ (ij,N )Z

∗ γ

γ∗1 − 1µ

θ2Mt/MτNt

¶γ∗1−1θ1−γ12

#µθt

MτNt

¶γ1−qµ

θ2Mt/MτNt

¶−γθ−γ12

µθt

MτNt

¶γ1In region III, i.e. when Mt/MτNt

≤ θ1, then

v(θt,Mt,MτNt) =

1

γ1 − 1q

µθt

MτNt

¶γ1θ−γ11

In region IV, i.e. when θ1 ≤Mt/MτNt≤ θ2, then

v(θt,Mt,MτNt) = ζ (ij,N )Z

∗ γ + γ∗1 − 1γ∗1 − 1

Mt

MτNt

µMt

θt

¶−γ1− q

µMt

θt

¶−γ1The stopping region is S = {(θt,Mt) : θ1 ≤ θ/MτNt

= M/MτNt≤ θ2 or θ/MτNt

≥ θ2}. The optimalstopping time is the first time t such that (θt,Mt) enters the stopping region.

Figure 7 gives a graphical depiction of all the possible regions. Clearly θt ≤Mt and we need only concern

ourselves with the region above the 45 degree line. The bold line depicts the boundary between the stopping

and the continuation region. Given that this is a two dimensional optimal stopping problem the decision

rule will be a mapping between (θt,Mt) and the discrete decision {stop - continue}. The depicted stopping

region is the set of points where the firm would choose to invest. The complement of this region is the set

of points where it would wait.

We are now ready to state and prove the key result:

Proposition 3 The function v given in Lemma 4 is the value function of problem (56) and the optimal

stopping policy given in Lemma 4 is optimal among all possible stopping policies

Proof. (Proposition 3) As a first step we show that the function v satisfies the following properties:

v(θs,Ms,MτNt) ≥ f(θt,Mt,MτNt

) (60)

where

f(θt,Mt;MτN ) ≡ ζ (ij,N )Z∗ θtMτN

µθtMt

¶−γ+ζ (ij,N )Z

∗ γ

γ∗1 − 1µ

θtMt

¶γ∗1−1 θtMτN

− q

µθtMt

¶−γvM (Ms,Ms,MτNt

) ≤ 0 for θt =Mt (61)

43

Page 44: Technological Growth, Asset Pricing, and Consumption Risk

NM

t

τ

θ

NMMt

τ

45

Region II

Region IV

Region III

Region I

Stopping Region

Continuation Region

θ

0)/;( 2 =NMMh t τθ

NM

t

τ

θ

NMMt

τ

45

Region II

Region IV

Region III

Region I

Stopping Region

Continuation Region

θ

0)/;( 2 =NMMh t τθ

Figure 7: Depiction of the various regions. The continuation region is separated from the stopping

region by the bold line.

Av(θs,Ms,MτNt) ≤ 0 (62)

whereAv = σ2

2 θ2vθθ+µθvθ−(ρ+ λ) v. Moreover, we shall demonstrate that v is continuous and differentiable

throughout its domain. Finally we will employ a verification Theorem for optimal stopping similar to

(Øksendal [2003]) to conclude.

Equation (60) holds by construction of the value function v(θs,Ms,MτNt). To see this note that according

to Lemma 4:

v(θs,Ms,MτNt) = max

θvτ θ(θt,Mt,MτNt

; θ) ≥

≥ vτ θ(θt,Mt,MτNt;

θtMτNt

) =

= f(θt,Mt,MτNt)

So we only need to check equation (61) and equation (62) in all the four regions and the smooth conditions.

In region I, θt/MτNt> θ2

¡Mt/MτNt

¢, and hence

vM (Mt,Mt,MτNt) = ζ (ij,N )Z

∗ γ

MτNt

µ1− γ∗1 − 1

γ∗1 − 1¶− γq

1

Mt

= −γq 1Mt

< 0

44

Page 45: Technological Growth, Asset Pricing, and Consumption Risk

Moreover

Av(θt,Mt,MτNt) = −ζ (ij,N ) Z

Z

θtMτNt

µθtMt

¶−γ− γλ

γ∗1 − 1ζ (ij,N )Z

∗ θtMτNt

µθtMt

¶γ∗1−1− q

Y

µθtMt

¶−γ=

"−ζ (ij,N ) Z

Z

θtMτNt

− γλ

γ∗1 − 1ζ (ij,N)Z

∗ θtMτNt

µθtMt

¶γ∗1+γ−1− q

Y

#µθtMt

¶−γ(63)

where

Z =−1

σ2

2 (γ − 1)γ + µ(1− γ)− ρ− λ

Y =1

σ2

2 γ2 − γ

¡µ− σ2

2

¢− ρ− λ

It is easy to check, by assumption 1− γ2 − γ > 0, that

Z =1

σ2

2 (γ1 + γ − 1) (1− γ2 − γ)> 0 (64)

Y =1

σ2

2 (γ1 + γ) (γ2 + γ)

It is easy to check that γ1γ2 = −2(ρ+λ)σ2 and γ1 + γ2 = 1− 2µσ2 . Hence,

1

Y=

γ + γ21− γ − γ2

γ1 + γ

γ1 + γ − 11

Z(65)

From equation (59), we obtain:

γζ (ij,N )Z∗

γ∗1 − 1µ

θ2Mt/MτNt

¶γ∗1+γ−1θ2 = (γ1 + γ − 1) θ2 ζ (ij,N )Z

γ∗1 − γ1− q

γ + γ1γ∗1 − γ1

(66)

Hence, by θt/MτNt> θ2

¡Mt/MτNt

¢, equation (65) and equation (66),

−ζ (ij,N) Z∗

Z

θtMτNt

− γλ

γ∗1 − 1ζ (ij,N )Z

∗ θtMτNt

µθtMt

¶γ∗1+γ−1− 1

Yq

≤ −ζ (ij,N) θ2Z∗

Z− λ

µ(γ1 + γ − 1) θ2 ζ (ij,N )Z

γ∗1 − γ1− q

γ + γ1γ∗1 − γ1

¶− 1

Yq

=

∙λ (γ1 + γ − 1)

γ1 − γ∗1− 1

Z

¸Z∗ζ (ij,N ) θ2 + q

∙−(γ + γ1)λ

γ1 − γ∗1− γ + γ21− γ − γ2

γ1 + γ

γ1 + γ − 11

Z

¸To arrive from equation (63) to equation (62), we only need to show that

0 ≥∙λ (γ1 + γ − 1)

γ1 − γ∗1− 1

Z

¸Z∗ζ (ij,N ) θ2 + q

∙−(γ + γ1)λ

γ1 − γ∗1− γ + γ21− γ − γ2

γ1 + γ

γ1 + γ − 11

Z

¸(67)

We first want to showλ (γ1 + γ − 1)

γ1 − γ∗1>1

Z(68)

45

Page 46: Technological Growth, Asset Pricing, and Consumption Risk

Actually, by equation (64), this is equivalent to showing that

λ (γ1 + γ − 1)γ1 − γ∗1

> (γ1 + γ − 1) (1− γ2 − γ)σ2

2(69)

which amounts to showing that

(1− γ2 − γ) (γ1 − γ∗1) <2λ

σ2= (−γ2 + γ∗1) (γ1 − γ∗1) (70)

Simplifying further, (70) becomes

−γ2 + γ∗1 > 1− γ2 − γ

γ∗1 > 1− γ

which is clearly true. Hence, we have proven equation (68). Then, since θ2 ≤ θ2 = q 1Z∗

1ζ(ij,N )

γ+γ1(γ1+γ−1) ,

Z > 0, and the assumption γ + γ2 − 1 < 0, we can show that∙λ (γ1 + γ − 1)

γ1 − γ∗1− 1

Z

¸Z∗ζ (ij,N ) θ2 + q

∙−(γ + γ1)λ

γ1 − γ∗1− γ + γ21− γ − γ2

γ1 + γ

γ1 + γ − 11

Z

¸≤

∙λ (γ1 + γ − 1)

γ1 − γ∗1− 1

Z

¸q

γ1 + γ

(γ1 + γ − 1) + q

∙−(γ + γ1)λ

γ1 − γ∗1− γ + γ21− γ − γ2

γ1 + γ

γ1 + γ − 11

Z

¸=

∙γ + γ2

γ + γ2 − 1− 1¸q1

Z

γ1 + γ

γ1 + γ − 1< 0

Hence, we showed equation (67), and hence, Av(θt,Mt,MτNt) ≤ 0.

In region II, Mt/MτNt≥ θ2 and θ2(Mt/MτNt

) ≥ θt/MτNt. A direct calculation yields

Av(θt,Mt,MτNt) = 0

since σ2

2 γ1 (γ1 − 1) + µγ1 − (ρ+ λ) = 0. It is equally straightforward to check that in region II, equation

(61) is satisfied automatically, since θt < Mt.

In region III, Mt/MτNt≤ θ1,

v(θt,Mt,MτNt) =

1

γ1 − 1q

µθt

MτNt

¶γ1θ−γ11

Since σ2

2 γ1 (γ1 − 1) + µγ1 − (ρ+ λ) = 0, we obtain

Av(θt,Mt;MτNt) = 0

and

vM (Mt,Mt;MτNt) = 0

46

Page 47: Technological Growth, Asset Pricing, and Consumption Risk

In region IV, θ1 ≤Mt/MτNt≤ θ2. Again, from

σ2

2 γ1 (γ1 − 1) + µγ1 − (ρ+ λ) = 0, we have

Av(θt,Mt,MτNt) = 0.

To verify equation (61),

vM (Mt,Mt,MτNt) = ζ (ij,N )Z

∗ γ + γ∗1 − 1γ∗1 − 1

(1− γ1)1

MτNt

+ γ1q1

Mt

=

∙ζ (ij,N )Z

∗ γ + γ∗1 − 1γ∗1 − 1

(1− γ1)Mt

MτNt

+ γ1q

¸1

Mt

≤∙ζ (ij,N )Z

∗ γ + γ∗1 − 1γ∗1 − 1

(1− γ1) θ1 + γ1q

¸1

Mt

=

∙(1− γ1) q

γ1γ1 − 1

+ γ1q

¸1

Mt

= 0

Having shown (60), (61), (62) in all regions of the state space, we are left with showing continuity and

continuous differentiability of v with respect to θt. Since both of these conditions are satisfied in the interior

of the four regions, it remains to check continuity and differentiability at the boundaries of regions I and II,

II and IV and IV and III. This is a matter of straightforward computation of left and right limits and we

leave the details of the computations out18 in order to save space.

Given the properties of v it is now straightforward to take any stopping time τ and use Ito’s formula to

obtain:

e−(ρ+λ)τv(θτ ,Mτ ;MτNt) = e−(ρ+λ)tv(θt,Mt;MτNt

) +

Z τ

t

Av(θs,Ms;MτNt)e−(ρ+λ)sds (71)

+

Z τ

t

σθsvθ(θs,Ms;MτNt)e−(ρ+λ)sdBs +

Z τ

t

vM (θs,Ms;MτNt)e−(ρ+λ)sdMs

From equation (60), (61) and (62), we have

e−(ρ+λ)tv(θt,Mt;MτNt) ≥ −Et

∙Z τ

t

Av(θs,Ms;MτNt)e−(ρ+λ)sds

¸+ Et

he−(ρ+λ)τv(θτ ,Mτ ;MτNt

)i

≥ Et

he−(ρ+λ)τv(θτ ,Mτ ;MτNt

)i≥ Et

he−(ρ+λ)τf(θτ ,Mτ ;MτNt

)i

That is,

v(θt,Mt;MτNt) ≥ Et

he−(ρ+λ)(τ−t)f(θτ ,Mτ ;MτNt

)i

Hence, the conjectured value function ν(θt,Mt;MτNt) is an upper bound for all value functions, and it is

obviously attainable by the proposed stopping rule. Hence it must be the value function of the optimal

stopping problem (56).

18Available upon request.

47

Page 48: Technological Growth, Asset Pricing, and Consumption Risk

This establishes that the stopping rule (48) is optimal. The next step is to establish that if firms follow

the threshold policies just described, then equation (49) is satisfied.

To see this let t∗ ≤ t be the first time such that θt∗ =Mt. Since in the proposed equilibrium firms plant

new trees only at19 θt =Mt, then there is no new tree planted between time t∗ and t. Hence,Z 1

0

Xj,t∗dj =

Z 1

0

Xj,tdj

Since

θt∗

Z 1

0

Xj,t∗dj ≤MCt ≤Mt

Z 1

0

Xj,tdt = θt∗

Z 1

0

Xj,t∗dj

it follows that

MCt =Mt

Z 1

0

Xj,tdt

Hence,Ct

MCt

=θtR 10Xj,tdt

Mt

R 10Xj,tdt

=θtMt

This concludes the proof of the main proposition.

Having constructed the equilibrium state price density, the rest of the verification that hCt,Kn,t,Hticonstitute an equilibrium in the sense of definition 1 is standard. The reader is referred to Basak [1999] and

the monograph of Karatzas and Shreve [1998] Chapter 4 for details.

A.1.2 Propositions and Proofs for section 4

Proposition 4 Let bγ1, bZ be given by

γ1 =

³σ2

2 − µ´+

q¡σ2

2 − µ¢2+ 2σ2(ρ+ λ

¡1−A

¢)

σ2bZ = − 1σ2

2 γ1(γ1 − 1) + µγ1 −£ρ+ λ(1− A)

¤and assume that:

γ1 > 1, bZ < 0

Then, the price of firm j in technological epoch N is given by (6) where

PAj,t = Z∗Xj,tθt

"1 +

γ

γ∗1 − 1µ

θtMt

¶γ+γ∗1−1#(72)

P oN,j,t = Z∗ANθt

"1

γ1 − 1µ

θtMt

¶γ+γ1−1µ Mt

MτN

¶γ1−1 ³ q

Z∗´ψ(ij,N )

−γ1#³1− 1{χN,j=1}

´(73)

P fN,t = Z∗ANθt

(λAZ

γ1 − 1

"µθtMt

¶γ+γ1−1− γ1 − 1

γ1 − 1µ

θtMt

¶γ+γ1−1#³ q

Z∗´µZ 1

0

ψ(i)−γ1di¶)

19Note that according to the proposed equilibrium all firms are always in Region III.

48

Page 49: Technological Growth, Asset Pricing, and Consumption Risk

The constants Z∗, γ∗1, γ1 are given in Proposition 1. Xj,t is given by (3), 1{χN,j=1} is the indicator

function used in equation (3) that takes the value 1 if firm j has planted a tree in the current epoch and 0

otherwise and ψ(ij,N ) is given by

ψ(ij,N ) ≡ Ξ

ζ(ij,N)=

q

Z∗γ1

γ1 − 1γ∗1 − 1

γ∗1 + γ − 11

ζ(ij,N )

Proof. (Proposition 4) Let PAj,t denote the value of asset in place, P

oN,j,t denote the value of current

options and P fN,t denote the value of all future update options for firm j. Then, by equation (52),

PAN,j,t =

Xj,tEt

£R∞t

e−ρ(s−t)Mγs θ

1−γs ds

¤Mγ

t θ−γt

= Xj,tZ∗∙θt +

γ

γ∗1 − 1M

1−γ−γ∗1t θ

γ+γ∗1t

¸(74)

from which (72) follows. Let τ∗N,j be optimal time for firm j to plant a new tree in the current epoch and

τN be the time when epoch N arrived. Since Mτ∗N,j = θτ∗N,j = ψ(ij,N )MτN , then the current growth option

has the following value,

P oN,j,t =

ANEt

³e−ρ(τ

∗N,j−t)Iτ∗N,j≤τNt+1

hζ (ij,N )

R∞τ∗N,j

e−ρ(s−τ∗N,j)Mγ

s θ1−γs ds− qMτNM

γτ∗N,j

θ−γτ∗N,j

i´Mγ

t θ−γt

³1− 1{χN,j=1}

´= ANEt

½e−(ρ+λ)(τ

∗N,j−t)

∙ζ (ij,N )Z

∗ γ + γ∗1 − 1γ∗1 − 1

ψ(ij,N )− q

¸MτN

¾M−γt θγt

³1− 1{χN,j=1}

´= ANq

1

γ1 − 1MτNM

−γt θγtEt

he−(ρ+λ)(τ

∗N,j−t)

i ³1− 1{χN,j=1}

´where the second to last equality follows from ψ(ij,N ) =

qZ∗

γ1γ1−1

γ∗1−1γ∗1+γ−1

1ζ(ij,N )

. Since (see page 197 of

Karatzas and Shreve [1991])

Et

he−(ρ+λ)(τ

∗N,j−t)

i=

µψ(ij,N )

θt/MτN

¶−γ1then, after some rearrangement, we obtain

P oN,j,t = ANq

1

γ1 − 1MτNM

−γt θγt

µψ(ij,N )

θt/MτN

¶−γ1 ³1− 1{χN,j=1}

´= Z∗ANθt

"1

γ1 − 1µ

θtMt

¶γ+γ1−1µ Mt

MτN

¶γ1−1 ³ q

Z∗´ψ(ij,N )

−γ1#³1− 1{χN,j=1}

´(75)

which is exactly equation (73).

Finally, since Mτ∗N,j = θτ∗N,j = ψ(ij,N )MτN , the value of all the future growth options is,

P fN,j,t =

E³Et

hP∞n=N+1 e

−ρ(τ∗n,j−t)Iτ∗n,j≤τn+1An³A(ij,n)

R∞τ∗n,j

e−ρ(s−τ∗n,j)Mγ

s θ1−γs ds− qMτnM

γτ∗n,j

θ−γτ∗n,j

´i´Mγ

t θ−γt

=

P∞n=N+1 A

nE³Et

he−ρ(τ

∗n,j−t)Iτ∗n,j≤τn+1

³A(ij,n)

R∞τ∗n,j

e−ρ(s−τ∗n,j)Mγ

s θ1−γs ds− qMτn

´i´Mγ

t θ−γt

49

Page 50: Technological Growth, Asset Pricing, and Consumption Risk

where E denote the expectation with respect to the unknown draws ij,n in future periods. To calculate

P fN,j,t, we first find the following conditional expectation for any n > N and any given ζ (ij,n) :

Et

"e−ρ(τ

∗n,j−t)Iτ∗n,j≤τn+1

Ãζ (ij,n)

Z ∞τ∗n,j

e−ρ(s−τ∗n,j)Mγ

s θ1−γs ds− qMτn

!#

= Et

"Eτ∗n,j

"e−ρ(τ

∗n,j−t)Iτ∗n,j≤τn+1

Ãζ (ij,n)

Z ∞τ∗n,j

e−ρ(s−τ∗n,j)Mγ

s θ1−γs ds− qMτn

!##

= Et

³e−(ρ+λ)(τ

∗n,j−τn)−ρ(τn−t)Mτn

´ ∙ζ (ij,n)Z

∗ γ + γ∗1 − 1γ∗1 − 1

ψ(ij,n)− q

¸= q

1

γ1 − 1Et

³e−(ρ+λ)(τ

∗n,j−τn)−ρ(τn−t)Mτn

´= q

1

γ1 − 1Et

õψ(ij,n)

θτn/Mτn

¶−γ1e−ρ(τn−t)Mτn

!= q

1

γ1 − 1ψ(ij,n)

−γ1Et

he−ρ(τn−t)θγ1τnM

1−γ1τn

iwhere the second to last equality follows from

Eτn

³e−(ρ+λ)(τ

∗n,j−τn)

´=

µψ(ij,n)

θτn/Mτn

¶−γ1τn ≥ t, and the property of iterated conditional expectation. Hence,

P fN,j,t =

P∞n=N+1 A

nE³q 1γ1−1ψ(ij,n)

−γ1Et

he−ρ(τn−t)θγ1τnM

1−γ1τn

i´Mγ

t θ−γt

=q

γ1 − 1µ

θtMt

¶γAN E

¡ψ(ij,n)

−γ1¢ ∞Xn=N+1

An−NEt

he−ρ(τn−t)θγ1τnM

1−γ1τn

i=

q

γ1 − 1µ

θtMt

¶γAN

µZ 1

0

ψ(i)−γ1di¶ ∞X

n=N+1

An−NEt

he−ρ(τn−t)θγ1τnM

1−γ1τn

i≡ q

γ1 − 1µ

θtMt

¶γAN

µZ 1

0

ψ(i)−γ1di¶· P f

N,j,t (θt,Mt)

where

P fN,j,t (θt,Mt) ≡

∞Xn=N+1

An−NEt

he−ρ(τn−t)θγ1τnM

1−γ1τn

i= (76)

= Et

Ãe−ρ(τN+1−t)

∞Xn=N+1

An−Nhe−ρ(τn−τN+1)θγ1τnM

1−γ1τn

i!(77)

Now, we only need to find P fN,j,t. It will be easiest to re-express P

fN,j,t in recursive form:

P fN,j,t (θt,Mt) = Et

³e−ρ(τN+1−t)

hAθγ1τN+1

M1−γ1τN+1

+ AP fN,j,t (θt,Mt)

i´(78)

We shall drop the subscripts in P fN,j,t to avoid excessive notation. In light of (78), the Bellman equation for

50

Page 51: Technological Growth, Asset Pricing, and Consumption Risk

P f is

0 =σ2

2θ2t P

fθθ + µθtP

fθ − (ρ+ λ) P f + λ

hAθ

γ1t M

1−γ1t + AP f

i=

σ2

2θ2t P

fθθ + µθtP

fθ −

£ρ+ λ− λA

¤P f + λA

hθγ1t M

1−γ1t

iThe solution to this second order ODE20 is

P fN,j,t (θt,Mt) = λAZθ

γ1t M

1−γ1t + θ

γ1t F (Mt)

where γ1 is the positive root of the following equation (note: γ1 > γ∗1 > γ1 )

σ2

2γ2 +

µµ− σ2

2

¶γ − £ρ+ λ− λA

¤= 0

and

Z =−1

σ2

2 γ1(γ1 − 1) + µγ1 −£ρ+ λ− λA

¤To determine F (Mt) we shall use a standard reflection condition at θt =Mt, namely:

P fM (θt,Mt) = 0 at θt =Mt

Employing this boundary condition yields

FM (Mt) = (γ1 − 1)λAZM−γ1t

Integrating:

F (Mt) = (γ1 − 1)λAZZ ∞Mt

M−γ1t

together with the assumption γ1 > 1 gives:

F (Mt) = −λAZ γ1 − 1γ1 − 1

M1−γ1t

That is,

P fN,j,t (θt,Mt) = λAZθ

γ1t M

1−γ1t − λAZ

γ1 − 1γ1 − 1

θγ1t M

1−γ1t

= λAZθγ1t M

1−γ1t

∙1− γ1 − 1

γ1 − 1θγ1−γ1t M

γ1−γ1t

¸(79)

Hence, the value of all the future growth options is

P fN,j,t =

q

γ1 − 1µ

θtMt

¶γAN

µZ 1

0

ψ(i)−γ1di¶· P f

N,j,t (θt,Mt) =

=q

γ1 − 1µ

θtMt

¶γAN

µZ 1

0

ψ(i)−γ1di¶λAZθ

γ1t M

1−γ1t

∙1− γ1 − 1

γ1 − 1θγ1−γ1t M

γ1−γ1t

¸(80)

= Z∗ANθt

(λAZ

γ1 − 1

"µθtMt

¶γ+γ1−1− γ1 − 1

γ1 − 1µ

θtMt

¶γ+γ1−1#³ q

Z∗´µZ 1

0

ψ(i)−γ1di¶)

(81)

20We have excluded solutions that explode near θt = 0.

51

Page 52: Technological Growth, Asset Pricing, and Consumption Risk

This completes the proof of the proposition.

Proof. (Lemma 1) The value of the entire stock market is:

PN,t =

Z 1

0

PN,j,tdj =

Z 1

0

PN,j,tdj =

Z 1

0

PAj,tdj +

Z 1

0

P oN,j,tdj + P f

N,t

Using Proposition 4, the first term is equal to:Z 1

0

PAj,tdj = Z∗Xtθt

"1 +

γ

γ∗1 − 1µ

θtMt

¶γ+γ∗1−1#

where Xt is given in (8). Similarly:Z 1

0

P oN,j,tdj = Z∗ANθt

"1

γ1 − 1µ

θtMt

¶γ+γ1−1µ Mt

MτN

¶γ1−1 ³ q

Z∗´ÃZ 1

KN,t

ψ(i)−γ1di

!#

Substituting these formulas into the definition of wo+ft , using the definition of xt in equation (28), and

simplifying gives the result with f given by:

f =

∙1 + γ

γ∗1−1³θtMt

´γ+γ∗1−1¸Γ

and Γ is given by:

Γ =

"1

γ1 − 1µ

θtMt

¶γ+γ1−1µ Mt

MτN

¶γ1−1 ³ q

Z∗´ÃZ 1

KN,t

ψ(i)−γ1di

!#

+λAZ

γ1 − 1

"µθtMt

¶γ+γ1−1− γ1 − 1

γ1 − 1µ

θtMt

¶γ+γ1−1#³ q

Z∗´µZ 1

0

ψ(i)−γ1di¶

Note that KN,t is a function of Mt

MτNonly, and hence Γ itself is a function of θt

Mt, Mt

MτN.

Proof. (Lemma 2) A feature of the model is that risk aversion is constant21 . Hence the Sharpe ratio is

a constant equal to γσ, and the instantaneous excess return on any asset is given by:

µ(j) − r = γσσj,t

where:

σj,t =

µσθtPN,j,t

¶∂PN,j,t∂θt

Hence in order to establish the result is suffices to compare the instantaneous volatility of assets in place,

current epoch growth options and future growth options. Using Proposition 4 and starting with the volatility

21The running maximum of a diffusion has no quadratic variation, and hence all of the quadratic variation of the

pricing kernel is driven by variations in θt and not Mt.

52

Page 53: Technological Growth, Asset Pricing, and Consumption Risk

of asset in place, we obtain:

σAj,t =

ÃσθtPAN,j,t

!∂PA

N,j,t

∂θt

= σ [w1 + (1− w1) (γ + γ∗1)]

= σ [(γ + γ∗1) + w1 (1− γ − γ∗1)] (82)

where

w1 =θt

θt +γ

γ∗1−1M1−γ−γ∗1t θ

γ+γ∗1t

=γ∗1 − 1

γ∗1 − 1 + γ³Mt

θt

´1−γ−γ∗1It is clear that 1 ≥ w1 ≥ 0. For the volatility of the current epoch growth option, we obtain similarly

σoj,t =

ÃσθtP oN,j,t

!∂P o

N,j,t

∂θt

= σ (γ + γ1) (83)

For the volatility of future options,

σfj,t =

Ãσθt

P fN,j,t

!∂P f

N,j,t

∂θt

= σ [w3 (γ + γ1) + (1− w3) (γ + γ1)]

= σ [(γ + γ1) + w3 (γ1 − γ1)] (84)

where

w3 =θγ+γ1t M

1−γ−γ1t

θγ+γ1t M

1−γ−γ1t − γ1−1

γ1−1θγ+γ1t M

1−γ−γ1t

=1

1− γ1−1γ1−1

³Mt

θt

´γ1−γ1Since γ1 ≥ γ1 and w3 ≤ 0, equation (83) and (84) implies

σfj,t ≤ σ (γ + γ1) < σ (γ + γ1) = σoj,t

To develop the relation between the volatility of assets in place and future growth options, we define the

function d (x) as follows,

d

µMt

θt

¶≡ σAj,t − σfj,t

σ

= [(γ + γ∗1) + w1 (1− γ − γ∗1)]− [(γ + γ1) + w3 (γ1 − γ1)]

= (γ∗1 − γ1) +γ∗1 − 1

γ∗1 − 1 + γ³Mt

θt

´1−γ−γ∗1 (1− γ − γ∗1)−γ1 − 1

γ1 − 1− (γ1 − 1)³Mt

θt

´γ1−γ1 (γ1 − γ1)

53

Page 54: Technological Growth, Asset Pricing, and Consumption Risk

It is clear that

d (1) = (γ∗1 − γ1) +γ∗1 − 1

γ∗1 − 1 + γ(1− γ − γ∗1)−

γ1 − 1γ1 − 1− (γ1 − 1)

(γ1 − γ1)

= 0

and

d0 (x) = − (γ∗1 − 1) γ(γ∗1 − 1 + γx1−γ−γ∗1 )2

(1− γ − γ∗1)2 − (γ1 − 1) (γ1 − 1)

[γ1 − 1− (γ1 − 1)xγ1−γ1 ]2(γ1 − γ1)

2

< 0

It follows that d (x) < 0 for all x > 1. That is, σAj,t ≤ σfj,t and the equality holds only if θt =Mt. Therefore,

we have

σAj,t ≤ σfj,t < σoj,t

and the equality holds only if θt =Mt.

54

Page 55: Technological Growth, Asset Pricing, and Consumption Risk

References

Abel, A. B. (1990): “Asset Prices under Habit Formation and Catching Up with the Joneses,” American

Economic Review, 80(2), 38—42.

Abel, A. B., and J. C. Eberly (2002): “Q for the Long Run,” mimeo. The Wharton School and Kellog.

(2003): “Investment, Valuation, and Growth Options,” mimeo. The Wharton School and Kellog.

(2004): “Q Theory Without Adjustment Costs Cash Flow Effects Without Financing Constraints,”

mimeo. The Wharton School and Kellog.

Atkeson, A., and P. J. Kehoe (1993): “Industry evolution and transition: the role of information

capital,” Federal Reserve Bank of Minneapolis, Staff Report: 162, 1993.

(1999): “Models of Energy Use: Putty-Putty versus Putty-Clay,” American Economic Review,

89(4), 1028—1043.

Bansal, R., R. Dittmar, and D. Kiku (2005): “Long Run Risks and Equity Returns,” Duke University,

Working Paper.

Bansal, R., R. Dittmar, and C. Lundblad (2004): “Consumption, Dividends and the Cross-Section of

Equity Returns,” forthcoming, Journal of Finance.

Bansal, R., and A. Yaron (2004): “Risks for the Long Run: A Potential Resolution of Asset Pricing

Puzzles,” Journal of Finance, 59(4), 1481—1509.

Basak, S. (1999): “On the Fluctuations in Consumption and Market Returns in the Presence of Labor

and Human Capital: An Equilibrium Analysis,” Journal of Economic Dynamics and Control, 23(7),

1029—1064.

Basu, S., J. G. Fernald, and M. S. Kimball (2002): “Are Technology Improvements Contractionary?,”

FRB of Chicago Working Paper No. 2004-20.

Berk, J. B., R. C. Green, and V. Naik (1999): “Optimal Investment, Growth Options, and Security

Returns,” Journal of Finance, 54(5), 1553—1607.

(2004): “Valuation and Return Dynamics of New Ventures,” Review of Financial Studies, 17(1),

1—35.

Caballero, R. J., and E. M. R. A. Engel (1991): “Dynamic (S, s) Economies,” Econometrica, 59(6),

1659—1686.

(1999): “Explaining Investment Dynamics in U.S. Manufacturing: A Generalized (S, s) Approach,”

Econometrica, 67(4), 783—826.

55

Page 56: Technological Growth, Asset Pricing, and Consumption Risk

Caballero, R. J., and R. S. Pindyck (1996): “Uncertainty, Investment, and Industry Evolution,”

International Economic Review, 37(3), 641—662.

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(2), 205—251.

Carlson, M., A. Fisher, and R. Giammarino (2004): “Corporate Investment and Asset Price Dynamics:

Implications for the Cross Section of Returns,” Journal of Finance, 59(6), 2577—2603.

Chan, Y. L., and L. Kogan (2002): “Catching Up with the Joneses: Heterogeneous Preferences and the

Dynamics of Asset Prices,” Journal of Political Economy, 110(6), 1255—1285.

Cochrane, J. H. (1996): “A Cross-Sectional Test of an Investment-Based Asset Pricing Model,” Journal

of Political Economy, 104(3), 572—621.

(2005): Asset pricing. Princeton University Press, Princeton and Oxford.

Comin, D., and M. Gertler (2003): “Medium Term Business Cycles,” C.V. Starr Center for Applied

Economics, New York University, Working Papers.

Cooper, I. (2004): “Asset Pricing Implications of Non-convex Adjustment Costs and Irreversibility of

Investment,” Journal of Finance, forthcoming.

Dupor, B., and W.-F. Liu (2003): “Jealousy and Equilibrium Overconsumption,” American Economic

Review, 93(1), 423—428.

Fama, E. F., and K. R. French (1992): “The Cross-Section of Expected Stock Returns,” Journal of

Finance, 47(2), 427—465.

(1995): “Size and Book-to-Market Factors in Earnings and Returns,” Journal of Finance, 50(1),

131—155.

Gala, V. D. (2005): “Investment and Returns,” Working Paper, Chicago GSB.

Gomes, J., L. Kogan, and L. Zhang (2003): “Equilibrium Cross Section of Returns,” Journal of Political

Economy, 111(4), 693—732.

Gourio, F. (2004): “Operating Leverage, Stock Market Cyclicality, and the Cross-Section of Returns,”

Working Paper, Chicago GSB.

Greenwood, J., and B. Jovanovic (1999): “The Information-Technology Revolution and the Stock

Market,” American Economic Review, 89(2), 116—122.

Greenwood, J., and M. Yorukoglu (1997): “1974,” Carnegie Rochester Conference Series on Public

Policy, 46, 49—95.

56

Page 57: Technological Growth, Asset Pricing, and Consumption Risk

Hamilton, J. D. (1994): Time series analysis. Princeton University Press, Princeton.

Hansen, L. P., J. C. Heaton, and N. Li (2005): “Consumption Strikes Back?: Measuring Long Run

Risk,” Working Paper, University of Chicago.

Helpman, E. (1998): General purpose technologies and economic growth. MIT Press, Cambridge and Lon-

don.

Jermann, U., and V. Quadrini (2002): “Stock Market Boom and the Productivity Gains of the 1990s,”

NBER Working Papers: 9034.

Jovanovic, B., and G. M. MacDonald (1994): “The Life Cycle of a Competitive Industry,” Journal of

Political Economy, 102(2), 322—347.

Jovanovic, B., and P. L. Rousseau (2003): “Two Technological Revolutions,” Journal of the European

Economic Association, 1(2-3), 419—428.

(2004): “Interest Rates and Initial Public Offerings,” NBER Working Papers: 10298.

Karatzas, I., and S. E. Shreve (1991): Brownian motion and stochastic calculus, vol. 113 of Graduate

Texts in Mathematics. Springer-Verlag, New York, second edn.

(1998): Methods of mathematical finance, vol. 39. Springer-Verlag.

Kogan, L. (2001): “An Equilibrium Model of Irreversible Investment,” Journal of Financial Economics,

62(2), 201—245.

(2004): “Asset Prices and Real Investment,” Journal of Financial Economics, 73(3), 411—431.

Larrain, B., and M. Yogo (2005): “Does Firm Value Move Too Much to be Justified by Subsequent

Changes in Cash Flow?,” Working Paper, The Wharton School.

Lettau, M., and S. C. Ludvigson (2005): “Expected Returns and Expected Dividend Growth,” Journal

of Financial Economics, 76(3), 583—626.

Menzly, L., T. Santos, and P. Veronesi (2004): “Understanding Predictability,” Journal of Political

Economy, 112(1), 1—47.

Novy-Marx, R. (2003): “An equilibrium model of investment under uncertainty,” mimeo. Chicago GSB.

Øksendal, B. (2003): Stochastic differential equations. Springer-Verlag.

Parker, J. A., and C. Julliard (2005): “Consumption Risk and the Cross Section of Expected Returns,”

Journal of Political Economy, 113(1), 185—222.

57

Page 58: Technological Growth, Asset Pricing, and Consumption Risk

Pastor, L., and P. Veronesi (2004): “Was There a Nasdaq Bubble in the Late 1990s?,” NBER Working

Papers: 10581.

(2005): “Technological Revolutions and Stock Prices,” CRSP Working Paper No. 606.

Pontiff, J., and L. D. Schall (1998): “Book-to-Market Ratios as Predictors of Market Returns,” Journal

of Financial Economics, 49(2), 141—160.

Vigfusson, R. J. (2004): “The Delayed Response To A Technology Shock. A Flexible Price Explanation,”

Working paper, The Federal Reserve Board.

Zhang, L. (2005): “The Value Premium,” Journal of Finance, 60(1), 67—103.

58