Vous êtes sur la page 1sur 34

Journal of Financial Economics 118 (2015) 211244

Contents lists available at ScienceDirect

Journal of Financial Economics


journal homepage: www.elsevier.com/locate/jfec

Regression-based estimation of dynamic asset


pricing models$
Tobias Adrian n, Richard K. Crump, Emanuel Moench
Federal Reserve Bank of New York, 33 Liberty Street, New York, NY 10045, United States

a r t i c l e i n f o abstract

Article history: We propose regression-based estimators for beta representations of dynamic asset pricing
Received 11 April 2013 models with an affine pricing kernel specification. We allow for state variables that are
Received in revised form cross-sectional pricing factors, forecasting variables for the price of risk, and factors that
19 June 2014
are both. The estimators explicitly allow for time-varying prices of risk, time-varying
Accepted 27 June 2014
Available online 30 July 2015
betas, and serially dependent pricing factors. Our approach nests the Fama-MacBeth two-
pass estimator as a special case. We provide asymptotic multistage standard errors
JEL classification: necessary to conduct inference for asset pricing tests. We illustrate our new estimators in
G10 an application to the joint pricing of stocks and bonds. The application features strongly
G12
time-varying, highly significant prices of risk that are found to be quantitatively more
C58
important than time-varying betas in reducing pricing errors.
& 2015 Elsevier B.V. All rights reserved.
Keywords:
Dynamic asset pricing
Fama-MacBeth regressions
Time-varying betas
Generalized method of moments
Minimum distance estimation
Reduced rank regression

1. Introduction prices of risk. The estimators and associated standard


errors are computationally as simple as Fama-MacBeth
Overwhelming evidence exists that risk premiums vary regressions, but they explicitly provide estimates of time-
over time (Campbell and Shiller, 1988; Cochrane, 2011). varying prices of risk, as well as estimates of the associated
Yet, widely used empirical asset pricing methods such as state variable dynamics. Our model combines key assump-
Fama and MacBeth (1973) two-pass regressions rely on the tions of the dynamic asset pricing models from fixed
assumption that prices of risk are constant. income applications with the computational ease of
This paper proposes regression-based estimators for Fama-MacBeth regressions that are popular in empirical
dynamic asset pricing models (DAPMs) with time-varying equity market research. The setup can also be viewed as a


We are grateful to the referee, Stijn Van Nieuwerburgh, for a number of helpful suggestions that substantially improved the paper. We would like to
thank Andrew Ang, Matias Cattaneo, Fernando Duarte, Darrell Duffie, Robert Engle, Arturo Estrella, Andreas Fuster, Eric Ghysels, Benjamin Mills, Monika
Piazzesi, Karen Shen, Michael Sockin, Jonathan Wright, seminar participants at the Federal Reserve Bank of New York, the National Bureau of Economic
Research (NBER) Summer Institute, and the Verein fr Socialpolitik for helpful comments and discussions. A special thanks goes to Wayne Ferson for his
detailed comments as our discussant at the NBER Asset Pricing meeting. Daniel Green, Ariel Zucker, and Benjamin Mills provided excellent research
assistance. The views expressed in this paper are those of the authors and do not necessarily represent those of the Federal Reserve Bank of New York or
the Federal Reserve System.
n
Corresponding author.
E-mail addresses: tobias.adrian@ny.frb.org (T. Adrian), richard.crump@ny.frb.org (R.K. Crump), emanuel.moench@ny.frb.org (E. Moench).

http://dx.doi.org/10.1016/j.jfineco.2015.07.004
0304-405X/& 2015 Elsevier B.V. All rights reserved.
212 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

reduced form representation of dynamic macro-finance pioneered by Robinson (1989). Kernel regressions have the
models with time-varying prices of risk. appealing feature of nesting least squares rolling window
We distinguish three different types of aggregate regressions which are often used in the empirical litera-
state variables: risk factors, price of risk factors, and ture (see, for example, Fama and French, 1997; Lewellen
factors that are both. By risk factors, we refer to variables and Nagel, 2006; among many others). In our implemen-
that are significant factors for the cross section of asset tation, however, we use a Gaussian kernel estimator with
returns, i.e. they have nonzero betas. By price of risk data-driven bandwidth choice following Ang and
factors, we refer to variables that significantly forecast Kristensen (2012).
the time series variation of excess returns but do not The affine price of risk specification we use closely
necessarily have nonzero betas. Prices of risk are resembles affine term structure models.1 Our approach
assumed to be affine functions of price of risk factors. thus lends itself to asset pricing applications across differ-
We show that by introducing this risk price specification ent asset classes. We present an empirical application for
into generic asset pricing models, one can derive simple the cross section of size-sorted equity portfolios and
regression-based estimators for all model parameters maturity-sorted Treasury portfolios. We show that a par-
that are consistent and asymptotically normal under simonious model with two pricing factors, two price of
mild conditions. risk factors, and one factor that serves both roles fits this
Our baseline estimator is a three-step regression that cross section of test assets very well on average, while, at
can be described as follows. In the first step, shocks to the the same time, giving rise to strongly significant time
state variables are obtained from a time series vector variation in risk premiums. We further find that allowing
autoregression (VAR). In the second step, asset returns for time variation in prices of risk is more important than
are regressed in the time series on lagged price of risk modeling time variation in factor risk exposures in terms
factors and the contemporaneous innovations to the cross of minimizing squared pricing errors of the model. In our
sectional pricing factors, generating predictive slopes and application, traditional estimation approaches such as the
risk betas for each test asset. In the third step, price of risk one by Fama and MacBeth (1973) and Ferson and Harvey
parameters are obtained by regressing the constant and (1991) imply substantially larger pricing errors than the
the predictive slopes from the time series regression on estimators we propose.
the betas cross-sectionally. We give asymptotic variance The remainder of the paper is organized as follows.
formulas that allow for conditional heteroskedasticity and Section 2 provides a discussion of the contribution of this
correct for the additional estimation uncertainty arising paper to the existing literature. We present the dynamic
from using generated regressors. asset pricing model in Section 3. We discuss estimation
We show that this three-step estimator coincides with and inference when betas are assumed to be constant in
the Fama-MacBeth estimator when two conditions are Section 4. In Section 4.1, we formally present the link of
met. First, state variables have to be uncorrelated across the dynamic asset pricing estimator to the static Fama-
time. Second, prices of risk have to be constant. Our MacBeth estimator, and we explain the contributions of
approach can thus be viewed as a dynamic version of the our results to the existing literature in detail. In Section 5,
Fama-MacBeth estimator, nesting the popular uncondi- we derive the corresponding estimator under the assump-
tional estimator as a special case. tion that betas vary over time. We illustrate our estimators
We also introduce an additional (quasi-) maximum in an empirical application in Section 6. Section 7
likelihood estimator (QMLE). This estimator is replacing concludes.
the third regression step with a simple eigenvalue decom-
position. The QMLE is asymptotically equivalent to the
three step regression estimator even in the case of condi- 2. Related literature
tional heteroskedasticity in the return errors. We show
that in our model generalized method of moments (GMM) Our approach can be seen as a generalization of the
and minimum distance (MD) estimation are exactly static Fama and MacBeth (1973) cross sectional asset
equivalent and that the QMLE is a special case of this pricing approach to dynamic asset pricing models. The
more general class of estimation approaches for certain empirical applications of the static Fama-MacBeth
choices of weighting matrix. approach are too numerous to list, but some of the seminal
While our main results are extensions of classic results works are Chen, Roll, and Ross (1986) and Fama and
in the cross-sectional pricing literature to a dynamic French (1992).
setting, we provide new interpretations of results in the Some previous authors have extended the Fama-
model when prices of risk are constant. For example, the MacBeth approach to conditional asset pricing models.
equivalence between GMM and MD estimation implies Ferson and Harvey (1991) use period-by-period Fama-
that the cross sectional T 2 statistic of Shanken (1985) could MacBeth regressions to obtain estimates of time-varying
be directly interpreted as a J-test for the moment restric- market prices of risk, which they then regress on lagged
tions of the static model. conditioning variables. They find evidence for predict-
We also extend the three-step regression estimator to able variation in prices of risk and associate most of the
the case where betas and the parameters in the vector
autoregression of the state variables are time-varying. We 1
For regression-based approaches to term structure models featur-
assume that these parameters evolve smoothly over time ing an exponentially affine pricing kernel, see Adrian, Crump, and
and estimate them using a kernel regression approach Moench (2013) and Abrahams, Adrian, Crump, and Moench (2014).
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 213

predictable variation in stock returns to time variation in observable variables, see, for example, Shanken (1990),
risk compensation instead of time variation in betas. Our Ferson and Harvey (1999), and, recently, Gagliardini,
estimation approach generalizes the one used in Ferson Ossola, and Scaillet (2014) and Chordia, Goyal, and
and Harvey (1991) by allowing for estimation in the Shanken (2013). A drawback to this approach is that it
presence of serially correlated pricing factors and expli- requires the correct specification for the functional form of
citly incorporating time variation of prices of risk. In the betas. As pointed out by Ghysels (1998) and Harvey
addition, we provide asymptotic standard errors for all (2001), models with misspecified betas often feature
parameters of the model taking into account the uncer- larger pricing errors than models with constant betas. In
tainty generated at each step of the estimation. contrast, the kernel estimator that we use imposes less
Jagannathan and Wang (1996), Lettau and Ludvigson structure than assuming a specific functional form for the
(2001), and others have used the Fama-MacBeth tech- parameters and, therefore, is likely more robust to mis-
nology to estimate scaled factor models. The beta repre- specification. Moreover, we show that our Gaussian kernel
sentations of such models are nested in our more estimator yields smaller pricing errors than simple rolling
general framework. Moreover, in contrast to our pro- window regressions for both specifications with constant
posed estimators, the scaled factor approaches typically and time-varying prices of risk.
do not explicitly provide estimates for the price of risk We provide a further comparison of our results to
parameters and the number of parameters grows quickly the existing literature throughout the remainder of
with the number of factors. the paper.
Our paper is further related to Balduzzi and Robotti
(2010), who estimate time-varying risk premiums for 3. Pricing kernel and return generating process
maximum-correlation portfolios, i.e., portfolios result-
ing from the projection of a candidate pricing kernel on Before describing the model, it is convenient to intro-
the set of test assets. Moreover, Gagliardini, Ossola, and duce the following notations that are used throughout the
Scaillet (2014) and Chordia, Goyal, and Shanken (2013) paper. The symbol  represents the Kronecker product;
present alternative estimation approaches for models vec, the vectorization operator. Im and n denote the
with time-varying risk premiums using Fama-MacBeth- m  m identity matrix and a n  1 column vector of ones,
type estimators when both the number of assets and respectively. Moreover, let 1 2  be the matrix formed
the number of time series observations tend to infinity. by appending the columns of the matrix 2 to the columns
Ang, Liu, and Schwarz (2010) study the implications for of the matrix 1. Finally, throughout the paper, equalities
efficiency of using individual stocks versus portfolios in involving conditional expectations are understood to hold
estimating cross-sectional pricing models. Finally, almost surely.
another strand of the literature investigates the impli- We assume that systematic risk in the economy is
cations of model misspecification in cross sectional captured by a K  1 vector of state variables Xt that follow
asset pricing models. For example, Kan, Robotti, and a stationary vector autoregression:
Shanken (2013) derive the asymptotic distribution of
the cross-sectional R2 and develop model comparison X t 1 X t vt 1 ; t 0; ; T  1; 1
tests which accommodate model misspecification. with initial condition X0. The dynamics of these state
Here, instead, we assume that the model is correctly variables can be assumed to be generated by an equili-
specified. brium model of the macroeconomy.
Our empirical application is closest to Ferson and The state variables can be risk factors, price of risk
Harvey (1991) and Campbell (1996), who use similar test factors, or both. By risk factors, we refer to variables that
assets and similar pricing factors in models with time- are significant factors for the cross section. By price of risk
varying and constant prices of risk, respectively. A number factors, we refer to variables that significantly forecast the
of recent papers estimate dynamic pricing kernels for the time variation of excess returns.2 While some state vari-
cross section of stocks and bonds (see, for example, ables act as both price of risk and risk factors, many
Mamaysky, 2002; Bekaert, Engstrom, and Grenadier, commonly used state variables act exclusively as one or
2010; Lettau and Wachter, 2011; Ang and Ulrich, 2012; the other. This setup thus nests that of Campbell (1996),
and Koijen, Lustig, and van Nieuwerburgh, 2013). What who argues that innovations in variables that have been
distinguishes our approach from that literature is the shown to forecast stock returns should be used in cross-
regression-based estimation methodology, which is simple sectional asset pricing studies.
to implement, is computationally robust, and allows for As a consequence, we partition the state variables
standard specification tests. We show that our empirical into three categories: X 1;t A RK 1 : risk factor only,
application features good pricing properties across stocks X 2;t A RK 2 : risk and price of risk factor, and X 3;t A RK 3 :
and bonds and that it implies notable time variation of price of risk factor only. In Section 6 we use all three
expected returns associated with highly significant types of factors in an application investigating the cross
dynamic price of risk parameters. Moreover, the dynamic
asset pricing model that we estimate yields substantially
2
smaller mean squared pricing errors than several alter- Variables which predict excess returns but are not contempora-
neously correlated with excess returns are sometimes referred to as
native models with constant prices of risk. unspanned factors. For applications to affine term structure models
Some prior literature on conditional factor pricing with unspanned factors see, for example, Joslin, Priebsch, and Singleton
models has assumed that betas are (linear) functions of (2012) or Adrian, Crump, and Moench (2013).
214 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

section of equity and bond returns. For simplicity of By definition of i;t ;


notation, we define 1   
" # " # " # i;t u;t C C t 1 ; Ri;t 1 F t i;t ; 12
X 1;t X 2;t v1;t
Ct ; Ft ; ut ; 2 so that
X 2;t X 3;t v2;t
0   0
Ri;t 1 i;t 0 1 F t i;t ut 1 ei;t 1 : 13
where Ct is for cross section and Ft is for forecasting. Let
K C K 1 K 2 , K F K 2 K 3 , and K K 1 K 2 K 3 . We The excess returns, Ri;t 1 , thus depend on the expected
0  
assume that excess return, i;t 0 1 F t , the component that is con-
     
E vt 1 F t 0; V vt 1 F t v;t ; 3 ditionally correlated with the innovations to the risk
0
factors, i;t ut 1 , and a return pricing error, ei;t 1 , that is
where F t denotes the information set at time t. We conditionally orthogonal to the risk factor innovations.
denote holding period returns in excess of the risk free Therefore, the innovations to the pricing factors Ct capture
rate of asset i by Ri;t 1 . We assume the existence of a systematic risk exposure, and the levels of the price of risk
pricing kernel M t 1 such that factors Ft are forecasting variables.
  
E M t 1 Ri;t 1 F t 0: 4 Previous approaches have been taken to model the
time variation in risk premiums in equity returns (for
Moreover, we assume that the pricing kernel has the example, in Gibbons and Ferson, 1985; Campbell, 1987;
linear form Ferson and Harvey, 1991; Lettau and Ludvigson, 2001;
  
M t 1  E M t 1 F t 0  1=2 among others). However, most, if not all, of these
    t u;t ut 1 ; 5
E Mt 1 F t approaches can be viewed as special cases of our more
general framework which has been derived from first
where t is the K C  1 vector of period t prices of risk and principles. Affine prices of risk are also commonly used
where the K C  K C matrix u;t is the conditional variance in the fixed income literature, see e.g., Duffee (2002), Dai
of ut 1 . It is important to point out that the above form and Singleton (2002), or Ang and Piazzesi (2003).
for the pricing  kernel incorporates that the covariance The system of equations (13) for i 1; ; N embeds the
 
C Ri;t 1 ; v3;t 1 F t 0 for all t. The same restriction is no-arbitrage restrictions that were derived from the form of
imposed in term structure models that feature the pricing kernel introduced in Eq. (5). Relative to a see-
unspanned factors. mingly unrelated regressions (SUR) model in which
As in Duffee (2002), we assume that prices of risk are 0
Ri;t 1 ai;t ci;t F t i;t ut 1 ei;t 1 , the assumption of no-
affine functions of the price of risk factors Ft, so that 0 0
arbitrage implies ai;t i;t 0 and ci;t i;t 1 . These are
 1=2   reduced rank restrictions resulting in a smaller number of
t u;t 0 1 F t ; 6
parameters to estimate. To the extent that the model is well
where 0 is a K C  1 vector, 1 is a K C  K F matrix, and specified, then the parameter restrictions imposed by no-
 
0 1 has full row rank. We then find the following arbitrage help in increasing the predictive accuracy for the
beta representation of expected returns: entire cross-section of excess returns. Hence, in our dynamic
   CM t 1 ; Ri;t 1 jF t  asset pricing model, a clear connection exists between the
E Ri;t 1 F t    
E M t 1 F t cross-sectional pricing performance and the predictive ability
0  1=2   of a given set of model factors.
t u;t C ut 1 ; Ri;t 1 jF t
 0  1   Standard, static cross-sectional asset pricing models
0 1 F t u;t C C t 1 ; Ri;t 1 jF t : 7 make two additional assumptions: 1 0 in Eq. (13), and
0 in Eq. (1) (see the reviews by Campbell, Lo, and
Thus,
   MacKinlay, 1997 and Cochrane, 2005). We consider these
0  
E Ri;t 1 F t i;t 0 1 F t ; 8 special cases in the following sections. However, the main
contribution of this paper is to study the dynamic case in
where i;t is a (time-varying) KC-dimensional exposure
which a0 and 1 a0.
vector,
While the focus of this paper is the estimation of the beta
1  
i;t u;t C C t 1 ; Ri;t 1 jF t 9 representation of dynamic asset pricing models, an extensive
literature estimates the stochastic discount factor (SDF) repre-
We can then decompose excess returns into an expected sentation using GMM (Hansen,
and an unexpected component:    1982). In that literature, the
expression E M t 1 Ri;t 1 F t 0 is estimated directly (see
0      
Ri;t 1 i;t 0 1 F t Ri;t 1  E Ri;t 1 F t : 10 Harvey, 1989, 1991). Singleton (2006) provides an overview of
  dynamic asset pricing estimators, Nagel and Singleton (2011)
The unexpected excess return Ri;t 1  E Ri;t 1 jF t can be provide a GMM estimator with an optimal weighting matrix,
further decomposed into a component that is condition- and Roussanov (2014) proposes a nonparametric approach to
ally correlated with the  innovations of the risk factors, estimating the SDF model.
ut 1 C t 1  E C t 1 jF t , and a return pricing error ei;t 1
that is conditionally orthogonal to the risk factor innova-
tions: 4. Estimation with constant betas
    
Ri;t 1  E Ri;t 1 jF t 0i;t C t 1 E C t 1 jF t
In this section, we assume that i;t i for all i and t,
ei;t 1 0i;t ut 1 ei;t 1 : 11 and we analyze an extension of the model with time-
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 215

varying i;t in Section 5. We can then stack this model as The expressions in Eq. (19) can be interpreted as a
three-step estimator in the following way. In the first
R B B1 F  BU E
0
0 T 14
step, shocks to the state variables are obtained from a
and time series vector autoregression. In the second step,
X X  V; 15 asset returns are regressed in the time series on lagged
 0 price of risk factors and the contemporaneous innova-
where R R1 RT  is N  T with Rt R1;t ; ; RN;t , tions to the cross-sectional pricing factors, generating
F  F 0 F T  1  is K F  T, U u1 uT  is K C  T, predictive slopes and risk betas for each test asset. In
 0
E e1 eT  is N  T with et e1;t ; ; eN;t , X X 1 X T  the third step, price of risk parameters are obtained by
is K  T, and V v1 vT  is K  T. Hereafter, we assume that regressing the constant and the predictive slopes from
N Z K C . The parameters of the return equation are the the time series regression on the betas cross-
stacked risk exposures  B, which  is a N  K C matrix with sectionally. This three-step estimator was initially pro-
rows composed of i : 1 r i rN and the prices of risk, . posed by Adrian and Moench (2008) in an application
We can nest the model in the SUR model, to affine term structure models with a linear pricing
R A0 0T A1 F  BU E AZ~ E; 16 kernel. In Section 4.1, we show that this estimator nests
  the two-pass regressions of Fama and MacBeth (1973),
~ 0 0
where A is a N  K C K F 1 matrix, Z T F  U is of
0
when 1 0 and 0. In Section 5, we further discuss
dimension K C K F 1  T, and the differences between our approach and the one
A0 B0 ; A1 B1 ; A A0 A1 B: 17 proposed in Ferson and Harvey (1991). Heuristically,
these authors first estimate t from cross-sectional
In practice, we do not observe U so that we replace it Fama-MacBeth regressions on time-varying betas and
with the residuals from OLS estimation of the VAR. The then by regressing t on a constant and lagged state
asymptotic variance formulas we provide in Theorem 1 variables.
incorporate the additional estimation uncertainty gen- The second regression-based approach is the following
erated by replacing U with U^ . In Appendix A, we provide MD procedure,
explicit instructions on how to construct estimators and

^
B^ md ; ^
md minQ B; ; A ols ; W
their associated standard errors. In Appendix B, we md
; 20
B;
discuss how to impose linear restrictions on the para-
meters B and and conduct inference on these where
restricted estimators. Here we focus on developing

intuition for the form of the estimators and discussing Q B; ; A^ ols ; W md T


their properties.
h i
0 0 0
0
 1  
 

Let Z^ T F 0 U^ and A^ ols RZ^ Z^ Z^ , and parti-  vec A^ ols  B I K C W md vec A^ ols  B I K C : 21
tion the estimator A^ ols as A^ 0;ols , A^ 1;ols , and B^ ols , respectively,
with associated heteroskedasticity-robust This estimator finds the closest approximation of the
pvariance
matrix

estimator V^ rob [so that V^ rob -p V rob and T vec A^ ols  A unconstrained estimator, A^ ols , to values of B; 0 , and 1,
-d N 0; V rob ]. which satisfy the restrictions in Eq. (17). This MD approach
Given this parameterization, two natural approaches turns out to be exactly equivalent to the GMM estimator in
can be taken to estimating the parameters B, 0, and 1. this model and, under certain choices of W md , nests the
The first is an indirect approach based on backing out 0 MLE if the error terms fet : 1 r t rT g are jointly Gaussian.4
and 1 via Specifically,
0 when

the weighting matrix is
    W md Z^ Z^  I N , then the solutions to Eq.  (21) are
 the
0 B0 WB  1 B0 WA0 ; 1 B0 WB  1 B0 WA1 ; 18 MLEs under the assumption that et  iid N 0; 2e  I N . We
for some positive-definite weight matrix W.3 When W I N estimators
as quasi-maximum likelihood esti-
label these
^
mators B^ qmle ;
this produces the regression-based counterpart to Eq. (18) qmle . Closed-form expressions for these

0
1 0 0
1 0 estimators are given in Appendix A. Specifically, these
^ 0;ols B^ ols B^ ols ^
B^ ols A^ 0;ols ; ^ ^
1;ols B ols B ols B^ ols A^ 1;ols : estimators replace the third regression step in the ordinary
least squares (OLS) estimation with a simple eigenvalue
19
decomposition.
We could consider alternative estimators that use data- In Theorem 1, we show that these two estimators
dependent weight matrices, but we prefer this formulation are asymptotically equivalent under our assumptions,
in conjunction with heteroskedasticity-robust standard as they both converge to the same limiting normal
errors to avoid taking a stance on the exact form of the distribution.
variance matrix of the return innovations.
Theorem 1. Under our assumptions,
p
d p
d
3
T vec ^  T vec ^
qmle  N 0; V ;
Here we assume that B is of full-column rank and is consequently N 0; V ;
ols
strongly identified. For cases in which B could be weakly identified see
Kleibergen (2009), Burnside (2010), Kleibergen and Zhan (2013), and
Burnside (2011). In cases of weak identification, the robust test statistics
4
of Kleibergen (2009) could be generalized to our setting. For weak In addition, this equivalence combined with the results of Andrews
identification robust inference in an SDF representation setting, see and Lu (2001) could be used to produce an intuitive model-selection
Gospodinov, Kan, and Robotti (2012). criterion to compare across specifications.
216 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

as T-1, where whether a specific element of is equal to zero, where



  0 0 1 EF t : 22
1
V FF  u H B; V rob H B; ;

0  
FF p lim F~  F~  =T  1, F~  T F 0 0 , and Theorem 2. Under our assumptions,
T-1
 h     1 0
   1 0
i

0 p ^ p ^
H B; I K F 1  B0 B B   B0 B B : d
T vec ols  N 0; V ;
d
T vec qmle  N 0; V ;

as T-1, where V is given in Appendix D.1.

The first term of V accounts for replacing the In Appendix D.1 we show that V is a simple
unobserved innovations U by estimated innovations. expression that invokes quantities that are known in
The second term accounts for all other sources of closed form and easy to compute. Using this result, we
estimation uncertainty, including that of using an can form a t-statistic of the null hypothesis that the
estimate of B to construct the estimator of . Relative sample average of the market price of risk for a given
to the existing literature, Theorem 1 provides a number pricing factor is equal to zero. This allows us to test
of insights. First, it extends feasible inference from the whether a given factor is unconditionally priced in the
static Fama-MacBeth approach that assumes 0 and cross section of test assets.
1 0 to the case with persistent factors and time-
varying prices of risk. Second, Theorem 1 provides a 4.1. Relation to Fama-MacBeth regressions
generalization of Theorem 1 of Shanken (1992), which
provides a correction for the uncertainty generated by Standard factor pricing models assume that prices of
estimating B to a setting with persistent factors and risk are constant and that the pricing factors are unfor-
time-varying prices of risk [under conditional homo- ecastable. Hence, the prevalent factor model used in the
0
skedasticity, i.e., when V rob plim Z^ Z^ =T  e literature implicitly assumes that data are generated by5
T-1
0 0
for a positive-definite variance matrix e ]. More gen- Ri;t 1 i 0 i vt 1 ei;t 1 23
erally, the results allow for conditionally heteroskedas- and
tic errors in the spirit of Theorem 1 of Jagannathan and
X t 1 vt 1 ; t 0; ; T 1; 24
Wang (1998), and so those results are extended to the
dynamic setting as well. Finally, we show the asympto- see, for example, Cochrane (2005, p. 276). This setup is
tic equivalence of the QML approach (a special case of nested in our model if 0 and 1 0. This model is
GMM/MD, as mentioned above) and the OLS approach most commonly estimated by the two-pass Fama-MacBeth
even under conditional heteroskedasticity, which is estimator (Fama and MacBeth, 1973) whose properties
also an extension of Theorem 4 of Shanken (1992) both have been studied by Shanken (1992), Jagannathan and
for constant and time-varying prices of risk. Wang (1998), and Shanken and Zhou (2007) among many
others. The Fama-MacBeth estimator for 0 is
^ ^ 0
1 0
Remark 1. (i) Although ols and qmle are asymptotically
FM
^ 0;ols B^ ols B^ ols B^ ols A^ 0;ols ; 25
equivalent, the associated estimators of B are generally
not. This is because the estimator B^ ols is not constructed where A0 is the estimated constant term from a contem-
under the restrictions in Eq. (17). However, with a simple poraneous regression of returns on de-meaned factors. For
additional step, we can construct an estimator of B based comparison with Theorem 1, note that under our assump-
on ^ ^
ols that is asymptotically equivalent to B qmle : tions it can be shown that

0 h

0 i  1 p FM
 
T ^ 0;ols  0 N 0; V FM
d
B^ 4ols R ^ F~ U^
ols  ^ ols F~  U^ ^ ols F~  U^ : ;
  FM FM  0
u H B; 0 V rob H B; 0 ;
V FM FM

Intuitively, B^ 4ols is the OLS estimator of B taking the   h 0 1 0


0    1 0
i
estimated prices of risk ^
ols as given.
HFM B; B B B  0  B0 B B ; 26
 
(ii) Under the assumption that et jF t  1  iid N 0; 2e  I N
^ where V FM is the probability limit of the
and all variables are X2-type variables, the estimators ols
rob

^ ^ ^ heteroskedasticity-robust variance matrix from a contem-


and qmle are asymptotically efficient. B qmle and B 4ols are poraneous regression of returns on factors and a constant.
also asymptotically efficient, although B^ ols is asymptoti-
Because we allow for conditional heteroskedasticity, the
cally efficient only when N K C .
variance matrix V FM is in the spirit of that obtained by
In traditional asset pricing models with constant prices Jagannathan and Wang (1998) when the risk-free rate is
of risk, the parameter 0 determines whether a risk factor observed. Similarly, the variance expression derived in
is priced in the cross section of test assets. However, when Shanken (1992) can be obtained by using V FM FM
with V rob
prices of risk are time varying, this parameter is no longer
of independent interest. Instead, to gauge whether differ- 5
Here we are assuming that the risk-free rate is observed, so the
ential exposures to a given pricing factor result in sig- model does not include the zero-beta rate. Similar results can be obtained
nificant spreads of expected excess returns, one has to test with the inclusion of a zero-beta rate.
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 217

formed under the assumption of conditionally homoske- As an exercise, consider the case in which the true data
dastic errors. generating process is governed by Eqs. (1) and (13) so that
FM
The analogous estimator, ^ 0;qmle , has received relatively the prices of risk vary over time but are mistakenly
less attention in the literature than its counterpart, derived assumed to be governed by Eqs. (27) and (28) and
under the assumption that et  iid N 0; e .6 As in the more estimated via two-pass Fama-MacBeth regressions. Inter-
FM FM estingly, it can be shown that in this case
general case above, ^ 0;ols and ^ 0;qmle are still asymptotically

p ^ FM d
p FM   T 0  N 0; V (see Theorem 2). Thus, the
equivalent so that T ^ 0;qmle  0 -d N 0; V FM even in
the presence of conditional heteroskedasticity. To our conventional estimator is consistent for the parameter .
knowledge, this has not previously been pointed out in However, Wald-type test statistics would commonly be
the literature. Following similar steps as in Appendix C, constructed using a plug-in version of the variance for-
mula of Shanken (1992), which under technical conditions,
even with the inclusion of a zero-beta rate, the direct 0 1

 
1
equivalence between the MD and GMM estimators (for converges in probability to v 1 v  B0 B
any choice of weight matrix) and the MLE (for specific  1
B0 e B B0 B . Comparing this expression and that of V
choices of weight matrix) can be established for the model
of Eqs. (23) and (24). Special cases of this result have been from Appendix D.1 shows that the bias of the standard
pointed out in the literature. Ahn and Gadarowski (1999) variance estimator depends on the values of , , and v.
discuss, and Kan and Chen (2005) show, the equivalence
between the MD estimator and the MLE. More recently, 5. Estimation with time-varying betas
Shanken and Zhou (2007) show the equivalence between
the GMM estimator and the MLE (see also Zhou, 1994; A large literature exists on estimating beta representa-
Kleibergen, 1998). tions of asset pricing models assuming that the betas vary
It follows from the equivalence between MD and GMM over time. For example, see Fama and MacBeth (1973),
estimation for the model of Eqs. (23) and (24) that the J- Ferson and Harvey (1991), and many more. In this section,
statistic is equivalent to the MD criterion function [i.e., Eq. we discuss estimation of our model in the case in which
(21)]. Thus, the cross-sectional T2 statistic of Shanken factor risk exposures as well as the parameters governing
(1985) [see Lewellen, Nagel, and Shanken (2010) for a the dynamics of the factors are time-varying. The model is,
detailed discussion of the test statistic], which corresponds therefore,
to the MD criterion function when there is an unknown 0 0 0
Ri;t 1 i;t 0 i;t 1 F t i;t ut 1 ei;t 1 29
zero-beta rate (evaluated at the two-pass estimators) may
be interpreted directly as a J-test of the moment restric- and
tions for the model. This is an intuitively appealing inter- X t 1 t t X t vt 1 : 30
pretation because the J-statistic is then a direct joint test of
the cross-sectional asset pricing restrictions imposed by To motivate our estimator consider the case in which the
the assumption of no-arbitrage. This is consistent with innovations
fut g

and
0
the betas are known. In addition, let
Lewellen, Nagel, and Shanken (2010), which emphasizes Bt 1;t ; ; N;t . Then, passing through the vectoriza-
the importance of analyzing the estimators of all the tion operator yields
0
 
parameters of the model instead of solely focusing on Rt 1 Bt ut 1 F~ t  Bt vec et 1 : 31
the price of risk. More generally, one key part of our
contribution is to extend the static setting discussed here From there it is easy to see that the associated estimator of
to the dynamic setting introduced in Section 3 without the price of risk is
compromising the simplicity of implementation that has tv

vec ~
made the Fama-MacBeth estimator so popular in the ols
applied finance literature. !
TX1
0

 1 TX
1

Some authors apply the Fama-MacBeth estimator in F~ t F~  B0 Bt


t t F~ t  B0 Rt 1  Bt ut 1 :
t
model specifications with constant prices of risk, where t0 t0

the pricing factors are given by the VAR(1) innovations of a 32


vector of state variables (see, for example, Chen, Roll, and
In practice, the estimator of Eq. (32) is infeasible without
Ross, 1986; Campbell, 1996; Petkova, 2006). These speci-
estimates of Bt, t, and t. Furthermore, without additional
fications thus rely on the return generating process:
assumptions, identification of these parameters would be
0 0 impossible as the number of parameters grows too quickly
Ri;t 1 i 0 i vt 1 ei;t 1 27
as T-1.
One approach to identify time variation in i;t that has
and been used in the literature is to posit that the parameters
i;t are (linear) functions of observable variables (see, for
X t 1 X t vt 1 ; t 0; ; T  1: 28
example, Shanken, 1990; Ferson and Harvey, 1999;
Gagliardini, Ossola, and Scaillet, 2014; and Chordia,
6
See, for example, Gibbons (1982), Kandel (1984), Roll (1985),
Goyal, and Shanken, 2013). However, a drawback to this
Shanken (1985, 1986), Kan and Chen (2005), Shanken and Zhou (2007), approach is that it requires the correct specification for the
and Kleibergen (2009), among others. functional form of the i;t . In fact, as pointed out by
218 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

Ghysels (1998) and Harvey (2001), among others, the beta precludes standard inference procedures without further
estimates obtained in this way are typically sensitive to the adjustment.
specification of the information set. As a consequence, the Eq. (29) is nested in a time-varying equivalent of the
magnitude of the resulting estimated pricing errors can SUR system discussed in Section 4. We solve the system by
vary substantially with the choice of conditioning vari- equation-by-equation weighted least squares regressions:
ables. Other limitations to this approach are that the !1
0 0
X
T   tv tv0
number of regressors can grow large and that ^
A ^
;A ^
; K s  t =T z z
0;i;t  1 1;i;t  1 i;t  1 h s s
commonly-used conditioning variables are available only s1
!
at low frequencies. X
T  
An alternative identifying assumption is that  Kh s t =T ztv
s Ri;s
s1
  
i;t i t T o1; t t =T o1; 34

 and
t t =T o1; 33 !1

0 X
T   0
^
^ t  1 ; Kb s t =T X~ s  1 X~ s  1
t 1
where all i , , and  are sufficiently smooth s1
!
functions to estimate the parameters nonparametrically. X
T  
Appendix D.2 provides some additional details about this  Kb s  t =T X~ s  1 X 0s ; 35
assumption and its implications.7 This assumption has the s1
 0 0 0  
appeal that it implies that the betas do not vary too much where ztvs 1; X s  1 ; C s and Kh x K x=h for some ker-
over short time periods, which is consistent with both nel function K and bandwidths h hT and b bT are
economic theory and prior empirical studies (see, for positive sequences that converge to zero. The set of regressors,
example, Braun, Nelson, and Sunier, 1995; Ghysels, 1998; ztv
t , is different than in the constant beta case in which
and Gomes, Kogan, and Zhang, 2003). Importantly, it estimated innovations, u^ t , were used instead of Ct to estimate
imposes less structure than assuming a precise functional the betas. When betas are time-varying, it is technically
form for the parameters and so is likely more robust to convenient to make this change as we can then directly rely
misspecification. Intuitively, the functional form assump- on results from Kristensen (2009).
tions in Eq. (33) imply that as T grows, the amount of local Intuitively, the kernel function in Eqs. (34) and (35) places
information about the function value increases. more weight on observations nearby and less weight on those
A number of different options exist for nonparametri- farther away, where the rate of decay is governed by the
cally estimating the ^ i;t . We follow Ang and Kristensen bandwidths h and b, respectively. Moreover, because we
(2012) and use kernel smoothing estimators. We can then smooth only in the time dimension, our approach does not
derive, at any point in time, an asymptotic distribution for suffer from the so-called curse of dimensionality. To choose
all parameters of our model, including the conditional the bandwidths we use a plug-in method developed in
betas and the price of risk parameters obtained from the Kristensen (2012) and Ang and Kristensen (2012). In
beta estimates. In addition to being more robust to mis- Appendix A.2, we provide more details on the implementa-
specification, kernel smoothing estimators have the tion of the bandwidth selection.
appealing feature that they nest, as a special case, rolling Given these first-stage estimates, the feasible estimator
window estimates of i;t , which are popular in the empiri- of is then
cal literature (for example, Chen, Roll, and Ross, 1986; !1
tv
1
TX
0 0

1
TX
0

Ferson and Harvey, 1991; Petkova and Zhang, 2005; among ^


vec
ols F~ t F~  B^ B^ t t F~ t  B^
t T t
many others). Rolling beta estimates are equivalent to t0 t0
using a uniform one-sided kernel instead of a Gaussian

 Rt 1  B^ t u^ t 1 ; 36
two-sided kernel, as we do here. The standard approach of
using backward-looking, five-year rolling regressions has where T is a positive sequence that satisfies T -0. This
two noteworthy drawbacks. First, for the estimator to be additional term guarantees the stability of the estimator by
consistent, the bandwidth sequence (i.e., the window) ensuring that the matrix is always invertible. It is straightfor-
needs to shrink to zero. However, the choice of five-year ward to show that when the betas and VAR coefficients no
windows is not data-dependent and so might not be longer time vary and T 0, then ^ tv is analytically equiva-
ols
appropriate for many applications (see Section 6 for ^
lent to from Section 4. We then have Theorem 3.
ols
further discussion). Second, the order of the smoothing
bias of the estimator for the betas and the price of risk Theorem 3. Under our assumptions,
parameters is larger for one-sided kernels. In fact, p tv
d  
^ 
T vec N 0; V tv
although the estimator of based on rolling regressions ols ;
(with appropriate data-dependent bandwidth choice) in as T-1, where
Eq. (36) below is consistent, a non-negligible bias term !1
Z 1  
V tv
f  B0 B d
7 0
See Robinson (1989). A number of other authors have used this "Z
assumption in conjunction with time-varying parameters. See Ang and 1

Kristensen (2012) for a lucid discussion about this approach to modeling  f 0 D0B z  1 DB f f
time-varying parameters. 0
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 219


 B0 e B d: rr
f^ i;t : i 1; ; N; t 1; ; Tg, which we stack as
Z 1 #
  rr
fB^ t : t 1; ; Tg. We then obtain the FM estimator of the
f  B0 B u B0 B d
0 constant price of risk parameter 0 from
Z !1
1   X
T rr0 rr
 1 rr0 rr
 f  B0 B d FM
^ 0 T  1 ^ t ; ^ t B^ t B^ t B^ t A^ 0;t ; 38
0
t1

rr
where A^ 0;t is the (stacked) estimated intercept from the
and z , f , u , and DB are defined in Appendix D.2.
rolling regressions, the analog to the constant beta case in
Despite the fact that ^ tv is based on estimates of , , Eq. (25). In practice, a five-year burn-in period is necessary
ols i;t t
and t, which converge at a rate slower than the para- to construct these estimators. For ease of notation, the
metric rate, our estimator of the price of risk achieves the equations presented in this subsection ignore this distinc-
parametric rate. This is an appealing feature as it means tion. As in the constant beta case, the estimator in Eq. (38)
that the additional flexibility we introduce in modeling the is derived from the asset pricing restriction that the
time variation in the betas and VAR coefficients does not intercept satisfies A0;t Bt 0 .
come at the cost of asymptotic efficiency. The intuition Ferson and Harvey (1991) have proposed to estimate
behind this result is that the additional averaging over time-varying prices of risk in conditional factor pricing
time to estimate accelerates the rate of convergence. models by first running Fama-MacBeth two-pass regres-
Furthermore, in the spirit of the comment in Remark 1, we sions as above and, subsequently, in a third estimation
can then reestimate the i;t from a kernel regression of step, regressing the obtained time series of market prices
^ tv F~ u^
Ri;t 1 on the sum ols t t 1 . Finally, in Appendix B, we of risk (^ t ) on one-month lagged predictor variables. This
discuss how to carry out restricted estimation and infer- estimator is similar in spirit to our estimator in which
ence for the price of risk parameter . market prices of risk are modeled as affine functions of a
The time variation in t and t implies that the mean of set of forecasting (X2 and X3) variables.
the factors is also shifting over time. The definition of To implement the FH estimator, we again use the
must be changed accordingly: innovations u^ t 1 as pricing factors but also control for
tX
1 the lagged values of the price of risk factors, Ft [i.e., Eq.
0 1  lim T  1 EF t : 37 (29)] to estimate the betas. We then estimate the price of
T-1
T risk parameters by regressing ^ t on a constant and the
We then have Theorem 4, an analogous result to Theorem 2, lagged price of risk factors, i.e.,
! !1
TX 1 TX
1
Theorem 4. Under our assumptions, FH 0 0
^ ^ F~ F~ t F~ : 39
p ^ tv d

t0
t1 t
t0
t

T ols  N 0; V tv ;

We compare the two estimators with the ones derived
tv
where V is defined in Appendix D.2. in this paper in terms of model-implied mean squared

pricing errors in Section 6.
The asymptotic variance, V tv , can be estimated in a

straightforward manner, and so inference on whether a
factor is priced on average can be conducted easily. 6. Empirical application

5.1. Comparison with estimators using rolling regressions In this section, we apply our estimation method to a
dynamic asset pricing model for equity and Treasury
In Section 6, we compare our results with the estima- returns. We choose test assets that have been studied
tors proposed by Fama and MacBeth (1973) and Ferson extensively in the empirical asset pricing literature to
and Harvey (1991) (hereafter FM and FH), both imple- illustrate the usefulness of the regression-based dynamic
mented using rolling regressions to obtain time-varying asset pricing approach. We show that a parsimonious
betas in the first estimation stage. To properly account for model with two pricing factors, two price of risk factors,
the persistent nature of (some) factors, we implement and one factor that is both fits the cross section of size-
these procedures using the estimated innovations u^ t 1 as sorted equity portfolios and constant maturity Treasury
pricing factors. This is in contrast to much of the empirical portfolios very well on average while, at the same time,
literature, which estimates betas using the level of the giving rise to strongly significant time variation in risk
pricing factors without controlling for lagged observations. premiums. We further show that allowing for time varia-
When the pricing factors are persistent, these estimates tion in factor risk exposures substantially improves the
are not consistent.8 Rolling regressions yield estimates precision of price of risk parameters. Finally, allowing for
time variation in prices of risk is more important than
modeling time variation in factor risk exposures for mini-
8
While the results they report are based on simple rolling regres- mizing squared pricing errors of the model. Traditional
sions without controlling for the potential persistence in the pricing
factors, Ferson and Harvey (1991) mention in Footnote 7 that their results
estimation approaches such as the one by Fama and
are robust to the estimation of rolling betas controlling for the lagged MacBeth (1973) and Ferson and Harvey (1991) imply
level of the pricing factors. substantially larger pricing errors than our estimator.
220 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

6.1. Data Section 4. The first panel reports the OLS estimates and the
second the QML estimates. In each panel, we provide the
We obtain ten size-sorted portfolios for US equities from estimated betas and associated standard errors for the
Ken French's online data library. We further use constant three cross-sectional pricing factors MKT, SMB, and TSY10.
maturity Treasury portfolios with maturities one, two, five, Several results are worth highlighting. First, the coeffi-
seven, ten, 20, and 30 years from the Center for Research in cients and standard errors implied by the OLS and the QML
Securities Prices (CRSP). We compute excess returns over the estimator are very similar. Hence, any discussion of esti-
one-month Treasury bill yield, which we also obtain from mated risk premiums does not qualitatively depend on the
French's website. Our sample spans the period 1964:01 choice of estimator of B. Second, while all size portfolios
2012:12 for a total of 588 monthly observations. significantly load on MKT and SMB, the Treasury portfolios
We use the following set of factors to price the joint do not. That is, Treasury portfolio returns do not contem-
cross section of equities and Treasuries. The excess poraneously co-move with shocks to the two equity
return on the value-weighted equity market portfolio pricing factors in the constant beta specification. The
(MKT) from CRSP and the small minus big (SMB) portfo- market betas of the size portfolios have the expected
lio from Fama and French (1993), and the ten-year magnitudes around one with relatively little dispersion.
Treasury yield (TSY10) serve as cross sectional pricing This is the well-known size effect: exposure to MKT does
factors. We obtain the first two factors from French's not explain the large spread between average excess
website, and the third from the Federal Reserve Statis- returns on small versus large market cap stocks. In con-
tical Release H.15. The first two factors explain a sub- trast, the risk exposures to SMB show a strong differential
stantial share of the variance of the size decile portfolio between the smallest and the largest size deciles. Finally,
returns. However, they are not usually considered to be while the Treasury portfolios do not load on the two
return forecasting variables. We, therefore, treat them as equity risk factors, the equity portfolios generally load
cross-sectional pricing factors and do not attribute to significantly on the ten-year Treasury yield factor.
them a role for explaining time variation in prices of risk. We now compare these estimates with those obtained
The ten-year Treasury yield can be considered a good assuming betas are time-varying. Fig. 1 provides plots of
proxy for the level of the term structure of Treasury factor risk exposures of two test assets, size5 and cmt10,
yields, which has been shown to be a priced factor in the for all three pricing factors in our model: MKT, SMB, and
cross section of Treasury returns (see e.g., Cochrane and TSY10. For each factor-asset pair we compare three differ-
Piazzesi, 2008; Adrian, Crump, and Moench, 2013). We ent beta estimates. The constant one (dash-dotted line) is
also allow this factor to determine time variation in obtained using the estimator in Section 4, the time-varying
factor risk premiums, as long-term Treasury yields have one (solid line) is obtained using the Gaussian kernel
been shown to contain predictive information for bond estimator with data-driven bandwidth choice discussed
and stock returns (see e.g., Keim and Stambaugh, 1986; in Section 5, and the five-year rolling window estimator
Campbell, 1987; Fama and French, 1989; Campbell and (dashed line) often used in the empirical asset pricing
Thompson, 2008). In addition to these three factors, we literature and also represents the first-stage estimates in
consider two price of risk factors: the term spread our implementation of the Fama and MacBeth (1973) and
between the yield on a ten-year Treasury note and the Ferson and Harvey (1991) estimators.
three-month Treasury bill (TERM) (also obtained from Several remarks are in order. First, in all cases, the time-
the Federal Reserve Statistical Release H.15), and the log varying beta estimates are centered around the constant
dividend yield (DY) of the Standard & Poor's (S&P) 500 estimates. Second, while the Gaussian kernel with data-
index from Haver Analytics. Both factors have previously driven bandwidth implies some variability in factor risk
been shown to predict equity and bond returns (see e.g., exposures, it features considerably less time variation in betas
Campbell and Shiller, 1988; Fama and French, 1989; than the five-year rolling beta estimator. In particular, for all
Campbell and Thompson, 2008; Cochrane, 2008) and factor-asset pairs, the latter implies betas with signs flipping
are, therefore, good proxies for time variation in risk multiple times across the sample period. At low frequencies,
premiums. In summary, in our model excess returns are however, the rolling beta estimates mimic the evolution of the
determined by risk exposures to MKT, SMB, and TSY10, Gaussian kernel-based betas. Moreover, despite the smooth
where the market prices of risk of these three pricing nature of Gaussian kernel estimates, their evolution over time
factors are assumed to vary over time as affine functions gives rise to some interesting observations. Most important,
of TSY10, TERM, and DY. the size5 portfolio's beta on the Treasury factor switches from
Given this set of test assets and pricing factors, the total a negative to a positive sign in the mid-1990s. Around the
number of risk exposure parameters to estimate is N  K C same time, the cmt10 portfolio's beta on the equity market
or 17  3 51. The number of market price of risk para- portfolio switches from a positive to a negative sign. Hence,
meters is K C  K F 1 or 3  4 12. our time-varying beta estimates replicate the empirical obser-
vation that the correlation between stock and bond returns
6.2. Empirical results has flipped signs sometime in the 1990s (see e.g. Baele,
Bekaert, and Inghelbrecht, 2010; Campbell, Sunderam, and
We start by discussing the estimates of factor risk Viceira, 2013; David and Veronesi, 2013). Another interesting
exposures assuming constant betas. Table 1 provides beta observation is that the beta of the ten-year constant maturity
estimates for all size and Treasury portfolio returns related Treasury return (cmt10) onto the ten-year Treasury yield
to the three risk factors, implied by the estimators in factor (TSY10) fluctuates quite substantially over time. Because
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 221

Table 1
Factor risk exposure estimates.
This table provides estimates of factor risk exposures from the constant-beta specification of the dynamic asset pricing model discussed in Section 6. It
reports ordinary least squares (OLS) estimates and quasi-maximum likelihood (QML) estimates. Asymptotic standard errors are provided in parentheses.
The pricing factors are MKT, the excess return on the Center for Research in Security Prices (CRSP) value-weighted equity market portfolio, SMB, the small
minus big portfolio both obtained from Ken French's website, and TSY10, the constant maturity ten-year Treasury yield from the Federal Reserve Statistical
Release H.15. The test assets are the ten size-sorted stock decile portfolios from Ken French's website (size1 size10), as well as constant maturity Treasury
returns for maturities ranging from one through 30 years (cmt1 cmt30). We obtain the latter from CRSP. Wald stats denote Wald tests for the joint
significance of all factor risk exposures associated with the respective pricing factor. LR stat is a likelihood ratio test for the joint significance of all factor
risk exposures across test assets and pricing factors in the model of Eq. (16) (see Kleibergen and Zhan, 2013). The sample period is 1964:012012:12. nnn
denotes significance at 1%, nn, significance at 5%, and n, significance at the 10% level.

MKT s:e:MKT SMB s:e:SMB TSY10 s:e:TSY10

OLS estimates

size1 0.851nnn (0.030) 1.171nnn (0.020) 0.109nnn (0.021)


size2 0.984nnn (0.019) 1.066nnn (0.018) 0.053nnn (0.017)
size3 1.007nnn (0.016) 0.898nnn (0.015)  0.111nnn (0.014)
size4 0.996nnn (0.007) 0.804nnn (0.004)  0.099nnn (0.005)
size5 1.018nnn (0.007) 0.662nnn (0.008) 0.036nnn (0.010)
size6 1.003nnn (0.017) 0.480nnn (0.021)  0.277nnn (0.045)
size7 1.031nnn (0.029) 0.362nnn (0.039)  0.160nnn (0.038)
size8 1.033nnn (0.032) 0.264nnn (0.035)  0.293nnn (0.029)
size9 0.995nnn (0.028) 0.067nnn (0.020)  0.429nnn (0.012)
size10 0.988nnn (0.005)  0.276nnn (0.007) 0.188nnn (0.009)
cmt1 0.004 (0.011) 0.009 (0.012)  1.043nnn (0.019)
cmt2 0.001 (0.025) 0.008 (0.320)  2.049nnn (0.204)
cmt5  0.006 (0.201)  0.003 (0.184)  4.226nnn (0.168)
cmt7  0.000 (0.149)  0.018 (0.160)  5.164nnn (0.138)
cmt10 0.007 (0.156)  0.015 (0.073)  6.165nnn (0.085)
cmt20 0.003 (0.129)  0.009 (0.111)  7.893nnn (0.117)
cmt30  0.027 (0.184)  0.004 (0.269)  8.581nnn (0.320)
Wald stat 182,785.308nnn (0.000) 81,346.613nnn (0.000) 30,021.782nnn (0.000)
LR stat 6,297.355nnn (0.000)

QML estimates

size1 0.852nnn (0.030) 1.174nnn (0.020) 0.091nnn (0.021)


size2 0.980nnn (0.020) 1.060nnn (0.018) 0.053nnn (0.017)
size3 1.007nnn (0.016) 0.898nnn (0.015)  0.082nnn (0.014)
size4 0.997nnn (0.007) 0.804nnn (0.004)  0.089nnn (0.005)
size5 1.019nnn (0.007) 0.664nnn (0.008) 0.039nnn (0.010)
size6 1.005nnn (0.017) 0.485nnn (0.020)  0.291nnn (0.045)
size7 1.031nnn (0.029) 0.363nnn (0.038)  0.185nnn (0.037)
size8 1.032nnn (0.032) 0.263nnn (0.035)  0.288nnn (0.029)
size9 0.994nnn (0.027) 0.066nnn (0.020)  0.417nnn (0.011)
size10 0.987nnn (0.005)  0.277nnn (0.007) 0.186nnn (0.009)
cmt1 0.004 (0.010) 0.009 (0.012)  1.048nnn (0.019)
cmt2 0.001 (0.025) 0.008 (0.318)  2.047nnn (0.204)
cmt5  0.006 (0.201)  0.003 (0.182)  4.217nnn (0.169)
cmt7  0.000 (0.147)  0.018 (0.158)  5.158nnn (0.137)
cmt10 0.007 (0.154)  0.015 (0.071)  6.151nnn (0.080)
cmt20 0.002 (0.121)  0.009 (0.106)  7.916nnn (0.114)
cmt30  0.027 (0.180)  0.004 (0.268)  8.579nnn (0.319)
Wald stat 184,171.026nnn (0.000) 80,463.712nnn (0.000) 31,121.825nnn (0.000)

the return on a bond is, to a first-order approximation, equal We next turn to a discussion of the estimated market
to minus its duration times the yield change, this time prices of risk. Table 2 provides estimates of the market
variation reflects the fact that the duration of longer-dated price of risk parameters 0 and 1 implied by three
Treasury securities has changed substantially over the 50 year different estimators. The second-to-last column provides
sample that we consider. In fact, duration was low in the late the average price of risk estimates for each factor and its
1970s and early 1980s, when rates were high, and has since asymptotic standard error as provided in Theorems 2 and
experienced a secular upward trend against the backdrop of 4, respectively. These statistics allow us to test whether a
falling rates. These dynamics are well captured by the time- given factor is priced on average in the cross section of test
varying beta estimates. Moreover, the five-year rolling assets. Finally, the last column provides a Wald statistic for
regression-based estimates mimic the evolution of time- a test of whether the coefficients in a particular row of 1
varying betas from the Gaussian kernel-based estimates with are jointly equal to zero. This statistic thus indicates
data-driven bandwidth choice well. For other asset-factor whether there is time variation in each of the factor risk
pairs, they appear too noisy. prices.
222 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

1.2 0.12

1.15 0.1

0.08
1.1

0.06
1.05
0.04
1
0.02
0.95
0

0.9
0.02

0.85 0.04

0.8 0.06
1970 1975 1980 1985 1990 1995 2000 2005 2010 1970 1975 1980 1985 1990 1995 2000 2005 2010

1 0.15

0.95 0.1

0.9 0.05

0.85
0
0.8
0.05
0.75
0.1
0.7
0.15
0.65
0.2
0.6

0.55 0.25

0.5 0.3
1970 1975 1980 1985 1990 1995 2000 2005 2010 1970 1975 1980 1985 1990 1995 2000 2005 2010

2.5 3.5

4
2
4.5
1.5
5
1
5.5

0.5 6

6.5
0
7
0.5
7.5
1
8

1.5 8.5
1970 1975 1980 1985 1990 1995 2000 2005 2010 1970 1975 1980 1985 1990 1995 2000 2005 2010

Fig. 1. Comparison of beta estimates. This figure provides plots of beta estimates obtained for different pairs of test assets and cross-sectional pricing
factors. t ; t shows time-varying betas estimated using the kernel regression approach presented in Section 5. 0 ; t denotes the constant beta estimate
obtained using the ordinary least squares (OLS) estimator described in Section 4. Rolling refers to the five-year rolling window estimate. size5 denotes the
fifth decile portfolio from the set of size-sorted stock portfolios from Ken French's website. cmt10 refers to the constant maturity Treasury returns for the
ten-year maturity, obtained from the Center for Research in Security Prices (CRSP). MKT, SMB, and TSY10 denote the value-weighted stock market portfolio
from CRSP, the small minus big portfolio from Fama and French (1993), and the ten-year Treasury yield from the Federal Reserve Statistical Release H.15.
The sample period is 1964:012012:12.

Table 2 reports estimates based on time-varying betas constant betas, respectively. The asymptotic standard
implied by a Gaussian kernel and Panels B and C show errors (and p-values in the case of the W 1 statistic) are
them for the three-step OLS and QML estimators under shown in parentheses. We make the following
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 223

Table 2
Price of risk estimates.
This table provides estimates of market price of risk parameters from the dynamic asset pricing model discussed in Section 6. Panel A reports ordinary
least squares (OLS) estimates for the specification with time-varying betas, and Panels B and C provide OLS and quasi-maximum likelihood (QML) estimates
for the specification with constant betas. The pricing factors are MKT, the excess return on the Center for Research in Security Prices (CRSP) value-weighted
equity market portfolio, SMB, the small minus big portfolio both obtained from Ken French's website, and TSY10, the constant maturity ten-year Treasury
yield from the Federal Reserve Statistical Release H.15. The price of risk factors are TSY10, TERM, the spread between the constant maturity ten-year
Treasury yield and the three-month Treasury bill, both obtained from the H.15 release, as well as DY, the log dividend yield obtained from Haver Analytics.
The first column, 0, gives the estimated constant in the affine price of risk specification for each pricing factor. The second through fourth columns provide
the estimated coefficients in the matrix 1, which determine loadings of prices of risk on the price of risk factors. The column provides an estimate of the
average price of risk as given in Eq. (22). The last column provides the Wald test statistic of the null hypothesis that the associated row is all zeros. The
sample period is 1964:012012:12. nnn denotes significance at 1%, nn, significance at 5%, and n, significance at the 10% level.

0 TSY10 TERM DY W 1

Panel A: Time-varying betas

MKT 0.062nnn  0.184nnn 0.302nnn 0.014nnn 6.797nn 25.328nnn


(0.017) (0.058) (0.088) (0.004) (2.785) (0.000)
SMB 0.054nnn  0.194nnn 0.099 0.011nnn 3.565 23.190nnn
(0.013) (0.044) (0.066) (0.003) (2.690) (0.000)
TSY10 0.004nnn  0.014nnn  0.046nnn 0.001nn  0.359 48.636nnn
(0.001) (0.005) (0.007) (0.000) (0.229) (0.000)

Panel B: Constant betas (OLS)

MKT 0.063nn  0.187n 0.301nn 0.014nn 6.067nnn 8.975nn


(0.028) (0.098) (0.147) (0.006) (1.487) (0.030)
SMB 0.054nn  0.192nnn 0.093 0.011nn 3.023 7.599n
(0.022) (0.073) (0.108) (0.005) (2.336) (0.055)
TSY10 0.004  0.013  0.050nnn 0.001  0.386nnn 21.237nnn
(0.002) (0.008) (0.012) (0.001) (0.085) (0.000)

Panel C: Constant betas (QMLE)

MKT 0.063nn  0.187n 0.301nn 0.014nn 6.066nnn 8.975nn


(0.028) (0.098) (0.147) (0.006) (1.424) (0.030)
SMB 0.054nn  0.192nnn 0.093 0.011nn 3.025 7.598n
(0.022) (0.073) (0.108) (0.005) (2.037) (0.055)
TSY10 0.004  0.013  0.050nnn 0.001  0.386nnn 21.237nnn
(0.002) (0.008) (0.012) (0.001) (0.103) (0.000)

observations. First, the estimated market price of risk from zero on average, it exhibits substantial time varia-
parameters are strikingly similar across the three estima- tion and fluctuates between positive and negative values
tors. This reinforces the above observation that the that are significantly different from zero. This is also
Gaussian kernel-based beta estimators with data-driven indicated by the Wald statistic W 1 for the rows of 1
bandwidth choice do not move sharply over time. Second, being jointly equal to zero, provided in the last column.
the price of risk parameters are estimated with much All three estimators suggest significant time variation in
greater precision in the time-varying beta case, as the the prices of risk on all three cross-sectional pricing
standard errors of most elements of 0 and 1 are factors of our model, including SMB.
substantially smaller in Panel A. Hence, because the price Looking at individual elements of 1, we find strong
of risk parameters are identified based on cross-sectional evidence for time variation in the prices of risk of MKT,
variation of the betas, allowing for time-varying risk SMB, and TSY10 as all but one element of the coefficient
exposures more precisely captures the dynamics of the matrix 1 are individually significant at least at the 10%
price of risk. Third, while the constant coefficients in the level. In particular, TSY10 affects the prices of risk of all
market prices of risk are all individually significant at the three factors with a negative sign. That is, higher long-
1% level across the three estimators, the average price of term interest rates drive down the price of risk for both
risk statistic discussed in Theorems 2 and 4 is statisti- equity and bond market factors. Third, while TERM does
cally different from zero only for MKT in the time-varying not significantly add to the variation in the price of SMB
beta case and for MKT and TSY10 in the constant beta risk, a high term spread strongly raises the price of MKT
case. Hence, according to all three estimators exposure to risk and reduces the price of TSY10 risk. Because equity
SMB risk is not unconditionally priced in our cross section portfolios load positively on MKT, this implies that a
of test assets. This is consistent with other studies positive term spread predicts higher expected excess
showing that SMB is not priced in the cross section of returns on stocks, in line with, e.g., Campbell (1987) and
size- and book-to-market-sorted equity portfolios (see, Fama and French (1989). Moreover, noting that the factor
for example, Lettau and Ludvigson, 2001). However, risk exposures of bond returns on TSY10 are negative, the
while the price of SMB risk is statistically not different latter finding is consistent with the evidence in, e.g.,
224 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

8 8

7 7

6 6

Fitted average returns


Fitted average returns

5 5

4 4

3 3

2 2

1 1

0
0
0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Realized average returns Realized average returns

8 8

7 7

6 6
Fitted average returns
Fitted average returns

5 5

4 4

3 3

2 2

1 1

0 0

0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8
Realized average returns Realized average returns

Fig. 2. Comparison of cross-sectional pricing properties. This figure provides plots of observed versus model-implied average excess returns on the set of
test assets estimated using four different approaches as discussed in Section 6. The upper-left graph reports results based on our benchmark specification
(t ; t ) with time-varying betas and time-varying prices of risk, estimated using the approach presented in Section 5. The upper-right graph shows the
unconditional fit of the specification with constant betas but time-varying prices of risk, estimated using the three-stage ordinary least squares (OLS)
estimator discussed in Section 4. The lower-left graph shows the average fit of the model estimated using the approach suggested in Ferson and Harvey
(1991), designated FH, which is based on time-varying betas estimated using five-year rolling window regressions. The lower-right graph presents results
for the Fama and MacBeth (1973), designated, FM, two-pass estimator that is also based on time-varying betas estimated using five-year rolling window
regressions but features constant prices of risk. We implement FM by treating the ten-year Treasury yield as a X1-type pricing factor and omitting the
dividend yield and the term spread as factors. All excess returns are stated in annualized percentage terms. The test assets are the ten size-sorted stock
decile portfolios from Ken French's website (size1 size10), as well as constant maturity Treasury returns for maturities ranging from one through 30
years (cmt1 cmt30), obtained from the Center for Research in Security Prices (CRSP). The plots are based on the OLS estimates of the model. The sample
period is 1964:012012:12.

Campbell and Shiller (1991), that a positive slope of the different model specifications and corresponding estima-
yield curve predicts higher future Treasury returns. Finally, tors. The upper-left graph shows the average model fit for
the log dividend yield DY has a positive impact on the the specification with both betas and market prices of risk
prices of risk of all three factors. This confirms previous time-varying, estimated with the Gaussian kernel-based
evidence, e.g. in Fama and French (1989), that the dividend estimator discussed in Section 5. The upper-right graph
yield predicts excess returns on stocks and bonds. displays the model fit for the specification with constant
Before diving into a more specific analysis of time betas, estimated using the three-step OLS regression
variation in risk premiums, we show the good perfor- approach outlined in Section 4. The lower two graphs
mance of our dynamic asset pricing model in explaining show average pricing errors implied by the Ferson and
average excess returns on size and Treasury portfolios. Harvey (1991) and Fama and MacBeth (1973) estimation
Fig. 2 shows average model-implied excess returns against approaches. While the former features time-varying and
average observed excess returns, as implied by four the latter constant prices of risk, both are based on betas
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 225

Table 3 the constant and the time-varying beta specification.


Mean squared pricing error comparison. In contrast, both the Ferson and Harvey (1991) and
This table compares mean squared pricing errors across various model
Fama and MacBeth (1973) estimators imply average fitted
estimation approaches for the asset pricing model discussed in Section 6.
Panel A reports, for each test asset, the mean squared pricing error implied by excess returns for the equity portfolios in our cross section
the various estimation approaches. t ; t denotes our benchmark specification that are all lower than the observed average excess
with both time-varying betas and market prices of risk and the betas being returns.
estimated using the approach discussed in Section 5. 0 ; t is a specification The model's ability to fit returns should be assessed not
with constant betas but time-varying prices of risk estimated using the
ordinary least squares (OLS) estimator discussed in Section 4. Columns 3
only on average but also at each point in time. Panel A of
(t ; 0 ) and 4 (0 ; 0 ) denote specifications with time-varying and constant Table 3 reports mean squared pricing errors for our model
risk exposures, respectively, and constant prices of risk. FH refers to the as implied by the different specifications and estimation
Ferson and Harvey (1991) estimator discussed in Section 5, which is based on approaches. Specifically, for each test asset i, we report the
time-varying betas estimated using five-year rolling window regressions.
quantity9
FM denotes the Fama and MacBeth (1973) two-pass estimator also based on
time-varying betas estimated using five-year rolling window regressions.
1 TX1
2
Mean squared pricing errors are stated in percentage terms. Panel B shows MSEi e^ : 40
the mean squared pricing errors of all model specifications relative to the T t 0 i;t 1
benchmark estimation. The test assets are the ten size-sorted stock decile
portfolios from Ken French's website (size1 size10), as well as constant The first column (t ; t ) shows our benchmark specifica-
maturity Treasury returns for maturities ranging from one through 30 years
(cmt1 cmt30), obtained from the Center for Research in Security Prices
tion with both time-varying betas and market prices of
(CRSP). The sample period is 1964:012012:12. risk and the betas being estimated using the approach
discussed in Section 5. The second column ( 0 ; t ) is a
t ; t 0 ; t t ; 0 0 ; 0 FH FM specification with constant betas but time-varying prices
of risk estimated using the OLS estimator discussed in
Panel A: Mean squared pricing errors
Section 4. Columns 3 (t ; 0 ) and 4 (0 ; 0 ) denote speci-
size1 5.87 6.13 7.07 7.06 6.35 6.34 fications with time-varying and constant risk exposures,
size2 2.77 2.80 3.49 3.49 3.25 3.31 respectively, and constant prices of risk. The fifth column
size3 1.96 2.00 2.80 2.80 2.45 2.53 (FH) provides the Ferson and Harvey (1991) estimator
size4 1.90 1.92 2.75 2.75 2.37 2.49
size5 1.69 1.72 2.49 2.49 2.16 2.26
discussed in Section 5, which is based on time-varying
size6 1.78 1.88 2.67 2.67 2.38 2.38 betas estimated using five-year rolling window regres-
size7 1.74 1.78 2.32 2.32 2.08 2.12 sions. Finally, the last column (FM) shows the Fama and
size8 1.51 1.52 1.96 1.96 1.79 1.81 MacBeth (1973) two-pass estimator based also on time-
size9 1.27 1.27 1.60 1.60 1.44 1.47
varying betas estimated using five-year rolling window
size10 0.33 0.33 0.58 0.58 0.48 0.54
cmt1 0.08 0.10 0.10 0.11 0.08 0.09 regressions. All mean squared pricing errors are stated in
cmt2 0.17 0.21 0.22 0.23 0.18 0.19 percent.
cmt5 0.35 0.38 0.44 0.44 0.38 0.40 The main result of the table is that none of the
cmt7 0.44 0.49 0.58 0.57 0.46 0.50 alternative estimation approaches generates mean
cmt10 0.43 0.61 0.71 0.74 0.55 0.58
cmt20 1.24 1.72 1.85 2.03 1.48 1.52
squared pricing errors that are smaller than those implied
cmt30 1.80 2.82 2.85 3.14 2.08 2.08 by the benchmark (t ; t ) specification for any of the test
assets. In particular, the specifications with constant prices
Average 1.49 1.63 2.03 2.06 1.76 1.80 of risk imply substantially larger pricing errors. The FH
estimator, which features time-varying prices of risk but
Panel B: Mean squared pricing errors relative to t ; t
betas estimated using five-year rolling window regres-
size1 1.04 1.20 1.20 1.08 1.08 sions, also produces pricing errors that substantially
size2 1.01 1.26 1.26 1.17 1.20 exceed those implied by our benchmark estimator. The
size3 1.02 1.43 1.43 1.25 1.29 relative performance of the various estimation approaches
size4 1.01 1.45 1.45 1.24 1.31
size5 1.02 1.47 1.47 1.28 1.34
can best be seen from mean squared error (MSE) ratios
size6 1.06 1.49 1.50 1.33 1.33 with respect to our benchmark estimation specification
size7 1.02 1.33 1.33 1.20 1.22 (t ; t ), provided in Panel B of Table 3. These ratios show
size8 1.01 1.29 1.30 1.19 1.19 that the benchmark specification outperforms the specifi-
size9 1.00 1.26 1.26 1.13 1.15
cation with time-varying prices of risk but constant betas,
size10 1.00 1.73 1.74 1.45 1.62
cmt1 1.27 1.26 1.32 1.03 1.05 substantially for the Treasury portfolios but, at most by a
cmt2 1.27 1.28 1.35 1.08 1.11 few percentage points for the size-sorted equity portfolios.
cmt5 1.09 1.26 1.26 1.09 1.15 This implies that in our model allowing for time variation
cmt7 1.10 1.31 1.30 1.05 1.13 in betas is relatively more important for Treasury returns.
cmt10 1.43 1.65 1.73 1.30 1.36
In contrast, allowing for time variation in prices of risk
cmt20 1.39 1.50 1.64 1.20 1.23
cmt30 1.56 1.58 1.74 1.15 1.15 dramatically reduces the mean squared pricing error, as
evidenced by the fact that both specifications with con-
Average 1.14 1.40 1.43 1.19 1.23 stant prices of risk (t ; 0 and 0 ; 0 ) imply MSEs that

estimated via five-year rolling regressions. The graphs 9


To ensure a fair comparison across estimators, we report mean
show that our joint dynamic asset pricing model fits the squared errors taken over the same sample period, thus taking into
cross section of average excess returns very well, in both account the trimming of data in the time-varying beta case.
226 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

40 0

30 5

20 10

10 15

0 20

10 25

20 30

30 35
1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010

20 45

15 50

10 55

5 60

0 65

5 70

10 75
1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010

Fig. 3. Price of MKT risk dynamics. This figure provides plots of the estimated time series of the price of MKT risk implied by the dynamic asset pricing
model with time-varying betas and prices of risk estimated using the approach in Section 5 and discussed in Section 6. The upper-left graph plots the price
of market risk along with its conditional 95% confidence interval. The remaining graphs provide the contributions of the three price of risk factors TSY10,
TERM, and DY to the dynamics of the price of market risk. All quantities are stated in annualized percentage terms. The sample period is 1964:012012:12.

exceed the benchmark specification between 20% and 74%. driven by time-varying prices of risk and, to a much
Turning to the last two columns, when betas are estimated smaller extent, by changes in the factor risk betas.
using five-year rolling regressions, allowing for prices of This finding is consistent with the results of Ferson and
risk to vary over time, as in the Ferson-Harvey estimator, Harvey (1991) and highlights the importance of using a
improves the model fit with respect to the Fama-MacBeth dynamic framework and an estimation approach consis-
estimator, but the difference is substantially smaller than tent with such a framework when testing asset pricing
when betas are estimated using Gaussian kernels. More models.
important, both the FH and FM estimators imply an We now turn to a characterization of the dynamics of
average MSE of 19% and 23% larger than that of our the price of risk. Fig. 3 provides a plot of the estimated
benchmark specification. Hence, estimators using rolling price of MKT risk implied by the model, as given by our
five-year window regressions perform substantially less benchmark estimation approach with time-varying betas
well than our estimator using time-varying betas obtained and time-varying prices of risk. The upper-left graph
from Gaussian kernel regressions.10 shows the time series evolution of the estimated price of
In sum, our results show that the time variation of risk along with its conditional 95% confidence interval. The
excess returns on stock and bond portfolios is mainly plot shows that the price of MKT risk is strongly time-
varying. While it has on average amounted to about 6%
10
over the past 50 years, there have been a few episodes in
For comparison, we consider estimators of the price of risk
parameters based on estimated betas using the level of price of risk
which the estimated price of market risk has been mark-
factors instead of innovations. However, not surprisingly, the results are edly negative. In particular, during the final two years of
very poor and so we omit them from the presented results. the dotcom bubble as well as in the two years before the
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 227

25 3

20
2
15
1
10

5 0

0 1

5
2
10

3
15

20 4
1965 1970 1975 1980 1985 1990 1995 2000 2005 2010 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010

Fig. 4. Time variation in the price of SMB and TSY10 risk. This figure provides plots of the estimated time series of the price of SMB and TSY10 risk implied
by the dynamic asset pricing model estimated using the method outlined in Section 5 and discussed in Section 6. The left panel plots the price of SMB risk
along with its conditional 95% confidence interval, and the right panel reports the price of TSY10 risk along with its conditional 95% confidence interval. All
quantities are stated in annualized percentage terms. The sample period is 1964:012012:12.

recent financial crisis, the estimated market risk premiums long-term Treasury risk generated strongly fluctuating risk
fell below zero, indicating that, according to our model, prices for stock portfolios over the last 50 years. While the
equity investors would have anticipated negative excess price of risk was mostly negative in the early part of the
returns on equity in these periods. sample, it flipped sign in the early 1980s and became
The remaining graphs in Fig. 3 show the contribution of negative again around the mid-1990s.
the three price of risk factors to these dynamics. Recall that, An important aspect of our modeling framework is that
in our model, t 0 1 F t , where Ft is the vector of price of we can use the dynamics of the pricing factors to predict
risk factors. Accordingly, the three graphs show the quan- expected excess returns further out than one month
tities 1j F jt , where 1j is the 1; j element of 1 and Fjt is the ahead. This is useful as it facilitates a quantitative analysis
jth factor in Ft. These graphs thus allow one to attribute the of risk premiums at longer-term investment horizons.
dynamics of the price of market risk to its various compo- Fig. 5 shows the model-implied expected excess return
nents. As an example, our model implies that the equity risk on the fifth size portfolio and the ten-year constant
premium was at an all-time high in the spring of 2009. maturity Treasury portfolio one year and five years into
Looking at the individual contributions of the three price of the future, as implied by our benchmark specification with
risk factors, this period was characterized by a combination time-varying betas and prices of risk. The charts indicate
of a very low ten-year Treasury yield, a relatively high term that the model-implied risk premiums feature sizable time
spread, and a fairly elevated dividend yield. variation. For the fifth size portfolio, they varied in a range
Fig. 4 shows the estimated time series of annualized from minus 15% to 30% at a one-year-ahead horizon and
prices of risk for the SMB and TSY10 factors along with between 2% and 12% at a five-year-ahead horizon. For the
their conditional 95% confidence intervals. Both series ten-year Treasury portfolio, the time variation of risk
exhibit substantial time variation. The price of SMB risk premiums is in a narrower range of around minus four
largely mimics the dynamics of the price of MKT risk, but it to slightly over 10% at the one-year horizon and between
has a somewhat lower average level. As shown in Table 2, slightly below zero and around 5% at the five-year horizon.
the average price of SMB risk is not significantly different Hence, our model and estimation approach predict mean-
from zero in our sample. However, as shown by its ingful variation of longer-term risk premiums, consistent
conditional 95% confidence interval, the price of SMB risk with the persistence of actual excess returns over long
has been significantly different from zero over various horizons. For comparison, we superimpose the corre-
subperiods in our sample. Turning to the evolution of the sponding long-horizon risk premiums implied by the
market price of TSY10 risk, shown in the right graphs of specification with time-varying betas but constant prices
Fig. 4, exposure to long-term Treasury risk was associated of risk. Not surprisingly, this specification implies only
with a positive price of risk for much of the period from minor time variation in risk premiums.
the beginning of the sample in 1963 through the early
1980s. However, around the time of the Volcker disinfla- 7. Conclusion
tion period, the price of TSY10 risk switched sign and has
since fluctuated around mostly negative values. The expo- Dynamic asset pricing models constitute the core of
sure of equity portfolios to the Treasury factor switched modern finance theory. Virtually all of the macro-finance
signs from negative to positive sometime in the mid- literature of recent decades is cast in dynamic terms,
1990s. Combined, these results imply that exposure to often giving rise to time-varying risk prices. Empirically,
228 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

30 30

25 25

20 20

15 15

10 10

5 5

0 0

5 5

10 10

15 15
1970 1975 1980 1985 1990 1995 2000 2005 2010 1970 1975 1980 1985 1990 1995 2000 2005 2010

10 10

8 8

6 6

4 4

2 2

0 0

2 2

4 4
1970 1975 1980 1985 1990 1995 2000 2005 2010 1970 1975 1980 1985 1990 1995 2000 2005 2010

Fig. 5. One-year and five-year risk premium dynamics. This figure provides plots of the estimated one- and five-year ahead expected excess returns for two
test assets implied by the dynamic asset pricing model with time-varying betas and prices of risk estimated using the approach in Section 5 and discussed
in Section 6. size5 denotes the fifth decile portfolio from the set of size-sorted stock portfolios from Ken French's website. cmt10 refers to the constant
maturity Treasury returns for the ten year maturity, obtained from the Center for Research in Security Prices (CRSP). All quantities are stated in annualized
percentage terms. The sample period is 1964:012012:12.

the time variation in prices of risk has been shown each test asset. In the third step, prices of risk are obtained
robustly (see, e.g., Campbell and Shiller, 1988; Cochrane, by either regressing the predictive slopes on the betas
2011). cross-sectionally or by an eigenvalue decomposition of
In this paper, we provide a unifying framework for the predictive slopes and betas. The three-step regression
estimating beta representations of generic dynamic asset estimator coincides with the estimator of Fama and
pricing models that impose cross-sectional no-arbitrage MacBeth (1973) when state variables are uncorrelated
restrictions and allow for betas to vary smoothly over time across time and prices of risk are constant. Our approach
and for prices of risk to vary with observable state thus nests the popular Fama-MacBeth two-pass estimator.
variables. We allow for state variables that are cross- All of the estimators presented in this paper are either
sectional pricing factors or forecasting variables for the directly or indirectly based on standard regression outputs.
price of risk, or both. Our estimation results show that all As a result, our estimation approach is computationally
three types of variables are empirically relevant. efficient and robust. We provide an application to the joint
Our regression-based estimation approach can be pricing of stocks and bonds, which features very good
explained as a three-step estimator. First, shocks to cross-sectional pricing properties with small average pri-
the state variables are obtained from a time series vector cing errors as well as strongly significant time variation of
autoregression. Second, asset returns are regressed on risk premiums. We find that the time variation in risk
lagged state variables and their contemporaneous innova- prices is more important than the time variation in betas
tions, generating predictive slopes and risk betas for for achieving good model fits.
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 229

Appendix A. Implementing the estimators

A.1. Constant betas

More concretely ^ 0;ols , ^ 0;qmle ,


^ ^
1;ols , and 1;qmle can be obtained by the following three steps.

0
0

1  0
1. Estimate the joint VAR in Eq. (1) via V^ X  ^ ols X~  where ^ ols X X~  X~  X~  and X~  T X 0 . Form U^ as the
0 0
K C  T matrix extracted from the first KC rows of V^ . Finally, construct the estimators ^ u U^ U^ =T and ^ FF F~  F~  =T .
0

0 1
2. Estimate A^ ols RZ^ Z^ Z^ and then form the heteroskedasticity robust standard errors

X !

T  0
 1
^ ^ 0 1 0
Z^ Z^
0
V^ rob T  ZZ  IN z^ t z^ t  e^ t e^ t  IN ;
t1

 0 0
where z^ t 1; F 0t  1 ; u^ t and e^ t Rt  A^ ols z^ t .

3. Estimate
0

1 0
0

1 0
^ 0;ols B^ ols B^ ols B^ ols A^ 0;ols ; ^ 1;ols B^ ols B^ ols B^ ols A^ 1;ols :
  0 0
Next, let L 1 K C , where i is the eigenvector associated with the ith largest eigenvalue of the matrix A^ ols Z^ Z^ A^ ols .
Then let

B^ qmle;0 L; D^ qmle;0 L0 A^ ols :


Define ^ ^
qmle;0 as the last KC columns of the matrix D qmle;0 . Then,

^
B^ qmle B^ qmle;0 ^ 1 D
D^ qmle ^
qmle;0 ; qmle;0 qmle;0 ;
^
and ^
qmle is the matrix formed from the first K F 1 columns of D qmle . Finally, construct the variance estimators


0
1
V^ ;ols ^ FF  ^ u H B^ ols ;
^ ^ ^ ^
ols V rob H B ols ; ols

and


0
1
V^ ;qmle ^ FF  ^ u H B^ qmle ;
^ ^ ^ ^
qmle V rob H B qmle ; qmle :

A.2. Time-varying betas


tv
^ can be obtained in three steps. The implementation requires choices of the trimming parameter T . In our empirical
application, we choose T 10  6 . In addition, to avoid boundary bias issues we drop the first and last 12 monthly
observations in our empirical application, following Ang and Kristensen (2012).

1. Estimate the time-varying joint VAR in Eq. (30). Assume t follows a polynomial of order P in t, i.e., regress X i;t 1 on
  0
t  X t  1 , where t 1; t; ; t P for i 1; ; K. Combine these coefficient estimates to form ^ t . In our application,
we choose P 6 following Ang and Kristensen (2012). Next, follow the steps in Ang and Kristensen (2009) and Kristensen
sr lr
(2012) to obtain the short-run and long-run bandwidth choices bi and bi for i 1; ; K. Then construct the estimator of
the ith row of t via
!1
h i X
T
st 0 XT
s t ~ 0
^ t  1 Kb X i;s X~ s  1 Kb X s  1 X~ s  1 ;
i;
s1
T s1
T
n o  0  
, X i;s is the ith element of Xs and X~ s  1 1; X 0s  1 . Here, Kx 2  1=2 exp x2 =2 . Then form v^ t by
sr lr
where b A bi ; bi

v^ t X t  ^ t  1 X~ t  1 . Finally, construct

X
T X
T
^ T 1 s t ~ 0 s t
x;t Kb X s  1 X~ s  1 ; ^ v;t T  1 Kb
0
v^ s v^ s ;
s1
T s1
T
where b bc is a common bandwidth choice. In our application, we use the average bandwidth chosen across the K equations.
230 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

2. Estimate the time-varying reduced-form return generating Eq. (29). Assume At follows a polynomial of order P in t, i.e.,
    0
regress Ri;t 1 on t  ztv
t , where t 1; t; ; t
P
for i 1; ; N. Combine these coefficient estimates to form A^ i;t . In
our application we choose P6 following Ang and Kristensen (2012). Next, follow the steps in Ang and Kristensen (2009)
sr lr
and Kristensen (2012) to obtain the short-run and long-run bandwidth choices hi and hi for i 1; ; N. Then, construct
the estimator of Ai;t via

!1
X
T
s  t tv tv0 XT
s  t tv
A^ i;t  1 Kh zs zs Kh zs Ri;s ;
s1
T s1
T
n o 0
0  
~
s X s  1 ; C s . Here, Kx 2
0  1=2
exp x2 =2 . Then, form e^ i;t Ri;t  A^ i;t  1 ztv
sr lr
where h A hi ; hi and ztv t . Finally,

construct

X
T X
T
^ T 1 s t ~ 0 s t
f ;t Kh F s  1 F~ s  1 ; ^ e;t T  1 Kh
0
e^ s e^ s ;
s1
T s1
T
where h hc is a common bandwidth choice. In our application, we use the average bandwidth chosen across the N
equations.
3. Estimate
!1
tv
1
TX
0 0

1
TX
0

^
vec F~ t F~ t  B^ t B^ t T  I K C K F 1 F~ t  B^ t Rt 1  B^ t u^ t 1 ;
ols
t0 t0
h i0
where B^ t ^ 1;t ^ N;t and ^ i;t are the last KC elements of A^ i;t from Step 2 using the long-run bandwidths hi for
lr

i 1; ; N. Finally, construct the variance estimators


tv tv tv
V^ V^ ;1 V^ ;2 ;
where
" #1
T
X

0
tv
V^ ;1 T  ^
 B^ t  1 B^ t  1
f ;t
t1
" #
T
X

tv0 tv 0
^
^ 1 ^ ^ ^ ^ ^ ^
f ;t ols DB z;t DB ols f ;t f ;t  B t  1 e;t B t  1
0

t1
" #1
T
X

0
 ^
 B^ t  1 B^ t  1
f ;t
t1

and
" #  1" #
T
X
T
X

0 0 0
tv
V^ ;2 T  ^
^ ^ ^
^ ^ ^ ^ ^
f ;t  B t  1 B t  1 f ;t  B t  1 B t  1 u;t B t  1 B t  1
t1 t1
" #1
T
X

0
 ^
B^ t  1 B^ t  1
f ;t  :
t1

Appendix B. Imposing restrictions on parameters

Although the classification of state variables into risk and price of risk factors allows for the specification of more
parsimonious models, situations still could exist in which one would like to impose zero (or other linear) restrictions to the
parameter of interest (or possibly to B). These restrictions could be most easily imposed by the following steps. Suppose,
 
without loss of generality,
 the
 restrictions are of the form H vec 0, whereH is a known q  K C K F 1 matrix with
0 0
rankH q, vecB ; vec , and the restrictions do not violate that rank B0 B K C . For example,
0

0 if one wanted to
impose the restriction that the second element of 0 is equal to zero, then H 00NK C 1 ; 0; 1; 0; 00 .
Let B^ and ^ , and the corresponding ^ , stand in for either the OLS or QMLE estimators introduced in this paper. Then the
restricted estimator can be found by

0
1
^ r min ^ W T ^ ^ W T 1 H0 HW T 1 H0 H ^ : 41
s:t: H vec 0
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 231

The optimal weight matrix is one that satisfies W T -p V  1 , as T-1, where V is the asymptotic variance of ^ . Under this
 
choice of weighting matrix with H vec 0,
p

d  1

T vec ^  N 0; V V H 0 HV H 0 HV ; 42

as T-1. In the case of B^ ols and ^ , V is


ols
" #
V B;ols Cols    0
V ; Cols 0NK F 1 I NK C V rob H B; ; 43
C0ols V
^
and V B;ols is the NK C  NK C bottom right sub-matrix of V rob . In the case of B^ qmle (or B^ 4ols ) and qmle , V is
" #
V B;qmle Cqmle
V ; 44
C0qmle V

   0    0
Cqmle HB B; V rob H B; ; V B;qmle HB B; V rob HB B; ; 45
and
   0   1 

HB B; FF u IK C ZZ  IN
 1
 
0
 FF u FF  B H B; ; 46
0
where ZZ p lim Z^ Z^ =T. Further details are provided in Appendix D.
T-1
In the case in which betas are time-varying we can follow similar steps. For the linear restriction H 0, where
 
vec , the restricted estimator can be written as
tv tv

tv

^ r vec ^ ols  W T 1 H0 HW T 1 H0 H vec ^ ols : 47


 tv   1
The optimal weight matrix is one that satisfies W T -p V as T-1. Under this choice of weighting matrix with H 0,
p tv
d  tv 0   1

T ^ r  N 0; V tv tv 0
 V H HV H HV tv
: 48

Appendix C. Preliminary results

Before proving Theorem 1, we provide some useful results on reduced rank regressions that are used throughout the
Appendix. We work in the generality of the model

Y t AX t FZ t t ; t 1; ; T; 49
 0 0 0
~
is1full rank. Let G A F  and Z t X t ; Z t . If
where ABD, B is n  k, D is k  m, k ominn; m, X t is m  1, Z t is p  1, and F

0 0
we stack the model, we have Y AX FZ E GZ~ E and G^ ols Y Z~ Z~ Z~ . Under the population moment condition
 
E Z~ t  t 0, the GMM objective function can be written as
!0 !
X
T   X
T  
T T 1 Z~ t  t W gmm T  1 Z~ t  t
t1 t1


0
0

0

T  vec Y  G^ ols Z~ G^ ols  G Z~ Z~ W gmm vec Y  G^ ols Z~ G^ ols  G Z~ Z~


1


0 0


0


T  vec G^ ols  G Z~ Z~ =T  I n W gmm Z~ Z~ =T  I n vec G^ ols  G ; 50


0


0

md
which is the MD criterion function with W Z Z =T  I n W~ ~ gmm ~ ~
Z Z =T  I n . Thus, the GMM and MD criterion
functions are one-to-one.  To show that ML is a special  case of MD/GMM, note that under the assumption
vecE  N 0; I T  2 I n the log-likelihood is B; D0 ; 2  nT2 log 2
2
 21 2 tr E 0 E . However,

0


0
0

trE 0 E tr Y  G^ ols Z~ Y  G^ ols Z~ tr G^ ols  G G^ ols  G Z~ Z~ ; 51


0

so that the ML, which solves minG trE 0 E is


the same as the

MD estimator with weight matrix Z~ Z~ =T  I n . Under the
general symmetric weight function W md W md 1  W md
2 with
2 3
md md
W W 1;12
W md 4 1;11 5;
1 52
W 1;12 W md
md0
1;22
232 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

and the normalization B0 W md


2 B I k (because B and D are not separately identified without further assumption), it can be
shown that the MD estimators are

 1=2 0


1
B^ md W md L; D^ md B^ md W md ^ F^ md F^ ols A^ ols  B^ md D
^ md W md W md
2 2 A ols ; 1;12 1;22 ; 53
 
where L 1 k and i is eigenvector associated with the ith largest eigenvalue of the matrix

1=2
1
1=2
W md A^ ols W md md md
W md ^0 md
2 1;11 W 1;12 W 1;22 1;21 A ols W 2 : 54

This follows because



0

minvec G^ ols  G W md1  W md


2 vec G^ ols  G
B;D;F
0

0

tr G^ ols W md ^ md
2 G ols W 1 tr G0 W md md
2 GW 1  2 tr G^ ols W md md
2 GW 1 : 55

We can ignore the first term as it is not a function of B, D, or F. If we fix A (i.e., B and D) and solve for F, then



1

1
F^ md A^ ols  A W md ^ md
1;12 F ols W 1;22 W md
1;22 F^ ols A^ ols BD W md md
1;12 W 1;22 ; 56
0
and plugging this back in and using the normalization B W md
2 B Ik we can obtain

0

min vec G^ ols G W md1  W md


2 vec G^ ols  G
B;D;F

1
0
tr D W md md md
1;11  W 1;12 W 1;22 W md0
1;12 D


1
 2  tr D W md W md
W md
W md0 ^ 0 W md B :
A 57
1;11 1;12 1;22 1;12 ols 2

^ md B^ 0 W md A^ ols . Plugging this back in yields the following maximization


Given B we can solve for D, which yields D md 2
problem:

1=2
1
1=2
max tr B~ W md
0
^
A W md
 W md
W md
W md0 ^ 0 W md
A B~
0
s:t: B~ B~ I k ; 58
2 ols 1;11 1;12 1;22 1;12 ols 2
B~

1=2
where B~ W md 2 B and the result follows.
Using these derivations, it is straightforward to form a bias-corrected estimator of for the bias induced by replacing
ut 1 by u^ t 1 . In particular, this bias arises because u^ t 1 is a function of X 1;t , which does not show up in the formulation for
returns in Eq. (13). The prescription to deal with this bias is simply to include X 1;t in the first-step regression (associated
with a full-rank coefficient matrix). The degree of the bias is affected by a subset of elements of , namely, those parameters
that designate how strong the predictive power of X1-type variables is for X1- and X2-type variables. The proofs of Theorems
1 and 2 can then be straightforwardly modified to provide appropriate limiting distributions for these estimators using the
results in this section and Appendix D.

Appendix D. Proofs

D.1. Constant betas

For the results in the constant-beta case, we make the following assumptions (in addition to those made in the main
text): (i) all eigenvalues of have modulus less than one; (ii) v;t v for
 0  all t and

v is positive definite; (iii) the initial
condition X0 is fixed; (iv) R0t ; vt is a stationaryergodic sequence with E  R0 ; vt 0 4 o 1; (v) the matrix B0 B has minimum
 t
eigenvalue bounded away from zero; and (vi) E ei;t vt v0t jF t  1 0 8 t and i 1; ; N.
All of these assumptions are standard except perhaps assumption (vi). Assumption (i) ensures that the dynamics of Xt are
stationary. From an economic perspective, this restriction rules out phenomena such as rational bubbles that would be
associated with exploding risk premiums. From a statistical point of view, the assumption allows us to avoid nonstandard
asymptotic arguments. Assumption (ii) is natural given that B does not time-vary in this case. Assumption (iii) ensures that
the influence
  of the initial condition is asymptotically negligible. Assumption (v) guarantees that the matrix B0 B satisfies
rank B0 B K C . Intuitively, we are assuming away the presence of redundant, uninformative or unspanned factors.
Assumption (vi) limits the degree of dependence between ei;t and vt and consequently simplifies our asymptotic  variance
0
formulas. To provide intuition for this assumption note that it would hold in the case that we assumed that R0t ; vt are
jointly independent and identically distributed, conditional on F t  1 , from an elliptically symmetric distribution. Under
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 233

these assumptions we have the following results


2 0
3
T  1=2  vec V X~  "  #!
6 0
7
6  1=2
6T
7 d
7
XX  v 0
~
 vec EX  7N 0; ; 59
6 0 V rob
4   5
 1=2
T  vec EV 0
p 0
and V^ rob V rob , where XX p lim X~  X~  =T .
T-1

D.1.1. Proof of Theorem 1 h i 0


1 0
^ . Let A^
We first show the result for ^ ^ ^ ^ ^ B^ ols A^ 01;ols . Then,
ols 01;ols A 1;ols A 2;ols so that ols B ols B ols

^ ols
0
1 0 0
0

1
B^ ols B^ ols B^ ols RM U^ F~  F~  M U^ F~ 
0
0

1 0
1 0 0
0

1
UM U^ F~  F~  M U^ F~  B^ ols B^ ols B^ ols EMU^ F~  F~  MU^ F~ 
0
1 0
0
1 0
0
0

1
 B^ ols B^ ols B^ ols B^ ols  B  B^ ols B^ ols B^ ols B^ ols  B UM U^ F~  F~  M U^ F~  ; 60
0

0 1

where M U^ I T  U^ U^ U^ U^ . Under our assumptions, the last term is op T  1=2


so that
p

T vec ^  T T T o 1; 61
ols 1 2 3 p

where

1

0 0
T1 F~  F~  =T  I K C vec T  1=2 U F~ ; 62

   1 0
p

T 2 I K F 1  B0 B B vec T A^ 01;ols  A ; 63

and
   1 0
p

B vec T B^ ols B :
0
T 3   B0 B 64
p

^
Under
our
assumptions,
0    T vec A ols A -d N 0; V rob and is asymptotically uncorrelated with
vec T  1=2 U F~ -d N 0; FF  u and so the result follows.
Now let us consider ^ md ^ ^0 md
qmle . By the derivations above (when F0) and with weight matrix W 1 Z Z =T and W 2 I N ,
^
then we can find B^ qmle and qmle as transformations of the KC eigenvectors associated with the largest eigenvalues of the
0
0
matrix A^ ols Z^ Z^ =T A^ ols (see Appendix A). By standard properties of MD estimators, we know that the asymptotic variance of

0
vec B^ qmle ; vec ^ is
qmle


 1

1
V B;qmle J 0md W md J md J 0md W md V rob W md J md J 0md W md J md ; 65

where
2

3
vec A^ ols  A vec A^ ols  A
J md 4  0 5
vecB0 vec
h  0   h i0 i
 I K C  I N  I K C K F 1  B I K C K F 1 0K C K F 1K 2 : 66
C

After incorporating the uncertainty from replacing U by U^ , it can then be shown that this yields
2    0    0 3
HB B; V rob HB B; HB B; V rob H B;
V B;qmle 4    0
   0 5; 67
H B; V rob HB B; FF 1  u H B; V rob H B;

where
   0   1 

HB B; FF u IK C ZZ  IN
 1
 
0
 FF u FF  B H B; ; 68
0 0
and ZZ plimT-1 Z^ Z^ =T plimT-1 Z~ Z~ =T.
234 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

D.1.2. Proof of Theorem 2


P b ^ ^
Let F EF t  and ^ F T  1 Tt 1 F t . Here we derive

 1the asymptotic distribution of the estimator 0 1 ^ F . We could
also estimate F by the last KF elements of I K  ^ ^ . These two approaches are asymptotically equivalent. Then,


 

b  ^ 0  0 ^ 1  1 F 1 ^ F  F op T  1=2 : 69
 0 ~ 0 
Define ~ F 1; 0F and 1 K C K 1 1 so that

 
 

b  ~ 0F  IK C vec ^  ~ 1 ^  op T  1=2 ; 70
PT
where ^ X T 1
t 1 X t . Note that
p 
T ^ X  X I K   1 T  1=2 V T op 1 71
and, from the proof of Theorem 1,
p


0
 p

vec T ^  1  I ~ T vec A^ ols  A op 1:


ols FF K C vec U F  H B; 72
Under our assumptions, the only covariance term arises from
p
0 p
0 X T
T X
0

T  1 vec V T = T vec U F~ = T T  1 F~ s  1  vt u0s : 73


s1t 1

For s a t, the sum converges in probability to zero under our assumptions so that
p b
d

T  N 0; V ; 74

where
   0 ~ h i0 0
V ~ 0F  I K C V ~ 0F  I K C 1 I K 
1
v I K   1 ~ 1 C C 0 ; 75
~ I   1 , and
C vu is formed from the first KC columns of the matrix v.
1 K vu

D.1.3. Derivations for Appendix B p

First we derive the asymptotic covariance matrix Cols . The asymptotic variance of T vec B^ ols  B is the bottom-right
block element of the matrix V rob and, from the proof of Theorem 1,
p


0
 p

vec T ^  1  I ~ T vec A^ ols  A op 1:


ols FF K C vec U F  H B; 76
   0
Thus, Cols 0NK F 1 I NK C V rob H B; . Next, we derive the asymptotic covariance matrix Cqmle . From the proof of
Theorem 1,
p

 p

vec T B^ qmle  B HB B; T vec A^ ols  A op 1;


   0
so that under our assumptions Cqmle HB B; V rob H B; .

D.2. Time-varying betas

For the results in the time-varying beta case, we proceed conditional on the realizations of the random processes 
and i  for i 1; ; N. However, we suppress these arguments in the expectation operator to simplify notation. To simplify
the notation in this Appendix we map, without loss of generality, i;t i;t 1 , t t 1 , and t t 1 . For the case in which
betas are time-varying it is more convenient to state the assumptions in terms of the linear model Y t;T At;T X t;T E t;T , t 1,
,T, which nests both Eqs. (29) and (30). Although this is a triangular array of models, we suppress the dependence on T for
  h 0
i h 0
i  
simplicity of notation. Finally, define z;t E ztv tv0
t zt , f ;t E F~ t  1 F~ t  1 , x;t E X~ t  1 X~ t  1 , e;t E et e0t , and
 0
v;t E vt vt , where u;t is the matrix formed from the first KC rows and columns of v;t . We make the following
assumptions.
h   i
(i) For all t Z1, supT Z 1 supt r T E J X 0t ; E 0t J 8 4 o1 for some 4 0 and is mixing where the mixing
coefficients
 
mT i sup sup PA \ B PAPB
 T r r T A A F  1 ;B A F 1
T i

satisfy mT i r mi, T Z 0, and the dominating sequence mi is geometrically decreasing. E t is a martingale


difference sequence with respect to F t X t ; E t  1 ; X t  1 ; E t  2 ; .
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 235

  
(ii) The observed data Y t ; X 0t : t 0; ; T have pbeen
symmetrically trimmed with positive trimming sequence
aT, which satisfies aT =bT -0, aT =hT -0, and T aT -0.
 
(iii) The sequence i;t satisfies  i;t i t=T o1 for i 1; ; N and similarly for t, x;t , e;t , and v;t for some
functions i ,     , x , e  and v , respectively. The elements of these functions are in
Cr 0; 1, the space of r times continuously differentiable functions, for some r Z2. For all A 0; 1, x , e ,
and v  are positive  definite with eigenvalues bounded and bounded away from zero. Finally,
sup0 r r 1  max  is bounded below one, where max  is the maximum eigenvalue of a matrix.
P   R1  
(iv) limT-1 T  1 Tt 1 f ;t  B0t Bt 0 f  B0 B d exists and is positive definite with all eigenvalues
bounded and bounded away from zero. Also,
T
X

lim T  1 f ;t 0 D0B z;t


1
DB f ;t f ;t  B0t e;t Bt
T-1
t1
Z 1

f 0 D0B z  1 DB f f  B0 e B d 77
0

and

X
T  
lim T  1 f ;t  B0t Bt u;t B0t Bt
T-1
t1
Z 1 
f  B0 B u B0 B d; 78
0
P
and limT-1 T  1 Tt 1
EX t  exist.
  
(v) X0 is fixed, and for 1 ri rN, E ei;t vt v0t F t  1 0.

(vi) The
 kernel K satisfies the following conditions. There exists B; L o1 such that either Kw 0 for kwk 4L and
Kw Kw0  rBkw  w0 k, or Kw is differentiable with Kw=wj rB and, for some 4 1,
   R
Kw=wj r kwk  for kwk ZL. Also, Kw rBkwk  for kwkZ L. Kw dw 1 and for some r Z 2,
R i R
w Kw dw 0 for i 1; ; r  1 and jwjr Kw dw o 1.
p .
(vii) The sequence T satisfies T T -0. The bandwidth sequence hT satisfies ThT -0, logT 2 .ThT -0, and
2r 2
 =
T  1=2 hT
1 2
-0 for some 4 0. The bandwidth sequence bT satisfies TbT -0, logT 2 TbT -0, and
2r 2
 1=2  1 =2
T bT -0 for some 4 0.

Assumptions (i)(iii) and (vi)(vii) are essentially the same as those of Kristensen (2012). The remaining assumptions are
tailored to our model specification. Following similar steps as in Section 3, the martingale difference assumption implies
 
that E Mt 1 Rt 1 es ; vs : s rt 0. Thus, these assumptions are consistent with the asset pricing restrictions discussed in
Section 3. When implementing the estimators introduced in this paper, different bandwidths should be used for each series
(see Appendix A.2), however, without loss of generality, the derivations rely on a common bandwidth choice hT and bT to
simplify the presentation. In addition, we suppress the dependence of the bandwidth sequences on T. Finally, define
  q
 
m   tr A0 A the matrix Euclidean norm.
i 1 Ai A1 A2 Am for a sequence of square matrices and A
To proceed, we make repeated use of the following two lemmas. The second lemma is Lemma B.11 of Kristensen (2012).
We restate it for convenience.

Lemma D.1. Under our assumptions,

X h i

T
st   0 0 log T
a ^ t  t T  1 Kb  s  t X~ s  1 X~ s  1 vs X~ s  1 x;t Op b
1 2r
Op ;
s1
hT bT

X

T
st   1 log T
A^ t At T  1  As At zs z0s es z0s z;t Op h
2r
b Kh Op ;
s1
T hT
  r!
   r logT
c sup A^ t  At  Op h Op ;
1rt rT hT
  r!
   r logT
d sup ^ t  t  Op b Op ;
1rt rT bT

uniformly over 1 r t r T.

Proof of Lemma D.1. Parts (a) and (b) follow by the same steps as in the proof of Theorem 2 in Ang and Kristensen (2009).
Parts (c) and (d) follow Kristensen (2009).
236 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

 
Lemma D.2. Assume that mt =2 o t  2 for some ; 40. Then, for any symmetric function T Y s ; Y t , the following
decomposition holds:
 1X T h i
T 2X
T Y s ; Y t T Y t  T RT ;
2 sot
Tt1 T
where
 1X
T    
T E T Y s ; Y t ; T y E T y; Y t ;
2 sot

and
h i
1=2
h 2 i
1=2 :
E R2T O sT;  T  1 =2 ; sT; sup E T Y s ; Y t 
sat

D.2.1. Proof of Theorem3


Throughout we use zt instead of ztv
t for simplicity of notation. We first find the asymptotic linear representation of
" #1
p tv
T
X

^
T vec T 1 ~ t  1 F~ 0
F  ^ 0 B^ t  I K K 1
B
ols t1 t T C F
t1
T
X

0
T  1=2 F~ t  1  B^ t vec Rt  B^ t u^ t : 79
t1

The first factor satisfies


" #1 Z 1
T
X
 
0 0
T 1
F~ t  1 F~ t  1  B^ t B^ t T f  B0 B d op 1: 80
t1

This follows because


 
 T
0

 1 X ~ 0 
T F t  1 F~ t  1  B^ t  Bt Bt 
 
t1
T 
X 
0 
~ 0  
rT  1 F t  1 F~ t  1  B^ t  Bt Bt 
t1

 XT  
  ~ ~0 
rC sup  B^ t  Bt  sup kBt k  T  1 t1
 F t  1 F t  1 
1rt rT 1rt rT

 X
T

  0
C sup  B^ t  Bt   T  1 tr F~ t  1 F~ t  1
1rt rT t1
r!
 r logT
Op h Op ; 81
hT
and, by similar steps,
T
X
0



0 log T
T 1 F~ t  1 F~ t  1  B^ t  Bt B^ t Bt
2r
Op h Op : 82
t1
hT
P 0

Thus we just need to deal with the term T  1=2 Tt 1 B^ t Rt  B^ t u^ t . Combining




 
 
Rt  B^ t u^ t B^ t F~ t  1  B^ t  Bt F~ t  1 ut  Bt u^ t  ut  B^ t Bt u^ t  ut et ; 83
and, by similar steps as above, that
 

 
  1=2 X ^ 0 ^ 
T
^ ~ 
T B t B t  Bt u t ut F t  1  op 1 84
 t1

yields
"Z #1
tv
1    
^  
vec f  B0 B d vec
ols T
0
"Z #1
1  
f  B B d
0
0
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 237

T
X
h

 i

0
T  1 F~ t  1  B^ t  B^ t Bt F~ t  1 ut et  Bt u^ t  ut op T  1=2 : 85
t1

The first term is op T  1=2 under our assumptions and so we only need deal with

X
T
0


 
0
T  1=2 B^ t  B^ t  Bt F~ t  1 ut et Bt u^ t ut F~ t  1
t1
X
T

 
0
T  1=2 B0t  B^ t Bt F~ t  1 ut et  Bt u^ t  ut F~ t  1 op 1; 86
t1

where the equality follows because


 
 T
0

  1=2 X ^ ^ ~ ~ 0 
T B t  Bt B t  Bt F t  1 F t  1  op 1; 87
 
t1

 
 T
0

  1=2 X ^ ^ ~ 0 
T B t  Bt B t  Bt ut F t  1  op 1; 88
 t1


 
 T
0 
  1=2 X ^ ~ 0 
T B t  Bt et F t  1  op 1; 89
 t1

and
 
 T
0   0 
  1=2 X ^ 
T B t  Bt Bt u^ t  ut F~ t  1  op 1; 90
 
t1

under our assumptions. Eqs. (87), (88), and (90) follow by similar steps as in Eq. (81) and Eq. (82). Equation (89) is

T
X
0 0
T  1=2 B^ t  Bt et F~ t  1
t1
X
T
0 0
T  1=2 D0B A^ t  At et F~ t  1
t1

X
T X
 1  0
T
st
T  3=2 Kh  D0B z;t zs z0s As  At 0 zs e0s et F~ t  1 op 1; 91
s1t 1
T
where DB is the K 1 K C  K C matrix, which satisfies At DB Bt and the second equality follows by Lemma D.1. To find the
order of this term, we follow similar steps as in Ang and Kristensen (2009). Thus, we need to find the order of two terms:

X
T
0
Kh 0  D0B z;t zt e0t et F~ t  1
1
T  3=2 92
t1

and
X s  t 0  1   0
T  3=2 Kh  DB z;t zs z0s As  At 0 zs e0s et F~ t  1 : 93
sat
T
For Eq. (92), note that
 
    0 
  3=2 X X
T T 
0  Kh 0   0 
Kh 0  D0B z;t zt e0t et F~ t  1  rC  T  3=2 zt et et F~ t  1 
1
T
 t1
 t1

1
X
T
r C  T  1=2 h  T 1 kzt k2 ket k2
t1

1
Op T  1=2 h ; 94

which is op 1 under our assumptions. For Eq. (93) note that


X s  t 0  1   0 X
2  T  3=2 Kh  DB z;t zs z0s As At 0 zs e0s et F~ t  1 T  3=2 0;T Y s ; Y t ; 95
sot
T sot

where

0;T Y s ; Y t 0;T Y s ; Y t 0;T Y t ; Y s 96


238 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

and

s t  1  0
0;T Y s ; Y t Kh  D0B z;t zs z0s As  At 0 zs e0s et F~ t  1 ; 97
T
 
where Y t et ; zt ; t with t Tt A 0; 1. We can, without loss of generality, proceed under the assumption that t  iid U 0; 1.
h i  r
Note first that E 0;T Y s ; Y t O h . Next, define y e; z; ,
h i h      1 0  0  0 i
E 0;T y; Y t E K h  t  D0B z t zz A A t ze0 et F~ t  1 0; 98

and
h i h       0  0i
E 0;T Y t ; y E K h  t  D0B z  1 zt z0t A t A zt e0t eF~
0  r  r
eF~  O h o h ; 99
so that, by Lemma D.2,
X   p
T  3=2 0;T Y s ; Y t Op hr T  R0;T : 100
sot
h
2 i1=2
The remainder term of the Hoeffding decomposition is R0;T Op T  1 =2 sups a t E 0;T Y s ; Y t  and, under

our assumptions,
h 2 i1=2

O h
 1 =2
supE 0;T Y s ; Y t  : 101
sat

Then we have

X
T

 
0
T  1=2 B0t  B^ t  Bt F~ t  1 ut et  Bt u^ t  ut F~ t  1 T tv tv tv
1 T 2 T 3 ; 102
t1

where

X
T

0
T tv
1 T
 1=2
B0t B^ t Bt F~ t  1 ut F~ t  1 ; 103
t1

X
T   0
T tv
2 T
 1=2
B0t Bt u^ t  ut F~ t  1 ; 104
t1

and

X
T
0
T tv
3 T
 1=2
B0t et F~ t  1 : 105
t1

T tv tv tv tv
3 is already simplified, so we need only simplify T 1 and T 2 . First consider T 1 :

X
T

0
T tv
1 T
 1=2
B0t B^ t Bt F~ t  1 ut F~ t  1
t1

X
T X
T
s t
 T  3=2 Kh 
s1t 1
T
  1  

B0tAs At zs z0s es z0s z;t DB DF;z zt Du vt z0t D0F;z op T  1=2


 1=2
T tv tv
1;1 T 1;2 op T ; 106

where the first equality follows by Lemma D.1, DF;z is the K F 1  K 1 K C matrix such that DF;z zt F~ t  1 , Du is the
K C  K matrix such that Du vt ut , and
XT 1  
T tv
1;1 T
 3=2
t1
Kh 0  B0t et z0t z;t DB DF;z zt Du vt z0t D0F;z 107
and
X s  t 0   1  
T tv
1;2 T
 3=2
Kh  Bt As  At zs z0s es z0s z;t DB DF;z zt Du vt z0t D0F;z : 108
sat
T
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 239

 tv 

T  Op h  1 T  1=2 by similar steps as above, and T tv is


1;1 1;2

X
T tv
1;2 T
 3=2
1;T Y s ; Y t ; 109
sot

where 1;T Y s ; Y t 1;T Y s ; Y t 1;T Y t ; Y s ,



s t   1
0
1;T Y s ; Y t  Kh  B0t As  At zs z0s es z0s z;t DB F~ t  1 ut F~ t  1 ; 110
T
 
Y t et ; zt ; t , and t Tt . Then, by Lemma D.2,
" #
p X
T
T tv
1;2 T  T 1 1;T Y t R1;T op 1; 111
t1
h i h i
where 1;T y E 1;T y; Y t E 1;T Y t ; y . We have
h i
E 1;T y; Y t
 
1    0      1  
E  Kh  t B t=T A  A t zz0 ez0 z t DB DF;z z t D0F;z
h
Z 1
 Kh  B0 A  Azz0 ez0 z  1 DB DF;z z D0F;z d
0
Z h  1
0  1
 K B  h A  A  hzz0 ez0 z  h DB DF;z z  hD0F;z d
1
 1h
 r  r
 B0 ez0 z  1 DB DF;z z D0F;z O h o h : 112
Similarly,
h i
E 1;T Y t ; y
h           i
E  Kh  t  B0 A t  A z t z  1 DB DF;z z Du v z0 D0F;z
Z 1
 
 Kh   B0 A  Az z  1 DB DF;z z Du v z0 D0F;z d
0
Z h  1
 
 Kh  B0 A h  Az hz  1 DB DF;z z Du v z0 D0F;z d
 1h  1
 r  r
O h o h : 113
Thus the contribution from this term is

X
T
1
T tv
1;2  T
 1=2
B0t et z0t z;t DB DF;z z;t D0F;z op 1; 114
t1
p
because T R1;T op 1 under our assumptions by similar steps as for R0;T . Next consider T tv
2 :

X
T   0 X
T
0
T tv
2 T
 1=2
B0t Bt u^ t  ut F~ t  1 T  1=2 B0t Bt DU ^ t  t X~ t  1 X~ t  1 DF;x ; 115
t1 t1

where DF;x is the K F 1  K 1 matrix such that DF;x X~ t  1 F~ t  1 . Then by Lemma D.1,

X
T X h i
T
s t 0  0 0 0
Bt Bt DU s  t X~ s  1 X~ s  1 vs X~ s  1 x;t X~ t  1 X~ t  1 D0F;x op 1
 3=2 1
T tv
2 T Kb
t 1s1
T
T tv tv
2;1 T 2;2 op 1; 116
where

X
T
0 0
Kb 0  B0t Bt DU vt X~ t  1 x;t X~ t  1 X~ t  1 D0F;x ;
 3=2 1
T tv
2;1 T 117
t1

and
X s  t 0 h  0 0
i 0
Bt Bt DU s  t X~ s  1 X~ s  1 vs X~ s  1 x;t X~ t  1 X~ t  1 D0F;x :
 3=2 1
T tv
2;2 T Kb 118
sat
T
240 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

 tv 

T  Op T  1=2 b  1 by similar steps as above, while


2;1

X  
T tv
2;2 T
 3=2
2;T Y 2;s ; Y 2;t ; 119
sot

where
     
2;T Y 2;s ; Y 2;t 2;T Y 2;s ; Y 2;t 2;T Y 2;t ; Y 2;s ; 120
h i
  st 0  0 0 0
Bt Bt DU s  t X~ s  1 X~ s  1 vs X~ s  1 x;t X~ t  1 X~ t  1 D0F;x ;
1
2;T Y 2;s ; Y 2;t Kb 121
T

and Y 2;t vt ; X~ t  1 ; t . Then, by Lemma D.2,


" #
p X
T  
T tv
2;2 T  T 1 2;T Y 2;t R2;t op 1; 122
t1
  h  i h  i

where 2;T y2 E 2;T y2 ; Y 2;t E 2;T Y 2;t ; y2 and y2 v; X~ ; . First,


h  i h    0   h   0 0
i   0
i
E 2;T y2 ; Y 2;t E Kb  t B t B t DU  t X~ X~ vX~ x t
1 ~
X t  1 X~ t  1 D0F;x
Z 1 h i
 0 0
Kb  B0 BDU  X~ X~ vX~ D0F;x d
0
Z b  1 h i
 0 0

1
0
K B  h B  hDU   h X~ X~ vX~ D0F;x d
 1b
0  r  r
B0 BDU vX~ D0F;x O b o b : 123
Similarly,
h  i h   h    0 0
i 0
i
E 2;T Y 2;t ; y2 E Kb  t B0 BDU t  X~ t  1 X~ t  1 vt X~ t  1 x  1 X~ X~ D0F;x
"Z #
1   0
Kb  B BDU  x x
0 1
d X~ X~ D0F;x
0
"Z #
b  1   0
K  B BDU  h  x  hx
0 1
d X~ X~ D0F;x
 1b  1
 r  r
O b o b : 124
Thus, the contribution from this term is

X
T
0  r  r
T tv
2;2 T
 1=2
B0t Bt DU vt X~ t  1 D0F;x O b o b ; 125
t1
p
because T :R2;t op 1 under our assumptions by similar steps as for R0;t . Thus,

p tv
Z  
1
^ 
T vec f  B0 B d
ols
"
XT

 
0 1
 T  1=2 DF;z I K 1 K C  z;t D0F;z D0B z;t  B0t vec et z0t
t1
#
X
T   0

T  1=2 DF;z  B0t Bt Du vec vt X~ t  1 op 1; 126


t1

and the result follows by Wooldridge and White (1988) and DF;z DB 0.

D.2.2. Proof of Theorem 4


By Theorem 3 and similar steps as in the proof of Theorem 2, we have that
! !
p ^ tv p tv
XT XT

^
T ols  T ols  T 1 ~
~ F;t 1 T  1=2
^ X;t  X;t op 1; 127
h i t1 t1

P
where ~ F;t E F~ t and X;t EX t . By Theorem 3, we only need focus on the expression T  1=2 Tt 1 ^ X;t  X;t . Recursive
substitution yields that
" !0 #
XT X T XT X
t1 t

0
T 1
EX t  T 1
t T 1
i s op T  1=2 ; 128
t1 t1 t1 s1 i s1
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 241

PT
with associated plug-in estimator, T  1 t1 ^ X;t . We aim to write
T
X
X
X  
T T
^ X;t  X;t T  1 w ^
T 1 t vec t  t T wt ^ t  t ;
1
129
t1 t1 t1
 
where w
t wt  t ;  t , wt wt  t , and  t 1 ; ; t  1 ; t 1 ; ; T and similarly for  t . We now need to

find the weights wt and wt . It is more straightforward to deal with the weights wt ,

X
T
  X
T   X
T  
T 1 wt ^ t  t T  1 ^ t  t T  1 ~ t ^ t  t ;
w 130
t1 t1 t1

where
!0
X
T 1
~ t w
w ~ t  t 02 ; 131
1 t 1 2 t 1

so that
!0

X
T 1
wt wt  t I K 02 132
1 t 1 2 t 1

with wT I K . Next we need to find w
t .
!0 !0 !0 !!
  X
t1 t1 X
T 1
w
0 0
t t; t 2 1  IK
2 ; 133
1 1 2 1 1 1 t 1 2 t 1

with w
1 0. Let
  h  i
wt wt  t ;  t wt  t w
t t; t : 134
Then by repeated applications of Lemma D.1, we have that
T
X
X
T

T  1=2 ^ X;t  X;t T  1=2 wt vec ^ t  t op 1: 135


t1 t1
By an additional application of Lemma D.1,

X
T

T  1=2 wt vec ^ t  t
t1
X
T X h  i

T
st  0 0
T  3=2 wt vecKb s  t X~ s  1 X~ s  1 vs X~ s  1 x;t
1
op 1; 136
t 1s1
T

under our assumptions. Let Y 4;t vt ; X~ t  1 ; 1 ; ; T and, following steps in the proof of Theorem 3,

X
T

T  1=2 wt vec ^ t  t T tv tv
4;1 T 4;2 op 1; 137
t1

where
XT 0

Kb 0wt vec vs X~ s  1 x;t


 3=2 1
T tv
4;1 T t1
138
and
X  
T tv
4;2 T
 3=2
4;T Y 4;s ; Y 4;t ; 139
sot
     
where 4;T Y 4;s ; Y 4;t 4;T Y 4;s ; Y 4;t 4;T Y 4;t ; Y 4;s and
h i

  st  0 0
4;T Y 4;s ; Y 4;t Kb wt vec s  t X~ s  1 X~ s  1 vs X~ s  1 x;t
1
: 140
T
 tv 
T  op 1 under our assumptions by similar steps as in the proof of Theorem 3. By Lemma D.2,
4;1
" #
p XT  
T tv
4;2 T  T 1
4;T Y 4;t R4;T op 1; 141
h t1 i h i

       
where 4;T Y 4;t E 4;T y4 ; Y 4;t E 4;T Y 4;t ; y4 and y4 v; X~ ; 1 ; ; T . If we define  t to be 1 ; ; T excluding t
and similarly for  t , then
h  i
E 4;T y4 ; Y 4;t
242 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

h   h   0 0
i  
i
E Kb t  t wt  t vec t  t X~ X~ vX~ x t  1
Z Z h  i  

1 1       0 0
Kb  t wt  t vec  t X~ X~ vX~ x t  1 dt d  t
0 0
"Z Z #Z
1 1   1   h   0 0
i  

wt  t d  t Kb  t vec  t X~ X~ vX~ x t  1 dt
0 0 0
"Z Z #Z
1 1   h  1 h  0 0
i

wt  t d  t
1
K vec   h X~ X~ vX~ x  h  1 d
0 0  1h
"Z Z #
1 1   0
 r  r
wt  t d  t vec vX~ x  1 O b o b : 142
0 0

Similarly, it can be shown that


h  i  r  r
E 4;T Y 4;t ; y4 O b o b 143
p
and that T  R4;T op 1 under our assumptions, so that

X
Z Z
T
0
1 1  
w t vec vt X~ t  1 x;t op 1;
1
T tv
4;2 T
 1=2
wt wt  t d  t : 144
t1 0 0
R1 R1
w t is a function of 0 d and 0 d and from Eqs. (132) and (133) we have that
!0

X
T 1 0 X
T t
m
w t IK 2 ; 145
1 t 1 2 t 1 m0

with w T I K and
!0 !0 !0 !!
  X
t 1 t1 0 X
T 1 0
w
t t; t 2 1  IK 2
1 1 2 1 1 1 t 1 2 t 1
!0 !!
X
t1
t  1  1 X
T
1  t
 IK 146
1 1 1 t 1

with w
1 0. Thus, we have that

p ^ tv
T ols 
! !
p tv
X
T X
T
 
T ^  T 1 ~ F;t ~
 1=2 1
w t x;t  I K vec vt x~ 0t  1 op 1; 147
ols 1 T
t1 t1

and so using Theorem 3 the asymptotic variance is


!0 ! !0 !0
X
T X
T
tv
V

lim T 1
~ F;t  IK C V tv
lim T 1
~ F;t  IK C
T-1 T-1
t1 t1
!
X
T

~
lim T  1
1 ~ 0 Ctv Ctv0 ;
w t x;t  v;t w 0t 148
1 1
T-1
t1

where
" !0 !#"Z #1
X
T 1  
Ctv

lim T  1 ~ F;t  IK C f  B0 B d
T-1 0
t1
" #
X
T   0 ~0
 lim T 1
DF;x  Bt Bt Du v;t w t 1 :
0
149
T-1
t1
T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244 243

References Fama, E.F., French, K.R., 1992. The cross-section of expected stock returns.
Journal of Finance 47, 427465.
Fama, E.F., French, K.R., 1993. Common risk factors in the returns on
Abrahams, M.G., Adrian, T., Crump, R.K., Moench, E., 2014. Decomposing stocks and bonds. Journal of Financial Economics 33, 356.
real and nominal yield curves. Staff Report 570, Federal Reserve Bank Fama, E.F., French, K.R., 1997. Industry costs of equity. Journal of Financial
of New York, New York, NY. Economics 43, 153193.
Adrian, T., Crump, R.K., Moench, E., 2013. Pricing the term structure with Fama, E.F., MacBeth, J.D., 1973. Risk, return, and equilibrium: empirical
linear regressions. Journal of Financial Economics 110, 110138. tests. Journal of Political Economy 113, 607636.
Adrian, T., Moench, E., 2008. Pricing the term structure with linear Ferson, W.E., Harvey, C.R., 1991. The variation of economic risk premiums.
regressions. Staff Report 340, Federal Reserve Bank of New York, Journal of Political Economy 99, 385415.
New York, NY. Ferson, W.E., Harvey, C.R., 1999. Conditioning variables and the cross
Ahn, S.C., Gadarowski, C., 1999. Two-pass cross-sectional regression of section of stock returns. Journal of Finance 54, 13251360.
factor pricing models: minimum distance approach. Working paper. Gagliardini, P., Ossola, E., Scaillet, O., 2014. Time-varying risk premium in
Andrews, D.W.K., Lu, B., 2001. Consistent model and moment selection large cross-sectional equity datasets. Working paper.
procedures for GMM estimation with application to dynamic panel Ghysels, E., 1998. On stable factor structures in the pricing of risk: do
data models. Journal of Econometrics 101, 123164. time-varying betas help or hurt?. Journal of Finance 53, 549573.
Ang, A., Kristensen, D., 2009. Testing conditional factor models. Working Gibbons, M.R., 1982. Multivariate tests of financial models: a new
paper. approach. Journal of Financial Economics 10, 327.
Ang, A., Kristensen, D., 2012. Testing conditional factor models. Journal of Gibbons, M.R., Ferson, W., 1985. Testing asset pricing models with
Financial Economics 106, 132156. changing expectations and an unobservable market portfolio. Journal
Ang, A., Liu, J., Schwarz, K., 2010. Using stocks or portfolios in tests of of Financial Economics 14, 217236.
factor models. Working paper. Gomes, J., Kogan, L., Zhang, L., 2003. Equilibrium cross section of returns.
Ang, A., Piazzesi, M., 2003. A no-arbitrage vector autoregression of term Journal of Political Economy 111, 693732.
structure dynamics with macroeconomic and latent variables. Journal Gospodinov, N., Kan, R., Robotti, C., 2012. Misspecification-robust infer-
of Monetary Economics 50, 745787. ence in linear asset-pricing models with irrelevant risk factors.
Ang, A., Ulrich, M., 2012. Nominal bonds, real bonds, and equity. Working Review of Financial Studies 27, 20972138.
paper. Hansen, L.-P., 1982. Large sample properties of generalized method of
Baele, L., Bekaert, G., Inghelbrecht, K., 2010. The determinants of stock moments estimators. Econometrica 50, 10291054.
and bond return comovements. Review of Financial Studies 23, Harvey, C.R., 1989. Time-varying conditional covariances in tests of asset
23742428. pricing models. Journal of Financial Economics 24, 289317.
Balduzzi, P., Robotti, C., 2010. Asset pricing models and economic risk Harvey, C.R., 1991. The world price of covariance risk. Journal of Finance
premia: a decomposition. Journal of Empirical Finance 17, 5480. 46, 111157.
Bekaert, G., Engstrom, E., Grenadier, S., 2010. Stock and bond returns with Harvey, C.R., 2001. The specification of conditional expectations. Journal
moody investors. Journal of Empirical Finance 17, 867894. of Empirical Finance 8, 573637.
Braun, P.A., Nelson, D.B., Sunier, A.M., 1995. Good news, bad news, Jagannathan, R., Wang, Z., 1996. The conditional CAPM and the cross-
volatility, and betas. Journal of Finance 50, 15751603. section of expected returns. Journal of Finance 51, 353.
Burnside, A.C., 2011. The cross-section of foreign currency risk premia and Jagannathan, R., Wang, Z., 1998. An asymptotic theory for estimating beta-
consumption growth risk: comment. American Economic Review 101, pricing models using cross-sectional regression. Journal of Finance 53,
34563476. 12851309.
Burnside, C., 2010. Identification and inference in linear stochastic Joslin, S., Priebsch, M., Singleton, K.J., 2012. Risk premiums in dynamic
discount factor models. NBER Working paper 16634. National Bureau term structure models with unspanned macro risks. Journal of
of Economic Research, Inc. Finance 69, 11971233.
Campbell, J.Y., 1987. Stock returns and the term structure. Journal of Kan, R., Chen, R., 2005. Finite sample analysis of two-pass cross-sectional
Financial Economics 18, 373399. regressions. Working paper.
Campbell, J.Y., 1996. Understanding risk and return. Journal of Political Kan, R., Robotti, C., Shanken, J., 2013. Pricing model performance and the
Economy 104, 572621. two-pass cross-sectional regression methodology. Journal of Finance
Campbell, J.Y., Lo, A.W., MacKinlay, A.C., 1997. The Econometrics of 68, 26172649.
Financial Markets. Princeton University Press, Princeton, NJ. Kandel, S., 1984. The likelihood ratio test statistic of meanvariance
Campbell, J.Y., Shiller, R.J., 1988. The dividendprice ratio and expecta- efficiency without a riskless asset. Journal of Financial Economics 13,
tions of future dividends and discount factors. Review of Financial 575592.
Studies 1, 195228. Keim, D.B., Stambaugh, R.F., 1986. Predicting returns in the stock and
Campbell, J.Y., Shiller, R.J., 1991. Yield spreads and interest rate move- bond markets. Journal of Financial Economics 17, 357390.
ments: a bird's eye view. Review of Economic Studies 58, 495514. Kleibergen, F., 1998. Reduced rank regression using GMM. In: Matyas, L.
Campbell, J.Y., Sunderam, A., Viceira, L.M., 2013. Inflation bets or deflation (Ed.), Generalised Method of Moments Estimation, Cambridge Uni-
hedges? the changing risks of nominal bonds. Working paper 09-088. versity Press, Cambridge, UK, pp. 171209.
Harvard Business School, Cambridge, MA. Kleibergen, F., 2009. Tests of risk premia in linear factor models. Journal
Campbell, J.Y., Thompson, S.B., 2008. Predicting excess stock returns out of Econometrics 149, 149173.
of sample: can anything beat the historical average?. Review of Kleibergen, F., Zhan, Z., 2013. Unexplained factors and their effects on
Financial Studies 21, 15091531. second pass r-squared's and t-tests. Working paper. Brown University,
Chen, N.-F., Roll, R., Ross, S.A., 1986. Economic forces and the stock Providence, RI.
market. Journal of Business 59, 383403. Koijen, R.S.J., Lustig, H.N., van Nieuwerburgh, S., 2013. The cross-section
Chordia, T., Goyal, A., Shanken, J., 2013. Cross-sectional asset pricing with and time-series of stock and bond returns. Working paper.
individual stocks: Betas versus characteristics. Working paper. Kristensen, D., 2009. Uniform convergence rates of kernel estimators with
Cochrane, J., 2005. Asset Pricing, revised edition Princeton University heterogeneous dependent data. Econometric Theory 25, 14331445.
Press, Princeton, NJ. Kristensen, D., 2012. Non-parametric detection and estimation of struc-
Cochrane, J., 2011. Discount rates. Journal of Finance 66, 10471109. tural change. Econometrics Journal 15, 420461.
Cochrane, J., Piazzesi, M., 2008. Decomposing the yield curve. Working Lettau, M., Ludvigson, S., 2001. Resurrecting the (C)CAPM: a cross-
paper. sectional test when risk premia are time-varying. Journal of Political
Cochrane, J.H., 2008. The dog that did not bark: a defense of return Economy 109, 12381287.
predictability. Review of Financial Studies 21, 15331575. Lettau, M., Wachter, J., 2011. The term structures of equity and interest
Dai, Q., Singleton, K., 2002. Expectations puzzles, time-varying risk premia, and rates. Journal of Financial Economics 101, 90113.
affine models of the term structure. Journal of Financial Economics 63, Lewellen, J., Nagel, S., 2006. The conditional CAPM does not explain asset-
415441. pricing anomalies. Journal of Financial Economics 82, 289314.
David, A., Veronesi, P., 2013. What ties return volatilities to price Lewellen, J., Nagel, S., Shanken, J., 2010. A skeptical appraisal of asset
valuations and fundamentals? Journal of Political Economy 121, pricing tests. Journal of Financial Economics 96, 175194.
682746. Mamaysky, H., 2002. Market prices of risk and return predictability in a
Duffee, G.R., 2002. Term premia and interest rate forecasts in affine joint stock-bond pricing model. Working paper.
models. Journal of Finance 57, 405443. Nagel, S., Singleton, K.J., 2011. Estimation and evaluation of conditional
Fama, E.F., French, K.R., 1989. Business conditions and expected returns asset pricing models. Journal of Finance 66, 873909.
on stocks and bonds. Journal of Financial Economics 25, 2349.
244 T. Adrian et al. / Journal of Financial Economics 118 (2015) 211244

Petkova, R., 2006. Do the Fama-French factors proxy for innovations in Shanken, J., 1990. Intertemporal asset pricing: an empirical investigation.
predictive variables? Journal of Finance 61, 581612. Journal of Econometrics 45, 99120.
Petkova, R., Zhang, L., 2005. Is value riskier than growth? Journal of Shanken, J., 1992. On the estimation of beta-pricing models. Review of
Financial Economics 78, 187202. Financial Studies 5, 133.
Robinson, P., 1989. Nonparametric estimation of time-varying parameters. Shanken, J., Zhou, G., 2007. Estimating and testing beta pricing models:
In: Hackl, P. (Ed.), Statistical Analysis and Forecasting of Economic alternative methods and their performance in simulations. Journal of
Structural Change, Springer-Verlag, Berlin, Germany, pp. 253264. Financial Economics 84, 4086.
Roll, R., 1985. A note on the geometry of Shanken's CSR T2 test for mean/ Singleton, K., 2006. Empirical Dynamic Asset Pricing. Princeton University
variance efficiency. Journal of Financial Economics 14, 349357. Press, Princeton, NJ.
Roussanov, N., 2014. Composition of wealth, conditioning information, Wooldridge, J.M., White, H., 1988. Some invariance principles and central
and the cross-section of stock returns. Journal of Financial Economics limit theorems for dependent heterogeneous processes. Econometric
111, 352380. Theory 4, 210230.
Shanken, J., 1985. Multivariate tests of the zero-beta CAPM. Journal of Zhou, G., 1994. Analytical GMM tests: asset pricing with time-varying risk
Financial Economics 14, 327348. premiums. Review of Financial Studies 7, 687709.
Shanken, J., 1986. Testing portfolio efficiency when the zero-beta rate is
unknown: a note. Journal of Finance 41, 269276.

Vous aimerez peut-être aussi