1. Introduction
The vector autoregressive model [VAR] and the dynamic factor model [DFM] are arguably among the most popular tools for analyzing economic and financial variables over time. Since the seminal contribution of
Sims (
1980), VARs have been theoretically extended and practically implemented to forecast, structurally analyze, and detect comovements in multivariate time series. DFMs were introduced more recently (
Forni et al., 2000;
Forni & Lippi, 2001;
Stock & Watson, 2002a,
2002b;
Bai & Ng, 2002;
Bai, 2003), but they rapidly contested the role of the workhorse in empirical macroeconomics.
The main reason for the success of the DFM is two-fold. First, it allows handling a much larger number of variables than those that are generally employed in traditional small-scale VARs, thus potentially boosting forecasting accuracy and solving the informational deficiency problems that arise in structural analyses when the agent’s information set is richer than the econometrician’s information set (see, e.g.,
Forni & Gambetti, 2014). Second, DFM allows for disentangling the shocks that drive the common component of high-dimensional time series and recovering the structural shocks from these common shocks only. Hence, in structural DFMs, the number of shocks is smaller than the number of variables (
Forni et al., 2009), which is in line with dynamic, stochastic general equilibrium models (see
Fernández-Villaverde et al., 2016 and the references therein) and, more generally, with the standard macroeconomic view that a low number of shocks drives aggregate fluctuations.
Efforts have recently been made to endow the VAR with the above-mentioned features of the DFM. On the one hand, shrinkage estimators have been proposed for medium–large VARs, both from a Bayesian perspective (e.g.,
Bańbura et al., 2010;
Koop, 2013;
Carriero et al., 2015) and from a classical standpoint (e.g.,
Hsu et al., 2008;
Kock & Callot, 2015;
Hecq et al., 2023). On the other hand, the multivariate autoregressive index (MAI) model—originally proposed by
Reinsel (
1983) as a convenient approach to dimension reduction in stationary VARs—has recently gained renewed attention.
1Late advances have shown that MAI and its variants allow for both forecasting variables and identifying shocks analogously to the DFM but without encountering some issues in model identification and statistical inference that characterize the latter, such as the requirement that the number of variables diverges at a given rate and the need for specific assumptions on both the correlation structure of the idiosyncratic components and the factor loadings (see, i.e.,
Bai, 2003;
Bai & Ng, 2006). Moreover, VARs with index structures have been shown to be able to accommodate features such as stochastic volatility (
Carriero et al., 2022) and time-varying parameters (
Cubadda et al., 2025), which are not easy to handle within the DFM framework. The MAI falls within reduced-rank VARs, a general class of models that include, as special cases, both the cointegrated VAR (see
Johansen, 1995 and the references therein) and the common serial correlation [CSC] models (see
Cubadda & Hecq, 2022b and the references therein). Although CSC models and MAI have similar mathematical formulations, their respective goals and properties are rather different; whereas the former are based on the existence of (possibly dynamic) linear combinations of autocorrelated time series that are white noise, VARs with an index structure assume that there is a limited number of channels through which information from the past is transmitted to the variables of interest. The main aim of this paper is twofold. First, recent developments in MAI are reviewed, such as the structural MAI (
Carriero et al., 2016), the vector heterogeneous index model for realized volatilities (
Cubadda et al., 2017), and augmentations of the original model with individual autoregressive structures (
Cubadda & Guardabascio, 2019), stochastic volatility (
Carriero et al., 2022), time-varying parameters (
Cubadda et al., 2025), high-dimensionality (
Cubadda & Hecq, 2022a), and cointegration (
Cubadda & Mazzali, 2024). Second, new results are provided in terms of representation theory for various models, and a novel modeling is proposed, namely the cointegrated index-augmented autoregressive model, which combines and extends the results in
Cubadda and Guardabascio (
2019) and
Cubadda and Mazzali (
2024).
This paper is organized as follows. Focusing on representation theory,
Section 2 reviews previous contributions and provides new insights into some of them.
Section 3 presents the new model and deals with its estimation, whereas some details of the estimation procedure are relegated to
Appendix A. Finally,
Section 4 provides some conclusions.
2. VAR Models with an Index Structure
In this section, we review models that are rooted in the original MAI formulation and provide new results with respect to representation theory for some of them. Analogies and differences with the DFM are discussed in detail. Estimation and identification issues are also covered.
2.1. The Structural Multivariate Autoregressive Index Model
Let us assume that the
n-vector time series
is generated by the following stationary VAR
model:
where
L is the lag operator,
;
is a vector or
n errors with
(positive definite) and finite fourth moments,
; and
is the natural filtration of the process
. For simplicity, deterministic elements are ignored.
Assumption 1. The following holds:where ω is a full-rank -matrix with , and is a -matrix for . Under Assumption 1, Model (
1) can be rewritten as
where linear combinations of
are called indexes. The MAI has at most
mean parameters, which implies a significant dimension reduction when
p is small with respect to
n.
2By premultiplying both sides of Equation (
2) with
, we get
which shows that the indexes follow a VAR
process and not a VARMA process, as is generally the case for linear combinations of elements of a VAR (see
Cubadda et al., 2009 and the references therein).
Remark 1. In view of Equations (2) and (3), the MAI resembles the exact DFM [EDFM] (see Lippi, 2019 and the references therein), but there are also some relevant differences. First, in the EDFM series, loads the factors contemporaneously and not only with lags. Second, the factors and the idiosyncratic terms in the EDFM are uncorrelated at any lag lead, whereas in the MAI, we have only for . Third, the contemporaneous variance matrix of the idiosyncratic terms in the EDFM is diagonal, whereas Σ is generally not diagonal. Placing emphasis on the analogies between MAI and EDFM,
Carriero et al. (
2016) propose identifying structural shocks as linear transformations of the index shocks only. Starting from the Wold representation of series
and inserting the decomposition of the identity matrix between
and
, as in
Centoni and Cubadda (
2003),
one obtains the following decomposition of series
where
,
,
,
, and
for
.
Since the shock is one of the indexes, may be interpreted as the common shock and as the common components of the series . Similarly, and can be labeled, respectively, as uncommon shocks and an uncommon component.
Interestingly, post-multiplying with
both sides of relation
we obtain
, which in turn implies that the Wold polynomial matrix of the MAI has the form
where
is an
-matrix for
.
Having substituted
in Equations (
6) and (
7) with the RHS of Equation (
8), we can finally prove the following proposition.
Proposition 1. In the MAI, the components of in (5) read as follows:where the uncommon component denotes n-dimensional white noise such that . Corollary 1. The indexes and the common component are linked through the relation , which trivially follows from Proposition 1.
Remark 2. In view of Proposition 1, the decomposition (5) has clear analogies with the analogous decomposition in the EDFM. However, differently from the idiosyncratic terms in the EDFM, the uncommon component is obviously cross-sectionally dependent.3 Carriero et al. (
2016) suggest recovering the structural shocks as linear transformations of the common shock
only. Hence, for most
, structural shocks can be recovered, as observed in DFMs and in dynamic stochastic general equilibrium models. In principle, all identification strategies that are available for structural VARs or structural DFMs (see
Stock & Watson, 2016 and the references therein) can be adopted.
On the estimation side,
Carriero et al. (
2016) prove that the iterative maximum likelihood procedure proposed by
Reinsel (
1983) is consistent when
. Moreover, they provide an MCMC algorithm for Bayesian estimation and show, by simulations, that the Bayesian approach outperforms the classical one when
. Finally, they document the practical value of the structural MAI by two empirical applications: the transmission mechanism of monetary policies and the propagation of demand and supply shocks.
2.2. The Vector Heterogeneous Autoregressive Index Model
The univariate heterogeneous AR model [HAR], originally proposed by
Corsi (
2009) is a popular tool for analyzing and forecasting daily realized volatility [RV] measures without resorting to more involved long-memory models. Technically speaking, the HAR is a constrained AR
model where the predictors are the first lags of the following: (i) the daily RV; (ii) the weekly (5 days) average of the daily RV; (iii) the monthly (22 days) average of the daily RV.
Cubadda et al. (
2017) propose a multivariate HAR for a set of
n daily RV measures,
, that is endowed with an index structure. In particular, the vector heterogeneous autoregressive index model [VHARI] reads as follows:
where
,
, and
denote, respectively, the time horizons of one day, one week, and one month such that
The VHARI enjoys two important properties that are not shared by alternative approaches, and they induce dimensional reductions in the vector HAR
4: First, the index
preserves the temporal cascade structure of the HAR model since
Second, pre-multiplying both sides of the VHARI by
yields the following:
which shows that the indexes follow a multivariate HAR model. In particular, when
, a univariate HAR model generated all the dynamics of the
n RVs.
On the estimation side,
Cubadda et al. (
2017) suggest using a switching algorithm [SA], an iterative method for the numerical maximization of the log-likelihood of complex models that has a long tradition in time series analysis (see
Boswijk & Doornik, 2004 and the references therein). In particular, the proposed SA requires the following steps:
Given an (initial) estimate of , maximize the conditional Gaussian likelihood , where .
Given the previously obtained estimates of A and , maximize the conditional likelihood .
Repeat steps 1 and 2 until numerical convergence occurs.
5
A key point of the above SA is that both steps 1 and 2 require running OLS regressions only. This feature provides the SA with several advantages over Newton-type optimization methods: computational simplicity, with no need for normalization conditions in
; explicit optimization at each step; and the ease of application of regularization schemes or linear restrictions on parameters (see
Cubadda and Guardabascio (
2019) for additional discussions).
Furthermore, when the SA is initialized with consistent estimates and is iterated sufficiently, the resulting estimator is asymptotically equivalent to the ML one (
Hautsch et al., 2023).
Cubadda et al. (
2017) show by simulation that the suggested SA performs well even when elements of
have a log-normal error distribution with GARCH variances.
Following
Patton and Sheppard (
2009),
Cubadda et al. (
2017) use a VHARI to build the optimal linear combination of ten different estimators of the volatility of the same market to evaluate its merits through an out-of-sample forecasting exercise. The VHARI model performs well, often outperforming previously existing methods.
2.3. The Index-Augmented Autoregressive Model
A possible limitation of MAI as a forecasting tool is that the only predictors of the series
for
are the lagged indexes, whereas the forecasts obtained through the DFM exploit information coming from the past of both factors and the series
itself (see the seminal contributions by
Stock and Watson (
2002a,
2002b)). Although the indexes may be interpreted as ’supervised’ factors that are constructed for emphasizing the comovements between the present and the past of the system, some variables are better predicted by their own lags rather than by any linear combination of all variables only.
In order to overcome such limitations,
Cubadda and Guardabascio (
2019) extended the basic MAI model by allowing individual AR structures for each element of
. Their key assumption is the following.
Assumption 2. It holdswhere is the generic element of the polynomial matrix , is the generic element of ω, and is the generic element of for , , . In other words, Assumption 2 states that there is a reduced number of channels p through which each variable is influenced by the past of other variables in the system, which is consistent with the common view that few shocks are responsible for most macroeconomic fluctuations.
Under Assumption 2 and using the reparametrization
, Model (
1) can be rewritten into the following index-augmented autoregressive model [IAAR]:
where
is a
diagonal matrix with
as a generic diagonal element, and for greater generality,
.
Remark 3. Since the number of parameters of Model (9) is equal to , it is necessary to impose proper upper bounds to either q or s to ensure that the MIAAR is more parsimonious than the VAR. To this end, it is easy to see that sufficient conditions are for or for any p and . However, in empirical applications, the estimated values of q are typically much smaller than n (see Cubadda & Guardabascio, 2019; Carriero et al., 2022). Remark 4. The individual forecasting equation of the IAAR readswhere is the i-th row of matrix . Equation (10) is entirely analogous to the individual forecasting equation of the DFM, with one important difference. Whereas factors are typically estimated using principal component methods, which aim to maximize the contemporaneous variability of series , the indexes in (10) are constructed explicitly by taking into account the covariability between each series and the lags of other elements of conditionally on the lags of the series . Remark 5. Interestingly, by the same argument underlying Proposition 1, we see that, differently from the MAI, , which implies, in view of Equation (7), that the uncommon component is generally autocorrelated in the case of the IAAR. Hence, the decomposition (5) for the IAAR closely resembles the analogous decomposition in the approximate DFM (see Lippi (2019) and the references therein). However, the estimation of the index does not require that , nor does it impose conditions on the autocorrelations and cross-correlations of the elements of or on the loading as in the approximate DFM. Cubadda and Guardabascio (
2019) proposed a two-step SA for the estimation of the IAAR, along with a variant where a
regularization scheme is applied in both steps. They show, by simulations, that the regularized version of the SA outperforms the standard one with
. Regarding model specification, they opt for the use of information criteria [IC], in line with previous contributions showing that IC outperforms likelihood ratio tests in the selection of reduced-rank VAR models (see, e.g.,
Gonzalo & Pitarakis, 1999;
Cavaliere et al., 2015;
Cavaliere et al., 2018). Finally, the IAAR proves to outperform well-known macroeconomic forecasting methods when applied to systems with
n ranging from 4 to 40.
Carriero et al. (
2022) endowed the IAAR with Stochastic Volatility [IAAR-SV] in the error
and offered Bayesian estimations using Markov Chain Monte Carlo [MCMC] techniques. Furthermore, they used (
4) to decompose the time-varying volatility
as follows:
Carriero et al. (
2022) applied the IAAR-SV to analyze the commonality in both levels and volatilities of inflation rates in several countries, and their main finding is that a substantial fraction of inflation volatility can be attributed to a global factor that also drives inflation levels and their persistence.
2.4. The Time-Varying Multivariate Autoregressive Index Model
A further step towards taking parameter instabilities over time into account was made by
Cubadda et al. (
2025), who proposed the following MAI with time-varying parameters and time-varying volatility [MAI-TVP-TVV]:
where
,
,
;
and
are independent at any lag and lead. Notice that it is assumed that the index loadings evolve over time as random walks, while the index weight
remain stable.
In order to overcome the computational limitation related to MCMC procedures,
Cubadda et al. (
2025) offer a hybrid estimation method that combines the SA, Kalman filter with forgetting factors (
Koop & Korobilis, 2014), and exponentially weighted moving average techniques (
Johansson et al., 2023) for the time-varying volatility.
An empirical application, where 25 US quarterly time series are used to forecast three key macroeconomic variables, shows that the MAI-TVP-TVV is one of the best models in a large set of competitors for all targets, improving upon its counterparts, especially for short horizons. Other interesting findings are that once the MAI is endowed with time-varying volatility [MAI-TVV], there are no clear improvements in adding time-varying parameters for point forecasting, but the MAI-TVP-TVV always outperforms the MAI-TVV in density forecasting.
2.5. The Dimension-Reducible VAR
Cubadda and Hecq (
2022a) studied the conditions under which the dynamics in a large-dimension VAR are entirely generated by a small-scale VAR. They show that such conditions are met when the coefficient matrices of the large VAR have the same common right space and a common left null space. This entails combing Assumption 1 with the following.
Assumption 3 is popularly known in time series econometrics as the CSC (see
Cubadda & Hecq, 2022b and the reference therein) given that
That is, there exist
linear combinations of variables
that are white noise, and as such, cyclical behavior cannot be exhibited.
Taking Assumptions 1 and 3 together leads to the dimension-reducible VAR model [DRVAR]:
where
is a
matrix for
.
Assuming, without loss of generality, that
and
, we can decompose series
as follows:
where
is the dynamic component, and
is the static one. Premultiplying both sides of DRVAR by
one obtains
where
, which shows that
is generated by a
q-dimensional VAR (
p) process.
By inserting the Wold representation of the dynamic component
in Equation (
12), it follows that
where
. Finally, by linearly projecting
on
, we obtain
with
, which can be inserted into Equation (
13) to obtain
where
and
for
.
Representation (
14) highlights that system dynamics are completely generated by common reduced form errors
. Consequently,
Cubadda and Hecq (
2022a) label
as the ignorable errors, as they are noise without structural interpretation. Since errors
and
are uncorrelated at any lead and lag, it is then possible to recover the structural shocks solely from the reduced form errors
of the common component
using any of the procedures that are commonly employed in structural VARs or structural DFMs (see
Stock and Watson (
2016) and the references therein).
In order to estimate the matrix
, one may rely on a nonparametric estimator proposed by
Lam et al. (
2011). The underlying intuition is that the matrix
lies in the space generated by the eigenvectors associated with the
q nonzero eigenvalues of the symmetric and semipositive definite matrix:
where
, and
is the autocovariance matrix of series
in lag
j. Under some regularity conditions, the matrix formed by the eigenvectors associated with the
q largest eigenvalues of the sample estimate of
M is a
-consistent estimator of
(up to an orthonormal transformation) when
q is fixed:
, and
for
, where
. Remarkably, the speed of convergence of the estimator, namely
, is the same as when the dimension
n is finite.
Moreover,
Cubadda and Hecq (
2022a) provide both the OLS and GLS estimators of the coefficient
in Equation (
11) and consistent information criteria for the selection of
q, and they show by simulations that the proposed methodology works well with the temporal and cross-sectional sizes that are typical in macroeconomics. Finally, the approach is applied to analyze a large set of US economic time series and to identify the shock that is responsible for most of the common volatility in the business cycle frequency band.
2.6. The Vector Error-Correction Index Model
The models considered so far do not explicitly deal with the possible presence of unit roots. Given that most macroeconomic and financial time series are characterized by stochastic trends, it is important to understand how a cointegrated VAR model can be augmented with an index structure.
Let us assume that series
follows the vector error-correction model [VECM]
where
and
are full-rank
(
) matrices such that
,
for
,
is non-singular, and
. Under such assumptions, it is well known that the elements of
are individually at most
and that they are jointly cointegrated with respect to an order 1 in the sense that
is
(see
Johansen (
1995) and the references therein).
To possibly reduce the number of parameters in the VECM,
Cubadda and Mazzali (
2024) made the following assumptions:
Assumption 4. For , the following holds:where ω is a full-rank matrix with , and A is a full-rank matrix. Assumption 5. The following holds:where γ is a full-rank matrix with . Under Assumptions 4 and 5, Model (
15) can be rewritten in the following vector error-correction index model [VECIM]:
where
is a full-rank
matrix (
), and
is an
matrix for
such that
. Notice that the cointegration matrix is given by
.
Interestingly, the indexes
themselves are generated by a
q-dimensional VECM:
where
, for
.
By first inserting the decomposition (
5) between
and
into the Wold representation of the first differences
:
and then further decomposing the common component
into permanent and transitory subcomponents as in
Centoni and Cubadda (
2003), we obtain the following:
where
Since errors
are the innovations of the common trends in the indexes
(see, e.g.,
Johansen, 1995) and errors
are such that
,
Cubadda and Mazzali (
2024) labeled
as the common permanent component and
as the common transitory component, whereas
is the uncommon component given that
and
.
Following a similar reasoning as the one leading to Proposition 1, post-multiplying, with
both sides of the relation
we again obtain
, which in turn implies that the Wold polynomial matrix of the VECIM has the same form as (
8). Finally, inserting (
8) in Equations (
17)–(
19), we can prove the following proposition:
Proposition 2. In the VECIM, the first differences of the components of in (16) readwhere the uncommon component is an n-dimensional random walk such that . Notice that Proposition 2 implies that Corollary 1 applies to the VECIM as well.
Remark 6. Given that the components in (16) are not correlated with each other at any lag and lead, the VECIM allows one to perform a structural analysis, taking advantage of the features of both the DFM—namely isolating shocks that are common among variables—and the VECM—namely disentangling shocks having transitory or permanent effects. For instance, one may identify the structural transitory shocks as and the impulse response functions as , where D is the matrix formed by the first r rows of , and C is a lower triangular matrix such that Since the first r rows of , being equal to C, form a lower triangular matrix, the usual interpretation of structural shocks obtained through a Cholesky factorization applies to .
Cubadda and Mazzali (
2024) offered a three-step SA for the estimation of the VECIM and proposed selecting the triple
in a unique search by IC. An extensive Monte Carlo study shows that the proposed methodology works reasonably well for
n, ranging from 6 to 18 when the model is identified by the Hannan–Quinn IC. Moreover, in an empirical application, they identified a shock that maximizes the variability of the common transitory component of unemployment at business cycle frequencies and another one that does the same, but for the common permanent component of unemployment. These two shocks are endowed with a neater economic interpretation than compared to a unique main business cycle shock identified according to
Angeletos et al. (
2020).
3. A New Proposal: The Cointegrated Index-Augmented Autoregressive Model
A possible limitation of the VECIM is that the uncommon component
is necessarily a random walk, which may be considered restrictive for some applications. For example,
Barigozzi et al. (
2021) proposed a DFM where the idiosyncratic components may be I(0) or I(1).
In order to overcome this issue, one can combine the VECIM with the IAAR. Formally, this involves using Assumption 5 along with the following one:
Assumption 6. For the VECM (15), the following holds:where is the generic element of the polynomial matrix for , , . Taking Assumptions 5 and 6, the model (
15) can be rewritten into the following cointegrated index-augmented auto-regressive model [CIAAR]:
where
is an
diagonal matrix with
as a generic diagonal element.
When the elements of series
are I(1), Model (
20) includes several earlier models for series
as special cases, as summarized in
Table 1.
Remark 7. Interestingly, by the same argument underlying Proposition 2, we see that, differently from the VECIM, , which implies, in view of Equation (19), that the first differences of the uncommon component are generally autocorrelated in the case of the CIAAR. The uncommon component is still stochastically singular with rank . Since system (20) has overall unit roots, while the common component has unit roots, uncommon component has unit roots (see Deistler & Wagner, 2017; Barigozzi et al., 2020 on the properties of singular I(1) stochastic processes). When , step 3 is clearly not needed, and steps 1 and 2 must be modified as follows:
- 1.1
Given (initial) estimates of
and
D, maximize
by estimating
A and
with OLS on the following model:
- 2.1
Given the previously obtained estimates of
A and
, maximize
by estimating
and
with OLS on the following model:
Finally, when , we can assume, without loss of generality, that . Then, step 3 is, again, not needed, whereas steps 1 and 2 must be modified as follows:
- 1.3
Given (initial) estimates of
and
D, maximize
by estimating
and
with OLS in the following model:
- 2.3
Given the previously obtained estimates of
and
, maximize
by estimating
and
with OLS in the following model:
The choice of initial values for the above procedures is discussed in
Appendix A, whereas the selection of the quadruple
can be carried out by IC sequentially or in a unique search, as suggested by
Cubadda and Mazzali (
2024).
4. Conclusions and Future Research Directions
The DFM and VAR are, arguably, among the most popular tools in macroeconometrics and financial econometrics. The two approaches should be considered complementary rather than substitutive, since each has its own merits. The MAI represents a link between these two methodologies: On the one hand, it is a VAR with a specific reduced-rank structure that alleviates the dimensionality problem; on the other hand, the MAI and its variants have several analogies with the DFM; in particular, they allow for identifying a small number of common reduced-form errors and for recovering structural shocks from those errors only.
However, the MAI is not affected by some theoretical limitations of the DFM, such as the requirement that the cross-sectional dimension diverges to infinity and the need for specific assumptions on the dynamic correlation structure of the idiosyncratic component and on the factor loadings. In a more practical perspective, VARs with an index structure can also handle features such as stochastic volatility (
Carriero et al., 2022) and time-varying parameters (
Cubadda et al., 2025), which are not easily accommodated in DFMs.
Recent developments in VAR models with index structures have considerably extended the original MAI formulation, endowing the model with individual autoregressive structures, stochastic volatility, time-varying parameters, high dimensionality, and cointegration. These extensions have proven to be useful tools for detecting common components, obtaining efficiency gains through the imposition of parameter restrictions, performing structural analysis, and boosting forecast accuracy.
Having reviewed most of the recent advances on the MAI and provided new insights on the representation theory underlying the IAAR and the VECIM, a new model, namely the CIAAR, was proposed along with an estimation procedure. The CIAAR extends previous contributions by allowing the VECIM for individual AR structures and the IAAR for cointegration.
There is plenty of room for future research that could be developed in at least three directions. First, the practical relevance of the CIAAR must be investigated both empirically and by simulations. Second, sparsity could be introduced in the MAI and its variants, employing regularized
regressions in the SA in place of OLS. This would open up the possibility of tackling both dimension reduction, through the index structure, and sparsity in the model coefficients, through Lasso and its variants. Third, the approaches considered in this survey could be applied to data with more elaborate dependence structures than vector time series, such as spatiotemporal processes or matrix–tensor time series. First contributions along these lines were provided by by
Pu et al. (
2025),
Wang et al. (
2022), and
Hecq et al. (
2024).
Funding
The financial support of MUR under the 20223725WE (PRIN 2022) grant is gratefully acknowledged.
Data Availability Statement
No new data were created or analyzed in this study.
Acknowledgments
Previous versions of this paper were presented at an intermediate workshop on methodological and computational issues in large-scale time series models for economics and finance in Messina, the Villa Mondragone time series symposium in honour of Marco Lippi in Monte Porzio Catone (Rome); the 11th ICEEE in Palermo; and the final workshop on methodological and computational issues in large-scale time series models for economics and finance in Monte Porzio Catone (Rome). The author thanks the participants, as well as three anonymous referees, for their helpful comments and suggestions. The usual disclaimers apply.
Conflicts of Interest
The author declares no conflicts of interest.
Appendix A
The choice of the initial values for an SA is important. Not only is an accurate initialization necessary to boost numerical convergence but the SA is also asymptotically equivalent to the ML one when the parameters to be initialized are consistently estimated (
Hautsch et al., 2023).
With reference to the SA in
Section 3, the initial values for
,
, and
D can be obtained as follows:
Use the usual Johansen procedure on the model (
15) and obtain estimates
,
, and
for
, where
.
Construct matrices for .
Construct the matrix .
Compute the singular-value decomposition , where the singular values are not increasingly ordered, and obtain as the matrix formed by the first q columns of V.
Compute the q-rank approximation of as , where is obtained from by setting the smallest singular values to 0.
Construct as the matrix formed by the first rows of .
Construct as a diagonal matrix with the diagonal equal to for .
The motivation for the above choices is twofold. First, the asymptotic distribution of the Johansen estimator of
is not affected by restrictions on the short-run parameters (
Johansen, 1995), which implies that
,
, and
are consistent, although inefficient, estimators of the associated parameters. Second, the right-singular vectors that correspond to the
q largest singular values of the matrix
consistently estimate
(see, e.g.,
Reinsel et al., 2022). By the same argument,
provides a consistent estimator of
. Finally, the consistency of
trivially follows from the ones of
and
.
Notes
| 1 | At the end of 2024, the annual citation rate of Reinsel ( 1983) in Scopus has increased by about 54% in the last 9 years, with the majority of recent citations coming from econometric journals. |
| 2 | Indeed, the matrix , once identified through normalizing restrictions, has free parameters. |
| 3 | Remarkably, when the factors in the EDFM are estimated by some principal components of series , the sample variance matrix of the estimated idiosyncratic component has a reduced rank as well. |
| 4 | The most obvious alternatives to the VHARI are likely multivariate principal component regression and reduced-rank regression. |
| 5 | |
References
- Angeletos, G. M., Collard, F., & Dellas, H. (2020). Business-cycle anatomy. American Economic Review, 110(10), 3030–3070. [Google Scholar] [CrossRef]
- Bai, J. (2003). Inferential theory for factor models of large dimensions. Econometrica, 71(1), 135–171. [Google Scholar] [CrossRef]
- Bai, J., & Ng, S. (2002). Determining the number of factors in approximate factor models. Econometrica, 70(1), 191–221. [Google Scholar] [CrossRef]
- Bai, J., & Ng, S. (2006). Confidence intervals for diffusion index forecasts and inference for factor-augmented regressions. Econometrica, 74(4), 1133–1150. [Google Scholar] [CrossRef]
- Bańbura, M., Giannone, D., & Reichlin, L. (2010). Large Bayesian vector auto regressions. Journal of Applied Econometrics, 25(1), 71–92. [Google Scholar] [CrossRef]
- Barigozzi, M., Lippi, M., & Luciani, M. (2020). Cointegration and error correction mechanisms for singular stochastic vectors. Econometrics, 8(1), 3. [Google Scholar] [CrossRef]
- Barigozzi, M., Lippi, M., & Luciani, M. (2021). Large-dimensional dynamic factor models: Estimation of impulse–response functions with I(1) cointegrated factors. Journal of Econometrics, 221(2), 455–482. [Google Scholar] [CrossRef]
- Boswijk, H., & Doornik, J. (2004). Identifying, estimating and testing restricted cointegrated systems: An overview. Statistica Neerlandica, 58(4), 440–465. [Google Scholar] [CrossRef]
- Carriero, A., Clark, T. E., & Marcellino, M. (2015). Bayesian VARs: Specification choices and forecast accuracy. Journal of Applied Econometrics, 30(1), 46–73. [Google Scholar] [CrossRef]
- Carriero, A., Corsello, F., & Marcellino, M. (2022). The global component of inflation volatility. Journal of Applied Econometrics, 37(4), 700–721. [Google Scholar] [CrossRef]
- Carriero, A., Kapetanios, G., & Marcellino, M. (2016). Structural analysis with multivariate autoregressive index models. Journal of Econometrics, 192(2), 332–348. [Google Scholar] [CrossRef]
- Cavaliere, G., De Angelis, L., Rahbek, A., & Taylor, A. (2015). A comparison of sequential and information-based methods for determining the co-integration rank in heteroskedastic VAR models. Oxford Bulletin of Economics and Statistics, 77(1), 106–128. [Google Scholar] [CrossRef]
- Cavaliere, G., De Angelis, L., Rahbek, A., & Taylor, A. (2018). Determining the cointegration rank in heteroskedastic VAR models of unknown order. Econometric Theory, 34(2), 349–382. [Google Scholar] [CrossRef]
- Centoni, M., & Cubadda, G. (2003). Measuring the business cycle effects of permanent and transitory shocks in cointegrated time series. Economics Letters, 80(1), 45–51. [Google Scholar] [CrossRef]
- Corsi, F. (2009). A simple approximate long-memory model of realized volatility. Journal of Financial Econometrics, 7(2), 174–196. [Google Scholar] [CrossRef]
- Cubadda, G., & Guardabascio, B. (2019). Representation, estimation and forecasting of the multivariate index-augmented autoregressive model. International Journal of Forecasting, 35(1), 67–79. [Google Scholar] [CrossRef]
- Cubadda, G., Guardabascio, B., & Grassi, S. (2025). The time-varying multivariate autoregressive index model. International Journal of Forecasting, 41(1), 175–190. [Google Scholar] [CrossRef]
- Cubadda, G., Guardabascio, B., & Hecq, A. (2017). A vector heterogeneous autoregressive index model for realized volatility measures. International Journal of Forecasting, 33(2), 337–344. [Google Scholar] [CrossRef]
- Cubadda, G., & Hecq, A. (2022a). Dimension reduction for high dimensional vector autoregressive models. Oxford Bulletin of Economics and Statistics, 84(5), 1123–1152. [Google Scholar] [CrossRef]
- Cubadda, G., & Hecq, A. (2022b). Reduced rank regression models in economics and finance. In Oxford handbook of economic forecasting. Oxford University Press. [Google Scholar]
- Cubadda, G., Hecq, A., & Palm, F. (2009). Studying co-movements in large multivariate data prior to multivariate modelling. Journal of Econometrics, 148(1), 25–35. [Google Scholar] [CrossRef]
- Cubadda, G., & Mazzali, M. (2024). The vector error correction index model: Representation, estimation and identification. Econometrics Journal, 27, 126–150. [Google Scholar] [CrossRef]
- Deistler, M., & Wagner, M. (2017). Cointegration in singular ARMA models. Economics Letters, 155(4), 39–42. [Google Scholar] [CrossRef]
- Fernández-Villaverde, J., Rubio-Ramírez, J., & Schorfheide, F. (2016). Solution and estimation methods for DSGE models. In J. Taylor, & H. Uhlig (Eds.), Handbook of macroeconomics (Vol. 2, pp. 527–724). North Holland. [Google Scholar]
- Forni, M., & Gambetti, L. (2014). Sufficient information in structural VARs. Journal of Monetary Economics, 66(1), 124–136. [Google Scholar] [CrossRef]
- Forni, M., Hallin, M., Lippi, M., & Reichlin, L. (2000). The generalized dynamic-factor model: Identification and estimation. Review of Economics and Statistics, 82(4), 540–554. [Google Scholar] [CrossRef]
- Forni, M., Hallin, M., Lippi, M., & Reichlin, L. (2009). Opening the black box: Structural factor models with large cross sections. Econometric Theory, 25(5), 1319–1347. [Google Scholar] [CrossRef]
- Forni, M., & Lippi, M. (2001). The generalized dynamic factor model: Representation theory. Econometric Theory, 17(6), 1113–1141. [Google Scholar] [CrossRef]
- Gonzalo, J., & Pitarakis, J. (1999). Dimensionality effect in cointegration analysis. In Cointegration, causality, and forecasting. A festschrift in honour of Clive WJ Granger (pp. 212–229). Oxford University Press. [Google Scholar] [CrossRef]
- Hautsch, N., Okhrin, O., & Ristig, A. (2023). Maximum-likelihood estimation using the zig-zag algorithm. Journal of Financial Econometrics, 21(4), 1346–1375. [Google Scholar] [CrossRef]
- Hecq, A., Margaritella, L., & Smeekes, S. (2023). Granger causality testing in high-dimensional VARs: A post-double-selection procedure. Journal of Financial Econometrics, 21(3), 915–958. [Google Scholar] [CrossRef]
- Hecq, A., Ricardo, I., & Wilms, I. (2024). Reduced-rank matrix autoregressive models: A medium n approach. Available online: https://arxiv.org/abs/2407.07973 (accessed on 20 January 2025).
- Hsu, N. J., Hung, H. L., & Chang, Y. M. (2008). Subset selection for vector autoregressive processes using Lasso. Computational Statistics & Data Analysis, 52(7), 3645–3657. [Google Scholar] [CrossRef]
- Johansen, S. (1995). Likelihood-based inference in cointegrated vector autoregressive models. Oxford University Press. [Google Scholar]
- Johansson, K., Ogut, M. G., Pelger, M., Schmelzer, T., & Boyd, S. (2023). A simple method for predicting covariance matrices of financial returns. Foundations and Trends in Econometrics, 12(4), 324–407. [Google Scholar] [CrossRef]
- Kock, A. B., & Callot, L. (2015). Oracle inequalities for high dimensional vector autoregressions. Journal of Econometrics, 186(2), 325–344. [Google Scholar] [CrossRef]
- Koop, G. M. (2013). Forecasting with medium and large Bayesian VARs. Journal of Applied Econometrics, 28(2), 177–203. [Google Scholar] [CrossRef]
- Koop, G. M., & Korobilis, D. (2014). A new index of financial conditions. European Economic Review, 71, 101–116. [Google Scholar] [CrossRef]
- Lam, C., Yao, Q., & Bathia, N. (2011). Estimation of latent factors for high-dimensional time series. Biometrika, 98(4), 901–918. [Google Scholar] [CrossRef]
- Lippi, M. (2019). Time-domain approach in high-dimensional dynamic factor models. In Oxford research encyclopedia of economics and finance. Oxford University Presss. [Google Scholar]
- Oberhofer, W., & Kmenta, J. (1974). A general procedure for obtaining maximum likelihood estimates in generalized regression models. Econometrica, 42(3), 579–590. [Google Scholar] [CrossRef]
- Patton, A., & Sheppard, K. (2009). Optimal combinations of realised volatility estimators. International Journal of Forecasting, 25(2), 218–238. [Google Scholar] [CrossRef]
- Pu, D., Fang, K., Lan, W., Yu, J., & Zhang, Q. (2025). Reduced rank spatio-temporal models. Journal of Business & Economic Statistics, 43(1), 98–109. [Google Scholar]
- Reinsel, G. (1983). Some results on multivariate autoregressive index models. Biometrika, 70(1), 145–156. [Google Scholar] [CrossRef]
- Reinsel, G., Velu, R., & Chen, K. (2022). Multivariate reduced-rank regression. Theory, methods and applications. Springer Nature. [Google Scholar]
- Sims, C. A. (1980). Macroeconomics and reality. Econometrica, 48(1), 1–48. [Google Scholar] [CrossRef]
- Stock, J. H., & Watson, M. W. (2002a). Forecasting using principal components from a large number of predictors. Journal of the American Statistical Association, 97(460), 1167–1179. [Google Scholar] [CrossRef]
- Stock, J. H., & Watson, M. W. (2002b). Macroeconomic forecasting using diffusion indexes. Journal of Business & Economic Statistics, 20(2), 147–162. [Google Scholar]
- Stock, J. H., & Watson, M. W. (2016). Dynamic factor models, factor-augmented vector autoregressions and structural vector autoregressions in macroeconomics. In J. Taylor, & H. Uhlig (Eds.), Handbook of macroeconomics (Vol. 2A, pp. 415–525). North Holland. [Google Scholar]
- Wang, D., Zheng, Y., Lian, H., & Li, G. (2022). High-dimensional vector autoregressive time series modeling via tensor decomposition. Journal of the American Statistical Association, 117(539), 1338–1356. [Google Scholar] [CrossRef]
Table 1.
Previous models as special cases of the CIAAR.
Table 1.
Previous models as special cases of the CIAAR.
| Model | Restrictions on Model (20) | N. of Restrictions |
|---|
| MAI | for
| |
| IAAR | | |
| VECIM | for | |
| Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).