Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

THIS ARTICLE HAS BEEN CORRECTED.

SEE LAST PAGE

Psychological Methods © 2012 American Psychological Association


2012, Vol. 17, No. 2, 176 –192 1082-989X/12/$12.00 DOI: 10.1037/a0027543

An SEM Approach to Continuous Time Modeling of Panel Data:


Relating Authoritarianism and Anomia
Manuel C. Voelkle Johan H. L. Oud
Max Planck Institute for Human Development, Berlin, Germany Radboud University Nijmegen

Eldad Davidov Peter Schmidt


University of Zurich Higher School of Economics (HSE)

Panel studies, in which the same subjects are repeatedly observed at multiple time points, are among the
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

most popular longitudinal designs in psychology. Meanwhile, there exists a wide range of different
This document is copyrighted by the American Psychological Association or one of its allied publishers.

methods to analyze such data, with autoregressive and cross-lagged models being 2 of the most well
known representatives. Unfortunately, in these models time is only considered implicitly, making it
difficult to account for unequally spaced measurement occasions or to compare parameter estimates
across studies that are based on different time intervals. Stochastic differential equations offer a solution
to this problem by relating the discrete time model to its underlying model in continuous time. It is the
goal of the present article to introduce this approach to a broader psychological audience. A step-by-step
review of the relationship between discrete and continuous time modeling is provided, and we demon-
strate how continuous time parameters can be obtained via structural equation modeling. An empirical
example on the relationship between authoritarianism and anomia is used to illustrate the approach.

Keywords: continuous time modeling, panel design, autoregressive cross-lagged model, longitudinal data
analysis, structural equation modeling

Supplemental materials: http://dx.doi.org/10.1037/a0027543.supp

How humans develop, how societies change over time, and what nesses. Roughly, two broad categories of methods can be distin-
factors affect these changes are fundamental research topics in guished: models in which time is considered explicitly and models in
psychology and the social sciences. However, while in the real which time is only considered implicitly. Hierarchical linear models
world most of these changes develop continuously, they usually (HLM; e.g., Raudenbush & Bryk, 2002) or latent growth curve
cannot be observed in a truly continuous manner. Rather, research- models (e.g., Bollen & Curran, 2006; Duncan, Duncan, & Strycker,
ers are forced to use “snapshots” of developmental processes in 2006) are typical examples of the former. In HLM time is explicitly
order to learn something about the underlying continuous time entered as a predictor into the model equation, while in latent growth
process and factors that possibly affect it. curve models (LGMs) time is represented by the factor loadings. In
Panel designs, in which the same subjects are repeatedly observed contrast, autoregressive and cross-lagged models are typical examples
across time, are typical examples of such “snapshots.” During the last of the latter, because here time is only considered implicitly by the
decades a number of different methods have been developed to order of the measurement occasions but not the exact time points or
analyze panel data, each associated with specific strengths and weak- time intervals between them (e.g., Finkel, 1995). As a consequence, it
is difficult to compare autoregressive and cross-lagged parameters
that are based on different time intervals. To illustrate the problems
This article was published Online First April 9, 2012. associated with models for longitudinal data that do not explicitly
Manuel C. Voelkle, Max Planck Institute for Human Development, account for time, the next three sections consider three examples of
Berlin, Germany; Johan H. L. Oud, Behavioural Science Institute, Radboud increasing complexity.
University Nijmegen, Nijmegen, the Netherlands; Eldad Davidov, Institute
for Sociology, University of Zurich, Zurich, Switzerland; Peter Schmidt,
Laboratory for Socio-Cultural Research, Higher School of Economics On Estimating and Comparing Autoregressive
(HSE), Moscow, Russia. Parameters
The work of the fourth author was supported by the Higher School of
Economics (HSE) Basic Research Program (International Laboratory for Imagine a researcher (Researcher 1) who is interested in the
Socio-Cultural Research). The data were collected within the GMF Project stability of physical well-being in adolescents. His long-term goal
on group-related enmity in Germany financed by the Volkswagen and the
may be to compare physical well-being in adolescents of different
Freudenberg Foundation from 2001 till 2011. We would like to thank Timo
von Oertzen for his helpful comments and stimulating discussions and Lisa
nationalities. For this purpose he has just finished a yearlong panel
Trierweiler for the English proof of the article. study on South African adolescents. For reasons of simplicity, let
Correspondence concerning this article should be addressed to Manuel us assume physical well-being represents a single factor measured
C. Voelkle, Max Planck Institute for Human Development, Lentzeallee 94, via self-report on a monthly basis (thus including the baseline
14195 Berlin, Germany. E-mail: voelkle@mpib-berlin.mpg.de measure, the study consists of 13 measurement occasions with a

176
CONTINUOUS TIME MODELING 177
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.
This document is copyrighted by the American Psychological Association or one of its allied publishers.

Figure 1. Autoregressive and cross-lagged parameter estimates of two studies on the relationship between
two constructs across T ⫽ 12 versus T ⫽ 6 intervals. All parameter estimates were constrained to equality
over time, and time intervals are assumed to be of equal length within each study (⌬t1, . . . , ⌬t12 ⫽ 1 month
in Study 1, in the upper half of the figure, and ⌬t1, . . . , ⌬t6 ⫽ 2 months in Study 2, in the lower
half).

monthly interval of ⌬ti ⫽ 1 month for all time intervals i ⫽ month during the first year and ⌬ti ⫽ 3 months during the second
1, . . . , T ⫽ 12).1 In order to estimate stability, he opts for a very year)? Even worse, Researcher 1 could decide to continue the
parsimonious model, which belongs to the most widely used study by handing out four identical questionnaires on physical
models for longitudinal data: the autoregressive model of order well-being and asking the adolescents to return them at any four
one (cf. Lütkepohl, 2005): different time points throughout the next year. In this case, inter-
vals would not only differ across time but also across individuals.
xi ⫽ axi⫺1 ⫹ wi. (1) How can we compute the stability (a) in such a design? Simply
Because we come back to this model several times in the follow- ignoring the issue of different time intervals by applying Equation
ing, it is worth having a closer look at it. In this simple form, 1 is certainly not a solution.
physical well-being (x) at any discrete measurement occasion (i) is
a function of the previous measurement occasion (xi⫺1 ), weighted On Interpreting Cross-Lagged Effects
by the autoregressive coefficient (a), and an error term (wi). If
there is perfect stability, a equals one, whereas a ⫽ 0 if physical Let us assume the two researchers are interested not just in the
well-being at any measurement occasion is completely indepen- stability of physical well-being but also in the relationship between
dent of previous well-being. Suppose the researcher observed a
stability coefficient of a ⫽ 0.64, indicating a moderate degree of
1
stability. Suppose further, another researcher (Researcher 2) did When discussing differences between discrete and continuous time
exactly the same study but with T ⫽ 6 and an interval of 2 months analyses, it is important to be precise with the notation. In the present
article we use t to indicate the exact time point of an observation. For
(⌬ti ⫽ 2). Researcher 2 observed a stability coefficient of a ⫽
example, x(t ⫽ 2011) means that variable x was observed in the year 2011,
0.42. How can we compare the two coefficients? Is stability higher or y(t ⫽ 105 ms) could mean that y was observed 105 ms after a stimulus
in Study 1 with a ⫽ 0.64 and ⌬ti ⫽ 1 month, or in Study 2 with onset. The unit of t, of course, depends on the object of research. In
a ⫽ 0.42 and ⌬ti ⫽ 2 months? contrast, i ⫽ 1, . . . , T, is an index denoting the rank of an observation in
As this brief example suggests, the autoregressive coefficient a a series of observations. In addition, we assume an initial measurement
(and error term w for that matter) depends on the length of the time occasion (i ⫽ 0), which is predetermined. Because a study with T ⫹ 1
interval ⌬ti between xi and xi⫺1 . This information, however, is measurement occasions has T intervals between them, we can use the same
missing in Equation 1, because the autoregressive model considers index i ⫽ 1, . . . , T to indicate the (rank of a) time interval. For example,
time only “implicitly” by accounting for the order of the measure- xi ⫽ 5 means that x was observed at the sixth (including the baseline
ment occasions but not for the length of the time intervals between measure at i ⫽ 0) measurement occasion. Of course the two notations may
also be combined, with ti indicating that time point t represents the (i ⫹
them.
1)th measurement occasion in this study. For example, x(ti ⫽ 5 ⫽ 2011)
The situation gets even more complicated if Researcher 1 de- means that x was observed in the year 2011, which constitutes the sixth
cides to extend his study by another year— but at a lower sampling measurement occasion in this study. The same logic applies to the intervals
rate of only four additional measurement occasions with an inter- between two adjacent time points: ⌬ti represents the ith interval of length
val of ⌬ti ⫽ 3 months between them. How can he estimate the ti – ti ⫺ 1. Note that the first interval is denoted ⌬t1 (and not ⌬t0 ), while the
stability of physical well-being (a) in such a design (i.e., ⌬ti ⫽ 1 first measurement of x is denoted x0.
178 VOELKLE, OUD, DAVIDOV, AND SCHMIDT

physical and social well-being. In particular, they want to know


whether physical well-being affects social well-being, whether
social well-being affects physical well-being, or whether there is a
reciprocal effect between the two constructs. Because the nature of
the research question precludes a randomized experiment, the
researchers choose a cross-lagged panel design as illustrated in
Figure 1. Cross-lagged panel designs are often employed to infer
causal relationships between variables (e.g., Granger, 1969). The
underlying idea is to compare the cross-lagged effect of one
construct P (physical well-being), measured at time point t on
another construct S (social well-being) measured at time point t ⫹
1, to the cross-lagged effect of S measured at t on P measured at
t ⫹ 1. Because temporally later events cannot cause earlier events,
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

panel designs are considered a more suitable way to test causal


This document is copyrighted by the American Psychological Association or one of its allied publishers.

relations than are cross-sectional designs (Finkel, 1995; Granger,


1969; Oud, 2007b; but see also Rogosa, 1980, for a critique on the
use of cross-lagged correlations).
The respective statistical model in Equation 2 is the multivariate
extension of the autoregressive model introduced in Equation 1:

x共ti兲 ⫽ A共⌬ti兲x共ti ⫺ ⌬ti兲 ⫹ w共⌬ti兲. (2)

However, there are two important differences: First, being a mul-


tivariate model, x共ti兲 represents no longer a single variable but a
V ⫻ 1 vector of outcome variables, with V being the number of
variables observed at each time point ti. In our case, V ⫽ 2, for
physical and social well-being. Likewise, A共⌬ti兲 is now a V ⫻ V
matrix relating the outcome variables over time. It contains the
autoregressive effects in the main diagonal and cross-lagged ef-
fects in the off-diagonals. Finally, w共⌬ti兲 is a V ⫻ 1 vector of
prediction errors, which are assumed to be uncorrelated over time.
Second—and more importantly—in contrast to Equation 1, Equa-
tion 2 makes it explicit that the autoregressive and cross-lagged
effects A共⌬ti兲 depend on the time interval ⌬ti between x共ti兲 and
x共ti ⫺ ⌬ti兲. The same applies to the error term w共⌬ti兲. Further-
more, the notation x共ti兲 points to the fact that although x may be
observed at any time point t, in an empirical study the time points
are always discrete, and thus the subscript i is added. Likewise, the
time interval ⌬ti between two discrete time points must also be
discrete with ⌬ti ⫽ ti ⫺ ti⫺1 . Note that just like a in Equation 1,
A共⌬ti兲 is a function of the time interval 共⌬ti兲, but apart from that the
underlying process is assumed to be constant over time. At this
point it is important to note that in Equation 2 we just highlight the
fact that A共⌬ti兲 and w共⌬ti兲 are actually functions of the time
intervals ⌬ti. The problem is that this information is not being used
in standard autoregressive and cross-lagged models, which simply
ignore time intervals. That is, in discrete time analysis Equation 2
is written simply as

Figure 2. A: Autoregressive parameters as a function of the time interval


between observations. B: Cross-lagged parameters as a function of the time
interval between observations. For a time interval of ⌬ti ⫽ 1, . . . , T ⫽ 1,
parameter estimates correspond to the discrete time effects observed by
Researcher 1, while parameter estimates correspond to the discrete time
effects observed by Researcher 2 for ⌬ti ⫽ 1, . . . , T ⫽ 2. C: Expected values
of authoritarianism (construct S) and anomia (construct P) as a function of
time.
CONTINUOUS TIME MODELING 179

xi ⫽ Axi⫺1 ⫹ wi. cross-lagged effects. Being surprised by the systematic nature of


the resulting plots, he starts to wonder whether his time travels
As illustrated in Figure 1, as long as the two researchers use were really necessary or whether he could have known this in
different time intervals (Researcher 1: ⌬ti ⫽ 1 month vs. Re- advance. Put more generally, given that the generating process
searcher 2: ⌬ti ⫽ 2 months), they arrive at different conclusions (i.e., the process by which the two constructs and their relationship
regarding not only the stability of physical (and social) well-being evolves over time) does not change, he wonders whether a single
but also the cross-lagged effects between the two constructs. For study would have been sufficient to compute this process.
both constructs, Researcher 1 observed larger stability coefficients As the reader may already suspect, the conjecture of Researcher
than did Researcher 2. In contrast, Researcher 2 observed stronger 1 is correct. From now on we refer to the generating process as the
cross-lagged effects, in particular with respect to the effect of continuous time model (Bergstrom, 1984, 1988). Continuous time
social well-being on physical well-being (a12 ⫽ 0.18 for ⌬ti ⫽ 1 models are models in which time is not entered explicitly as an
month vs. a12 ⫽ 0.27 for ⌬ti ⫽ 2 months). In other examples, the explanatory variable but only as an index variable (see footnote 1).
relative size of the cross-lagged effects may even reverse, suggest- In contrast to discrete time models, however, the index variable
ing a change in the “causal” direction of effects. Based on such may take on any continuous set of values.2 In the following we
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

divergent results, it is easy to imagine the fierce debate in the


This document is copyrighted by the American Psychological Association or one of its allied publishers.

demonstrate how to estimate a continuous time model based on a


scientific community on the true nature of the relationship between discrete time panel study (without time traveling). Once the con-
two (or more) constructs. tinuous time model is known, we can solve all of the problems
Of course, the alert reader has long realized that there is no point mentioned in the earlier examples.
in comparing parameter estimates that are based on different time However, before we start introducing the approach, a word of
intervals. In this example, the underlying process that generated caution is necessary: Time traveling is not possible! In the real
the results of the two studies is exactly the same. The example is world, we cannot repeat the same study with different intervals and
also quite realistic—in fact, the parameters were produced by the be sure that (apart from the intervals) nothing changes. Rather, we
same model that also underlies the empirical example on the have to assume that there is one underlying continuous time model
relationship between authoritarianism and anomia presented later (i.e., a single generating process). If we are willing to make this
in this article. The observed differences in discrete time parameter assumption, we can compute how different constructs and the
estimates are solely due to the fact that the two researchers used relations among them appear for any arbitrary time interval. If we
different time intervals in their studies. In order to find this out, are not willing to make this assumption, the only way to find out
however, we need to derive parameter estimates that are indepen- is to conduct a new study for each time interval we are interested
dent of the time intervals used in any specific study. Before we in. The degree to which it is reasonable to make this assumption
demonstrate how this is done, let us consider a final—and more depends on the research question at hand and is ultimately up to
abstract— example. the researcher to decide. If we are willing to assume that there is
one generating process underlying the relationship between phys-
On Finding the Generating Process ical well-being and social well-being, continuous time modeling
allows us to compare the results of studies that have been con-
Let us assume Researcher 1 has the extraordinary ability to
ducted with different time intervals, permits the computation of
time-travel. After having completed his study on the relationship
effects in studies with varying time intervals (within the study),
between physical and social well-being with a time interval of
and may even allow us to inter- or extrapolate to other time
⌬ti ⫽ 1 month, he travels back in time and repeats the study with
intervals. If we are not willing to make this assumption, there is no
a time interval of ⌬ti ⫽ 0.5 months. As we have seen earlier, he
point in comparing parameter estimates of different studies or
obtained
parameter estimates obtained at different time intervals within the


0.64
A共⌬ti ⫽ 1兲 ⫽ 0.03
0.18
0.89 冊 same study, regardless of the statistical method that is used. Put
more generally, we have to assume that the object of research (to
identify the generating process) is independent of the method (any
in his first attempt. When redoing the study, after traveling back in specific study with specific time intervals).
time, he obtained


0.80
A共⌬ti ⫽ 0.5兲 ⫽ 0.02
0.10
0.94 .冊 The Present Article
Obviously, what is needed is an approach that brings time back
Because he traveled back in time, except for the different time into discrete time models where it is otherwise considered only
intervals, everything else (i.e., the true relationship between implicitly. This can be done by using stochastic differential equa-
physical and social well-being) was exactly the same as in the
first study. Having obtained these different parameter estimates, 2
he becomes curious about how the parameters will change if he One can also conceive of such a continuous time model as a dynamical
system. At each point in time, the system has a specific configuration, but
repeats the process with many different intervals. So he travels
time itself never acts as an explanatory variable in the model. Instead, the
back and forth in time, let us say 1,000 times, and records the model itself is an explanatory model. Arguably, this is often a more
parameter estimates for different intervals between ⌬ti ⫽ 0 months realistic view of the world compared with models that include time ex-
and ⌬ti ⫽ 10 months. He then plots the parameter estimates plicitly as a predictor (e.g., HLMs or LGMs). For example, when using an
against the different time intervals and obtains the results shown in LGM to study learning, we typically do not assume that time “causes”
Figure 2A for the autoregressive effects and Figure 2B for the learning, although we mathematically model it this way.
180 VOELKLE, OUD, DAVIDOV, AND SCHMIDT

tions, and it is the goal of the present article to introduce contin- subscript i indicates that although time itself is continuous (t),
uous time modeling based on stochastic differential equations to a observations are necessarily taken at discrete time points (ti). For
broader psychological audience. Having said that, little of what is the moment let us further assume that all variables are in deviation
being presented in the following is actually new in the sense that form, so that there is no need for an intercept term in Equation 2.
the method did not exist before. Quite the contrary: Continuous This assumption is relaxed later on. Finally, w共⌬ti兲 is a V ⫻ 1
time models based on differential equations have been around vector of error terms. Using SEM, x does not have to be directly
since Isaac Newton in the 17th century. Likewise, stochastic observed but may be latent (cf. McArdle, 2009). From a substan-
differential equations are well established in other disciplines, such tive point of view, this typically requires measurement invariance
as physics or engineering, and many mathematical problems as- over time (Meredith, 1993) but also opens up some flexibility with
sociated with them were resolved in the first half of the 20th respect to the measurement error (co)variance structure.
century. However, with few rather technical exceptions (e.g., Oud
& Delsing, 2010; Oud & Jansen, 2000; Singer, 1998), continuous The Continuous Time Auto- and Cross-Effects Model
time models are virtually absent in the psychological literature, and
The continuous time model underlying Equation 2 is now
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

even experienced quantitative psychologists routinely rely on dis-


This document is copyrighted by the American Psychological Association or one of its allied publishers.

crete time models in the assumed absence of better alternatives. introduced in a stepwise fashion. To facilitate understanding we
Accordingly, our goals in the present article are (1) to introduce start with an intuitive approach to the basic idea, before deriv-
psychologists to continuous time modeling and (2) to demonstrate ing the exact model and explicating the relationship between
how continuous time models may overcome many of the problems continuous and discrete time parameters. For reasons of read-
of discrete time analyses. Furthermore, we would like to (3) ability and comprehension, we do not discuss the stochastic
facilitate the use of continuous time models and (4) enable readers error term until later on.
to understand and critically evaluate the outcomes of continuous
time analyses. To achieve the latter two goals, we present contin- An Intuitive Approach to Continuous Time Modeling
uous time models within the general framework of structural
equation modeling (SEM), which most readers are familiar with, Consider again our introductory example on estimating and
and supplement the article by computer code that does the analysis comparing autoregressive parameters in which two researchers
and is available for download from the accompanying online investigated the stability of physical health using an autoregressive
Supplemental Materials. Finally, we provide an example of a model. Because the time intervals differed, we were hesitant (and
continuous time model using an empirical data set. rightly so) to compare the resulting autoregressive effects directly,
We proceed in the following order: First, we briefly review the but we wondered whether stability is higher in Study 1 with a ⫽
conventional autoregressive cross-lagged model for discrete mea- 0.64 and ⌬ti ⫽ 1 month or in Study 2 with a ⫽ 0.42 and
surement. Second, the underlying continuous time model is intro- ⌬ti ⫽ 2 months. An intuitively appealing solution to this
duced in a stepwise fashion. Third, a number of important exten- problem could be to compute the difference between x(ti)
sions are introduced, including the continuous time stochastic error and x(ti – ⌬ti) and divide this difference by the length of the time
process and continuous time intercepts. Fourth, after a short tech- interval (⌬ti). This “normalizes” the change from one measure-
nical summary of the relationship between the continuous and ment occasion to the next with respect to the length of the time
discrete time model, which is aimed at the mathematically more interval between them and gives us the rate of change: the so-
advanced audience, the model is translated into the commonly called difference quotient. Predicting this normalized difference
known SEM framework. Fifth, an example of the relationship instead of x(ti) results in the difference equation
between authoritarianism and anomia is provided to illustrate the
approach. Finally, after considering further extensions and sug- ⌬x共ti兲
⫽ Aⴱ x共ti ⫺ ⌬ti兲 (3)
gesting additional reading, we conclude with a discussion of the ⌬ti
advantages and limitations of the method.
with ⌬x(ti) ⫽ x(ti) – x(ti – ⌬ti). The original autoregressive matrix
A共⌬ti兲 and Aⴱ are related by
The Discrete Time Autoregressive
A共⌬ti兲 ⫺ I
Cross-Lagged Model Aⴱ ⫽ or inversely A共⌬ti兲 ⫽ Aⴱ ⌬ti ⫹ I, (4)
⌬ti
The discrete time autoregressive model, with or without cross-
lagged effects, is one of the most often used methods for the with I being an identity matrix. In contrast to the autoregressive
analysis of change in psychology (Hertzog & Nesselroade, 2003; cross-lagged matrix A共⌬ti兲, the new matrix Aⴱ is rendered rela-
Jöreskog, 1979; McArdle, 2009). The basic model has already tively independent of the time interval and can, therefore, be
been introduced in Equation 2, which is compared across studies with different observation intervals. In-
deed, Aⴱ is already a crude approximation of the underlying
x共ti兲 ⫽ A共⌬ti兲x共ti ⫺ ⌬ti兲 ⫹ w共⌬ti兲. continuous time model (i.e., the generating process). In order to
better distinguish between continuous time and discrete time pa-
As defined before, x(ti) is a V ⫻ 1 vector of outcome variables, rameters, in the following we speak of auto-effects and cross-
with V being the number of variables observed at each time point. effects in the continuous case, compared with autoregressive and
The V ⫻ V matrix A共⌬ti兲 relates the outcome variables over time. cross-lagged effects in the discrete case. The new matrix Aⴱ
It is a function of the time interval 共⌬ti兲, but apart from that it is contains the continuous time auto-effects in the main diagonal and
assumed to be time-invariant (but see Oud & Jansen, 2000). The cross-effects in the off-diagonals. Having computed the difference
CONTINUOUS TIME MODELING 181

equation for both studies, it is now possible to compare the point (t) and are no longer restricted to the discrete time points
strength of the effects. The continuous time auto-effects for con- actually observed in a study. Just as in Equation 3, the unknown
struct P in Study 1 are Aⴱ ⫽ (0.64⫺1)/1 ⫽ ⫺0.36, whereas in derivative
Study 2 they are Aⴱ ⫽ (0.42 ⫺ 1)/2 ⫽ ⫺0.29. Note that due to the
subtraction of I in the numerator, autoregressive parameters be- dx共t兲
tween 0 and 1 become negative when translated into continuous dt
time auto-effects. In our example we see that— other than sug-
can then be predicted by x共t兲, weighted by the so-called drift
gested by the autoregressive effects (0.64 in Study 1 and 0.42 in
matrix A:
Study 2)—the difference between the two auto-effects (– 0.36 in
Study 1 and – 0.29 in Study 2) is much smaller, and the effect dx共t兲
appears to be even weaker (i.e., more negative) in Study 1 com- ⫽ Ax共t兲. (5)
dt
pared with Study 2. By simply ignoring the different time inter-
vals, we would have come to exactly the opposite conclusion. Equation 5 is a (nonstochastic) differential equation because the
derivative of x共t兲 is a function of x共t兲 itself. Fortunately, it is not a
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

Although this brief example involves only autoregressive effects


This document is copyrighted by the American Psychological Association or one of its allied publishers.

and no cross-lagged effects, the matrix notation in Equation 3 complicated differential function, so it is possible to quickly iden-
indicates that it generalizes to all possible autoregressive and tify the only function x(t) that satisfies Equation 5 as
cross-lagged models with any arbitrary number of variables and
time points. The approach is not only intuitively appealing but may x共t兲 ⫽ eA·共t⫺t0兲 x共t0 兲, (6)
sometimes be a convenient way to translate discrete time param- with x共t0 兲 representing the vector of (exogenous) outcome vari-
eters (i.e., A共⌬ti兲) into continuous time parameters (i.e., Aⴱ ) in ables at initial time point t0 . More precisely, Equation 6 is the
order to compare effects across studies with different observation unique solution of Equation 5. Proof of this relationship is given in
intervals. But most importantly, it is easy to implement. In a first Appendix A.
step, any conventional program can be used to fit a standard At some point, however, the continuous time coefficients in drift
autoregressive cross-lagged model, while in a second step the matrix A have to be related to the discrete time autoregressive
parameters can be transformed into continuous time parameters parameters (i.e., A共⌬ti兲 in Equation 2). This is done by setting
using Equation 4. This two-step procedure has been termed the Equation 6 equal to Equation 2 (see Appendix B). For starting
indirect approach by Hamerle, Nagl, and Singer (1991). value x共t0 兲 ⫽ x共ti ⫺ ⌬ti兲 and 共t ⫺ t0 兲 ⫽ ⌬ti, this allows us to
Unfortunately, the intuitive approach is associated with at least two express the exact relationship between discrete and continuous
serious shortcomings, so despite the intuitive appeal, its use is gen- time as
erally discouraged. First, the relationship A共⌬ti兲 ⫽ Aⴱ⌬ti ⫹ I is only
a crude approximation of the relationship between the true continuous A共⌬ti兲 ⫽ eA·⌬ti. (7)
time matrix (A)—the so-called drift matrix, which is introduced in the
next paragraph—and the discrete time autoregressive matrix A共⌬t i兲. As before, A共⌬ti兲 contains all autoregressive and cross-lagged
Second, the indirect approach requires the time intervals within parameters of the discrete time model, while drift matrix A con-
a study to be of equal length to enable equality constraints tains the corresponding auto- and cross-effects of the underlying
among parameters of different intervals. Consequently, the in- continuous time model. As apparent from Equation 7, the relation-
tuitive approach is at most an imprecise ad hoc method to ship between the two is not linear, but follows a highly nonlinear
compare autoregressive effects across studies with different matrix exponential function, which causes the sometimes rather
time intervals between but equal time intervals within. From a paradoxical relationships between discrete and continuous time.
didactical perspective, however, it prepares the ground for In contrast to the intuitive approach, Equation 7 gives the exact
introducing the exact relationship between continuous time and relationship between autoregressive cross-lagged matrix A共⌬ti兲
discrete time modeling because, in principle, continuous time and continuous time drift matrix A. The relationship of this exact
modeling follows the same logic. However, since the more approach to the intuitive approach becomes clear when expressing
accurate exact approach (introduced in the next section) is Equation 7 in power series expansion. The exponential function is
available, we strongly discourage the use of the intuitive ap- commonly defined as
proach in practice.


ak 1 1
ea ⫽ ⫽ 1 ⫹ a ⫹ a2 ⫹ a3 ⫹ . . . .
An Exact Direct Approach to Continuous Time k! 2! 3!
k⫽0

Modeling
Thus, we can also express Equation 7 in matrix notation as
By accounting for the length of the time interval (⌬ti) when
estimating parameters (Aⴱ ), the difference equation (Equation 3) 1 2 1
A共⌬ti兲 ⫽ eA·⌬t i ⫽ I ⫹ A·⌬ti ⫹ A · ⌬t2i ⫹ A3 · ⌬t3i . . . .
brings time back into autoregressive cross-lagged panel models. 2! 3!
The time intervals, however, remain discrete as indicated by sub-
(8)
script i. As a thought experiment, we could imagine what would
happen if we make the time intervals in Equation 3 smaller and Written in this form, it is obvious that Aⴱ in the intuitive approach
smaller (i.e., ⌬ti 3 0). Mathematically speaking, this corresponds (see Equation 4) accounts for just the first part of the entire power
to taking the derivative of x共t兲 with respect to time. Note that— series (dashed part in Equation 8). This is the reason why the
being a thought experiment—we can now refer to the exact time intuitive approach is only a crude approximation of the exact
182 VOELKLE, OUD, DAVIDOV, AND SCHMIDT

approach just introduced. The logic underlying the intuitive ap- letter w) in Equation 2. The continuous time error process W共t兲 is
proach, which does not need differential calculus, and the exact the limiting form of the discrete time random walk, better known
approach, involving the matrix exponential function, is the same. as Wiener process or Brownian motion. The Wiener process has
Likewise, Aⴱ and A can be interpreted analogously. three important defining properties: First, it has independent in-
Having introduced the exact relationship between discrete and crements with distribution N共0,⌬t兲 over interval ⌬t. Second, its
continuous time modeling, the question remains how to estimate initial value W共t0 兲 ⫽ 0, and third, W(t) is continuous (both almost
the continuous time drift matrix (A) based on discrete time inter- surely). As is the case for the discrete time error process, variances
vals in a given panel study. As before, it is tempting to adopt the larger or smaller than 1 are possible by premultiplication with the
indirect approach, that is, to estimate the discrete time parameters Cholesky triangle G. This also allows the specification of possible
(A共⌬ti兲) in a first step, and solve Equation 7 for A in a second step. covariances among the prediction errors. Because the error cova-
However, while the linear approximation is no longer an issue with riance matrix Q—which is also referred to as the diffusion ma-
this exact approach, the second problem of how to deal with trix—and G contain the same information, for any given G, Q is
parameter constraints for different time intervals remains (Hamerle also known, and vice versa (Q ⴝ GGⴕ).
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

et al., 1991). Thus, an approach is needed that constrains the Unfortunately, taking the derivative of the Wiener process W共t兲
This document is copyrighted by the American Psychological Association or one of its allied publishers.

discrete time parameters to the underlying continuous time param- is not as simple as taking the derivative of x共t兲 in Equation 5. At
eters directly during estimation. Because of the simultaneous es- this point we do not want to go into mathematical details, but
timation of discrete and continuous time parameters, we speak of roughly speaking the problem is that in discrete time it is easy to
a direct approach in contrast to the two-step indirect method. define a random walk as a process of adding independent incre-
Before having a closer look at the direct approach and how it can ments, which are randomly drawn from N(0,I). However, by
be used to translate discrete time models into continuous time, making the increments smaller and smaller (i.e., by approximating
however, we must first consider the stochastic error term, which the derivative in continuous time with ⌬t 3 0), we eventually end
has been deliberately ignored so far. up with an infinite variance. One could think of it in terms of test
theory, where making a test longer reduces its error variance, while
A More Realistic View of the World: making the test shorter increases its error variance. If we could
Introducing Errors make the test infinitesimally small (i.e., if a single item would not
be the smallest unit), its error variance would go to infinity. This
Consider again discrete time Equation 2. It has been shown that
situation has given rise to a substantial amount of mathematical
a simple but crude and problematic way to translate discrete time
research. To avoid the derivative, the stochastic differential equa-
parameters into a continuous time framework is to compute the
tion is often formulated in differential form by multiplying both
difference equation (Equation 3). In principle, the same has to be
sides by dt and is then interpreted as a stochastic integral equation.
done with the error process. In discrete time, the error process is
Even though the resulting integral still does not follow the normal
assumed to follow a random walk through time. As the name
rules of integration, the good news is that mathematics solved this
already suggests, random walk refers to a process where the value
problem a long time ago, and it can be shown that for initial value
at each time point [w共ti兲] is an additive function of the value of the
x(t0) and any time interval (⌬t ⫽ t ⫺ t0) between x(t0) and x(t) the
immediately prior time point [共w共ti ⫺ ⌬ti兲] and an additional
solution of
component e [i.e., w共ti兲 ⫽ w共ti ⫺ ⌬ti兲 ⫹ e]. If for all
⌬ti ⫽ 1 all e are drawn from a standard multivariate normal dx共t兲 dW共t兲
distribution with N(0,I), then for ⌬ti ⱖ 1 all successively nonover- ⫽ Ax共t兲 ⫹ G (10)
dt dt
lapping increments ⌬w共ti兲 ⫽ w共ti兲 ⫺ w共ti ⫺ ⌬ti兲 are independent
with covariance matrix ⫽ ⌬tiI. Without changing the nature of the is
process, error variances other than 1 over ⌬ti ⫽ 1, or correlated error
terms, can be obtained by premultiplying ⌬w共ti兲 with G, where G is
the Cholesky triangle of the desired covariance matrix Q (i.e.,
Q ⫽ GG⬘). Thus, we could complement Equation 3 by adding the
x共t兲 ⫽ eA·共t⫺t0兲 x共t0 兲 ⫹ 冕 t

t0
eA·共t⫺s兲 GdW共s兲 (11)

error term with

冋冕 册冕
⌬w共ti兲
G . t t
⌬ti cov eA·共t ⫺ s兲GdW共s兲 ⫽ eA·共t⫺s兲 QeA⬘·共t⫺s兲 ds
All problems associated with the intuitive (difference) approach t0 t0

再 冎
would, of course, remain the same.
As demonstrated in the previous section, the exact drift matrix
⫽ irow A⫺1
# 关e
A# ·共t⫺t0 兲
⫺ I兴rowQ
A is obtained by computing the derivative of x共t兲 with respect to
time (see Equation 5). In the same manner, the continuous time
error process W共t兲 can be put in derivative form: for Q ⫽ GGⴕ and A# ⫽ A 䊟 I ⫹ I 䊟 A. Note that the first part
dW共t兲 of Equation 11 corresponds to Equation 6 (see also the proof of the
G . (9) first part in Appendix A). The variable of integration (s) has been
dt
chosen in order not to confuse it with the upper limit of integration
Capital letter W is used to denote the continuous time error process t. Via the operator row, the elements of matrix Q are put row-wise
and to distinguish it from the discrete time error term (lowercase into a column vector, while irow represents the inverse operation
CONTINUOUS TIME MODELING 183

(i.e., putting the elements back into a matrix). 䊟 denotes the duced in a stepwise fashion throughout the previous parts of this
Kronecker product. Equations 10 and 11 are described in more article, is
detail in Appendix C, while we refer the reader to Arnold (1974,
pp. 128 –134), Ruymgaart and Soong (1985, pp. 80 –99), Oud and dx共t兲 dW共t兲
⫽ Ax共t兲 ⫹ b ⫹ G . (14)
Jansen (2000), or Singer (1990) for mathematical details. Because dt dt
of the stochastic component of the error term, Equation 10 is called
Step 3: For x共t兲 ⫽ x共t0 兲 and any time interval (⌬t ⫽ t ⫺ t0 )
a stochastic differential equation.
between x共t0 兲 and x共t兲, the solution of Equation 14 is

Introducing Intercepts
The basic stochastic differential model introduced in Equation
x共t兲 ⫽ eA·共t⫺t0兲 x共t0 兲 ⫹ A⫺1 关eA·共t⫺t0兲 ⫺ I兴b ⫹ 冕 t

t0
eA·共t⫺s兲 GdW共s兲

10 can be extended in various ways. In particular, until now all


variables were assumed to be in deviation form; thus the model did (15)
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

not permit nonzero mean trajectories. This assumption can be with


relaxed by adding a V ⫻ 1 continuous time intercept vector b to
This document is copyrighted by the American Psychological Association or one of its allied publishers.

Equation 10. Then solving for initial value x共t兲 ⫽ x共t0 兲 allows us to
write the expected value of x共t兲 at any time point t as

E关x共t兲兴 ⫽ eA·共t⫺t0兲 E关x共t0 兲兴 ⫹ A⫺1 关eA·共t⫺t0兲 ⫺ I兴b.


cov
冋冕 t

t0
册冕
eA·共t⫺s兲 GdW共s兲 ⫽
t

t0
eA·共t⫺s兲 QeA⬘·共t⫺s兲 ds

再 冎
(12)

Because in general the corresponding elements in drift matrix A ⫽ irow A⫺1


# 关e
A# ·共t⫺t0 兲
⫺ I兴rowQ
become negative for autoregressive parameters between 0 and 1,
eA·共t⫺t0兲 approaches zero for increasing time intervals. Thus E关x共t兲兴
approaches ⫺A⫺1b as 共t ⫺ t0 兲 3 ⬁. The final mean vector for Q ⫽ GGⴕ and A# ⫽ A 䊟 I ⫹ I 䊟 A.
(⫺A⫺1b), to which the process eventually converges, represents Step 4: For x共t0 兲 ⫽ x共ti ⫺ ⌬ti兲 and 共t ⫺ t0 兲 ⫽ ⌬ti, Equation
the so-called (stable) equilibrium position. Just like in any ordinary 13 can be set equal to Equation 15, which yields the relationships
regression analysis, it is possible to account for different mean between continuous and discrete time parameters. Having identi-
trajectories of different groups by replacing the V ⫻ 1 vector b by fied these relationships, the discrete time model can be expressed
the product of a V ⫻ R matrix B and an R ⫻ 1 vector of R as a function of the underlying continuous time parameters:
exogenous variables u (cf. Oud & Delsing, 2010). In principle, the
variables in vector u may either be continuous or represent dummy x共ti兲 ⫽ eA·⌬tix共ti ⫺ ⌬ti兲 ⫹ A⫺1 关eA·⌬ti ⫺ I兴b ⫹ w共⌬ti兲 (16)
variables (e.g., to allow different mean trajectories for men and with covariance matrix Q共⌬ti兲 as defined earlier.
women). As before, we assume that b and Bu do not vary across Step 5: All that remains to be done is to estimate the parameters
time (but see Oud & Jansen, 2000). Due to lack of space, however, in Equation 16, and there are different ways to do so (Oud &
group differences are not considered any further in the present Singer, 2008). One way is to use SEM, a well-established and
article. convenient way to obtain maximum likelihood parameter esti-
mates if it is possible to reformulate Equation 16 as a structural
On the Relationship Between Continuous and equation model and minimize the well-known function
Discrete Time: A Summary
F ML ⫽ log兩⌺兩 ⫹ tr共S⌺⫺1 兲 ⫺ log兩S兩 ⫺ 共V ⫹ 1兲 (17)
So far, we have introduced the logic and rationale of continuous
with V denoting the number of observed variables, S being the
time modeling in a stepwise fashion by aiming at readers who are
observed, and ⌺ the model implied augmented moment matrix.
new to continuous time modeling. The present section integrates
The reformulation of Equation 16 as a structural equation model is
the previous parts in a compact—and mathematically explicit—
demonstrated in the next paragraph.
form. Essentially, continuous time modeling can be summarized in
five steps: First, the discrete time model is formulated as usual.
Second, the derivative with respect to time is computed, resulting Continuous Time Modeling in SEM
in a stochastic differential equation. Third, the stochastic differen- In SEM one commonly distinguishes between a measurement
tial equation of Step 2 is solved for any arbitrary time interval. part as defined in Equation 18 and a structural part as shown in
Fourth, discrete time parameters are constrained according to Step Equation 19 (see also Bollen, 1989; e.g., Jöreskog, 1973; Mayr,
3 during the estimation process. Fifth, the model is estimated. For Erdfelder, Buchner, & Faul, 2007):
this we formulate it as a structural equation model. In the following
we provide a short summary of these steps. y ⫽ ⌳␩ ⫹ ␧ with cov共␧兲 ⫽ ⌰ (18)
Step 1: The complete discrete time model corresponds to Equa-
tion 2, augmented by the intercept vector b共⌬ti兲: ␩ ⫽ B␩ ⫹ ␨ with cov共␨兲 ⫽ ⌿. (19)

x共ti兲 ⫽ A共⌬ti兲x共ti ⫺ ⌬ti兲 ⫹ b共⌬ti兲 ⫹ w共⌬ti兲. (13) In the measurement model, vector y contains the manifest (i.e.,
directly observed) variables, which are related to the latent factors
Step 2: Taking the derivative with respect to time, the corre- in vector ␩, weighted by the factor loading matrix ⌳, plus the
sponding stochastic differential equation, which has been intro- corresponding measurement error vector ␧, with covariance matrix
184 VOELKLE, OUD, DAVIDOV, AND SCHMIDT

⌰. In the structural model, the variables of interest ␩ are related to sary constraints can be formulated within the SEM framework, the
each other via matrix B. The prediction errors are contained in next section provides an empirical example of the approach.
vector ␨ with covariance matrix ⌿. For reasons of simplicity, we
do not consider the measurement model in the present article, so
that we can ignore Equation 18 by setting ␩ ⫽ y. The approach, An Empirical Example: Relating Authoritarianism
however, generalizes readily to more complex models including and Anomia
latent variables. As usual, we assume that ␧ is uncorrelated with ␨
Over the past 5 decades, the theoretical concepts of authoritar-
and ␩ and that E关␧兴 ⫽ E关␨兴 ⫽ 0. For more detailed information
ianism and anomia have played an important role in sociology and
we refer the reader to Bollen (1989; see pp. 14 and 20 for model
social psychology. To date, most researchers (e.g., Alba, Schmidt,
definition and standard assumptions). As mentioned earlier, mea-
& Wasmer, 2004; Altemeyer, 1996; Lutterman & Middleton,
surement invariance is important when tracking latent constructs
1970; Scheepers, Felling, & Peters, 1992; Stenner, 1997) have
over time (cf. Meredith, 1993; Vandenberg, 2002). Once ⌳, ⌰, B,
agreed that authoritarianism reflects (a) an individual preference
and ⌿ have been defined, it is easy to derive the model implied
for submission under authorities (authoritarian submission), (b) a
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

covariance matrix (Bollen, 1989; p. 325) and estimate parameters


strict orientation along the perceived conventions of the ingroup
This document is copyrighted by the American Psychological Association or one of its allied publishers.

by minimizing Equation 17. So all that needs to be done is to


(authoritarian conventionalism), and (c) aggressive stances toward
define ␩, the matrix of regression coefficients B, and the vector of
outgroups (authoritarian aggression). Anomia has been defined by
error terms ␨, respectively, the covariance matrix ⌿, in terms of
Srole (1956) as consisting of five subdimensions: (a) political
Equation 16.
powerlessness, (b) social powerlessness, (c) generalized socioeco-
For T ⫹ 1 time points with V constructs at each occasion (each
nomic retrogression, (d) normlessness and meaninglessness, and
possibly measured by different indicators), vector ␩ consists of
(e) social isolation. The direction of the causal relation between the
共T ⫹ 1兲 · V distinct elements. However, if we want to permit
two constructs, however, is still controversial. First, it was hypoth-
nonzero mean trends, ␩ must be extended by an additional element
esized that anomia leads to authoritarianism (Merton, 1949; Srole,
(the unit variable that has 1 for all sample units), so that
1956), because it was assumed that individuals who feel normless
␩⬘ ⫽ 共关x共t0 兲兴⬘ 关x共t1 兲兴⬘ 关x共ti兲兴⬘ . . . 关x共ti⫽T兲兴⬘ 1兲⬘ and meaningless adopt authoritarian attitudes in order to regain
orientation in an environment that is perceived as increasingly
with all vectors within ␩ as defined before. Accordingly, the complex and irritating. This view, however, was challenged by an
vector of error terms is alternative explanation proposed by McClosky and Schaar (1965),
who suggested that authoritarianism causes anomia. According to
␨⬘ ⫽ 共关x共t0 兲 ⫺ ␮x共t0兲 兴⬘ 关w共⌬t1 兲兴⬘ 关w共⌬ti兲兴⬘ . . . 关w共⌬ti⫽T兲兴⬘ 1兲⬘, McClosky and Schaar, certain personality characteristics as re-
flected by authoritarianism lead to anomia, because the narrow-
with ␮x共t0兲 being the V ⫻ 1 mean vector of the V constructs mindedness of authoritarian people confines their opportunities for
observed at the first time point, and w共⌬ti兲 being a vector of social interactions with others (e.g., Schlueter, Davidov, &
discrete time error terms as defined earlier. Schmidt, 2007). Both positions share the view that the two con-
Constraining the discrete time parameters in Equation 2 to the structs are rather stable over time, even though authoritarianism is
underlying continuous time parameters according to Equation 16, regarded as somewhat more stable than anomia, because it repre-
the two matrices B and ⌿ become sents a personality characteristic in the broadest sense. In contrast,
anomia, which reflects an attitude of disorientation, is more sus-


0 0 ... 0 0 ␮x共t 0兲


ceptible to displaying changes in the relative position of individ-
eA·⌬t1 ⫺1
0 0 0 A 关eA·⌬ti ⫺ I兴b uals over time.
In our illustrative application we take up the controversy and try
B⫽ 0 eA·⌬ti 0 0 A⫺1 关eA·⌬ti ⫺ I兴b to answer the question of whether anomia leads to authoritarianism
· ·· .
· ·
· (supporting the view of Merton, 1949; Srole, 1956) or whether
0 0 eA·⌬ti⫽T 0 A⫺1 关eA·⌬ti⫽T ⫺ I兴b authoritarianism leads to anomia (supporting the view of McClo-
0 0 ... 0 0 0 sky & Schaar, 1965). Furthermore, we are interested in the stability
and of the two constructs over time. Although no claim is made with
respect to causal relationships in a strict sense, which would


⌽关x共t0 兲兴 require an experimental design, panel data offer a good (and


oftentimes the only) opportunity to come close to the experimental
0 Q共⌬t1 兲 ideal (Finkel, 1995).
⌿⫽ 0 0 Q共⌬ti兲
· ·· .
· ·
· Sample and Measurement Instruments
0 0 0 Q共⌬ti⫽T兲
0 0 0 0 1 Data were taken from a recent panel study of the German
⌽关x共t0 兲兴 is the covariance matrix of the (exogenous) constructs at general population 16 years of age and older without an immigra-
the first occasion. We do not assume any prediction error at the tion background (see Heitmeyer, 2004). Computer-assisted inter-
first occasion. views were conducted at five points of measurement in 2002,
Having introduced the relationship between discrete and con- 2003, 2004, 2006, and 2008. Note that the first three assessment
tinuous time parameters and having demonstrated how the neces- waves were 1 year apart, while the last two measurement occasions
CONTINUOUS TIME MODELING 185

Table 1
Anomia and Authoritarianism Measured During Five Years

Anomia Authoritarianism

Date N M SD N M SD

2002 2,721 2.50 0.80 2,722 2.84 0.68


2003 1,175 2.70 0.82 1,176 2.85 0.67
2004 826 2.81 0.78 826 2.83 0.67
2006 1,024 2.63 0.80 1,298 2.92 0.81
2008 560 2.48 0.80 1,047 2.70 0.82

took place after a 2-year interval. No measurements were obtained variables only. This was done for reasons of simplicity. As dem-
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

in 2005 and 2007. onstrated in the theoretical part of this article, the model can be
This document is copyrighted by the American Psychological Association or one of its allied publishers.

Authoritarianism was measured by four items that were se- easily extended to any arbitrary number of indicators at each time
lected from an authoritarianism scale used in previous German point. Maximum likelihood parameter estimates are provided in
studies (Schmidt, Stephan, & Herrmann, 1995). Items were Table 2. As expected, with an autoregressive coefficient of 0.869,
presented on a 4-point rating scale providing response options authoritarianism is a very stable construct. Likewise, with an
from 1 (agree totally) to 4 (do not agree at all). The original autoregressive coefficient of 0.589, the stability of anomia is
response options were recoded, so that higher values indicate somewhat lower but still reasonably high. Both cross-lagged ef-
higher agreement. The item were “In order to preserve law and fects are significant, but the effect of authoritarianism on anomia
order, it is necessary to act harder against outsiders,” “One (0.202) is much higher than the effect of anomia on authoritarian-
should punish criminal acts harder,” “One should be obedient ism (0.033). Being a standard structural equation model, all pa-
and respectful to authorities,” and “One should be grateful to rameters can be interpreted as usual, so we do not go into details
leaders who tell us what to do.” The average of the four items at this point.
was used for all subsequent analyses. Anomia was measured by Instead, we take up the question raised at the beginning of the
the following three 4-point rating scale items: “Everything has article on how to compare the observed effects to effects of other
become so much in disarray that one does not know where one studies. Suppose another researcher had conducted a similar study
actually stands,” “Matters have become so difficult these days but used different time intervals and thus obtained different pa-
that one does not know what is going on,” and “People were rameter estimates. Are the differences solely due to the different
better off in the past.” Just as for authoritarianism, the average time intervals, or do parameters differ irrespective of the length of
rating of all three items was computed. the time interval? With the present discrete time model we cannot
Table 1 contains some descriptive information and an overview answer this question. Furthermore, in the present analysis we
of the sample size across the five observation waves. A total of simply ignored the fact that the first two time intervals were 1 year,
2,722 persons participated in the study. Response rates were 43% whereas the last two time intervals were 2 years. Even if time
in the second wave, 30% in the third wave, 48% in the fourth intervals differ only slightly, in some cases this may lead to
wave, and 21% in the last wave. Although the loss of subjects is completely wrong results and conclusions when ignored, while in
substantial, it is typical for longitudinal surveys like the present other situations different time intervals may have only a minor
one. Full information maximum likelihood was used to deal with effect on results. In order to find out, however, we have to use
missing values. continuous time modeling.3

Discrete Time Autoregressive Cross-Lagged Model Continuous Time Auto- and Cross-Effects Model With
To investigate the causal relationship between anomia and au- Unequal Intervals
thoritarianism, a standard discrete time autoregressive cross-
lagged model was fitted to the data. Figure 3 shows a pictorial As described earlier, the matrix exponential relationship
representation of the model. As discussed before, this model does A共⌬ti兲 ⫽ eA·⌬ti between the discrete time matrix A共⌬ti兲 and the
not account for the fact that the time intervals between observation continuous time matrix A lies at the heart of the exact approach
waves differed and will, therefore, yield incorrect results. How- (see Equation 7). To our knowledge, at present Mx and OpenMx
ever, since we are interested in the differences between the (cor- are the only SEM programs that allow the use of such nonlinear
rect) continuous time model and the (incorrect) discrete time constraints on matrices. Thus, while the results of the discrete time
model, let us start with a discrete time model. model in the previous section can be obtained by using any
The model contains 14 parameters to be estimated: Two autore-
gressive and two cross-lagged effects, two prediction error vari- 3
As pointed out by an anonymous reviewer, an alternative approach to
ances, two intercepts, and one prediction error covariance. All account for (few) unequal intervals is the use of phantom variables
parameters are constrained to equality over time. In addition, the (McArdle, 2009; Rindskopf, 1984). This allows equality constraints on the
two means, two variances, and the covariance of the latent mea- discrete time parameters, even if some time intervals differ within a study.
sures at the first time point were freely estimated. All measurement All other problems associated with discrete time analyses, however, re-
errors were set to zero, thus reducing the analysis to manifest main.
186 VOELKLE, OUD, DAVIDOV, AND SCHMIDT
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.
This document is copyrighted by the American Psychological Association or one of its allied publishers.

Figure 3. Bivariate discrete time autoregressive cross-lagged model of authoritarianism (au) and anomia (an).
Squares represent observed (manifest) variables, circles/ellipses represent latent variables. For reasons of
simplicity, all measurement errors were fixed to zero in the present example (dashed circles), making it a model
with only manifest variables. However, as the figure illustrates, the model can be easily extended to latent
variables. The triangle to the right represents the constant 1. Its path coefficients (single-headed arrows) represent
the means or intercepts of the variables in question. Path coefficients associated with dashed lines are all fixed
to 1, while path coefficients associated with solid lines are freely estimated but are usually constrained to other
parameters as described in the text. Double-headed arrows indicate covariances.

common SEM program, we used OpenMx (Boker et al., 2011) for Comparing A共⌬ti ⫽ 1兲 to the autoregressive and cross-lagged
estimating the continuous time parameters. OpenMx is an open parameters of the discrete time model in Table 2, we find that the
source R-based (R Development Core Team, 2011) SEM program correct autoregressive parameters are somewhat higher (anomia:
that is freely available. All program scripts for the analyses in the 0.643 vs. 0.589; authoritarianism: 0.893 vs. 0.869) than the dis-
present article are available for download from the online Supple- crete time parameters. In contrast, for ⌬ti ⫽ 2,
mental Materials.
Results of the continuous time auto- and cross-effects model
with unequal intervals are given in Table 3. Most importantly, the
eA·⌬ti ⫽ e冉
⫺0.447
0.043
0.232
⫺0.117 冊 ·2 ⫽ A共⌬ti ⫽ 2兲 ⫽ 冉 0.419
0.050
0.271

0.804 ,
drift matrix A is
the effects are lower (anomia: 0.419 vs. 0.589; authoritarianism:

A⫽ 冉
⫺0.447 0.232
0.043 ⫺0.117 . 冊 0.804 vs. 0.869). Obviously, by ignoring the length of the time
intervals, the parameters obtained via a standard autoregressive
cross-lagged analysis are nonlinear mixtures of the parameters
In fact, we used this same drift matrix to construct the introductory obtained for ⌬ti ⫽ 1 and ⌬ti ⫽ 2. Even though the difference in
example on the relationship between physical and social well- the stability of anomia (0.643 for ⌬ti ⫽ 1 vs. 0.419 for
being. Having computed the parameters of the underlying contin- ⌬ti ⫽ 2) is already substantial, parameters may differ even more
uous time model, it becomes possible to compute the correspond- for more complex designs with larger differences in time intervals.
ing discrete time parameters at any arbitrary point in time (see In these situations, parameters of standard discrete time models
Equation 16). For example, computing the discrete time autore- can no longer be interpreted in any meaningful way.
gressive and cross-lagged effects for time interval ⌬ti ⫽ 1 we Having obtained the continuous time parameters, we may
obtain now also inter- or extrapolate to any time interval of interest—

冉 0.643 冊
provided that such inter- or extrapolation is meaningful on
eA·⌬t i ⫽ e冉 冊 ·1 ⫽ A共⌬ti ⫽ 1兲 ⫽
⫺0.447 0.232 0.176
⫺0.117 substantive grounds. Because the drift matrix is identical to the
0.893 .
0.043
0.033
CONTINUOUS TIME MODELING 187

Table 2 hypothesis of McClosky and Schaar (1965) that it is more likely


Parameter Estimates of the Discrete Time Autoregressive Cross- that authoritarianism causes anomia than vice versa.
Lagged Model of Anomia and Authoritarianism Across Four
Time Intervals Extensions and Further Reading
Parameter Estimate SE Because the primary goal of this article is to introduce contin-
uous time modeling based on stochastic differential equations to a
Autoregressive effects
broader psychological audience, we limited ourselves to the con-
aanan 0.589ⴱⴱ 0.014
aauau 0.869ⴱⴱ 0.009 tinuous time model in its basic form. In recent years, however, the
Cross-lagged effects basic approach has been extended in various ways, and ongoing
aanau 0.202ⴱⴱ 0.015 research promises further advancements. In this section we want to
aauan 0.033ⴱⴱ 0.009 briefly mention some of these extensions and developments.
Latent intercepts
ban 0.532ⴱⴱ 0.041 Obviously, all formulae provided in the present article general-
bau 0.289ⴱⴱ 0.026 ize readily to multiple parallel processes with none, some, or all
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

Residuals auto- and cross-effects being freely estimated. Also, multiple in-
This document is copyrighted by the American Psychological Association or one of its allied publishers.

var(wan) 0.346ⴱⴱ 0.008 dicators may be used at each time point as long as measurement
var(wau) 0.174ⴱⴱ 0.004
invariance can be guaranteed (Meredith, 1993). In addition, it is
cov(wauan) 0.026ⴱⴱ 0.004
Initial measurement occasion straightforward to include predictors and random subject effects
M(ant0) 2.503ⴱⴱ 0.015 (so-called traits; cf. Oud & Jansen, 2000). Using SEM to estimate
M(aut0) 2.843ⴱⴱ 0.013 continuous time parameters allows us to make use of the full
var(ant0) 0.633ⴱⴱ 0.017 flexibility of modern latent variable models (B. O. Muthén, 2002),
var(aut0) 0.458ⴱⴱ 0.012
cov(ant0, aut0) 0.245ⴱⴱ 0.011 including a range of different estimators, different link functions
between indicators and latent constructs, or multiple group analy-
Note. Full information maximum likelihood: ⫺2 log(L) ⫽ 23,073.60. ses, to name just a few. The approach has also been extended to
an ⫽ anomia; au ⫽ authoritarianism; anau ⫽ regression of anomia on time-varying drift matrices (Oud & Jansen, 2000). Furthermore,
authoritarianism; auan ⫽ regression of authoritarianism on anomia; var ⫽
the general idea of continuous time modeling is not limited to
variance; cov ⫽ covariance.
ⴱⴱ
p ⬍ .01.

Table 3
drift matrix of the introductory example on the relationship Parameter Estimates of the Continuous Time Auto- and
between physical and social well-being, the relationship be- Cross-Effects Model for Unequal Intervals
tween drift matrix A and the autoregressive cross-lagged pa-
rameters A共⌬t i兲 is depicted in Figure 2. The only difference is Continuous time parameter
that time intervals are now in years rather than months. Au- Parameter estimate SE
thoritarianism is represented by construct S and anomia by Drift matrix (A)
construct P. Figure 2A shows the autoregressive coefficients of Auto-effects
anomia and authoritarianism, Figure 2B the cross-lagged ef- aanan ⫺0.447ⴱⴱ 0.020
fects, and Figure 2C the expected values of authoritarianism aauau ⫺0.117ⴱⴱ 0.009
Cross-effects
and anomia. For ⌬t i ⫽ 0, the values in Figure 2C correspond aanau 0.232ⴱⴱ 0.018
to the descriptive means of the first measurement occasion in aauan 0.043ⴱⴱ 0.010
the discrete time model (cf. Equation 12). Probably most strik- Continuous time intercepts (b)
ing is the effect of the choice of time interval on the autore- ban 0.536ⴱⴱ 0.046
gressive effects: Based on the discrete time parameters, anomia bau 0.220ⴱⴱ 0.022
Diffusion matrix (Q)
and authoritarianism would both appear to be fairly stable qanan 0.473ⴱⴱ 0.016
constructs in a study with time intervals of 1 year. In contrast, qauau 0.154ⴱⴱ 0.004
in a study based on 6-year intervals, the stability of anomia qanau ⫽ qauan ⫺0.005 (ns) 0.005
would be expected to be close to zero (aanan ⫽ 0.06), whereas Initial measurement occasion
M(ant0) 2.503ⴱⴱ 0.015
the autoregressive effect of authoritarianism would still be M(aut0) 2.843ⴱⴱ 0.013
substantial (aauau ⫽ 0.55; see Figure 2A). Likewise, in a study var(ant0) 0.633ⴱⴱ 0.017
with half-year intervals one would likely conclude that neither var(aut0) 0.458ⴱⴱ 0.012
of the two constructs has a strong effect on the other, while a cov(ant0, aut0) 0.245ⴱⴱ 0.011
study with a time interval of 4 years lends strong support to the Model fita
⫺2 log(L) 23,415.93
hypothesis that authoritarianism causes anomia (see Figure 2B). df 13,361
Without information on continuous time parameters one would
be left with such contradictory results. By knowing the under- Note. Time intervals are ⌬t1 ⴝ ⌬t2 ⴝ 1 and ⌬t3 ⴝ ⌬t4 ⴝ 2. an ⫽ anomia;
lying drift matrix, however, it becomes readily apparent that au ⫽ authoritarianism; anau ⫽ effect of authoritarianism on anomia;
auan ⫽ effect of anomia on authoritarianism; var ⫽ variance; cov ⫽
authoritarianism is not only a more stable construct (– 0.117 vs. covariance.
– 0.447) but also has a stronger effect on anomia (0.232) than a
Full information/raw data maximum likelihood estimation.
ⴱⴱ
the other way around (0.043). Thus, our results support the p ⬍ .01.
188 VOELKLE, OUD, DAVIDOV, AND SCHMIDT

standard (vector) autoregressive and cross-lagged models, as used gressive and cross-lagged models being two of the most well
in this article, but applies to most longitudinal models in the social known representatives. Unfortunately, these methods consider
sciences that consider time only implicitly by accounting for the time only implicitly, by accounting for the order of measurement
order of measurement occasions but not for the length of the occasions but not for the length of the time intervals between them.
intervals between them. To some degree this is also true for hybrid As illustrated by three brief examples at the beginning of the
models, such as the autoregressive latent trajectory model (Bollen article, this is highly problematic. First, parameters of standard
& Curran, 2004; Curran & Bollen, 2001; Delsing & Oud, 2008) or autoregressive models cannot be compared across studies with
growth curve models with time-varying covariates (Bollen & different time intervals. Second, it is difficult to estimate and
Curran, 2006). Likewise, continuous time models are not limited to compare parameter estimates (e.g., stability coefficients) that are
panel data but apply equally to time series data of single subjects based on different time intervals within the same study. Third, in
(cf. Molenaar & Campbell, 2009). At present, most commonly cross-lagged studies, the relative size of the lagged effect of one
used psychological methods for the analysis of (individual) time variable A on another variable B, and vice versa, is highly depen-
series simply ignore the length of the time intervals between dent on the time interval (see Figure 2B). In some cases the effects
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

observations. This is particularly true when using lagged (block- may even reverse, leaving the researcher with the paradoxical
This document is copyrighted by the American Psychological Association or one of its allied publishers.

Toeplitz) covariance matrices to fit P-technique models (Cattell, situation that for one time interval A “causes” B, while for another
Cattell, & Rhymer, 1947; Molenaar & Nesselroade, 2009), dy- time interval B “causes” A. Finally, a discrete time model is
namic factor analytic models (Molenaar, 1985), or recently devel- inherently bound to the time intervals in a given study. It tells us
oped unified SEM (Gates, Molenaar, Hillary, Ram, & Rovine, little about the generating process that caused the data independent
2010; Kim, Zhu, Chang, Bentler, & Ernst, 2007). While in prin- of the specific time intervals a researcher happens to have used.
ciple all of these models can be extended to account for different Continuous time models on the basis of stochastic differential
time intervals, to our knowledge this has not yet been done. equations overcome these limitations. Although these models have
However, given that parameter estimates are inherently bound to been known for several decades, they are virtually absent from the
the length of the time intervals, future research should focus on psychological literature. Accordingly, it was the purpose of the
extending continuous time modeling to these approaches as well. present article to introduce psychologists to continuous time mod-
Finally, it should be mentioned that various approximations eling by providing a step-by-step introduction to the approach. In
have been developed to avoid the matrix exponential function in short, the idea is to take the derivative of a continuous time process
Equation 6 (cf. Bergstrom, 1988; Oud, 2007b; Oud & Delsing, with respect to time. By solving the resulting differential equation,
2010). These are all based on different approximations of the the relationship between discrete and continuous time parameters
power series expansion in Equation 8. For example, the intuitive can be computed. Knowing this relationship, it becomes possible
approach used at the beginning of the article to introduce the basic to constrain the parameters of a discrete time model for any
idea of continuous time modeling is one such approximation— arbitrary ⌬ti to the underlying continuous time parameters when
albeit not a good one. One advantage of the approximations is that estimating the model. This way one obtains both (discrete and
many of them can be implemented in standard SEM packages (we continuous time) parameter sets directly during estimation. Once
provide some example Mplus [L. K. Muthén & Muthén, 1998 – the continuous time parameters are known, we can easily solve all
2010] code using an approximate approach in the online Supple- of the problems mentioned earlier.
mental Materials). Given that the OpenMx syntax provided along Although there are different ways to formulate and estimate
with this article is free of charge and offers the exact solution, this continuous time models (e.g., via filter techniques; cf. Oud &
advantage seems negligible. However, in combination with a re- Singer, 2008), in this article we used SEM. With SEM we can not
cently proposed oversampling technique (Singer, 2012), the ap- only capitalize on the full flexibility of general latent variable
proximations can be an efficient way to avoid estimation problems modeling (e.g., B. O. Muthén, 2002), we also used an approach
associated with procedures based on the eigenvalue decomposition that is familiar to most psychologists. In particular, the same
of the drift matrix A. This is particularly true when working with assumptions and limitations (e.g., in terms of the number of
complex eigenvalues. In the present article we limited ourselves to subjects, number of variables, or distributional properties) apply to
asymptotically stable, nonoscillating models, that is, models with the models discussed in the present article as to any other structural
negative and real-valued eigenvalues of A. By allowing complex equation model. By minimizing Equation 17, we obtain maximum
eigenvalues, however, continuous time modeling can also be used likelihood parameter estimates as well as the likelihood of the data
to estimate (possibly coupled and/or damped) oscillating processes given the entire model [the –2 log(L) value returned by the syntax
(e.g., Oud, 2007a; Oud & Folmer, 2011; Singer, 2012; Voelkle & provided with this article corresponds to minus two times the log
Oud, in press). Last but not least, we did not consider individually of the likelihood reported by most other SEM programs]. Com-
varying time intervals. The use of oversampling to estimate oscil- paring nested models via the likelihood ratio statistic lets users
lating and nonoscillating continuous time models with individually conduct significance tests on any parameter or combination of
varying time intervals is discussed by Voelkle and Oud (in press). parameters they are interested in, as well as compute the overall
goodness-of-fit indices. This topic has been extensively discussed
Discussion and Conclusions in the literature (e.g., Bollen, 1989; Bollen & Long, 1993; Hu &
Bentler, 1998, 1999; Hu, Bentler, & Kano, 1992; Marsh, 2004).
Despite the fact that most real-world phenomena change con- The probably biggest drawback of continuous time modeling is
tinuously over time, usually few discrete measurement occasions that— compared with other methods for the analysis of change
are available to infer the underlying process. For this purpose, a currently used by psychologists—the mathematics behind it may
number of different methods have been developed, with autore- appear somewhat daunting. Granted, this is true to some degree,
CONTINUOUS TIME MODELING 189

but in the present article we showed that the basic idea underlying when we want to compare parameters that have been obtained in
continuous time modeling is actually quite simple. Furthermore, a study with ⌬ti ⫽ 1 month to parameters that have been obtained
with the relevant and freely available computer code (download- in a study with ⌬ti ⫽ 2 months, as illustrated in our first introduc-
able from the online Supplemental Materials) at hand, users do not tory example. Without inter- or extrapolating the findings of one of
have to worry about the correct implementation of the most com- the two studies to the time interval of the other, no comparisons
plicated Equations 14, 15, and 16 but may simply specify their can be made and no cumulative knowledge can be generated. In
standard discrete time structural equation model and the software other situations it seems better to avoid such comparisons from the
returns the continuous time parameter estimates. For the models beginning. For example, relating a study on emotional stability at
discussed in the present article, differences in computation times a level of minutes to a study on emotional stability over years may
are also negligible. The continuous time model in our empirical not seem like a reasonable thing to do, even though—from a
example took about 9 s to be estimated on a standard PC. How- mathematical point of view— continuous time modeling would
ever, due to the complex (matrix exponential) parameter con- allow us to do so.
straints, the optimization process is more susceptible to noncon- What does this all mean for applied quantitative research in
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

vergence compared with simpler structural equation models. As psychology? Science progresses by cumulating evidence. For this
This document is copyrighted by the American Psychological Association or one of its allied publishers.

optimizers in SEM packages are continuously being improved, in purpose, it is crucial to be able to compare findings of studies that
the near future this may no longer be a problem. For the time investigate the same phenomenon but use different time intervals.
being, however, finding a converging model by using the direct Likewise, it must be possible to compare parameter estimates that
exact approach depends on good starting values. Should users were obtained at different time intervals within the same study.
experience problems in finding a converging model, we recom- With conventional standard autoregressive and cross-lagged mod-
mend fitting a discrete time model first and applying the simple els—which belong to the most widely used longitudinal research
Equation 4 to derive starting values, which should usually suffice. methods in psychology—this is not possible. Continuous time
Another alternative is the use of oversampling (Singer, 2012), with modeling overcomes these limitations. We hope that the present
which, in our experience, starting values are less of a problem. The introduction stimulates researchers to apply the approach to their
use of oversampling to estimate coupled (damped and undamped) own data and thus help to produce cumulative knowledge.
oscillators, as well as individually varying time intervals, is dis-
cussed in Voelkle and Oud (in press). References
To illustrate the approach, we provided an empirical example on Alba, R., Schmidt, P., & Wasmer, M. (Eds.). (2004). Germans or foreign-
the relationship between authoritarianism and anomia. In the ex- ers? Attitudes toward ethnic minorities in post-reunification Germany.
ample, two competing theories on the relationship between the two New York, NY: Palgrave Macmillan.
constructs were compared. As expected, anomia and authoritari- Altemeyer, B. (1996). The authoritarian specter. Cambridge, MA: Harvard
anism were both found to be fairly stable over time, with author- University Press.
itarianism showing a slightly stronger (i.e., less negative) auto- Arnold, L. (1974). Stochastic differential equations. New York, NY:
effect than anomia (– 0.117 vs. – 0.447). In addition, we found that Wiley.
Bergstrom, A. R. (1984). Continuous time stochastic models and issues of
although there was a small but significant continuous time effect of
aggregation over time. In Z. Griliches & M. D. Intriligator (Eds.),
anomia on authoritarianism (0.043), the effect of authoritarianism Handbook of econometrics (Vol. 2, pp. 1145–1212). Amsterdam, the
on anomia was much larger (0.232), lending support to the theory Netherlands: Elsevier Science.
of McClosky and Schaar (1965). Bergstrom, A. R. (1988). The history of continuous-time econometric
Let us finish the article with a word of caution and a more models. Econometric Theory, 4, 365–383.
general comment. First the word of caution: When should discrete Boker, S. M., Neale, M., Maes, H., Wilde, M., Spiegel, M., Brick,
time analysis be preferred over continuous time analysis? From a T., . . . Fox, J. (2011). OpenMx: An open source extended structural
mathematical point of view, the quick answer to this question is equation modeling framework. Psychometrika, 76, 306 –317.
never. The continuous time model contains exactly the same Bollen, K. A. (1989). Structural equations with latent variables. New
information as the discrete time model and more. It accounts not York, NY: Wiley.
Bollen, K. A., & Curran, P. J. (2004). Autoregressive latent trajectory
only for the order of measurement occasions but also for the time
(ALT) models: A synthesis of two traditions. Sociological Methods &
intervals between them. Thus, knowing the continuous time pa- Research, 32, 336 –383. doi:10.1177/0049124103260222
rameters, it is easy to reconstruct the discrete time parameters. The Bollen, K. A., & Curran, P. J. (2006). Latent curve models: A structural
opposite is not true: Knowing the discrete time parameters may tell equation perspective. Hoboken, NJ: Wiley.
us little about the underlying continuous time model (i.e., the Bollen, K. A., & Long, S. J. (Eds.). (1993). Testing structural equation
generating process). From an applied perspective, however, it may models. Newbury Park, CA: Sage.
sometimes be that there is no point in interpreting continuous time Cattell, R. B., Cattell, A. K. S., & Rhymer, R. M. (1947). P-technique
parameters. This may be the case because the process actually demonstrated in determining psychophysical source traits in a normal
develops in discrete time steps and/or because the length of time individual. Psychometrika, 12, 267–288. doi:10.1007/BF02288941
intervals (within and across studies) does not vary and— being a Curran, P. J., & Bollen, K. A. (2001). The best of both worlds: Combining
autoregressive and latent curve models. In L. Collins & A. G. Sayer
constant— contains no information. In these situations one may as
(Eds.), New methods for the analysis of change (pp. 107–135). Wash-
well use a discrete time model. More importantly, the user must be ington, DC: American Psychological Association.
careful when using continuous time parameters to inter- or extrap- Delsing, M. J. M. H., & Oud, J. H. L. (2008). Analyzing reciprocal
olate to discrete time points that have not been observed. No matter relationships by means of the continuous-time autoregressive latent
which statistical method is being used, this is always a dangerous trajectory model. Statistica Neerlandica, 62, 58 – 82. doi:10.1111/j.1467-
thing to do. Sometimes we have no other option, for example, 9574.2007.00386.x
190 VOELKLE, OUD, DAVIDOV, AND SCHMIDT

Duncan, T. E., Duncan, S. C., & Strycker, L. A. (2006). An introduction to multivariate time series. Psychometrika, 50, 181–202. doi:10.1007/
latent variable growth curve modeling: Concepts, issues, and applica- BF02294246
tions (2nd ed.). Mahwah, NJ: Erlbaum. Molenaar, P. C. M., & Campbell, C. G. (2009). The new person-specific
Finkel, S. E. (1995). Causal analysis with panel data. Thousand Oaks, CA: paradigm in psychology. Current Directions in Psychological Science,
Sage. 18, 112–117. doi:10.1111/j.1467-8721.2009.01619.x
Gates, K. M., Molenaar, P. C. M., Hillary, F. G., Ram, N., & Rovine, M. J. Molenaar, P. C. M., & Nesselroade, J. R. (2009). The recoverability of
(2010). Automatic search for fMRI connectivity mapping: An alternative P-technique factor analysis. Multivariate Behavioral Research, 44, 130 –
to Granger causality testing using formal equivalences among SEM path 141. doi:10.1080/00273170802620204
modeling, VAR, and unified SEM. NeuroImage, 50, 1118 –1125. doi: Muthén, B. O. (2002). Beyond SEM: General latent variable modeling.
10.1016/j.neuroimage.2009.12.117 Behaviormetrika, 29, 81–117. doi:10.2333/bhmk.29.81
Granger, C. W. J. (1969). Investigating causal relations by econometric Muthén, L. K., & Muthén, B. O. (1998 –2010). Mplus users’s guide (6th
models and cross-spectral methods. Econometrica, 37, 424 – 438. doi: ed.). Los Angeles, CA: Muthén & Muthén.
10.2307/1912791 Oud, J. H. L. (2007a). Comparison of four procedures to estimate the
Hamerle, A., Nagl, W., & Singer, H. (1991). Problems with the estimation damped linear differential oscillator for panel data. In K. van Montfort,
of stochastic differential equations using structural equation models. J. H. L. Oud, & A. Satorra (Eds.), Longitudinal models in the behavioral
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.
This document is copyrighted by the American Psychological Association or one of its allied publishers.

Journal of Mathematical Sociology, 16, 201–220. doi:10.1080/ and related sciences (pp. 19 –39). Mahwah, NJ: Erlbaum.
0022250X.1991.9990088 Oud, J. H. L. (2007b). Continuous time modeling of reciprocal relation-
Heitmeyer, W. (Ed.). (2004). Deutsche Zustände, Folge 3 [Current state in ships in the cross-lagged panel design. In S. M. Boker & M. J. Wenger
Germany, Series 3]. Frankfurt am Main, Germany: Suhrkamp. (Eds.), Data analytic techniques for dynamical systems (pp. 87–129):
Hertzog, C., & Nesselroade, J. R. (2003). Assessing psychological change Mahwah, NJ: Erlbaum.
in adulthood: An overview of methodological issues. Psychology and Oud, J. H. L., & Delsing, M. J. M. H. (2010). Continuous time modeling
Aging, 18, 639 – 657. doi:10.1037/0882-7974.18.4.639 of panel data by means of SEM. In K. van Montfort, J. H. L. Oud, & A.
Hu, L.-T., & Bentler, P. M. (1998). Fit indices in covariance structure Satorra (Eds.), Longitudinal research with latent variables (pp. 201–
modeling: Sensitivity to underparameterized model misspecification. 244). New York, NY: Springer.
Psychological Methods, 3, 424 – 453. Oud, J. H. L., & Folmer, H. (2011). Modeling oscillation: Approximately
Hu, L.-T., & Bentler, P. M. (1999). Cutoff criteria for fit indexes in or exactly? Multivariate Behavioral Research, 46, 985–993.
covariance structure analysis: Conventional criteria versus new alterna- Oud, J. H. L., & Jansen, R. A. R. G. (2000). Continuous time state space
tives. Structural Equation Modeling, 6, 1–55. modeling of panel data by means of SEM. Psychometrika, 65, 199 –215.
Hu, L.-T., Bentler, P. M., & Kano, Y. (1992). Can test statistics in doi:10.1007/BF02294374
covariance structure analysis be trusted? Psychological Bulletin, 112, Oud, J. H. L., & Singer, H. (2008). Continuous time modeling of panel
351–362. data: SEM versus filter techniques. Statistica Neerlandica, 62, 4 –28.
Jöreskog, K. G. (1973). A general method for estimating a linear structural R Development Core Team. (2011). R: A language and environment for statistical
equation system. In A. S. Goldberger & O. D. Duncan (Eds.), Structural computing. Vienna, Austria: R Foundation for Statistical Computing.
equation models in the social sciences (pp. 85–112). New York, NY: Raudenbush, S. W., & Bryk, A. S. (2002). Hierarchical linear models: Applica-
Seminar Press. tions and data analysis methods (2nd ed.). Thousand Oaks, CA: Sage.
Jöreskog, K. G. (1979). Statistical estimation of structural models in Rindskopf, D. (1984). Using phantom and imaginary latent variables to
longitudinal development investigations. In J. R. Nesselroade & P. B. parameterize constraints in linear structural models. Psychometrika, 49,
Baltes (Eds.), Longitudinal research in the study of behavior and devel- 37– 47. doi:10.1007/BF02294204
opment (pp. 303–352). New York, NY: Academic Press. Rogosa, D. R. (1980). A critique of cross-lagged correlation. Psychological
Kim, J., Zhu, W., Chang, L., Bentler, P. M., & Ernst, T. (2007). Unified Bulletin, 88, 245–258. doi:10.1037/0033-2909.88.2.245
structural equation modeling approach for the analysis of multisubject, Ruymgaart, P. A., & Soong, T. T. (1985). Mathematics of Kalman-Bucy
multivariate functional MRI data. Human Brain Mapping, 28, 85–93. filtering. Berlin, Germany: Springer.
doi:10.1002/hbm.20259 Scheepers, P., Felling, A., & Peters, J. (1992). Anomie, authoritarianism,
Lütkepohl, H. (2005). New introduction to multiple time series analysis. and ethnocentrism: Update of a classic theme and an empirical test.
Berlin, Germany: Springer. Politics and the Individual, 2, 43– 60.
Lutterman, K. G., & Middleton, R. (1970). Authoritarianism, anomia, and Schlueter, E., Davidov, E., & Schmidt, P. (2007). Applying autoregressive
prejudice. Social Forces, 48, 485– 492. doi:10.2307/2575572 cross-lagged and latent growth models to a three-wave panel study. In K.
Marsh, H. W. (2004). In search of golden rules: Comment on hypothesis- van Montfort, J. H. L. Oud, & A. Satorra (Eds.), Longitudinal models in
testing approaches to setting cutoff value for fit indexes and dangers in the behavioral and related sciences (pp. 315–336). Mahwah, NJ: Erl-
overgeneralizing Hu and Bentler⬘s (1999) findings. Structural Equation baum.
Modeling, 11, 320 –341. Schmidt, P., Stephan, K., & Herrmann, A. (1995). Entwicklung einer
Mayr, S., Erdfelder, E., Buchner, A., & Faul, F. (2007). A short tutorial of Kurzskala zur Messung von Autoritarismus [The development of a brief
GPower. Tutorials in Quantitative Methods for Psychology, 3, 51–59. scale for the measurement of authoritarianism]. In G. Lederer & P.
McArdle, J. J. (2009). Latent variable modeling of differences and changes Schmidt (Eds.), Autoritarismus und Gesellschaft. Trendanalysen und
with longitudinal data. Annual Review of Psychology, 60, 577– 605. vergleichende Jugenduntersuchungen 1945–1993 [Authoritarianism and
doi:10.1146/annurev.psych.60.110707.163612 society: Trend analyses and comparative youth studies 1945–1993] (pp.
McClosky, H., & Schaar, J. H. (1965). Psychological dimensions of an- 221–227). Opladen, Germany: Leske & Budrich.
omy. American Sociological Review, 30, 14 – 40. doi:10.2307/2091771 Singer, H. (1990). Parameterschätzung in zeitkontinuierlichen dynamis-
Meredith, W. (1993). Measurement invariance, factor analysis and factorial chen Systemen [Parameter estimation in continuous time dynamic sys-
invariance. Psychometrika, 58, 525–543. tems]. Konstanz, Germany: Hartung-Gorre.
Merton, R. K. (1949). Social structure and anomie: Revisions and exten- Singer, H. (1998). Continuous panel models with time dependent param-
sions. In R. Anshen (Ed.), The family (pp. 226 –257). New York, NY: eters. Journal of Mathematical Sociology, 23, 77–98. doi:10.1080/
Harper Brothers. 0022250X.1998.9990214
Molenaar, P. C. M. (1985). A dynamic factor model for the analysis of Singer, H. (2012). SEM modeling with singular moment matrices Part II: ML-
CONTINUOUS TIME MODELING 191

estimation of sampled stochastic differential equations. Journal of Mathematical improvement in measurement invariance methods and procedures.
Sociology, 36, 22–43. doi:10.1080/0022250X.2010.532259 Organizational Research Methods, 5, 139 –158. doi:10.1177/
Srole, L. (1956). Social integration and certain corollaries: An exploratory study. 1094428102005002001
American Sociological Review, 21, 709–716. doi:10.2307/2088422 Voelkle, M. C., & Oud, J. H. L. (in press). Continuous time modeling
Stenner, K. (1997). Societal threat and authoritarianism: Racism, intoler- with individually varying time intervals for oscillating and nonoscil-
ance, and punitiveness in America, 1960 –1994. Ann Arbor, MI: UMI. lating processes. British Journal of Mathematical and Statistical
Vandenberg, R. J. (2002). Toward a further understanding of and Psychology.

Appendix A

Solution of Equation 5
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

冋 册
This document is copyrighted by the American Psychological Association or one of its allied publishers.

dx共t兲
Proof that the unique solution of Equation 5 ⫽ Ax共t兲 for initial value x共t0 兲 ⫽ x0 over any time
dt
interval 共⌬t ⫽ t ⫺ t0 兲 between x共t0 兲 and x共t兲 is given by Equation 6 [x共t兲 ⫽ eA·共t⫺t0兲 x共t0 兲].
The proof consists of first showing that the derivative of Equation 6 is Equation 5 and, second, that taking
the initial value in Equation 6 gives x共t0 兲 ⫽ x0 .
1. Taking the derivative of Equation 6 gives

dx共t兲
⫽ AeA·共t⫺t0兲 x共t0 兲 ⫽ Ax共t兲.
dt
The first rewrite in this equation follows because by definition



共At兲k 1 1
e A·t
⫽ ⫽ I ⫹ At ⫹ 共At兲2 ⫹ 共At兲3 ⫹ . . .
k! 2! 3!
k⫽0

and therefore

deA·t
dt
1
2!
1
3!
1
2!
1

⫽ 0 ⫹ A ⫹ A2 t ⫹ A3 t2 ⫹ A4 t3 ⫹ . . . ⫽ A I ⫹ At ⫹ 共At兲2 ⫹ 共At兲3 ⫹ . . . .
3! 冊
The second rewrite applies Equation 6.
2. For t ⫽ t0 , Equation 6 gives

eA·0 x共t0 兲 ⫽ x共t0 兲 ⫽ x0 .

Appendix B

Relationship Between Discrete and Continuous Time Effects

Setting Equation 6 关x共t兲 ⫽ eA·共t⫺t0兲 x共t0 兲兴 equal to Equation 2 [x共ti兲 ⫽ A共⌬ti兲x共ti ⫺ ⌬ti兲] for any arbitrary
time point x共t兲 ⫽ x共ti兲, initial time point x共t0 兲 ⫽ x共ti ⫺ ⌬ti兲, and time interval 共t ⫺ t0 兲 ⫽ ⌬ti, we obtain

eA·⌬ti x共t0 兲 ⫽ A共⌬ti兲x共t0 兲,


then becoming Equation 7:

A共⌬ti兲 ⫽ eA·⌬t i.

(Appendices continue)
192 VOELKLE, OUD, DAVIDOV, AND SCHMIDT

Appendix C

Relationship Between Discrete and Continuous Time Error Covariance Matrix

Taking the integral of the error part in Equation 10 G 冋 dW共t兲


dt 册
yields

冕 t

t0
G
dW共s兲
ds
ds ⫽ 冕t

t0
GdW共s兲.

The variable of integration s is only symbolic and replaces t in order not to confuse it with the upper limit
of integration. Unfortunately, the integral cannot be defined as an ordinary Riemann integral but can be solved
as a Wiener stochastic integral or, more generally, as an Itô stochastic integral with its own rules of integration
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.

(e.g., see Arnold, 1974, pp. 128 –134). Note that G is independent of time in the differential equation
This document is copyrighted by the American Psychological Association or one of its allied publishers.

(Equation 10), but its effect varies over the interval in the solution. In particular, it is multiplied by the matrix
exponential derived earlier. Thus we write

冕t

t0
eA·共t⫺s兲 GdW共s兲.

If the drift matrix is zero 共A ⫽ 0兲, this reduces to

冕 t

t0
GdW共s兲.

As it is usually the case (e.g., in any structural equation model), the covariance matrix of the error terms
corresponds to the expected value of the outer product of the error vectors. In our case this would correspond
to

cov
冋冕t

t0
册 冋冉冕
eA·共t⫺s兲 GdW共s兲 ⫽ E
t0
t
eA·共t⫺s兲 GdW共s兲 · 冊 冉冕t

t0
冊册
eA·共t⫺s兲 GdW共s兲 ⬘ .

The expectation of the product of the two integral forms can be written as one integral form, so that the error
covariance matrix corresponds to

冕t

t0
eA·共t⫺s兲 QeA⬘·共t⫺s兲 ds ⫽ irow A⫺1
# 再 冋
eA #·共t⫺t0兲 ⫺ I rowQ 册 冎
for Q ⫽ GG⬘ and A# ⫽ A 䊟 I ⫹ I 䊟 A. Mathematical details can be found in Arnold (1974, pp. 66 – 67)
and Singer (1990).

Received October 22, 2009


Revision received October 13, 2011
Accepted January 24, 2012 䡲
This article is intended solely for the personal use of the individual user and is not to be disseminated broadly.
This document is copyrighted by the American Psychological Association or one of its allied publishers.

Correction to Voelkle, Oud, Davidov, and Schmidt (2012)

In the article “An SEM Approach to Continuous Time Modeling of Panel Data: Relating Author-
itarianism and Anomia,” by Manuel C. Voelkle, Johan H. L. Oud, Eldad Davidov, and Peter
Schmidt (Psychological Methods, Advance online publication. April 9, 2012. doi:10.1037/
a0027543), the supplemental materials link was missing. All versions of this article have been
corrected.

DOI: 10.1037/a0029251

You might also like