bayesian data analysis with the bivariate hierarchical ... · bayesian data analysis with the...

15
Full Terms & Conditions of access and use can be found at http://www.tandfonline.com/action/journalInformation?journalCode=hmbr20 Download by: [KU Leuven University Library] Date: 22 February 2016, At: 08:08 Multivariate Behavioral Research ISSN: 0027-3171 (Print) 1532-7906 (Online) Journal homepage: http://www.tandfonline.com/loi/hmbr20 Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx & Joachim Vandekerckhove To cite this article: Zita Oravecz, Francis Tuerlinckx & Joachim Vandekerckhove (2016) Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model, Multivariate Behavioral Research, 51:1, 106-119, DOI: 10.1080/00273171.2015.1110512 To link to this article: http://dx.doi.org/10.1080/00273171.2015.1110512 View supplementary material Published online: 16 Feb 2016. Submit your article to this journal Article views: 13 View related articles View Crossmark data

Upload: others

Post on 19-Feb-2020

21 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

Full Terms & Conditions of access and use can be found athttp://www.tandfonline.com/action/journalInformation?journalCode=hmbr20

Download by: [KU Leuven University Library] Date: 22 February 2016, At: 08:08

Multivariate Behavioral Research

ISSN: 0027-3171 (Print) 1532-7906 (Online) Journal homepage: http://www.tandfonline.com/loi/hmbr20

Bayesian Data Analysis with the BivariateHierarchical Ornstein-Uhlenbeck Process Model

Zita Oravecz, Francis Tuerlinckx & Joachim Vandekerckhove

To cite this article: Zita Oravecz, Francis Tuerlinckx & Joachim Vandekerckhove (2016) BayesianData Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model, MultivariateBehavioral Research, 51:1, 106-119, DOI: 10.1080/00273171.2015.1110512

To link to this article: http://dx.doi.org/10.1080/00273171.2015.1110512

View supplementary material

Published online: 16 Feb 2016.

Submit your article to this journal

Article views: 13

View related articles

View Crossmark data

Page 2: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH, VOL. , NO. , –http://dx.doi.org/./..

Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck ProcessModel

Zita Oravecza, Francis Tuerlinckxb, and Joachim Vandekerckhovec

aThe Pennsylvania State University; bUniversity of Leuven; cUniversity of California, Irvine

KEYWORDSIntensive longitudinal dataanalysis; dynamicalmodeling;Ornstein-Uhlenbeck;Bayesian modeling;individual differences

ABSTRACTIn this paper, we propose amultilevel processmodeling approach to describing individual differencesin within-person changes over time. To characterize changes within an individual, repeatedmeasuresover time are modeled in terms of three person-specific parameters: a baseline level, intraindividualvariation around the baseline, and regulatory mechanisms adjusting toward baseline. Variation dueto measurement error is separated from meaningful intraindividual variation. The proposed modelallows for the simultaneous analysis of longitudinal measurements of two linked variables (bivariatelongitudinal modeling) and captures their relationship via two person-specific parameters. Relation-ships between explanatory variables and model parameters can be studied in a one-stage analysis,meaning that model parameters and regression coefficients are estimated simultaneously. Mathe-matical details of the approach, including a description of the core process model—the Ornstein-Uhlenbeck model—are provided. We also describe a user friendly, freely accessible software programthat provides a straightforward graphical interface to carry out parameter estimation and inference.The proposed approach is illustrated by analyzing data collected via self-reports on affective states.

Introduction

Recent advances in social science data collection strate-gies have led to a proliferation of data sets that consistof long chains of longitudinal measurements taken fromdifferent persons. For example, the widely used meth-ods of experience sampling (Csikszentmihalyi & Larson,1987), or the more general ecological momentary assess-ments (Stone & Shiffman, 1994), provide researcherswith a wide variety of measurements in natural set-tings. Such data often require complex statistical analy-ses. A new field, called intensive longitudinal data analy-sis (ILD; see, e.g., Mehl & Conner, 2012; Walls & Schafer,2006) has emerged to meet this demand. Its strategiesfocus on analyzing temporal data of several participantswith an emphasis on capturing interindividual variationsin terms of parameters describing intraindividual vari-ability. Unpacking underlying characteristics and pro-cesses related to intraindividual variability has crucialimportance in many domains, including developmen-tal research (see, e.g., Ram & Gerstorf, 2009), person-ality psychology, and emotion research (see, e.g., Kup-pens, Oravecz, & Tuerlinckx, 2010; Kuppens, Stouten, &Mesquita, 2009).

CONTACT Zita Oravecz [email protected] The Pennsylvania State University, Health and Human Developments Building, State College, PA .Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/hmbr

Supplemental data for this article can be accessed on the publisher’s website.

Wepropose characterizing longitudinalmeasurementsfrom an individual in terms of parameters of theOrnstein-Uhlenbeck (OU; Oravecz & Tuerlinckx, 2011;Oravecz, Tuerlinckx, & Vandekerckhove, 2011; Uhlen-beck & Ornstein, 1930) process. Using a univariate OUprocess to model within-person change over time in onelongitudinal variable enables us to describe dynamic char-acteristics such as intraindividual variation and dynamicstability maintenance processes, such as regulation andadaptation. Extending this framework to two longitudinalvariables within person, a bivariate OU process can addi-tionally capture coupled within-person variation in termsof cross-effect parameters, such as covariation.

Consider an experience-sampling study that aims tostudy affective instability (see, e.g., the study describedin the Application section). In such studies, participants(commonly more than 20) are semirandomly prompted,for example, through a mobile app to report their lev-els of arousal (activation) and valence (pleasantness) atthe moment of the prompt in the midst of their dailyactivities. This integral blend of pleasure and arousal isoften labeled core affect (Russell, 2003) in the emotion lit-erature, and change over time in terms of self-reportedcore affect has been the focus of several studies (see, e.g.,

© Taylor & Francis Group, LLC

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 3: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH 107

Barrett, 2004; Kuppens, Tuerlinckx, Russell, & Barrett,2013). Recently, neural correlates of core affect have alsobeen found (Wilson-Mendenhall, Barrett, & Barsalou,2013).

The resulting data set would naturally have differentnumbers of observations per participant, taken at dif-ferent timepoints (unbalanced, unequally spaced data).Commonly usedmodels to analyze these data often resortto models that assume equal spacing of measurements;for example, discrete time series models (Bolger & Lau-renceau, 2013; Walls & Schafer, 2006). Not only can dis-cretization bias inference (Delsing, Oud, & Bruyn, 2005),but it turns out that unequally sampling designs mightbe more advantageous when the sampling rate is low(Voelkle &Oud, 2013), whichmight occur in experience-sampling studies when researchers try to decrease bur-den on the participants. The proposedOUmodel assumesthat the underlying change mechanisms behind the coreaffect self-reports take place in continuous time, and theunequally spaced and unbalanced observed data are sam-ples from this process; therefore, it is especially well fitfor analyzing intensive longitudinal data fromexperience-sampling studies.

The latent OU process we propose for modelingobserved data can be characterized in terms of the follow-ing parameters: Each participant can be described with abaseline (which is a baseline core affect in our example)and regulatory mechanisms with different levels of inten-sity to regulate toward this baseline. Around this base-line people exhibit different levels of intraindividual vari-ation, which in the proposed model is separated frommeasurement error variation through state space model-ing (Fahrmeir & Tutz, 2001). That is to say that the OUframework allows us to decomposemanifest variability inthe raw self-reports of pleasantness and activation scoresinto psychologically meaningful parameters. Psychologi-cally meaningful means that our model parameters canbe interpreted as directly representing psychological con-cepts such as intraindividual affect variation and regula-tion. Moreover, synchronicity in changes between activa-tion and pleasantness levels along with concurrence inregulatory dynamics are captured through two person-specific model parameters. The baseline, variation, regu-latory mechanisms, and synchronicity are parameters ofthe OUmodel and can be considered as meaningful indi-cators of affective system quality.

In the proposed OU framework, all of these indica-tors describing the within-person change can be made afunction of time-invariant covariates (TICs). These canbe any explanatory variables that are considered relativelystable over time. For example, we hypothesized that aperson’s tendency for rumination (for its measurement,see Trapnell & Campbell, 1999) might be connected tothe self-regulation parameter of our model; therefore, we

regressed this parameter (and other model parametersas well) on this covariate. Moreover, the baseline levelsof pleasantness and activation for each individual can beadjusted as a function of time-varying covariates (TVC),therefore allowing for example adaptationmechanisms toenter into the model. For example, actual measurementtime can be turned into a TVC to investigate whether (andhow) the baseline changes over time. To conclude, theproposed framework enables the researcher to approach amultifaceted substantive problemwith a realistic and nec-essary level of complexity.

The proposed model also carries several desirablecharacteristics from the statistical inference perspective.Most important, inference in all parameters is performedin a single step. Traditional approaches (such as the max-imum likelihood framework) routinely derive point esti-mates for parameters (e.g., for intraindividual variance)and then in a second step link these point estimatesto covariates (e.g., regressing intraindividual variation inaffect on neuroticism scores). This approach is problem-atic as relying on point estimates neglects the error thatis in the parameter estimates. In the one-step approach,process model parameters, regression terms, and errorterms are all estimated simultaneously, providing a prin-cipled way of propagating error in the parameter esti-mates. Finally, implementing parameter estimation in theBayesian statistical framework results in probability dis-tributions for each model parameter. This allows us toevaluate likely values of the model parameters in proba-bilistic terms, such as how likely it is that a parameter islarger than 0 or that it is within a certain range.

The Ornstein-Uhlenbeck process model shows corre-spondence to othermodeling techniques often utilized formodeling for ILD. It is similar to the traditional bivariatelinear mixed models (LMM; see, e.g., MacCallum, Kim,Malarkey, & Kiecolt-Glaser, 1997) in the sense that themean structure (baseline for OU) can be made a func-tion of TICs and TVCs. However, there are several dis-tinctions as well. First, the proposed bivariate OU modelassumes that dynamics occur on the latent level; therefore,the intraindividual variation and autocorrelation struc-ture are modeled on the latent level. Moreover, whereasan autoregressive error term can be added to LMMs, theyare typically not allowed to vary across individuals nor canthey be made a function of covariates. The OU model’sself-regulation parameter controls for the autocorrelationin the changes and can be made person specific. Besidesthe correspondence with the LMM framework, the OUmodel falls into the class of dynamical models termed asstochastic differential equations (SDEs; see, e.g., Chow,Ferrer, & Nesselroade, 2007; Molenaar & Newell, 2003;Oud, 2007; Oud & Jansen, 2000), which extend ordinarydifferential equations (ODEs; e.g., the oscillator model:Chow, Ram, Boker, Fujita, & Clore, 2005; the reservoir

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 4: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

108 Z. ORAVECZ ET AL.

model: Deboeck & Bergeman, 2013) in allowing for pro-cess noises or uncertainties in how the latent processeschange over time. When compared to oscillatory modelsof change, the OU model does not reinforce an oscilla-tory pattern on the dynamics itself but can include cyclicchanges in baseline levels (as TVCs), while at the sametime capturing stochastic variation that is separate frommeasurement noise. The proposed modeling frameworkalso extends other current SDE models in the psycho-metric literature or their discrete-time difference equa-tion and time series counterparts (e.g., Browne & Nessel-roade, 2005; Chow, Ho, Hamaker, & Dolan, 2010; Wang,Hamaker, &Bergeman, 2012) by also allowing for randomeffects, or between-person variations, in terms of mean-ingful process model parameters (baseline, intraindivid-ual variation, self-regulation, and synchronicity), as wellas ways to incorporate the effects of time-varying covari-ates on these parameters.

Besides expounding on the technical account of theproposed model and illustrating its advantages, we aim toprovide guidelines on how model fitting can be done in apractical sense. This stems from the recognition that ourproposed approach does not represent the mainstream ofmethods used in the field of intensive longitudinal dataanalysis. We introduce the basic notions of Bayesian dataanalysis, including posterior predictivemodel checks. Theresearch tool to carry out inference, the Bayesian hier-archical Ornstein-Uhlenbeck modeling (BHOUM) pro-gram will also be discussed. BHOUM is a user-friendlyparameter estimation engine with a graphical user inter-face. It is a stand-alone program and can be downloaded(optionally with its MATLAB source code) from the firstauthor’s website (www.zitaoravecz.net). The focus of thisarticle is on carrying out the data analysis and the inter-pretation of the results, and we further refer prospectiveusers to the detailed User’s Guide on the BHOUM soft-ware (available on the journal’s website as supplementalmaterial).

Investigating temporal dynamics in terms of processmodel parameters has potential applications in manyareas. One example is core affect, defined above. Kuppenset al. (2010) formulated the DynAffect theory that linkedgeneral characteristics of core affect changes to Ornstein-Uhlenbeck process parameters such as attraction point orbaseline, intraindividual variation, and regulatory force.We chose the DynAffect framework to demonstrate dataanalysis with the hierarchical OU model, and we willreanalyze data from Kuppens et al. (2010), Study 1. Ourapproach goes several steps further than the original anal-ysis as we will introduce time-varying and time-invariantcovariates in a one-stage analysis. Moreover, we studythe cross-effect parameters of the two dimensions of coreaffect.

The bivariate hierarchical Ornstein-Uhlenbeckmodel

Model specification: The bivariateOrnstein-Uhlenbeck state spacemodel

The core of the proposed model is the Ornstein-Uhlenbeck stochastic process, first described by twoDutch scientists Leonard Ornstein and George EugeneUhlenbeck (Uhlenbeck & Ornstein, 1930). The OU pro-cess characterizes the within-person dynamics on thelatent level; for example, the underlying changes in one’score affect. The process can be seen as a continuoustime analogue of a discrete-time first-order autoregressive(AR1) process. In an AR1 process we regress the currentposition of the process on its previous position one timeunit earlier. Similarly, the current position of the OU pro-cess depends on its previous position, but instead of onetime unit of difference, the elapsed time between the twopositions can take any positive value. This idea is math-ematically formalized by a differential equation, whichdescribes the rate of change in the process level over anychosen amount of time. The OU process is also perturbedby some noise; therefore, its mathematical formula is astochastic differential equation, defined as follows:

d�(t ) = B(μ − �(t ))dt + �dW(t ) (1)

Equation (1) is referred to as the “dynamics (or state)equation” in the state-space modeling framework. Letus expound on Equation (1): �(t ) (2 × 1) is a two-dimensional latent random variable; for example, levels ofpleasantness and activation at time t; d�(t ) (2 × 1) rep-resents the change in these levels with respect to time t ,and the right side of the equation shows that this is partlydetermined by the distance between the current positionof the process �(t ) (2 × 1), from the baseline, denotedby μ (2 × 1).1 The level of self-regulation is expressedthrough the 2 × 2 regulatory force (or drift or dampen-ing) matrix, B. The other factor governing change in thelatent process is the second term on the right side of Equa-tion (1),�dW(t ), which represents the stochastic compo-nent of the process. The termW(t ) stands for the positionof a Wiener process (also known as Brownian motion)at time t : this process evolves in continuous time, and itsposition is uninfluenced by its previous positions, mean-ing that it follows a random trajectory. Practically speak-ing, the dW(t ) term adds random variation (noise) to thesystem. Finally, the effect of this is scaled by the diffusionmatrix � (2 × 2) (details follow).

As a result of the mean-reverting specification (as shown in Equation []) theOUprocess does not have an ever-expanding variance expectation as in basicrandomwalk processes.

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 5: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH 109

Integrating over the transition equation, Equation (1),results in the position equation:

�(t ) = e−Bm�(t − m) + μ(1 − e−Bm)

+�e−Bm∫ t

t−meBudW(u),

which shows the position of the process after elapsedtime m. The last term on the left-hand side is a so-calledstochastic integral. Stochastic integrals cannot be solvedby regular calculus, but require special approaches, suchas the Ito calculus. The Ito calculus extends the methodsof regular calculus to the domain of stochastic processes.The solution in our case (Dunn & Gipson, 1977) leads tothe following equation:

�(t + m) | �(t ) ∼ N2(μ + e−Bm(�(t ) − μ),

� − e−Bm�e−BTm), (2)

resulting in an equation that describes the conditional dis-tribution of the position of the process after elapsed timem. Notation T stands for transpose of a matrix.

Equations (2) and (3) additionally feature thematrix�,which is the stationary covariancematrix of the process—that is, the variance of the process run for an infinitelylong time. � is related to the diffusion matrix � anddrift matrix B through the following equation (see, e.g.,Gardiner, 1986, p. 110):

��T = B� + �BT. (3)

Equation (3) demonstrates that the scale of the diffu-sion process can be partitioned into a dampening con-tribution of the mean-reversion process (governed by theregulatory force matrix B) and the stationary covariance.This is particularly useful since this reparameterizationallows us to express the process in terms of psychologi-cally meaningful parameters. Finally, coupled influencesare captured by the off-diagonal elements of � (covaria-tion) and B (synchronicity in self-regulation).

As can be seen, over time the OU process continu-ously drifts toward the baseline due to themean-revertingdynamics. In addition, there is a stochastic input term thatinfluences the change trajectory. Psychological processesfor which this type of perturbation and mean reversioncan be an appropriate model include emotion, mood andaffect regulation (Gross, 2002), semantic foraging (i.e.,search in semantic memory; see in Hills, Jones, & Todd,2012), and so on.

The empirical measurements, Y(t ) (e.g., pleasantnessand activation self-reports), are typically discrete. There-fore, we link the observed discrete data to the continu-ous underlying state (or levels) of the process by addingsome measurement error. In the state space modelingframework, the equation describing this idea is called the“observation equation.” We map the latent dynamics to

the manifest data through the following specifications:

Y(t ) = �(t ) + ε(t ). (4)

The measurement error is represented by ε(ts), which isdistributed according to a bivariate normal distributionwith expectation (0, 0)T and covariance matrix �ε . Next,we expand this basic state space model with a hierarchicalstructure to be able to fit a multilevel OU model to inten-sive longitudinal data.

Hierarchical extension: The bivariatemultilevel OUmodel for intensive longitudinal data

A typical structure for an intensive longitudinal dataset would be the following: Longitudinal variables fora person p (p = 1, . . . , P) are measured at np time-points: tp1, tp2, . . . , tps, . . . , tp,np . We jointly model twodata points for each person p at timepoint tps, denoted asY (tps) = (Y1(tps),Y2(tps))T. The index s denotes the sthmeasurement occasion of that individual. In the hierar-chical Ornstein-Uhlenbeck (HOU)model we assume thatthese observations are functions of a latent underlyingstate denoted as �(tps) = (�1(tps), �2(tps))T and somemeasurement error.

As proposed, the underlying latent states (e.g., changesin one’s core affect) are assumed to be governed by a two-dimensional OU process. For simplicity, we use only theindices p and swhen denoting parameters or data that arerelated to the specific observation at tps. Then an HOUmodel for a single person p can be written as

Y ps = �ps + εps, (5)

where Y ps is a shorthand for Y (tps) and stands for theobserved random vector; �ps denotes the latent state (ortrue score, shorthand for �(tps)); and εps stands for themeasurement error with the distributional assumption:εps

iid∼ N2(0, �ε ). Based on Equation (2), the latent under-lying level of the bivariate process for person p at timepoint s can be written as

�ps|�p,s−1 ∼ N2(μps + e−Bp(tps−tp,s−1)(�p,s−1 − μps),

�p − e−Bp(tps−tp,s−1)�pe−BTp (tps−tp,s−1)

). (6)

Parameter μps is the person-specific bivariate baseline,which can also be referred to as home base, bivariateattractor, or attraction point, and is somewhat similarto a mean vector. It can change over time as a functionof time-varying covariates. For example, baseline valenceand arousal levels can change as function of actual timeduring the daily cycle. Variation around the baseline ismodeled through�p, which is a person-specific intraindi-vidual 2-by-2 covariance matrix. In terms of valence andarousal, the two diagonal elements represent variation in

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 6: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

110 Z. ORAVECZ ET AL.

these, while the off-diagonals capture the covariation inthe changes.

The model assumes that there is always some level ofattraction, or regulation over time toward the baselinelevel, and the dynamics of this aremodeled through the 2-by-2 regulatory force matrix Bp. Following our core affectexample, for each person arousal and valence levels can beregulated with different intensity, represented by the twodiagonal elements of the Bp matrix, and the cross effect ofthese dynamics, the off-diagonal of Bp, is also a person-specific parameter.

Finally, for the the first observation, �p1, it is assumedthat�p1 ∼ N2(μp, �p). This is because Equation (2) can-not be used for the first observation as there is no previ-ous position to condition on. Therefore, we assume thatthe first observation is simply a function of the person’sbaseline level and intraindividual variance.

To summarize, the hierarchical Ornstein-Uhlenbeckmodel represents a process modeling approach to mod-eling change in two longitudinal measures. As explained,the changes over time in the two manifest longitudinalvariables are governed by an underlying latent process(the OU process). Since we also assume that changes inthe two longitudinal variables are interlinked, this processmodel approach can be visualized by picturing a latentprocess evolving over time among the two dimensions,which are being defined by our two longitudinal variablesof interest. The actual longitudinal measurements we col-lect are samples from the current position of this latentprocess evolving in the two-dimensional space, and thesemeasurements are perturbed by measurement error. TheOU model based dynamics on the person-level (or Level1) are described by Equation (6). Given the hierarchi-cal nature of our model, bivariate longitudinal data foreach person is described by a unique set of OU processparameters, coming from joint Level 2 (population) dis-tributions. Next we expound upon the parameterizations,define Level 2 distributions, and introduce time-varyingand time-invariant covariates.

The two-dimensional baseline as a functionof time-varying and time-invariant covariates

The latent baseline levels (or attraction point parame-ter) μps can be made a function of person-specific time-varying and person-specific time-invariant covariates. Letus assume that we measure time-invariant covariate j,for person p x jp, which could be for example their ten-dency to ruminate. We can have k TICs measured, ( j =1, . . . , k), and the TICs can be collected into a vectorof length k + 1, denoted as xp = (xp0, xp1, xp2, . . . , xpk)T,with xp0 = 1. Even if there is no time-invariant covariateinformation, we assume an intercept in the model.

Regarding the time-varying aspect, suppose that wemeasure covariate z for person p, and z = 1, . . . ,E, thenthe vector zps = (zps1, . . . , zpsE )T collects all these values.No intercept is introduced in the vector zps. The index sindicates that values may change from one observationpoint to the next. A natural candidate for a TIC to regressbaseline levels of valence and arousal is the time of theself-report; for example, we expect some people to showlow levels of pleasantness and activation in the morning.

The Level 2 distribution (distribution on the “popula-tion” level) of μps with regression on the time-invariantand time-varying covariates and allowing for a person-specific random deviation can be written as follows:

μps ∼ N2(�pμzps + Aμxp, �μ

), (7)

where the covariance matrix �μ is defined as follows:

�μ =[

σ 2μ1

σμ1μ2

σμ1μ2 σ 2μ2

]. (8)

The matrices �pμ and Aμ are parameter matrices ofdimension 2E × P and 2 × (k + 1), respectively, contain-ing the regression weights for the time-varying and thetime-invariant covariates.

The intraindividual covariancematrix as a functionof time-invariant covariates

The matrix �p stands for the stochastic or intraindividual2 × 2 covariance matrix

�p =[

γ1p γ12pγ12p γ2p

]. (9)

Its diagonal elements (i.e., γ1p and γ2p) determine theintraindividual variances in the two measured longitudi-nal variables, and the off-diagonals can be decomposedinto γ12p = ργp

√γ1pγ2p, where ργp is the cross correlation

of the observations. Since the diagonal elements (γ1p andγ2p) are variances, they are constrained to be positive. Forcomputational convenience, we log transform these vari-ances so that they take values on the real line. Then wespecify their Level 2 distributions as normal distributionsof these log-transformed values. For γ1p, that is

log(γ1p

) ∼ N(xTpαγ1, σ

2γ1

).

The mean of this distribution is modeled via the prod-uct of time-invariant covariates and their correspondingregression weights. More specifically, all TICs are col-lected in the vector xTp , which has k + 1 components. Thefirst element of this vector is a constant 1, representingan intercept, and in case there are no TICs added, themean of the distribution reduces to a simple Level 2 meanthat is the same for all persons. The vector αγ1 containsthe (fixed) regression coefficients for the covariates. The

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 7: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH 111

parameter σ 2γ1is the residual variance in the random log

variance of the first dimension, after having taken thecovariates into account. If only the intercept is present inthe model, σ 2

γ1reflects the total amount of interindivid-

ual variability in the log variance of the first dimension. Asimilar logic applies in the modeling of γp2.

The cross correlation ργp is bounded between −1 and1. By taking advantage of the Fisher z transformationF(ργp ) = 1

2 log1+ργp

1−ργp, we can transform its values to the

real line:

F(ργp ) ∼ N(xTpαργ

, σ 2ργ

),

with epγ2 ∼ N(0, σ 2ργ

). The density of the original ργp

can be derived by applying a transformation-of-variablestechnique (see, e.g., Mood, Graybill, & Boes, 1974), butit is not a common density function. Again, αργ

con-tains k + 1 regression weights; xTp contains the k covari-ate values for person p with 1 for the intercept; andσ 2

ργrepresents interindividual variation in terms of cross

correlation.

The regulatory force as a function of time-invariantcovariates

The regulatory force or centralizing tendency is parame-terized by thematrixBp, which is decomposed in the samemanner as the covariance matrix �p in Equation (9), sothat it stays positive definite. Matrix Bp has to stay posi-tive definite by definition to ensure that there is always anadjustment toward the baseline and never away from it.This implies that the process is stable and stationary.

The elements of the person-specific matrix Bp areassumed to come from Level 2 distributions that aredefined in the same manner as for �p and can be madethe function of time-invariant covariates in the samemanner. This way Bp contains two centralizing tenden-cies, one for each dimension (i.e., β1p and β2p), and astandardized cross-centralizing tendency parameter (ρβp)that represents the concurrence in regulatory dynamics.These parameters control the strength and the directionof the self-regulation toward the baseline. As β1p andβ2p go toward zero (i.e., no self-regulation), the OU pro-cess approaches a Brownian motion process; that is, acontinuous-time random walk process. When the twoparameters become very large and tend toward infinity,the OU process becomes a white noise process.

Bayesian statistical inference in the HOUmodel

We implemented parameter estimation for the hierar-chical OU model by taking advantage of Bayesian sta-tistical methods. The Bayesian approach features two

main advantages in our current settings. First, parametersin this framework have probability distributions, whichoffers an intuitively appealing way of describing uncer-tainty and knowledge about the parameters. Second, thereare distinct computational advantages: the use of MarkovchainMonte Carlo (MCMC)methods sidesteps the high-dimensional integration problem over the numerous ran-dom effect distributions.

When carrying out Bayesian data analysis, we use thesestochastic numerical integration methods to sample fromthe posterior density of the parameters. The posterior den-sity is the conditional density function of the parame-ters given the data, and it is directly proportional to theproduct of the likelihood of the data (given the param-eters) and the prior distribution of the parameters.2 Theprior distribution incorporates prior knowledge about theparameters, and if there is none, it is can be set to a vague(diffuse) distribution. The BHOUM toolbox follows thisphilosophy: all priors are set to be vague. In addition,the more data one acquires, the less influential the priorbecomes on the posterior as its shape is overwhelmed bythe tighter shape of the likelihood.

Markov chain Monte Carlo methods are general-purpose algorithms for sampling from the high-dimensional posterior of the presented model. MCMCalgorithms perform iterative sampling during whichvalues are drawn from approximate distributions that areimproved in each step, in such a way that they convergeto the targeted posterior distribution. After a sufficientlylarge number of iterations, one obtains a Markov chainwith the posterior distribution as its equilibrium distri-bution, and the generated samples are random drawsfrom the posterior distribution. Summary statistics of thegenerated samples can then be used to characterize theposterior distribution (i.e., to estimate its mean, variance,mass over a certain interval, etc.) More details aboutthe Bayesian methodology and MCMC can be foundin Gelman, Carlin, Stern, and Rubin (2013) and Robertand Casella (2004). For the HOU model there is noclosed-form analytical solution for the main parametersof interest; therefore, high-dimensional numerical inte-gration is required to calculate posterior point estimates.With MCMC methods we can solve this problem whileavoiding having to explicitly calculate p(Y ).

In theBHOUMtoolbox, a specificMCMCalgorithm—the Metropolis-within-Gibbs sampler—is implementedto estimate the HOU model parameters. In this algo-rithm, alternating conditional sampling is performed: Theparameter vector is divided into subparts (a single ele-ment or a vector), and in each iteration the algorithm

Formally, p(ξ|Y ) ∝ p(Y |ξ)p(ξ), where ξ stands for the vector of all param-eters in themodel. The normalization constant, p(Y ), whereY stands for thedata, doesnot dependon theparameter and is thereforenot consideredhere.

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 8: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

112 Z. ORAVECZ ET AL.

draws a new sample from the conditional distribution ofeach subpart given all the other parameters anddata; theseconditional distributions are called full conditional distri-butions. In our application, several such Markov chainsare initiated from different starting values in order toexplore the posterior distribution and avoid local optima.The BHOUM toolbox offers a default convergence checkusing the Gelman-Rubin R̂ statistic (for more informa-tion, see Gelman, Carlin, Stern, & Rubin, 2004).

Data: Experience-sampling study on core affect

Study settings

In this section we provide a description of how to usethe BHOUM software through analyzing data from anexperience-sampling study. The corresponding data setwas collected at the University of Leuven (Belgium) andcontains repeated measurements of 79 university stu-dents’ pleasantness and activation levels (i.e., their coreaffect).

Per the principles of the experience-sampling design,measurements were made in the participants’ naturalenvironments: They carried a Tungsten E2 palmtop com-puter that was programmed to beep at semirandom timesduring waking hours over 14 consecutive days.When sig-naled by a beep, the participants were asked to mark theirposition on a 99 × 99 core affect grid with unpleasant–pleasant feelings forming the horizontal dimension andarousal–sleepiness the vertical.

Moreover, several dispositional questionnaires wereadministered to measure a range of covariates in the par-ticipants. These variables were neuroticism and extraver-sion (part of the five-factor model of personality, or BigFive; Costa & McCrae, 1992; for the current study atranslated version was used, see in Hoekstra, Ormel, &De Fruyt, 1996), positive and negative affect (PA andNA; Hoeksma, Oosterlaan, Schipper, & Koot, 1988), self-esteem (and self-esteem variability; Rosenberg, 1989),satisfaction with life (Diener, Emmons, Larsen, & Grif-fin, 1985), reappraisal and suppression (Gross & John,2003), and rumination (Trapnell & Campbell, 1999).These covariates were used as time-invariant covariates inthe analysis that follows.

Summary of the proposed data-analytical approach

Although several HOU models were fit to the data setin Kuppens et al. (2010), none of those models involvedcovariates. That is to say, so far all analyses were per-formed in two stages: OU parameters were estimated, andcorrelation coefficients (in the classical sense) were cal-culated between the person-specific Bayesian posterior

point estimates and the covariate scores from the dispo-sitional questionnaires. In the current analysis, the latentOUparameters are regressed on the time-invariant dispo-sitional measures described at the same time as the latentdynamical process model parameters are estimated. Thisway, uncertainty in the parameter estimates is directlyaccounted for in the results so that the analysis avoids gen-erated regressor bias (Pagan, 1984). In addition, as part ofthe same analysis we incorporate time-varying covariateson the baseline, thereby further improving the accuracyof the parameter estimation.

Methods: Analyzing data with the hierarchicalOUmodel

The BHOUM toolbox contains several functions to dealwith various aspects of Bayesian statistical inference.BHOUM is primarily intended to be used as a stand-alone software program (noMATLAB license is required)through a graphical user interface (GUI).3 While no cod-ing is required from the user’s part, all MATLAB scriptsare available for download.

Parameter estimation

In the current analysis we model pleasantness and acti-vation levels of 79 people from the described experience-sampling study with a hierarchical OU process. All latentprocess parameters (baseline, intraindividual variation,regulation, cross effects) are modeled as functions of10 time-invariant covariates: neuroticism, extraversion,positive affect, negative affect, self-esteem, within-personstandard deviation of self-esteem, satisfaction with life,reappraisal, suppression, and rumination. Moreover, cir-cadian rhythm in the core affect baseline is modeled interms of linear and quadratic time effects.

Running BHOUMtoolbox.exe displays a user-friendlyData reader GUI that allows the researcher to load thedata and specify which variables are chosen to be partof the analysis. For the data format needed to use theBHOUM program, please consult the supplementaryonline files.

Once the required data have been input, the user canmove to the next window (Model specifier) where modeland sampling algorithm specifications can be set. Thedefault model is the one described in the previous section.In this fully specified model, all process model parame-ters are random effects. This way, the means of the twodimensions (μ1p and μ2p), the corresponding stochastic

The stand-alone BHOUM version with the accompanying free MATLAB Com-piler Runtime (MCR) has been tested for Windows bit and bit. If usersdo not want to install MCR because they have a MATLAB license already, thatMATLAB should be run in bit mode.

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 9: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH 113

variances (γ1p and γ2p), and the cross correlation (ργp),as well as the two regulatory force (or centralizing ten-dencies: β1p and β2p) parameters and their cross effect(ρβp) are allowed to be person-specific and regressed ontime-invariant covariates if anywas previously loaded andselected in the Data reader window. Alternative mod-els are offered as well, which are simplified versions ofthe default fully person-specific HOU model. For exam-ple, the measurement error can be removed from themodel. Another option is assuming that the changes intwo dimensions are independent; that is, all ργp and ρβp

are equal to 0. By using this option we can also modelonly one longitudinal variable by inputting the same vari-able twice (i.e., choosing twice the same column namein Data reader, OU process dimensions) and then choos-ing the “Independent dimensions” option in the Modelspecifier. The program will then fit two independent one-dimensional HOU models.

The Model specifier window allows setting the prop-erties of the Markov chain Monte Carlo sampling algo-rithm. Some default values are preset, and these will pro-vide sufficient exploration of the posterior distribution inmost cases. The properties are the following: (1) num-ber of posterior samples (per chain, same for each chain),used for posterior inference, (2) length of some necessaryadaptive period (the burn-in) preceding the samples setin box (1), (3) the number of chains that are run fromdifferent starting values to explore the posterior density,and (4) thinning factor. The thinning option is primarilyimplemented for computer memory capacity considera-tions. Because of high within-chain autocorrelation, someparameter estimates might require long chains to be runto explore the posterior density. By thinning these longchains, we store only every xth value, where x equals theinput of the Thin field.

It is good practice to report the setting of these for val-ues when reporting the results of the analysis. For the cur-rent analysis, we set four chains each consisting of 3,000iterations thinned by factor 3, following an adaptationperiod of 2,000 iterations resulting in a final total of 12,000posterior samples (4× 3,000) for each parameter.We alsoenabled the option to calculate the Deviance Informa-tion Criterion (DIC; Spiegelhalter, Best, Carlin, & Linde,2002), for use in later model comparison.

When the iterations are finished, two new windowspop up: the Result browser and a noninteractive tablethat gives a summary of the posterior statistics of themost important parameters. This window shows the pos-teriormeans, standard deviations, and percentiles of theseparameters.Moreover, it provides information about con-vergence by displaying R̂ statistics (Gelman et al., 2004),effective number of samples (the number of independentsamples, computed by using the total number of posterior

samples and ameasure of their mutual dependence wheremore dependent samples count as fewer, while entirelyindependent samples count fully), and sample sizes.

The Result browser window offers several ways toexplore the results. By default, the interface shows a warn-ing if convergence is not reached for all parameters interms of any(R̂) > 1.1, and graphical tools are includedto explore the posterior samples of the parameters. For thecurrent analysis, theMCMCprocedure convergedwith allR̂ statistics lower than 1.1.

Posterior predictive checks

Moreover, there are two posterior predictive checks(PPCs) implemented in the program. Both of them arebased on generating new data sets based on the full poste-rior distribution of the parameters and comparing certainproperties of the observed and generated data sets. Thefirst check assesses the similarity between the observedand replicated trajectories: It computes the degree of over-lap between the observed and simulated trajectories bycalculating the correlation between the frequencies withwhich the observed data fall in a certain area in a two-dimensional space and the average frequency with whichthey fall in that area across replicated data sets. The result-ing measure is a correlation coefficient averaged over par-ticipants. The correlationwas 0.86 for the current data set,slightly improved fit compared to the samemeasure (0.80)reported in Kuppens et al. (2010), in the analysis withoutcovariates.

The second PPC indicates whether the observed andreplicated trajectories are similar in terms of turningangles. A turning angle is a clockwise angle between twoline segments that connect three subsequent points intime in the two-dimensional space created by the twolongitudinally measured variables. We average over allturning angles person-wise, resulting in a person-specificaverage turning angle value.We calculate this measure forreplicated data sets, and based on these, a 95% predic-tion interval is established for every person. The programreturns which proportion of the observed average turn-ing angles falls within this interval. With respect to thismeasure, the current analysis showed that 95% of the gen-erated person-specific average turning angles fell withinthe 95% prediction interval, showing an adequate fit ofthe HOUmodel (for more details on posterior predictivechecks, see Gelman, Meng, & Stern, 1996).

Results

General characteristics

Table 1 shows Level 2 results in terms of posterior meanestimates and 95% posterior credible intervals (PCIs).

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 10: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

114 Z. ORAVECZ ET AL.

Table . Summary of the results from the hierarchical Ornstein-Uhlenbeck model on Level .

PosteriorModel parameter Description mean 95% PCI

Valenceα

μ1Baseline . . .

σ 2μ1

Inter-individual variation in baseline . . .

δLμ1Linear time-effect − . − . .

δQμ1Quadratic time-effect . . .

e(αγ1

) Intra-individual variance . . .

e(αβ1

) Self-regulation . . .

σ 21ε Measurement error . . .

Arousalα

μ2Baseline . . .

σ 2μ2

Inter-individual variation in baseline . . .

δLμ2Linear time-effect . . .

δQμ2Quadratic time-effect − . − . − .

e(αγ2

) Intra-individual variance . . .

e(αβ2

) Self-regulation . . .

σ 22ε Measurement error . . .

Cross effectsσ

μ1μ2Covariance between the baseline levels − . − . − .

αργ

Cross-correlation − . − . − .

αρβ

Self-regulation correlation − . − . .

Note. The e(.) stands for the expected value of that parameter on the normal scale (these parameters were estimated using the log scale).The first column displays the mathematical notation of the parameter; the second column describes the function of the parameter. The third column displays themean of the posterior distribution corresponding to the parameter; columns four and five summarize the lower and upper limits of a symmetric % posteriorcredibility interval (PCI).

As can be seen, the baseline core affect is rather pleas-ant (αμ1 = 5.7833) and not particularly aroused (αμ1 =4.4786 on a measurement scale that ranged from 0.1 to9.99). Note that this baseline point was allowed to changeas a function of measurement time nested in the diur-nal cycle, specifically in terms of linear and quadratictime-varying covariates centered around noon, meaningthat we allow for each person’s attraction point to varywith the time of day. The average pleasantness and acti-vation feelings in this analysis correspond to the base-line core affect at noon. The black lines in the two panelsof Figure 1 represent the average (across persons) diur-nal pattern, based on the posterior mean estimates ofthe linear and quadratic time effects in the valence andin the arousal dimensions. For the valence dimension,the linear time effect has a very low magnitude (δLμ1 =−0.0871), and its 95% PCI is rather wide, meaning thatthe valence baseline did not change as a linear functionof time of day. However, there was a small quadratictime effect (δQμ1 = 0.0043) with a comparatively nar-row PCI that suggests that on average there was a smallquadratic trend in the valence baseline position, whichis somewhat noticeable on the black line in Figure 1, leftpanel. With respect to the black line in the right panel,there is a more remarkable quadratic trend in the Level2 mean arousal change over time. Indeed, with respectto arousal, both linear and quadratic effects have rela-tively large magnitudes (δLμ2 = 0.9334, δQμ2 = −0.0303)

with comparatively narrow PCIs, (0.7343, 1.1351) and(−0.0366, −0.0241), respectively. In both plots, the graylines correspond to the person-specific diurnal profiles inthe baseline levels. There appears to be large variation inthese profiles, especially in terms of intercept, and thereseems to be more between-person variability with respectto arousal (σ 2

μ2= 0.8212) than with respect to valence

(σ 2μ1

= 0.5270).The average intraindividual variance was higher in the

arousal dimension (αγ2 = 3.8756) than in the valence(αγ1 = 2.9777) dimension. The intraindividual variancesare large compared to the measurement errors (σ 2

1ε =0.2197, σ 2

2ε = 0.4646), showing that the latent processexplains a large part of the variation in the data—people’smovements through affect space appear to bewelldescribed by an OU process.

The magnitude of the regulatory force is larger in thearousal than in the valence dimension. Thismeans that onaverage, people return to their baseline faster when theirarousal level fluctuates than when their valence does. Thisis especially interesting since the results described alsoshow that there is more variability in the arousal dimen-sion. Together, these findings demonstrate that these twodynamical aspects are distinct, and more variation in theobserved data does not necessarily mean the lack of self-regulation. Also, there was no evidence that regulatorydynamics in valence and arousal are systematically relatedto one another, as αρβ

was practically zero.

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 11: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH 115

Figure . Daily patterns in baseline core affect dimensions. The left panel shows the pleasantness, the right panel the activation dimen-sions. The grey lines correspond to expected person-specific trajectories based on OU model parameter estimates. The black lines showthe expected Level (population) trajectory based on the corresponding parameter estimates in the two core affect dimensions.

The current analysis revealed links between the instan-taneous changes in valence and arousal (average crosscorrelation αργ

= −0.2342). This indicates that changesin the valence dimensions were likely accompanied withchanges in the arousal dimension, in the opposite direc-tion, and vice versa. This finding suggests, for example,that when people were aroused, their valence was likelyto drop slightly. A somewhat related effect was shownby the covariance between the baseline levels: σμ1μ2 =−0.2678 indicated that people who had higher arousalbaseline tended to have lower valence baseline and viceversa.

Finally, Figure 2 provides yet another example of theevaluation of the model fit (see also posterior predictivechecks before). While a stochastic model such as the OUmodel cannot be expected to fit the data perfectly, the pri-mary utility of a process model lies in its ability to cap-ture and quantify those qualitative aspects of the data thatare relevant for psychological interpretation. The figureshows data from two participants (on the left) and foursets of model-generated data for each (on the right). Inboth cases, and in general for our participants, the datagenerated by the model strongly resemble the observeddata.

Results on time-invariant covariates

All eight person-specific process model parameters wereregressed on 10 time-invariant covariates; that is, a multi-ple regressionmodel with 10 predictors was fitted for each

of the eight parameters.4 Based on the posterior samples,Table 2 displays the result on the regression coefficientswhose 95% PCI did not contain 0: These were the effectsfor which the magnitude was relatively high and the cor-responding 95%PCIs were comparatively narrow, provid-ing substantial evidence that the latent process parametersdiffered markedly as a function of these covariates.

As expected, positive affect was positively related tothe valence baseline point: People who frequently experi-enced positive affect tended to feel more pleasant on aver-age. However, with respect to the baseline, there was onlyone more remarkable covariate: The lack of ruminationstrategy for controlling emotional experience predicted amore aroused baseline level.

With respect to intraindividual core affect variation,only the within-person variability in the measurement ofself-esteem showed a marked effect:5 people with morevariable self-esteem had higher levels of variation in theircore affect in general. This way, an important cogni-tive/evaluative aspect (how one thinks of oneself) wasconnected to affect variation.

Possibly the most compelling aspect of HOU modelanalysis concerns the regulatorymechanism and the crosseffects. We would like to point out that self-regulation inthe model refers to a stronger mean-reverting tendency.

Given the simultaneous inclusion of the predictors, the estimated regressionweights of each predictor here are conditional on the other predictors in themodel (i.e., they indicate effects over andabove thoseof theotherpredictors.)

Note that, while the self-esteem measure was collected at every measure-ment occasion, the person-specific variance in self-esteem is time-invariant.

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 12: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

116 Z. ORAVECZ ET AL.

Figure . An illustration of the qualitative aspects of our data that are captured by the Ornstein-Uhlenbeck model. In the top left, the datafrom one participant are plotted. This participant has a home base in an area of low pleasantness but high activation (i.e., upset/distress),has medium variation in pleasantness but is stable in activation, and has medium levels of self-regulation. Each of the four panels in thetop right contain data generated from the model using this participant’s parameters. The data in the bottom left are from a participantwith a home base in an area of high pleasantness and medium activation. Volatility is low in pleasantness but high in activation, and self-regulation is average. In both the top rowand the bottom row, and in general for our participants, themodel recreations of participant datawell capture the salient qualities of the real data. In the two panels on the left side, the numbers next to μ (baseline), γ (intraindividualvariation), and β (regulation) correspond to the point estimates in pleasantness and in the activation dimensions with respect to theseparameters.

That is to say that its desirability might depend on theactual baseline level. As can be seen from Table 2, most ofthe credible effects actually relate to these aspects. First,people with higher neuroticism scores showed lower lev-els of valence self-regulation, while extroverts had lowerlevels of arousal self-regulation. Higher negative affectand self-esteem variability scores predicted better self-regulation of core affect. This suggests that people whofrequently experienced negative emotions and fluctua-tions while reflecting on their own worth, showed higherlevels of affect regulation. This brings up the questionwhether pathologies that are associated with negativeaffect and self-doubtingmight have underlying dynamical

characteristics where negative baseline associated withstrong self-regulation lead to pathological consequences.While this is only theoretical at this point as the cur-rent study is not conclusive (NA and self-esteem variabil-ity did not show remarkable association in this study),it appears to be a promising question for further explo-ration. Finally, from the three emotion self-regulationstrategies (reappraisal, thought suppression, and rumi-nation), only reappraisal predicted better self-regulation,and only in the arousal dimension. In fact, the othertwo strategies (thought suppression and rumination) areconsidered to be maladaptive when it comes to emotionself-regulation (see, e.g., thought suppression in Wegner

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 13: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH 117

Table . Summary of the regression weights with a % posterior credible interval not containing .

Model Displayed name Posterior % Posteriorparameter in BHOUM Description Covariate mean credible interval

Valenceα

μ1PAalpha_Mu_ Attractor Positive affect . . .

αγ1SESD

alpha_gamma_ Variation Self-esteem variability . . .

αβ1N

alpha_beta_ Self-regulation Neuroticism − . − . − .

αβ1NA

alpha_beta_ Self-regulation Negative affect . . .

αβ1SESD

alpha_beta_ Self-regulation Self-esteem variability . . .

Arousalα

μ2RUMalpha_Mu_ Attractor Rumination − . − . − .

αγ2SESD

alpha_gamma_ Variation Self-esteem variability . . .

αβ2E

alpha_beta_ Self-regulation Extraversion − . − . − .

αβ2NA

alpha_beta_ Self-regulation Negative affect . . .

αβ2SESD

alpha_beta_ Self-regulation Self-esteem variability . . .

αβ2RE

alpha_beta_ Self-regulation Reappraisal . . .

Cross effectsα

ργRUM alpha_rho_gamma_ Cross correlation Rumination . . .

Note. Model parameters refer to the regression weights. For example, αμ1PA is the regression weight for positive affect relating to the valence baseline (μ1).

The first columndisplays themathematical notation of theOUparameter; the second column shows theway they are displayed in theBayesianhierarchical Ornstein-Uhlenbeck modeling (BHOUM) toolbox’s output window. The third and fourth columns contain verbal descriptions of which model parameter was regressed onwhich explanatory variable. The fifth column displays the mean of the posterior distribution corresponding to the parameter; columns and summarize thelower and upper limits of a symmetric % posterior credibility interval (PCI).

& Zanakos, 1994, and rumination in Nolen-Hoeksema,2000).

It is interesting to note the discrepancies between ourcurrent results and those obtained from the original two-stage analysis reported in Kuppens et al. (2010). The moststriking differences are in the valence dimension. Withrespect to the baseline, Kuppens et al. (2010) reportedsignificant correlations with neuroticism, extraversion,positive affect, negative affect, and satisfaction with life.The current analysis only found positive affect a mean-ingful covariate. While the directions of the regressionweights for the aforementioned covariates were the samein the current analysis as well, their posterior credibilityintervals were comparatively wide to draw any conclu-sions.With respect to intraindividual variability, we foundonly self-esteem variability as a reliable covariate, whilein terms of traditional correlation measures in Kuppenset al. (2010), not only self-esteem variability, but also self-esteem, negative affect, and neuroticism were significant.Finally, Kuppens et al. (2010) did not note any significantcorrelations with respect to valence self-regulation, whileour analysis showed that neuroticism, negative affect, andself-esteem variability all have predictive power.

These differences serve to highlight the importance ofhandling of parameter uncertainty across model compo-nents: While the original two-stage analysis disregardedeach parameter’s estimation uncertainty (by collapsing anentire posterior distribution into a single measurementpoint), our analysis was able to account for the posterioruncertainty in each parameter individually. As a result,outlying parameter estimates that may have driven a two-stage correlation might be down-weighted to make the

correlation disappear. Alternatively, parameters central inthe distribution might be down-weighted, bringing a pre-viously unobserved correlation to the surface. The prop-agation of uncertainty in parameter estimates is a con-siderable advantage of the hierarchical Bayesian approachapplied broadly.

Discussion

The HOU process model is a psychometric modelingtool that can be applied to various phenomena thatare assumed to change dynamically over time. Throughan example application, we demonstrated how vari-ous aspects of the temporal change mechanism can beexplored, including associations between parameters ofintraindividual change and exploratory variables. Wenote that if—as in our case—these specific relation-ships between model parameters and predictors were nothypothesized a priori, such results should be consideredexploratory, and it is important to replicate them in a con-firmatory setting before they can give rise to strong theo-retical claims.

Intensive longitudinal data are expected to becomemore common with the increased availability of tech-nology for collecting ecological repeated measures data.Methods for the analysis of such data are therefore of greatinterest to both methodologists and applied researchers.Substantial contributions of the HOU model to emotionand personality psychology involve separating substan-tively different mechanisms underlying observed scores.For example, variability measured through experience-sampling studies can be decomposed into measurement

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 14: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

118 Z. ORAVECZ ET AL.

error and person-specific dynamical patterns in terms ofintra-individual variance and self-regulation. Moreover,the bivariate aspect of the framework allows us to takedependency between two longitudinally measured vari-ables into account, along with studying interindividualdifferences in terms of synchronicity parameters.

We further demonstrated how individual differencecan be explained through the addition of meaningfulcovariate covariates. The ability to regress model param-eters onto covariates in a single step increased the accu-racy of the estimated regression coefficients. We expectthat these desirable properties, together with a user-friendly parameter estimation implementation, will facil-itate application of the model among applied researchers.

Finally, we would like to address the question of studydesign. The presented model is most useful for inten-sive longitudinal data: data from several individuals, withmore than a handful of data points each. Ideally, datahave some degree of variance: The model is not ideal formeasurements that only take one or two different values.If data do not contain enough information for efficientestimation of the model, convergence issues and/or largeuncertainties in the parameter estimates may occur.

Article information

Conflict of Interest Disclosures: Each author signed a form fordisclosure of potential conflicts of interest. No authors reportedany financial or other conflicts of interest in relation to the workdescribed.Ethical Principles: The authors affirm having followed profes-sional ethical guidelines in preparing this work. These guide-lines include obtaining informed consent from human partici-pants,maintaining ethical treatment and respect for the rights ofhuman or animal participants, and ensuring the privacy of par-ticipants and their data, such as ensuring that individual partic-ipants cannot be identified in reported results or from publiclyavailable original or archival data.Funding: The research reported in this paper was sponsored inpart by Belgian Federal Science Policy within the framework ofthe Interuniversity Attraction Poles program IAP/P7/06 (FT),in part by the grant GOA/15/003 from the University of Leu-ven (FT), in part by the grant G.0806.13 from the ResearchFoundation—Flanders (FT), in part by the grant #48192 fromThe John Templeton Foundation (ZO and JV) and in part by thegrant #1230118 from the National Science Foundation’s Meth-ods, Measurements, and Statistics panel (JV).Role of the Funders/Sponsors: None of the funders or spon-sors of this research had any role in the design and conduct ofthe study; collection, management, analysis, and interpretationof data; preparation, review, or approval of the manuscript; ordecision to submit the manuscript for publication.Acknowledgements: The ideas and opinions expressed hereinare those of the authors alone, and endorsement by the authors’institutions or The John Templeton Foundation is not intendedand should not be inferred. The authors are grateful to ChelseaMuth, Marlies Houben, Madeline Pe and Peter Kuppens for

their beta-testing efforts of BHOUM as well as for their help-ful comments on the manuscript.

References

Barrett, L. F. (2004). Feelings or words? Understanding thecontent in self-report ratings of experienced emotion.Journal of Personality and Social Psychology, 87(2), 266.doi:10.1037/0022-3514.87.2.266

Bolger, N., & Laurenceau, J. (2013). Intensive longitudinal meth-ods. New York, NY: Guilford.

Browne, M. W., & Nesselroade, J. R. (2005). Representing psy-chological processes with dynamic factor models: Somepromising uses and extensions of autoregressive movingaverage time series models. In A. Maydeu-Olivares & J. J.McArdle (Eds.), Contemporary psychometrics: A Festschriftfor Roderick P. McDonald (pp. 415–452). Mahwah, NJ: Erl-baum.

Chow, S.-M., Ferrer, E., & Nesselroade, J. R. (2007). Anunscented Kalman filter approach to the estimation of non-linear dynamical systems models. Multivariate BehavioralResearch, 42(2), 283–321. doi:10.1080/00273170701360423

Chow, S.-M., Ho, M.-H. R., Hamaker, E. J., & Dolan,C. V. (2010). Equivalences and differences betweenstructural equation and state-space modeling frame-works. Structural Equation Modeling, 17, 303–332.doi:10.1080/10705511003661553

Chow, S.-M., Ram, N., Boker, S. M., Fujita, F., & Clore, G.(2005). Emotion as a thermostat: Representing emotion reg-ulation using a damped oscillator model. Emotion, 5, 208–225. doi:10.1037/1528-3542.5.2.208

Costa, P. T., & McCrae, R. R. (1992). Normal personalityassessment in clinical practice: The NEO personality inven-tory. Psychological assessment, 4(1), 5. doi:10.1037/1040-3590.4.1.5

Csikszentmihalyi, M., & Larson, R. (1987). Validity andreliability of the experience sampling method. TheJournal of Nervous and Mental Disease, 175, 526–536.doi:10.1097/00005053-198709000-00004

Deboeck, P. R., & Bergeman, C. (2013). The reservoir model:A differential equation model of psychological regulation.Psychological Methods, 18(2), 237. doi:10.1037/a0031603

Delsing, M. J. M. H., Oud, J. H. L., & Bruyn, E. E. J.(2005). Assessment of bidirectional influences betweenfamily relationships and adolescent problembehavior.Euro-pean Journal of Psychological Assessment, 21, 226–231.doi:10.1027/1015-5759.21.4.226

Diener, E., Emmons, R. A., Larsen, R. J., & Griffin, S. (1985).Satisfactionwith life scale. Journal of Personality Assessment,49, 71–75. doi:10.1207/s15327752jpa4901-13

Dunn, J. E., & Gipson, P. S. (1977). Analysis of radio teleme-try data in studies of home range. Biometrics, 33, 85–101.doi:10.2307/2529305

Fahrmeir, L., &Tutz, G. (2001).Multivariate statisticalmodellingbased on generalized linear models. New York, NY: SpringerVerlag.

Gardiner, C. W. (1986). Handbook of stochastic methods forphysics, chemistry and the natural sciences (2nd ed.). Berlin,Germany: Springer-Verlag.

Gelman, A., Carlin, J. B., Stern, H. S., & Rubin, D. B. (2004).Bayesian data analysis (2nd ed.). Boca Raton, FL: Chapman& Hall.

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6

Page 15: Bayesian Data Analysis with the Bivariate Hierarchical ... · Bayesian Data Analysis with the Bivariate Hierarchical Ornstein-Uhlenbeck Process Model Zita Oravecz, Francis Tuerlinckx

MULTIVARIATE BEHAVIORAL RESEARCH 119

Gelman, A., Carlin, J. B., Stern, H. S., Dunson, D. B., Vehtari, A.,& Rubin, D. B. (2013). Bayesian data analysis (3rd ed.). BocaRaton, FL: Chapman & Hall/CRC.

Gelman, A., Meng, X.-L., & Stern, H. (1996). Posterior predic-tive assessment of model fitness via realized discrepancies.Statistica sinica, 6(4), 733–760.

Gross, J. J. (2002). Emotion regulation: Affective, cognitive,and social consequences. Psychophysiology, 39(03), 281–291. doi:10.1037/0022-3514.85.2.348

Gross, J. J., & John, O. P. (2003). Individual differences in twoemotion regulation processes: Implications for affect, rela-tionships, and wellbeing. Journal of Personality and SocialPsychology, 85, 348–362. doi:10.1037/0022-3514.85.2.348

Hills, T. T., Jones, M. N., & Todd, P. M. (2012). Optimal forag-ing in semantic memory. Psychological Review, 119(2), 431.doi:10.1037/a0027373

Hoeksma, J. B., Oosterlaan, J., Schipper, E., & Koot, H. (1988).Development and validation of brief measures of positiveand negative affect: The PANAS scale. Journal of Personalityand Social Psychology, 54, 1063–1070.

Hoekstra, H. A., Ormel, J., & De Fruyt, F. (1996). NEO PI-R,NEO FFI Big Five Persoonlijkheidsvragenlijsten: Handleiding[NEO PI-R, NEO FFI Big Five Personality Questionnaire:Manual]. Lisse, The Netherlands: Swets & Zeitlinger B.V.

Kuppens, P., Oravecz, Z., & Tuerlinckx, F. (2010). Feelingschange: Accounting for individual differences in the tem-poral dynamics of affect. Journal of Personality and SocialPsychology, 99, 1042–1060. doi:10.1037/a0020962

Kuppens, P., Stouten, J., &Mesquita, B. (2009). Individual differ-ences in emotion components and dynamics: Introductionto the special issue. Cognition & Emotion, 23, 1249–1258.doi:10.1080/02699930902985605

Kuppens, P., Tuerlinckx, F., Russell, J. A., & Barrett, L. F.(2013). The relation between valence and arousal insubjective experience. Psychological Bulletin, 139(4), 917.doi:10.1037/a0030811

MacCallum, R. C., Kim, C., Malarkey, W. B., & Kiecolt-Glaser,J. K. (1997). Studying multivariate change using multilevelmodels and latent curve models. Multivariate BehavioralResearch, 32(3), 215–253. doi:10.1207/s15327906mbr3203-1

Mehl, M. R., & Conner, T. S. (2012).Handbook of research meth-ods for studying daily life. New York, NY: Guilford Press.

Molenaar, P. C. M., & Newell, K. M. (2003). Direct fit of atheoretical model of phase transition in oscillatory fingermotions. British Journal of Mathematical and Statistical Psy-chology, 56, 199–214. doi:10.1348/000711003770480002

Mood, A. M., Graybill, F. A., & Boes, D. C. (1974). Introduc-tion to the theory of statistics. New York, NY: McGraw-Hill.

Nolen-Hoeksema, S. (2000). The role of rumination indepressive disorders and mixed anxiety/depressive symp-toms. Journal of Abnormal Psychology, 109(3), 504.doi:101037/10021-843X.109.3.504

Oravecz, Z., & Tuerlinckx, F. (2011). The linear mixedmodel and the hierarchical Ornstein-Uhlenbeck model:Some equivalences and differences. British Journal ofMathematical and Statistical Psychology, 64, 134–160.doi:10.1348/000711010X498621

Oravecz, Z., Tuerlinckx, F., & Vandekerckhove, J. (2011). Ahierarchical latent stochastic differential equation modelfor affective dynamics. Psychological Methods, 16, 468–490.doi:10.1037/a0024375

Oud, J. H. L. (2007). Comparison of four procedures to estimatethe damped linear differential oscillator for panel data. In J.Oud & A. Satorra (Eds.), Longitudinal models in the behav-ioral and related sciences. Mahwah, NJ: Lawrence ErlbaumAssociates. pp. 19–39.

Oud, J. H. L., & Jansen, R. A. R. G. (2000). Continuous time statespace modeling of panel data by means of SEM. Psychome-trika, 65(2), 199–215. doi:10.1007/BF02294374

Pagan, A. (1984). Econometric issues in the analysis of regres-sions with generated regressors. International EconomicReview, 25, 221–247. doi:10.2307/2648877

Ram, N., & Gerstorf, D. (2009). Time-structured and netintraindividual variability: Tools for examining the devel-opment of dynamic characteristics and processes. Psychol-ogy and Aging, 24(4), 778. doi:10.1037/a0017915

Robert, C. P., & Casella, G. (2004).Monte Carlo statistical meth-ods. New York, NY: Springer.

Rosenberg, M. (1989). Society and the adolescent self-image.Middletown, CT: Wesleyan University Press.

Russell, J. A. (2003). Core affect and the psychological con-struction of emotion. Psychological Review, 110, 145–172.doi:10.1037/0033-295X.110.1.145

Spiegelhalter, D. J., Best, N. G., Carlin, B. P., & Linde, A. van der.(2002). Bayesianmeasures ofmodel complexity andfit (withdiscussion). Journal of the Royal Statistical Society, Series B,6, 583–640. doi:10.1111/1467-9868.00353

Stone, A. A., & Shiffman, S. (1994). Ecological momentaryassessment in behavioral medicine. Annals of BehavioralMedicine, 16, 199–202. doi:10.1023/A:1023657221954

Trapnell, P. D., & Campbell, J. D. (1999). Private self-consciousness and the five-factormodel of personality: Dis-tinguishing rumination from reflection. Journal of Person-ality and Social Psychology, 76, 284–304. doi:10.1037/0022-3514.76.2.284

Uhlenbeck, G. E., & Ornstein, L. S. (1930). On the the-ory of Brownian motion. Physical Review, 36, 823–841.doi:10.1103/PhysRev.36.823

Voelkle, M. C., & Oud, J. H. (2013). Continuous time mod-elling with individually varying time intervals for oscil-lating and non-oscillating processes. British Journal ofMathematical and Statistical Psychology, 66(1), 103–126.doi:10.1111/j.2044-8317.2012.02043.x

Walls, T. A., & Schafer, J. L. (2006).Models for intensive longitu-dinal data. New York, NY: Oxford University Press.

Wang, L. P., Hamaker, E., & Bergeman, C. (2012). Inves-tigating inter-individual differences in short-term intra-individual variability. Psychological Methods, 17(4), 567.doi:10.1037/a0029317

Wegner, D. M., & Zanakos, S. (1994). Chronic thoughtsuppression. Journal of Personality, 62(4), 615–640.doi:10.1111/j.1467-6494.1994.tb00311.x

Wilson-Mendenhall, C. D., Barrett, L. F., & Barsalou, L. W.(2013). Neural evidence that human emotions share coreaffective properties. Psychological Science, 24(6), 947–956.doi:10.1177/0956797612464242

Dow

nloa

ded

by [

KU

Leu

ven

Uni

vers

ity L

ibra

ry]

at 0

8:08

22

Febr

uary

201

6