Skip to main content

Computational Models for Affect Dynamics

  • Chapter
  • First Online:
Affect Dynamics

Abstract

Computational models of affect dynamics are ubiquitous. These models are appropriate for either exploring intensive longitudinal data or testing theories about affect dynamics. In this chapter, we give a brief overview of some of the computational models that have been applied in the field of affect dynamics, focusing on both discrete-time and continuous-time models. The emphasis of this chapter lies on describing the core ideas of the models and how they can be interpreted. At the end, we provide references to other important topics for the interested reader.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Chapter
USD 29.95
Price excludes VAT (USA)
eBook
USD 149.00
Price excludes VAT (USA)
Softcover Book
USD 199.99
Price excludes VAT (USA)
Hardcover Book
USD 199.99
Price excludes VAT (USA)

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

Notes

  1. 1.

    Because of this, we present network models in the section on autoregressive models. This was a practical choice, and we do not mean to imply that network models are always autoregressive in nature. In fact, most networks are not (e.g., Ising models; Kruis & Maris, 2016).

  2. 2.

    Note that we specifically talk about reinforcement learning in the context of emotion dynamics: This class of model is applicable to many more subjects, like decision-making, conditioning, and learned behavior (see Sutton & Barto, 2018).

  3. 3.

    To remain in line with the mathematical notations of this chapter, we changed the notation of this model (see Rutledge et al., 2014).

  4. 4.

    Importantly, this implies that vector fields are deterministic—they show what the model would expect if there were no perturbations to the system.

  5. 5.

    In the application that we do not discuss, the OU model was only part of a series of equations (see Pellert et al., 2020; Schweitzer & Garcia, 2010).

  6. 6.

    The damped linear oscillator is an example of a second-order differential equation, where speed and location of a variable y at time t are both related to changes in speed over time (i.e., acceleration; speeding up or slowing down over time).

  7. 7.

    If this is the case, we are able to predict the emotional behavior of an individual for eternity, as the system is deterministic.

References

  • Adolf, J. K., Voelkle, M. C., Brose, A., & Schmiedek, F. (2017). Capturing context-related change in emotional dynamics via fixed-moderated time series analysis. Multivariate Behavioral Research, 52, 499–531. https://doi.org/10.1080/00273171.2017.1321978

    Article  Google Scholar 

  • Albers, C. J., & Bringmann, L. F. (2020). Inspecting gradual and abrupt changes in emotion dynamics with the time-varying change point autoregressive model. European Journal of Psychological Assessment, 36, 492–499. https://doi.org/10.1027/1015-5759/a000589

    Article  Google Scholar 

  • Allen, B. D., & Carifio, J. (1995). Nonlinear analysis: Catastrophe theory modeling and Cobb’s Cusp Surface Analysis Program. Evaluation Review, 19, 64–83. https://doi.org/10.1177/0193841X9501900103

    Article  Google Scholar 

  • Ariens, S., Ceulemans, E., & Adolf, J. K. (2020). Time series analysis of intensive longitudinal data in psychosomatic research: A methodological overview. Journal of Psychosomatic Research, 137, 110191.

    Article  Google Scholar 

  • Arlot, S., & Celisse, A. (2010). A survey of cross-validation procedures for model selection. Statistics Surveys, 4, 40–79. https://doi.org/10.1214/09-SS054

    Article  Google Scholar 

  • Bengtsson, T., & Cavanaugh, J. E. (2006). An improved Akaike information criterion for state-space model selection. Computational Statistics & Data Analysis, 50, 2635–2654. https://doi.org/10.1016/j.csda.2005.05.003

    Article  Google Scholar 

  • Bennett, D., Davidson, G., & Niv, Y. (2020). A model of mood as integrated advantage. Retrieved from https://10.31234/osf.io/dzsme

  • Bennett, D., & Niv, Y. (2018, June 7). Opening Burton’s clock: Psychiatric insights from computational cognitive models. Retrieved from https://10.31234/osf.io/y2vzu

  • Bergmeir, C., Hyndman, R. J., & Koo, B. (2018). A note on the validity of cross-validation for evaluating autoregressive time series prediction. Computational Statistics & Data Analysis, 120, 70–83. https://doi.org/10.1016/j.csda.2017.11.003

    Article  Google Scholar 

  • Bergstrom, A. R. (1984). Continuous-time stochastic models and issues of aggregation over time. In Z. Griliches & M. D. Intriligator (Eds.), Handbook of econometrics. Elsevier Science.

    Google Scholar 

  • Boker, S. M. (2002). Consequences of continuity: The hunt for intrinsic properties within parameters of dynamics in psychological processes. Multivariate Behavioral Research, 37, 405–422. https://doi.org/10.1207/S15327906MBR3703_5

    Article  Google Scholar 

  • Boker, S. M., & Graham, J. (1998). A dynamical systems analysis of adolescent substance abuse. Multivariate Behavioral Research, 33, 479–507. https://doi.org/10.1207/s15327906mbr3304_3

    Article  Google Scholar 

  • Boker, S. M., & Laurenceau, J.-P. (2006). Dynamical systems modeling: An application to the regulation of intimacy and disclosure in marriage. In T. A. Walls & J. L. Schafer (Eds.), Models for intensive longitudinal data. Oxford University Press.

    Google Scholar 

  • Boker, S. M., & Nesselroade, J. R. (2002). A method for modeling the intrinsic dynamics of intraindividual variability: Recovering the parameters of simulated oscillators in multi-wave panel data. Multivariate Behavioral Research, 37, 127–160. https://doi.org/10.1207/S15327906MBR3701_06

    Article  Google Scholar 

  • Boker, S. M., Staples, A. D., & Hu, Y. (2016). Dynamics of change and change in dynamics. Journal of Person-Oriented Research, 2, 34–55. https://doi.org/10.17505/jpor.2016.05

    Article  Google Scholar 

  • Boker, S. M., Neale, M. C., Maes, H. H., Wilde, M. J., Spiegel, M., Brick, T. R., Estabrook, R., Bates, T. C., Mehta, P., von Oertzen, T., Gore, R. J., Hunter, M. D., Hackett, D. C., Karch, J., Brandmaier, A. M., Pritikin, J. N., Zahery, M., Kirkpatrick, R. M., Wang, Y., et al. (2020). OpenMx 2.17.4 user guide.

    Google Scholar 

  • Bolger, N., Davis, A., & Rafaeli, E. (2003). Diary methods: Capturing life as it is lived. Annual Review of Psychology, 54, 579–616. https://doi.org/10.1146/annurev.psych.54.101601.145030

    Article  Google Scholar 

  • Bonsall, M. B., Wallace-Hadrill, S. M. A., Geddes, J. R., Goodwin, G. M., & Holmes, E. A. (2012). Nonlinear time-series approaches in characterizing mood stability and mood instability in bipolar disorder. Proceedings of the Royal Society B, 279, 916–924. https://doi.org/10.1098/rspb.2011.1246

    Article  Google Scholar 

  • Booij, S. H., Wigman, J. T. W., Jacobs, N., Thiery, E., Derom, C., Wichers, M., & Oravecz, Z. (2020). Cortisol dynamics in depression: Application of a continuous-time process model. Psychoneuroendocrinology, 115. https://doi.org/10.1016/j.psyneuen.2020.104598

  • Borsboom, D., & Cramer, A. O. J. (2013). Network analysis: An integrative approach to the structure of psychopathology. Annual Review of Clinical Psychology, 9, 91–121. https://doi.org/10.1146/annurev-clinpsy-050212-185608

    Article  Google Scholar 

  • Borsboom, D., Cramer, A. O. J., Schmittmann, V. D., Epskamp, S., & Waldorp, L. J. (2011). The small world of psychopathology. PLoS One, 6, e27407. https://doi.org/10.1371/journal.pone.0027407

    Article  Google Scholar 

  • Box, G. E. P. (1976). Science and statistics. Journal of the American Statistical Association, 71, 791–799.

    Article  Google Scholar 

  • Box, G. E. P. (1979). Robustness in the strategy of scientific model building. In R. L. Launer & G. N. Wilkinson (Eds.), Robustness in statistics. Academic.

    Google Scholar 

  • Box, G. E. P. (1966). Use and abuse of regression. Technometrics, 8(4), 625–629.

    Article  Google Scholar 

  • Box, G. E. P., & Jenkins, G. M. (1970). Time series analysis: Forecasting and control. Holden-Day.

    Google Scholar 

  • Breiman, L. (2001). Statistical modeling: The two cultures. Statistical Science, 16, 199–231.

    Article  Google Scholar 

  • Brickman, P., & Campbell, D. T. (1971). Hedonic relativism and planning the good society. In M. H. Appley (Ed.), Adaptation-level theory. United States of America, Academic Press.

    Google Scholar 

  • Bringmann, L. F., Ferrer, E., Hamaker, E. L., Borsboom, D., & Tuerlinckx, F. (2018). Modeling nonstationary emotion dynamics in dyads using a time-varying vetor-autoregressive model. Multivariate Behavioral Research, 53, 293–314. https://doi.org/10.1080/00273171.2018.1439722

    Article  Google Scholar 

  • Bringmann, L. F., Vissers, N., Wichers, M., Geschwind, N., Kuppens, P., Peeters, F., Borsboom, D., & Tuerlinckx, F. (2013). A network approach to psychopathology: New insights into clinical longitudinal data. PLoS One, 8, e60188. https://doi.org/10.1371/journal.pone.0060188

    Article  Google Scholar 

  • Brown, N. J. L., Sokal, A. D., & Friedman, H. L. (2013). The complex dynamics of wishful thinking: The critical positivity ratio. American Psychologist, 68, 801–813. https://doi.org/10.1037/a0032850

    Article  Google Scholar 

  • Bulteel, K., Tuerlinckx, F., Brose, A., & Ceulemans, E. (2016). Using raw var regression coefficients to build networks can be misleading. Multivariate Behavioral Research, 51, 330–344. https://doi.org/10.1080/00273171.2016.1150151

    Article  Google Scholar 

  • Butner, J., Amazeen, P. G., & Mulvey, G. M. (2005). Multilevel modeling of two cyclical processes: Extending differential structural equation modeling to nonlinear coupled systems. Psychological Methods, 10, 159–177. https://doi.org/10.1037/1082-989X.10.2.159

    Article  Google Scholar 

  • Cabrieto, J., Adolf, J., Tuerlinckx, F., Kuppens, P., & Ceulemans, E. (2018). Detecting long-lived autodependency changes in a multivariate system via change point detection and regime switching models. Scientific Reports, 8. https://doi.org/10.1038/s41598-018-33819-8

  • Chen, M., Chow, S.-M., & Hunter, M. D. (2018). Stochastic differential equation models with timevarying parameters. In K. van Montfort, J. H. L. Oud, & M. C. Voelkle (Eds.), Continuous time modeling in the behavioral and related sciences.

    Google Scholar 

  • Chen, Y., & Zhang, S. (2020). A latent Gaussian process model for analysing intensive longitudinal data. The British Journal of Mathematical and Statistical Psychology, 73, 237–260. https://doi.org/10.1111/bmsp.12180

    Article  Google Scholar 

  • Chow, S.-M., Lu, Z., & Zhu, H. (2016). Fitting nonlinear ordinary differential equation models with random effects and unknown initial conditions using the stochastic approximation expectation maximization (SAEM) algorithm. Psychometrika, 81, 102–134. https://doi.org/10.1007/s11336-014-9431-z

    Article  Google Scholar 

  • Chow, S.-M., Ram, N., Boker, S. M., Fujita, F., & Clore, G. (2005). Emotion as a thermostat: Representing emotion regulation using a damped oscillator model. Emotion, 5, 208–225. https://doi.org/10.1037/1528-3542-5.2.208

    Article  Google Scholar 

  • Chow, S.-M., Witkiewitz, K., Grasman, R. P. P. P., & Maisto, S. A. (2015). The cusp catastrophe model as cross-sectional longitudinal mixture structural equation model. Psychological Methods, 20, 142–164. https://doi.org/10.1037/a0038962

    Article  Google Scholar 

  • Chow, S.-M., Ho, M.-H. R., Hamaker, E. L., & Dolan, C. V. (2010). Equivalence and differences between structural equation modeling and state-space modeling techniques. Structural Equation Modeling, 17(2), 303–332.

    Article  Google Scholar 

  • Cobb, L., & Watson, B. (1980). Statistical catastrophe theory: An overview. Mathematical Modelling, 1, 311–317.

    Article  Google Scholar 

  • Congard, A., Dauvier, B., Antoine, P., & Gilles, P.-Y. (2011). Integrating personality, daily life events and emotion: Role of anxiety and positive affect in emotion regulation dynamics. Journal of Research in Personality, 45, 372–384. https://doi.org/10.1016/j.jrp.2011.04.004

    Article  Google Scholar 

  • Cunningham, W. A., Dunfield, K. A., & Stillman, P. E. (2013). Emotional states from affective dynamics. Emotion Review, 5, 344–355. https://doi.org/10.1177/1754073913489749

    Article  Google Scholar 

  • Daros, A. R., Daniel, K. E., Boukhechba, M., Chow, P. I., Barnes, L. E., & Teachman, B. A. (2019). Relationship between trait emotion dysregulation and emotional experiences in daily life: An experience sampling study. Cognition and Emotion, 34, 743–755. https://doi.org/10.1080/02699931.2019.1681364

    Article  Google Scholar 

  • de Haan-Rietdijk, S., Voelkle, M. C., Keijsers, L., & Hamaker, E. L. (2017). Discrete-vs. continuous time modeling of unequally spaced experience sampling method data. Frontiers in Psychology, 8, 1849.

    Article  Google Scholar 

  • Deboeck, P. R. (2013). Dynamical systems and models of continuous time. In T. D. Little (Ed.), The Oxford handbook of quantitative methods in psychology. Oxford University Press.

    Google Scholar 

  • Deboeck, P. R., & Bergeman, C. S. (2013). The reservoir model: A differential equation model of psychological regulation. Psychological Methods, 18, 237–256. https://doi.org/10.1037/a0031603

    Article  Google Scholar 

  • Deboeck, P. R., & Preacher, K. J. (2016). No need to be discrete: A method for continuous time mediation analysis. Structural Equation Modeling: A Multidisciplinary Journal, 23, 61–75. https://doi.org/10.1080/10705511.2014.973960

    Article  Google Scholar 

  • Dickey, D. A., & Fuller, W. A. (1979). Distribution of the estimators for autoregressive time series with a unit root. Journal of the American Statistical Association, 74(366a), 427–431.

    Article  Google Scholar 

  • Diener, E., & Iran-Nejad, A. (1986). The relationship in experience between various types of affect. Journal of Personality and Social Psychology, 50, 1031–1038.

    Article  Google Scholar 

  • Doll, B. B., Simon, D. A., & Daw, N. D. (2012). The ubiquity of model-based reinforcement learning. Current Opinion in Neurobiology, 22, 1075–1081. https://doi.org/10.1016/j.conb.2012.08.003

    Article  Google Scholar 

  • Driver, C. C., Oud, J. H. L., & Voelkle, M. C. (2017). Continuous time structural equation modeling with R package ctsem. Journal of Statistical Software, 77. https://doi.org/10.18637/jss.v077.i05

  • Driver, C. C., & Voelkle, M. C. (2018a). Hierarchical Bayesian continuous time dynamic modeling. Psychological Methods, 23, 774–799. https://doi.org/10.1037/met0000168

    Article  Google Scholar 

  • Driver, C. C., & Voelkle, M. C. (2018b). Understanding the time course of interventions with continuoustime dynamic models. In K. van Montfort, J. H. L. Oud, & M. C. Voelkle (Eds.), Continuous time modeling in the behavioral and related sciences.

    Google Scholar 

  • Dunn, E. C., Nishimi, K., Gomez, S. H., Powers Lott, A., & Bradley, B. (2018). Developmental timing of trauma exposure and emotion dysregulation in adulthood: Are there sensitive periods when trauma is most harmful? Journal of Affective Disorders, 227, 869–877. https://doi.org/10.1016/j.jad.2017.10.045

    Article  Google Scholar 

  • Ekman, P. (1992). An argument for basic emotions. Cognition and Emotion, 6, 169–200.

    Article  Google Scholar 

  • Eldar, E., & Niv, Y. (2015). Interaction between emotional state and learning underlies mood instability. Nature Communications, 6. https://doi.org/10.1038/ncomms7149

  • Eldar, E., Rutledge, R. B., Dolan, R. J., & Niv, Y. (2015). Mood as representation of momentum. Trends in Cognitive Sciences, 20, 15–24. https://doi.org/10.1016/j.tics.2015.07.010

    Article  Google Scholar 

  • Epskamp, S. (2020a). Psychometric network models from time series and panel data. Psychometrika, 85, 206–231. https://doi.org/10.1007/s11336-020-09697-3

    Article  Google Scholar 

  • Epskamp, S., Waldorp, L. J., Mottus, R., & Borsboom, D. (2018). The Gaussian graphical model in cross-sectional and time-series data. Multivariate Behavioral Research, 53, 453–480. https://doi.org/10.1080/00273171.2018.1454823

    Article  Google Scholar 

  • Epskamp, S. (2020b). graphicalVAR: Graphical VAR for experience sampling data. R package version 0.2.4. Retrieved from https://CRAN.R-project.org/package=graphicalVAR

  • Epskamp, S., Cramer, A. O. J., Waldorp, L. J., Schmittmann, V. D., & Borsboom, D. (2012). qgraph: Network visualizations of relationships in psychometric data. Journal of Statistical Software, 48(4), 1–18.

    Article  Google Scholar 

  • Farrell, S., & Lewandowsky, S. (2010). Computational models as aids to better reasoning in psychology. Current Directions in Psychological Science, 19, 329–335. https://doi.org/10.1177/0963721410386677

    Article  Google Scholar 

  • Feldman Barrett, L. (2009). Variety is the spice of life: A psychological construction approach to understanding variability in emotion. Cognition and Emotion, 23, 1284–1306. https://doi.org/10.1080/02699930902985894

    Article  Google Scholar 

  • Flay, B. R. (1978). Catastrophe theory in social psychology: Some applications to attitudes and social behavior. Behavioral Science, 23, 335–350.

    Article  Google Scholar 

  • Fredrickson, B. L., & Losada, M. F. (2005). Positive affect and the complex dynamics of human flourishing. American Psychologist, 60, 678–686. https://doi.org/10.1037/0003-066X.60.7.678

    Article  Google Scholar 

  • Fried, E. I., van Borkulo, C. D., Cramer, A. O. J., Boschloo, L., Schoevers, R. A., & Borsboom, D. (2017). Mental disorders as networks of problems: A review of recent insights. Social Psychiatry and Psychiatric Epidemiology, 52, 1–10. https://doi.org/10.1007/s00127-016-1319-z

    Article  Google Scholar 

  • Frijda, N. H. (2007). The laws of emotion. Routledge.

    Google Scholar 

  • Gelman, A. (2006). Multilevel (hierarchical) modeling: What it can and cannot do. Technometrics, 48, 432–435. https://doi.org/10.1198/004017005000000661

    Article  Google Scholar 

  • Gelman, A., & Hill, J. (2006). Data analysis using regression and multilevel/hierarchical models. Cambridge University Press.

    Book  Google Scholar 

  • Greene, T., Gelkopf, M., Fried, E. I., Robinaugh, D. J., & Pickman, L. L. (2020). Dynamic network analysis of negative emotions and DSM-5 posttraumatic stress disorder symptom clusters during conflict. Journal of Traumatic Stress, 33, 72–83. https://doi.org/10.1002/jts.22433

    Article  Google Scholar 

  • Groen, R. N., Ryan, O., Wigman, J. T. W., Riese, H., Penninx, B. W. J. H., Giltay, E. J., Wichers, M., & Hartman, C. A. (2020). Comorbidity between depression and anxiety: Assessing the role of bridge mental states in dynamic psychological networks. BMC Medicine, 18. https://doi.org/10.1186/s12916-020-01738-z

  • Gross, J. J. (2015). Emotion regulation: Current status and future prospects. Psychological Inquiry, 26, 1–26. https://doi.org/10.1080/1047840X.2014.940781

    Article  Google Scholar 

  • Guastello, S. J. (2014). Nonlinear dynamical models in psychology are widespread and testable. American Psychologist, 69, 628–629. https://doi.org/10.1037/a0036980

    Article  Google Scholar 

  • Guthier, C., Dormann, C., & Voelkle, M. C. (2020). Reciprocal effects between job stressors and burnout: A continuous-time meta-analysis of longitudinal studies. Psychological Bulletin, 146, 1146–1173. https://doi.org/10.1037/bul0000304

    Article  Google Scholar 

  • Hamaker, E. L., Ceulemans, E., Grasman, R. P. P. P., & Tuerlinckx, F. (2015). Modeling affect dynamics: State of the art and future challenges. Emotion Review, 7, 316–322. https://doi.org/10.1177/1754073915590619

    Article  Google Scholar 

  • Hamaker, E. L., & Wichers, M. (2017). No time like the present: Discovering the hidden dynamics in intensive longitudinal data. Current Directions in Psychological Science, 26, 10–15. https://doi.org/10.1177/0963721416666518

    Article  Google Scholar 

  • Hamilton, J. D. (1994a). State-space models. In R. F. Engle & D. L. McFadden (Eds.), Handbook of econometrics. Elsevier. https://doi.org/10.1016/S15734412(05)80019-4

    Chapter  Google Scholar 

  • Hamilton, J. D. (1994b). Time series analysis (Vol. 2). Princeton University Press.

    Google Scholar 

  • Hamilton, J. D. (2010). Regime switching models. In S. N. Durlauf & L. E. Blume (Eds.), Macroeconometrics and time series analysis. Springer.

    Google Scholar 

  • Hartelman, P. A. I., van der Maas, H. L. J., & Molenaar, P. C. M. (1998). Detecting and modelling developmental transitions. British Journal of Developmental Psychology, 16, 97–122.

    Article  Google Scholar 

  • Harvey, A. C. (1989). Forecasting, structural time series models and the Kalman filter. Cambridge university press.

    Google Scholar 

  • Hektner, M. J., Schmidt, A. J., & Csikszentmihalyi, M. (2006). Experience sampling methods: Measuring the quality of everyday life (1st ed.). SAGE Publications.

    Google Scholar 

  • Hu, Y., Boker, S., Neale, M., & Klump, K. L. (2014). Coupled latent differential equations with moderators: Simulation and application. Psychological Methods, 19, 56–71. https://doi.org/10.1037/a0032476

    Article  Google Scholar 

  • Hunter, M. D. (2018). State space modeling in an open source, modular, structural equation modeling environment. Structural Equation Modeling, 25, 307–324. https://doi.org/10.1080/10705511.2017.1369354

    Article  Google Scholar 

  • Hyndman, R. J., Ahmed, R. A., Athanasopoulos, G., & Shang, H. L. (2011). Optimal combination forecasts for hierarchical time series. Computational Statistics and Data Analysis, 55, 2579–2589. https://doi.org/10.1016/j.csda.2011.03.006

    Article  Google Scholar 

  • Hyndman, R. J., & Koehler, A. B. (2006). Another look at measures of forecast accuracy. International Journal of Forecasting, 22, 679–688. https://doi.org/10.1016/j.ijforecast.2006.03.001

    Article  Google Scholar 

  • Jekel, M. (2019). Empirical content as a criterion for evaluating models. Cognitive Processing, 20, 273–275. https://doi.org/10.1007/s10339-019-00913-2

    Article  Google Scholar 

  • Kellert, S. H. (2001). Extrascientific uses of physics: The case of nonlinear dynamics and legal theory. Philosophy of Science, 68, S455–S466.

    Article  Google Scholar 

  • Koopmans, M. (2011). Epilogue: Psychology at the edge of chaos. In S. J. Guastello, M. Koopmans, & D. Pincus (Eds.), Chaos and complexity in psychology: The theory of nonlinear dynamical systems. Cambridge University Press.

    Google Scholar 

  • Kruis, J., & Maris, G. (2016). Three representations of the Ising model. Scientific Reports, 6. https://doi.org/10.1038/srep34175

  • Kuppens, P., Champagne, D., & Tuerlinckx, F. (2012a). The dynamic interplay between appraisals and core affect in daily life. Frontiers in Psychology, 3. https://doi.org/10.3389/fpsyg.2012.00380

  • Kuppens, P., Oravecz, Z., & Tuerlinckx, F. (2010). Feelings change: Accounting for individual differences in the temporal dynamics of affect. Journal of Personality and Social Psychology, 99, 1042–1060. https://doi.org/10.1037/a0020962

    Article  Google Scholar 

  • Kuppens, P., Tuerlinckx, F., Russell, J. A., & Feldman-Barrett, L. (2012b). The relation between valence and arousal in subjective experience. Psychological Bulletin, 139, 917–940. https://doi.org/10.1037/a0030811

    Article  Google Scholar 

  • Kuppens, P., & Verduyn, P. (2015). Looking at emotion regulation through the window of emotion dynamics. Psychological Inquiry, 26, 72–79. https://doi.org/10.1080/1047840X.2015.960505

    Article  Google Scholar 

  • Kuppens, P., & Verduyn, P. (2017). Emotion dynamics. Current Opinion in Psychology, 17, 22–26. https://doi.org/10.1016/j.copsyc.2017.06.004

    Article  Google Scholar 

  • Kuranova, A., Booij, S. H., Menne-Lothmann, C., Decoster, J., van Winkel, R., Delespaul, P., De Hert, M., Derom, C., Thiery, E., Rutten, B. P. F., Jacobs, N., van Os, J., Wigman, J. T. W., & Wichers, M. (2020). Measuring resilience prospectively as the speed of affect recovery in daily life: A complex systems perspective on mental health. BMC Medicine, 18. https://doi.org/10.1186/s12916-020-1500-9

  • Larsen, R. J., Augustine, A. A., & Prizmic, Z. (2009). A process approach to emotion and personality: Using time as a facet of data. Cognition and Emotion, 23, 1407–1426. https://doi.org/10.1080/02699930902851302

    Article  Google Scholar 

  • Larson, R., & Csikszentmihalyi, M. (1983). The experience sampling method. In H. T. Reis (Ed.), New directions for methodology of social and behavioral sciences (Vol. 15). Jossey-Bass.

    Google Scholar 

  • Loossens, T., Mestdagh, M., Dejonckheere, E., Kuppens, P., Tuerlinckx, F., & Verdonck, S. (2020). The affective ising model: A computational account of human affect dynamics. PLoS Computational Biology, 16, e1007860. https://doi.org/10.1371/journal.pcbi.1007860

    Article  Google Scholar 

  • Loossens, T., Dejonckheere, E., Tuerlinckx, F., Verdonck, S. (2021). Informing VAR(1) with qualitative dynamical features improves predictive accuracy. Psychological Methods.

    Google Scholar 

  • Losada, M. (1999). The complex dynamics of high performance teams. Mathematical and Computer Modelling, 30, 179–192. https://doi.org/10.1016/S0895-7177(99)00189-2

    Article  Google Scholar 

  • Luce, R. D. (1995). Four tensions concerning mathematical modeling in psychology. Annual Review of Psychology, 46, 1–26.

    Article  Google Scholar 

  • Lütkepohl, H. (2005). New introduction to multiple time series analysis. Springer.

    Book  Google Scholar 

  • Lütkepohl, H., & Xu, F. (2012). The role of the log transformation in forecasting economic variables. Empirical Economics, 42(3), 619–638.

    Article  Google Scholar 

  • Mason, L., Eldar, E., & Rutledge, R. B. (2017). Mood instability and reward dysregulation—A neurocomputational model of bipolar disorder. JAMA Psychiatry, 74, 1275–1276.

    Article  Google Scholar 

  • Masson, M. E. J. (2011). A tutorial on a practical Bayesian alternative to null-hypothesis significance testing. Behavioral Research Methods, 43, 679–690. https://doi.org/10.3758/s13428-0100049-5

    Article  Google Scholar 

  • Moors, A., & Fischer, M. (2019). Demystifying the role of emotion in behavior: Toward a goal-directed account. Cognition & Emotion, 33, 94–100. https://doi.org/10.1080/02699931.2018.1510381

    Article  Google Scholar 

  • Myin-Germeys, I., Kasanova, Z., Vaessen, T., Vachon, H., Kirtley, O., Viechtbauer, W., & Reiningshaus, U. (2018). Experience sampling methodology in mental health research: New insights and technical developments. World Psychiatry, 17, 123–132. https://doi.org/10.1002/wps.20513

    Article  Google Scholar 

  • Myung, J. (2003). Tutorial on maximum likelihood estimation. Journal of Mathematical Psychology, 47, 90–100. https://doi.org/10.1016/S0022-2496(02)00028-7

    Article  Google Scholar 

  • Navarro, D. J., Pitt, M. A., & Myung, I. J. (2004). Assessing the distinguishability of models and the informativeness of data. Cognitive Psychology, 49, 47–84. https://doi.org/10.1016/j.cogpsych.11011.001

    Article  Google Scholar 

  • Nelson, B., McGorry, P. D., Wichers, M., Wigman, J. T. W., & Hartmann, J. A. (2017). Moving from static to dynamic models of the onset of mental disorder: A review. JAMA Psychiatry, 74, 528–534. https://doi.org/10.1001/jamapsychiatry.2017.0001

    Article  Google Scholar 

  • Oravecz, Z., & Brick, T. R. (2019). Associations between slow- and fast-timescale indicators of emotional functioning. Social Psychological and Personality Science, 10, 864–873. https://doi.org/10.1177/1948550618797128

    Article  Google Scholar 

  • Oravecz, Z., Tuerlinckx, F., & Vandekerckhove, J. (2011). A hierarchical latent stochastic differential equation model for affective dynamics. Psychological Methods, 16, 468–490. https://doi.org/10.1037/a0024375

    Article  Google Scholar 

  • Oravecz, Z., Wood, J., & Ram, N. (2018). On fitting a continuous-time stochastic process model in the Bayesian framework. In K. van Monfort, J. H. L. Oud, & M. C. Voelkle (Eds.), Continuous time modeling in the behavioral and related sciences. Springer.

    Google Scholar 

  • Ou, L., Hunter, M. D., & Chow, S.-M. (2019). What’s for dynr: A package for linear and nonlinear dynamic modeling in R. The R Journal, 11. https://doi.org/10.32614/RJ-2019-012

  • Oud, J. H. L. (2007). Continuous time modeling of reciprocal relationships in the cross-lagged panel design. In S. M. Boker & M. J. Wenger (Eds.), Notre dame series on quantitative methodology: Data analytic techniques for dynamical systems. Lawrence Erlbaum Associates Publishers.

    Google Scholar 

  • Oud, J. H. L., & Jansen, R. A. R. G. (2000). Continuous time state space modeling of panel data by means of SEM. Psychometrika, 65, 199–215. https://doi.org/10.1007/BF02294374

    Article  Google Scholar 

  • Pellert, M., Schweighofer, S., & Garcia, D. (2020). The individual dynamics of affective expressions on social media. EPJ Data Science, 9. https://doi.org/10.1140/epjds/s13688-019-0219-3

  • Popper, K. (1959). The logic of scientific discovery. Routledge.

    Google Scholar 

  • Reed, R. G., Barnard, K., & Butler, E. A. (2015). Distinguishing emotional coregulation from codysregulation: An investigation of emotional dynamics and body weight in romantic couples. Emotion, 15(1), 45–60. https://doi.org/10.1037/a0038561

    Article  Google Scholar 

  • Rindskopf, D. (1984). Using phantom and imaginary latent variables to parameterize constraints in linear structural models. Psychometrika, 49(1), 37–47.

    Article  Google Scholar 

  • Roberts, D. R., Bahn, V., Ciuti, S., Boyce, M. S., Elith, J., Guillera-Arroita, G., Hauenstein, S., LahozMonfort, J. J., Schröder, B., Thuiller, W., Warton, D. I., Wintle, B. A., Hartig, F., & Dormann, C. F. (2017). Cross-validation strategies for data with temporal, spatial, hierarchical, or phylogenetic structure. Ecography, 40, 913–929. https://doi.org/10.1111/ecog.02881

    Article  Google Scholar 

  • Rojas, R. (1996). Neural networks: A systematic introduction. Springer.

    Book  Google Scholar 

  • Rutledge, R. B., Moutoussis, M., Smittenaar, P., Zeidman, P., Taylor, T., Hrynkiewicz, L., Lam, J., Skandali, N., Siegel, J. Z., Ousdal, O. T., Prabhu, G., Dayan, P., Fonagy, P., & Dolan, R. J. (2017). Association of neural and emotional impacts of reward prediction errors with major depression. JAMA Psychiatry, 74, 790–797. https://doi.org/10.1001/jamapsychiatry.2017.1713

    Article  Google Scholar 

  • Rutledge, R. B., Skandali, N., Dayan, P., & Dolan, R. J. (2014). A computational and neural model of momentary subjective well-being. Proceedings of the National Academy of Sciences, 111(33), 12252–12257. https://doi.org/10.1073/pnas.1407535111

    Article  Google Scholar 

  • Ryan, O., Kuiper, R. M., & Hamaker, E. L. (2018). A continuous time approach to intensive longitudinal data: What, why, and how? In K. van Montfort, J. H. L. Oud, & M. C. Voelkle (Eds.), Continuous time modeling in the behavioral and related sciences. Springer.

    Google Scholar 

  • Sacharin, V., Sander, D., & Scherer, K. R. (2012). The perception of changing emotion expressions. Cognition & Emotion, 26, 1273–1300. https://doi.org/10.1080/02699931.2012.656583

    Article  Google Scholar 

  • Sander, D., Grandjean, D., & Scherer, K. R. (2005). A systems approach to appraisal mechanisms in emotion. Neural Networks, 18, 317–352. https://doi.org/10.1016/j.neunet.2005.03.001

    Article  Google Scholar 

  • Santangelo, P. S., Limberger, M. F., Stiglmayr, C., Houben, M., Coosemans, J., Verleysen, G., Kuppens, P., Tuerlinckx, F., Vanpaemel, W., & Ebner-Priemer, U. W. (2016). Analyzing subcomponents of affective dysregulation in borderline personality disorder in comparison to other clinical groups using multiple e-diary datasets. Borderline Personality Disorder and Emotion Dysregulation, 3, 5. https://doi.org/10.1186/s40479-016-0039-z

    Article  Google Scholar 

  • Scargle, J. D. (1981). Studies in astronomical time series analysis: Modeling random processes in the time domain. Astrophysical Journal Supplement Series, 45, 1–71.

    Article  Google Scholar 

  • Scherer, K. R. (2000). Emotions as episodes of subsystem synchronization driven by nonlinear appraisal processes. In M. Lewis & I. Granic (Eds.), Emotion, development, and self-organization: Dynamic systems approaches to emotional development (Cambridge Studies in Social and Emotional Development). Cambridge University Press. https://doi.org/10.1017/CBO9780511527883.005

    Chapter  Google Scholar 

  • Scherer, K. R. (2005). What are emotions? And how can they be measured? Social Science Information, 44, 695–729. https://doi.org/10.1177/0539018405058216

    Article  Google Scholar 

  • Schimmack, U. (2001). Pleasure, displeasure, and mixed feelings: Are semantic opposites mutually exclusive? Cognition and Emotion, 15, 81–97. https://doi.org/10.1080/02699930126097

    Article  Google Scholar 

  • Schwarz, G. (1978). Estimating the dimension of a model. The Annals of Statistics, 6, 461–464.

    Article  Google Scholar 

  • Schweitzer, F., & Garcia, D. (2010). An agent based model of collective emotions in online communities. The European Physical Journal, 77, 533–545. https://doi.org/10.1140/epjb/e2010-00292-1

    Article  Google Scholar 

  • Seeley, S. H., Garcia, E., & Mennina, D. S. (2015). Recent advances in laboratory assessment of emotion regulation. Current Opinion in Psychology, 3, 58–63. https://doi.org/10.1016/j.copsyc.2015.02.009

    Article  Google Scholar 

  • Smaldino, P. E. (2017). Models are stupid, and we need more of them. In R. R. Vallacher, S. J. Read, & A. Nowak (Eds.), Computational social psychology. Routledge.

    Google Scholar 

  • Smith, K. E., Crosby, R. D., Wonderlich, S. A., Forbush, K. T., Mason, T. B., & Moessner, M. (2018). Network analysis: An innovative framework for understanding eating disorder psychopathology. International Journal of Eating Disorders, 51, 214–222. https://doi.org/10.1002/eat.22836

    Article  Google Scholar 

  • Steele, J. S., & Ferrer, E. (2011). Latent differential equation modeling of self-regulatory and coregulatory affective processes. Multivariate Behavioral Research, 46, 956–984. https://doi.org/10.1080/00273171.2011.625305

    Article  Google Scholar 

  • Steele, J. S., Gonzales, J. E., & Ferrer, E. (2018). Uses and limitations of continuous-time models to examine dyadic interactions. In K. van Montfort, J. H. L. Oud, & M. C. Voelkle (Eds.), Continuous time modeling in the behavioral and related sciences.

    Google Scholar 

  • Strogatz, S. (2018). Nonlinear dynamics and chaos: With applications to physics, biology, chemistry, and engineering. CRC Press.

    Book  Google Scholar 

  • Sussmann, H. J., & Zahler, R. S. (1978). A critique of applied catastrophe theory in the behavioral sciences. Behavioral Science, 23, 383–389.

    Article  Google Scholar 

  • Sutton, R. S., & Barto, A. G. (2018). Reinforcement learning: An introduction. MIT Press.

    Google Scholar 

  • Tashman, L. J. (2000). Out-of-sample tests of forecasting accuracy: An analysis and review. International Journal of Forecasting, 16, 437–450.

    Article  Google Scholar 

  • Thagard, P., & Nerb, J. (2002). Emotional gestalts: Appraisal, change, and the dynamics of affect. Personality and Social Psychology Review, 6, 274–282.

    Article  Google Scholar 

  • Thom, R. (1975). Structural stability and morphogenesis (D. H. Fowler, Trans.) [(Original work published in 1972)]. W. A. Benjamin.

    Google Scholar 

  • Tong, H. (2011). Threshold models in time series analysis—30 years on. Statistics and its Interface, 4(2), 107–118.

    Article  Google Scholar 

  • Uhlenbeck, G. E., & Ornstein, L. S. (1930). On the theory of Brownian motion. Physical Review, 46, 823–841.

    Article  Google Scholar 

  • van de Leemput, I. A., Wichers, M., Cramer, A. O. J., Borsboom, D., Tuerlinckx, F., Kuppens, P., van Nes, E. H., Viechtbauer, W., Giltay, E. J., Aggen, S. H., Derom, C., Jacobs, N., Kendler, K. S., van der Maas, H. L. J., Neale, M. C., Peeters, F., Thiery, E., Zachar, P., & Scheffer, M. (2014). Critical slowing down as early warning for the onset and termination of depression. Proceedings of the National Academy of Sciences, 111, 87–92. https://doi.org/10.1073/pnas.1312114110

    Article  Google Scholar 

  • Vandekerckhove, J., Matzke, D., & Wagenmakers, E.-J. (2015). Model comparison and the principle of parsimony. In J. R. Busemeyer, Z. Whang, J. T. Townsend, & A. Eidels (Eds.), The Oxford handbook of computational and mathematical psychology. Oxford University Press.

    Google Scholar 

  • Vanhasbroeck, N., Devos, L., Pessers, S., Kuppens, P., Vanpaemel, W., Moors, A., & Tuerlinckx, F. (2021). Testing a computational model of subjective well-being: A preregistered replication of Rutledge et al. (2014). Cognition and Emotion, 35(4), 822–835. https://doi.org/10.1080/02699931.2021.1891863.

  • Velicer, W. F., & Molenaar, P. C. (2012). Time series analysis for psychological research. In Handbook of psychology, second edition. Society. https://doi.org/10.1002/9781118133880.hop202022

    Chapter  Google Scholar 

  • Verduyn, P., Van Mechelen, I., Tuerlinckx, F., Meers, K., & Van Coillie, H. (2009). Intensity profiles of emotional experience over time. Cognition and Emotion, 23, 1427–1443. https://doi.org/10.1080/02699930902949031

    Article  Google Scholar 

  • Villano, W. J., Otto, A. R., Ezie, C. E. C., Gillis, R., & Heller, A. S. (2020). Temporal dynamics of real world emotions are more strongly linked to prediction error than outcome. Journal of Experimental Psychology: General, 149, 1755–1766. https://doi.org/10.1037/xge0000740

    Article  Google Scholar 

  • Vinckier, F., Rigoux, L., Oudiette, D., & Pessiglione, M. (2018). Neuro-computational account of how mood fluctuations arise and affect decision-making. Nature Communications, 9. https://doi.org/10.1038/s41467-018-03774-z

  • Voelkle, M. C., Ebner, N. C., Lindenberger, U., & Riediger, M. (2013). Here we go again: Anticipatory and reactive mood responses to recurring unpleasant situations throughout adulthood. Emotion, 13, 424–433. https://doi.org/10.1037/a0031351

    Article  Google Scholar 

  • Voelkle, M. C., & Oud, J. H. L. (2013). Continuous time modelling with individually varying time intervals for oscillating and non-oscillating processes. British Journal of Mathematical and Statistical Psychology, 66, 103–126. https://doi.org/10.1111/j.2044-8317.2012.02043.x

    Article  Google Scholar 

  • Wagenmakers, E.-J., & Farrell, S. (2004). AIC model selection using Akaike weights. Psychonomic Bulletin & Review, 11, 192–196.

    Article  Google Scholar 

  • Wagenmakers, E.-J., Molenaar, P. C. M., Grasman, R. P. P. P., Hartelman, P. A. I., & van der Maas, H. L. J. (2005). Transformation invariant stochastic catastrophe theory. Physica D, 211, 263–276. https://doi.org/10.1016/j.physd.2005.08.014

    Article  Google Scholar 

  • Wagenmakers, E.-J., Ratcliff, R., Gomez, P., & Iverson, G. J. (2004). Assessing model mimicry using the parametric bootstrap. Journal of Mathematical Psychology, 48, 28–50. https://doi.org/10.1016/j.jmp.2003.11.004

    Article  Google Scholar 

  • Watts, D. J., & Strogatz, S. H. (1998). Collective dynamics of ‘small-world’ networks. Nature, 393, 440–442.

    Article  Google Scholar 

  • Wehrens, R., Putter, H., & Buydens, L. M. C. (2000). The bootstrap: A tutorial. Chemometrics and Intelligent Laboratory Systems, 54, 35–52.

    Article  Google Scholar 

  • Wichers, M., Smit, A. C., & Snippe, E. (2020). Early warning signals based on momentary affect dynamics can expose nearby transitions in depression: A confirmatory single-subject time-series study. Journal for Person-Oriented Research, 6, 1–15. https://doi.org/10.17505/jpor.2020.22042

    Article  Google Scholar 

  • Wood, J., Oravecz, Z., Vogel, N., Benson, L., Chow, S.-M., Cole, P., Conroy, D. E., Pincus, A. L., & Ram, N. (2018). Modeling intraindividual dynamics using stochastic differential equations: Age differences in affect regulation. Journal of Gerontology: Psychological Sciences, 73, 171–184. https://doi.org/10.1093/geronb/gbx013

    Article  Google Scholar 

  • Zeeman, E. C. (1976). Catastophe theory. Scientific American, 234, 65–83. https://doi.org/10.1038/scientificamerican0476-65

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Niels Vanhasbroeck .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Properties of the VAR

1.1 Properties of the AR Model

Given the AR model (repeated here):

$$ {y}_j=\delta +\varphi {y}_{j-1}+{\varepsilon}_j. $$

we can define some properties of the process. These properties are defined and mathematically derived below.

Predictions.Given a first observation y0 collected at time t0, we are able to predict the next measurement y1, as:

$$ {\displaystyle \begin{array}{cc}\left\langle {y}_1\right.\left|{y}_0\right\rangle & =\delta +\varphi {y}_0+{\varepsilon}_1\\ {}& =\delta +\varphi {y}_0\end{array}} $$

where denotes the time-dependent expected value (i.e., E[.]). Note that the innovations do not play a role in the expectation of y1, given that their expected value is equal to 0.

Using the same principle, we can also make predictions about observations further in the future. For instance, the expectation of y2 conditional on the observation y0 is given by:

$$ {\displaystyle \begin{array}{cc}\left\langle {y}_2\right.\left|{y}_0\right\rangle & =\delta +\varphi {y}_1\\ {}& =\delta +\varphi \left(\delta +\varphi {y}_0\right)\\ {}& =\left(1+\varphi \right)\delta +{\varphi}^2{y}_0\end{array}} $$

In general, the prediction of a future observation yj conditional on y0 is:

$$ \left\langle {y}_j\right.\left|{y}_0\right\rangle =\left(\sum \limits_{k=0}^{j-1}{\varphi}^k\right)\delta +{\varphi}^j{y}_0 $$
(10.7)

Baseline. Since the magnitude of φj shrinks as j increases, in the long-time limit, it holds that:

$$ li{m}_{j\to \infty}\left(\sum \limits_{j-1}^{k=0}{\varphi}^k\right)=\frac{1}{1-\varphi} $$

As a result, the predictions ⟨yj| y0⟩ converge towards a fixed point:

$$ {\displaystyle \begin{array}{cc} li{m}_{j\to \infty}\left(\left\langle {y}_1\right.\left|{y}_0\right\rangle \right)& = li{m}_{j\to \infty}\left(\left(\sum \limits_{k=0}^{j-1}{\varphi}^k\right)\delta +{\varphi}^j{y}_0\right)\\ {}& =\frac{1}{1-\varphi}\delta +0\\ {}& =\frac{\delta}{1-\varphi}\\ {}& =\mu \end{array}} $$
(10.8)

This fixed point μ can be considered the emotional baseline (i.e., the dotted line in Fig. 10.1) and represents the emotional state to which the emotional state is expected to evolve to. As such coincides with the end state of a regulation process (provided nothing happens to disrupt the regulation process).

It also represents the state that will be visited the most by the individual over longer periods of time. For that reason, it coincides with the mean of the distribution of observations {yj | j,…,N} for sufficiently large N. Because of this, the baseline is also referred to as the stationary mean. The term stationary is used to stress that the baseline is independent of time.

When an AR(1) process is only observed during a short period of time during which the emotional state is still relaxing (i.e., converging) towards the baseline, then the mean of the observations will differ from the stationary mean. Only when measurements have been collected for a sufficiently long period of time will the mean of the data distribution coincide with the stationary mean.

Uncertainty. Until now, we were only concerned with point-predictions of future observations. However, we can also compute the uncertainty that is associated with these predictions. For this, we realize that the observation y1 is normally distributed with mean δ + φy0(the prediction) and variance σε2:

$$ {y}_1\mid {y}_0\sim N\left(\delta +\varphi {y}_0,{\sigma}_{\varepsilon}^2\right) $$

Because of stochasticity, uncertainty about predictions typically grows the further in the future you go. It can be shown that the future observation yj, given observation y0, is normally distributed with the mean being the point-prediction in Eq. (10.7) and variance given by:

$$ {\sigma}_j^2=\sum \limits_{j-1}^{k=0}{\varphi}^{2k}{\sigma}_{\varepsilon}^2 $$
(10.9)

where in the long-time limit:

$$ li{m}_{j\to \infty}\left(\sum \limits_{j-1}^{k=0}{\varphi}^{2k}\right)=\frac{1}{1-{\varphi}^2} $$

so that the variance of the uncertainty distribution in the long-time limit converges to

$$ {\sigma}^2=\frac{\sigma_{\varepsilon}^2}{1-{\varphi}^2} $$
(10.10)

Like the stationary mean, this variance is time-independent and thus called the stationary variance.

Autocovariance. An AR model relies on the assumption that measurements yj at time tj are related to measurements yj − 1 at time tj − 1, i.e. that there is a time-dependence between measurements. The extent to which this relationship holds is expressed by the autocovariance. The autocovariance at lag-p σp is defined as:

$$ {\sigma}_p=\left\langle \left({y}_{j+p}-\mu \right)\right.\left.\left({y}_j-\mu \right)\right\rangle $$

To compute the autocovariance of the AR process, we first reformulate the model in terms of the baseline μ. To do so, we substitute δ for (1 − φ)μ (see Eq. (10.8)) to obtain:

$$ {\displaystyle \begin{array}{c}{y}_j=\left(1-\varphi \right)\mu +\varphi {y}_{j-1}+{\varepsilon}_j\\ {}=\mu -\varphi \mu +\varphi {y}_{j-1}+{\varepsilon}_j\end{array}} $$

Then, by rearranging the terms, we can write

$$ {y}_j-\mu =\varphi \left({y}_{j-1}-\mu \right)+{\varepsilon}_j. $$

Setting the innovations to zero (they do not correlate with anything), we find (see Eq. (10.7))

$$ {\displaystyle \begin{array}{cc}{\sigma}_p& =\left\langle \left({y}_{j+p}-\mu \right)\right.\left.\left({y}_j-\mu \right)\right\rangle \left(\mathrm{Def}.\mathrm{autocovariance}\right)\\ {}& =\left\langle {\varphi}^p\right.\left({y}_j-\mu \right)\left.\left({y}_j-\mu \right)\right\rangle \left(\mathrm{Generalization}\kern0.17em \mathrm{previous}\kern0.17em \mathrm{property}\right)\\ {}& =\left\langle {\varphi}^p{\left({y}_j-\mu \right)}^2\right\rangle \\ {}& ={\varphi}^p{\sigma}^2\end{array}} $$
(10.11)

Here we have used the fact that the centered variable yj − μ have the same stationary variance (Eq. (10.10)) as the variable yj themselves. If we standardize the measurements so that σ2 = 1, we obtain the autocorrelation:

$$ \rho (p)={\varphi}^p $$

From this expression, it can be seen that the autoregressive coefficient φ of the AR model corresponds to the autocorrelation between measurements at lag 1.

1.2 Properties of the VAR Model

We can generalize the properties of the AR model to fit the d-dimensional VAR model (repeated here):

$$ {\displaystyle \begin{array}{c}{\boldsymbol{y}}_j=\boldsymbol{\delta} +\varPhi {\boldsymbol{y}}_{j-1}+{\boldsymbol{\varepsilon}}_j\\ {}{\boldsymbol{\varepsilon}}_j\sim N\left(\mathbf{0},{\varSigma}_{\varepsilon}\right)\end{array}} $$

Predictions. Just like for the AR model, the prediction of a future observation conditional on the observation y0 is given by (see Eq. (10.7))

$$ \left\langle {y}_j|{\mathrm{y}}_0\right\rangle =\left(\sum \limits_{k=0}^{j-1}{\varPhi}^k\right)\boldsymbol{\delta} +{\varPhi}^j{\boldsymbol{y}}_0 $$
(10.12)

Importantly, this equation results in a vector that contains all expectation values for all d variables of the model.

Baseline. Using a similar reasoning as for the AR model (see Eq. (10.8)), but this time using matrices instead of scalars, it can be shown that the predictions of the VAR model Eq. (10.12) converge to the baseline:

$$ \boldsymbol{\mu} ={\left({I}_d-\varPhi \right)}^{-1}\boldsymbol{\delta} $$

where Id is the d-dimensional identity matrix.

Uncertainty. An expression similar to Eq. (10.9) can be obtained for the growing uncertainty of the VAR model:

$$ {\varSigma}_j=\sum \limits_{j-1}^{k=0}{\varPhi}^k{\varSigma}_{\varepsilon}{\left({\varPhi}^k\right)}^T $$

For stable transition matrices Φ, this covariance matrix becomes constant in the long-time limit. This stationary covariance is given by

$$ \varSigma =\sum \limits_{\infty}^{k=0}{\varPhi}^k{\varSigma}_{\varepsilon}{\left({\varPhi}^k\right)}^T $$

and is a solution of the discrete-time Lyapunov equation

$$ \varSigma -\varPhi \varSigma {\varPhi}^T={\varSigma}_{\varepsilon} $$

Given the transition matrix Φ and the covariance Σε of the innovations, this Lyapunov equation enables us to compute the stationary covariance without having to compute an infinite sum.

Autocovariance. The autocovariance of the VAR model is similar to the autocovariance of the AR (see Eq. (10.11), namely

$$ {\displaystyle \begin{array}{cc}{\varSigma}_p& =\left\langle \left({y}_{j+p}-\mu \right){\left({y}_j-\mu \right)}^T\right\rangle \\ {}& ={\varPhi}^p\varSigma \end{array}} $$

Appendix 2: Autocorrelation of Bivariate VAR

If we take a bivariate VAR model with the intercepts δ = 0, then we can compute the autocovariance as:

$$ {\displaystyle \begin{array}{cc}{\sigma}_{tt-1}& =\left({y}_{1t}{y}_{1t-1}\right)\\ {}& =\left({\delta}_1+{\varphi}_{11}{y}_{1t-1}+{\varphi}_{12}{y}_{2t-1}+{\varepsilon}_{1t}\right){y}_{1t-1}\\ {}& ={\delta}_1{y}_{1t-1}+{\varphi}_{11}{y}_{1t-1}^2+{\varphi}_{12}{y}_{2t-1}{y}_{1t-1}+{\varepsilon}_{1t}\\ {}& ={\delta}_1{y}_{1t-1}+{\varphi}_{11}{y}_{1t-1}^2+{\varphi}_{12}{y}_{2t-1}{y}_{1t-1}+{\varepsilon}_{1t}\\ {}& =0+{\varphi}_{11}{\sigma}_1^2+{\varphi}_{12}{\sigma}_{12}+0\\ {}& ={\varphi}_{11}{\sigma}_1^2+{\varphi}_{12}{\sigma}_{12}\end{array}} $$

We can compute the autocorrelation as:

$$ {\displaystyle \begin{array}{cc}{\rho}_{tt-1}& =\frac{\sigma_{tt-1}}{\sigma_{1t}{\sigma}_{1t-1}}\\ {}& =\frac{\sigma_{tt-1}}{\sigma_1^2}\\ {}& =\frac{\varphi_{11}{\sigma}_1^2+{\varphi}_{12}{\sigma}_{12}}{\sigma_1^2}\\ {}& ={\varphi}_{11}+\frac{\varphi_{12}{\sigma}_{12}}{\sigma_1^2}\end{array}} $$

More generally, it holds that for a variable yi:

$$ {\displaystyle \begin{array}{cc}{\sigma}_{tt-1}& ={y}_{it}{y}_{it-1}\\ {}& =\left({\delta}_i+{\varphi}_{ii}{y}_{it-1}+\sum \limits_{\begin{array}{c}{i}^{\prime}=1\\ {}{i}^{\prime}\ne i\end{array}}^d{\varphi}_{i{i}^{\prime}}{y}_{i^{\prime}t-1}+{\varepsilon}_{it}\right){y}_{it-1}\\ {}& ={\varphi}_{ii}{\sigma}_i^2+\sum \limits_{\begin{array}{c}{i}^{\prime}=1\\ {}{i}^{\prime}\ne i\end{array}}^d{\varphi}_{i{i}^{\prime}}{\sigma}_{i{i}^{\prime}}\end{array}} $$

and:

$$ {\rho}_{tt-1}={\varphi}_{ii}+\sum \limits_d^{\begin{array}{c}{i}^{\prime}=1\\ {}{i}^{\prime}\ne i\end{array}}\frac{\varphi_{i{i}^{\prime}}{\sigma}_{i{i}^{\prime}}}{\sigma_i^2} $$

Rights and permissions

Reprints and permissions

Copyright information

© 2021 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Vanhasbroeck, N., Ariens, S., Tuerlinckx, F., Loossens, T. (2021). Computational Models for Affect Dynamics. In: Waugh, C.E., Kuppens, P. (eds) Affect Dynamics. Springer, Cham. https://doi.org/10.1007/978-3-030-82965-0_10

Download citation

Publish with us

Policies and ethics