469
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Applying dynamic systems theory and complexity theory methods in psychotherapy research: A systematic literature review

ORCID Icon, ORCID Icon & ORCID Icon
Pages 828-844 | Received 31 Jan 2023, Accepted 21 Aug 2023, Published online: 31 Aug 2023

Abstract

Objective

Dynamic systems theory and complexity theory (DST/CT) is a framework explaining how complex systems change and adapt over time. In psychotherapy, DST/CT can be used to understand how a person's mental and emotional state changes during therapy incorporating higher levels of complexity. This study aimed to systematically review the variability of DST/CT methods applied in psychotherapy research.

Methods

A primary studies search was conducted in the EBSCO and Web of Knowledge databases, extracting information about the analyzed DST/CT phenomena, employed mathematical methods to investigate these phenomena, descriptions of specified dynamic models, psychotherapy phenomena, and other information regarding studies with empirical data (e.g., measurement granularity).

Results

After screening 38,216 abstracts and 4,194 full texts, N = 41 studies published from 1990 to 2021 were identified. The employed methods typically included measures of dynamic complexity or chaoticity. Computational and simulation studies most often employed first-order ordinary differential equations and typically focused on describing the time evolution of client-therapist dyadic influences. Eligible studies with empirical data were usually based on case studies and focused on data with high time intensity of within-session dynamics.

Conclusion

This review provides a descriptive synthesis of the current state of the proliferation of DST/CT methods in the psychotherapy research field.

Clinical or methodological significance of this article: This review describes the use of mathematical methods in understanding the complexity of psychotherapy dynamics. These methods could be useful tools to build detailed theories about how therapy progresses, including how different aspects change as therapy goes on, which might produce beneficial insight for both therapists and clients. These theories could show for instance how things shift from one phase to another in a particular client's system. However, proving these ideas with real therapy data is tricky due to the limited information that could be gathered during sessions and the unpredictable nature of how the change is measured.

Psychotherapy is a complex activity that includes an ongoing and multifaceted interaction of many variables evolving over time. Both clinical experience and research studies inform us that there are phenomena in the therapeutic process, such as sudden gains (Olthof et al., Citation2020a), that defy being modeled in a straightforward and linear manner. Change in psychotherapy is often nonlinear (Guastello et al., Citation2008), and linear models represent rough approximations that cannot do justice to the intricate nature of the therapeutic process. Therefore, several authors have suggested treating psychotherapy as a complex dynamic system that develops over time (e.g., Gelo & Salvatore, Citation2016; Schiepek et al., Citation1997).

General systems theory (GST) is an overarching interdisciplinary theory dealing with the study of systems—models where the whole is composed of interconnected parts (Boulding, Citation1956). GST proposes that systems have common properties across different scientific fields and can be studied generally (von Bertalanffy, Citation1968). The development of the systems and their components is studied in a subfield dynamic systems theory (DST). DST methods usually describe the development of complex systems in terms of sets of difference (discrete) or differential (continuous) equations. Most suitable differential equations do not have analytical solutions readily available and require the use of advanced mathematical tools to be workable (Strogatz, Citation2018).

An important group of models derived from GST are the complex systems, the subject of complexity theory (CT), and its related fields (such as chaos theory). While stable systems react to changes in conditions (i.e., perturbation) by returning to a previous state, complex systems behave adaptively, self-organizing into new emerging states (so-called phase transition; Strathern & McGlade, Citation2014). Chaos theory represents a bridge between deterministic and stochastic conceptions of reality studying unpredictable systems with sensitive dependence on initial conditions which can occur even for deterministic and relatively simple models (Sivakumar, Citation2004).

The most common paradigm in dealing with complexity is reductionism—attempting to decrease the complexity by studying the parts of the system separately or simplifying the relationships between them (e.g. linearization). While this approach led to several great scientific advances, especially in natural sciences (e.g. classical mechanics in physics), it can lead to a critical loss of information when dealing with more complex systems and their emerging behavior. Evidence of these irreducible complex systems can be found both in natural (Strogatz, Citation2018) and social sciences (Mellado et al., Citation2022) creating a need for adoption and focus on complexity theory.

From the DST/CT perspective, psychotherapy can be described via several distinctive but universally recognized complex dynamic system features that include (a) a large number of interrelated components or agents (e.g., therapeutic factors, interacting structures) that can produce emerging self-organizing functioning at the whole-system level; (b) nonlinear temporal dynamics and self-emerging structures while adapting to the environment; and (c) occasionally stochastic nature and an emergent structure that cannot always be traced back to the initial elements of the system (McGill et al., Citation2020). The eventual outcome of the treatment process may vary greatly even with a slight change in the initial conditions (Schiepek et al., Citation2017).

In psychotherapy, studying change processes through repeated measurements of variables is essential. This can reveal DST/CT phenomena such as critical fluctuations, sudden gains or losses, and potential chaotic behavior in the therapeutic systems. Macroscopic patterns may emerge, influenced by order and control parameters (the former are measurable quantities characterizing emerging patterns, and the latter are parameters influencing the system’s behavior), which are not visible at the microscopic level. This self-organization process, a core concept of synergetics (Haken & Tschacher, Citation2017), involves the system evolving towards preferred or stable states represented by attractors. These states are shaped by control and order parameters, guiding the system's dynamics and potentially changing attractors over time (Schiepek et al., Citation2017). It should also be noted that not all psychotherapeutic change processes must be complex in nature. The “true” behavior may be more regular, linear, and deterministic. However, while some of these systems may be deterministic (i.e., fully predictable from the initial conditions), others may include a stochastic component (i.e., randomness inherent in the system behavior).

In their literature review, Mellado et al. (Citation2022) summarized results obtained from studies employing DST/CT and the DST/CT methods used to operationalize therapeutic change. However, their study (1) was limited to empirical studies only and excluded simulation and mathematical studies (i.e., studies based on simulated data only and studies that only specified a mathematical model), and (2) they did not focus on the specification of the mathematical models used in the primary studies.

Study Aims

A systematic review can bridge the gap between theory and practice, synthesizing literature and providing an overview of DST/CT methods in psychotherapy research. Traditional linear models often fall short of capturing the complexity of psychotherapeutic processes. By reviewing DST/CT methods, we can identify approaches that better capture this complexity within one easily accessible resource for researchers, practitioners, or students, distilling complex mathematical concepts into a cumulative format.

The present systematic review built on Mellado et al.’s (Citation2022) study and aimed to synthesize DST/CT methods employed in existing studies to model empirical or simulated time-series data in the context of psychotherapy. Other descriptive data extracted from each included study were utilized to provide more information about the specific context of the use of these methods. Moreover, the preregistration and online supplemental materials (https://osf.io/k3845/) accompanying this review include detailed and more precise descriptions of methods (e.g., search and selection process, all extracted information) and results (e.g., phenomena, specifications of mathematical methods).

Methods

Selection of Primary Studies

This review adheres to the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) guidelines (Moher et al., Citation2015). We used the following databases to search for articles in 06/2021: Academic Search Ultimate, Academic Search Complete, PsycINFO, SocINDEX with Full Text, PsycARTICLES, eBook Academic Collection, MathSciNet, and Web of Science. See Supplement 1 for the search string and decisions made during the search process. We considered studies eligible if they (1) employed any DCT/CT modeling approach (only “hard systems” were included, cf. Carey et al., Citation2015; studies describing qualitative, action-based methodologies termed “soft systems” and studies using only noise reduction or information processing methods were excluded); (2) focused on psychotherapy and counseling phenomena (we excluded studies focusing on psychopharmacotherapy, biological, chemical, and brain research, as well as purely mathematical studies); (3) included either real or simulated data, or at least a precise mathematical model (i.e., we excluded theoretical studies); and (4) focused on adult humans (i.e., we excluded studies focused on children or animals).

To reduce the potential for publication bias, we included gray literature (i.e., unpublished or nonpeer-reviewed studies accessible in EBSCO Open Dissertations). While no restrictions were placed on the date of publication, only articles written in English were included.

The screening process was divided into two phases: abstract and full-text screening. The abstract screening was conducted by the first author. During this phase, only articles that failed to meet the inclusion criteria were excluded; all remaining articles were kept for full-text inspection. The full-text screening was conducted by the first two authors (each screened half of the articles independently; ambiguous articles were discussed by both authors until a consensus was reached).

Data Extraction and Synthesis

We extracted four broad arrays of information from the primary studies: (1) basic descriptive data; (2) analytical/mathematical methods and models used; (3) psychotherapeutic/psychological phenomena that were analyzed; (4) information about empirical data (if collected in the study), and (5) auxiliary information. See Supplement 2 for the detailed list of all aspects that were coded (note that we use a different “numeral system” in the supplementary materials). The extracted information was coded into a spreadsheet file (see Supplement 3). Data extraction was conducted by the first and second authors discussing each included study until a consensus was reached. Extracted content (i.e., topics emerging from the primary studies) was later clustered into more general categories based on content similarity (using a qualitative content analysis)—usually, the resulting categories were highly interwoven and the following division was rather artificial (mainly regarding DST/CT phenomena). During the data synthesis, extracted information was aggregated using a narrative and interpretative synthesis to gain insight across studies for which aggregation was possible and relevant.

Results

Out of the 38,216 articles identified during the search process, we excluded 34,022 articles in the abstract screening phase and an additional 4,040 articles in the full-text screening phase. This resulted in 143 articles that met the initial inclusion criteria. Since this considerably exceeded the number of studies suitable for a narrative and interpretive synthesis, we further decided to exclude all studies that did not include any psychotherapy intervention directly (n = 102), resulting in a final number of 41 studies (see ). See Supplement 4 for the flow diagram.

Table I. Descriptive information about all studies included in the review.

Descriptive Information

The included studies were published between 1990 and 2021 with a median of 2015. Interestingly, the publication dates have a bimodal distribution, as no studies were published between 2000 and 2005. All studies focused on the application of models to time-series data. Out of the total sample of 41 studies, 24 used their own empirical data, nine studies used only an empirical example from a previous study, and eight studies were purely in silico studies. One study used simulations without properly specifying the mathematical model (Laramée et al., Citation2016, #14). Six of the empirical studies also included simulations. Two studies were neither empirical nor a simulation study but were focused on the specification of the mathematical models only (Haken & Tschacher, Citation2017, #10; Tschacher et al., Citation2015a, #38). Twenty-one of the empirical studies follow the procedure of specifying a research question first, then collecting data, and finally selecting a suitable DST/CT method to answer the research question. However, the remaining 12 studies specified a theoretical and mathematical model first and then used empirical data to showcase the model. Moreover, out of the 21 “data first, model after” studies, 7 used DST/CT tools only to create variables for further linear, noncomplex analyses, such as means-comparison (Helmich et al., Citation2020, #11; Lichtenberg & Knox, Citation1991, #18; Olthof et al., Citation2020a, #19, Citation2020b, #20; Schiepek et al., Citation2014, #28; Stapelberg, Citation2016, #36; Tschacher et al., Citation1998, #40).

Furthermore, 21 studies contained comprehensive theoretical and conceptual explanations of complexity-related concepts or methods used, while the remaining 20 were written in a more applied manner already assuming certain prior knowledge of DST/CT by the reader. See and Supplement 5 for more detailed descriptive information about all included studies (specifically see Supplement 5a for a distribution of studies divided according to countries of origin and year of publication and Supplement 5b for a figure showing in which journals were the included studies published).

Analytical/Mathematical Methods and Models

Analyzed complexity phenomena

Temporal dynamics are a necessary prerequisite for any dynamic system (Ryo et al., Citation2019). Several authors described the temporal evolution of the system variables in the included studies and/or focused on specific trajectories of the system (n = 13) or focused on the decrease in the amplitude of the time signal oscillation (n = 4) due to the decay of change over time or due to external influences, such as intrapersonal damping.

Self-organization is a topic in complex dynamic systems and represents the spontaneous emergence of organized behavior in response to certain conditions (Heylighen, Citation2008). Eight studies directly targeted self-organization, five studies focused on the identification of stable attractors (i.e., a tendency of the system to remain within a certain range of values even after perturbation), and four studies focused on the identification of the emergence of order from disorder unfolding in time (e.g., embedded patterns of order within and across psychotherapeutic sessions as a counterpart to complexity; Heylighen, Citation2008).

Emergence is a term tightly interconnected with self-organization (being a subset of it). In this review, we divided emergence into several categories: the emergence of phases with possible phase transitions (n = 12), the emergence of critical instability and sudden shifts, such as detecting early warning signals (n = 11), the emergence of synchronous states, such as the degree of synchrony between the client and the therapist (n = 7), and the emergence of chaotic patterns with sensitivity to initial conditions (n = 6). For more detailed information about the extracted DST/CT phenomena and the number of studies that used them, see Supplement 6 (text) and for tabulated data specifically Supplement 6a.

Mathematical methods

Of the 41 studies, 22 employed purely deterministic DST/CT methods, 14 combined deterministic approaches with some stochastic elements, and two studies used only stochastic methods (Lichtenberg & Knox, Citation1991, #18; Stapelberg, Citation2016, #36).

As far as the particular techniques used to analyze the DST/CT phenomena are considered, these included agent-based models (n = 2, see Supplement 7a), methods for modeling stochastic processes (i.e., spectral analysis, n = 4; Shannon entropy, n = 4; autocorrelations, n = 5, see Supplement 7b), methods measuring dynamic complexity (n = 10, see Supplement 7c), and measuring chaoticity (n = 4, see Supplement 7d). Agent-based models were used to simulate interactions between individual agents in a given system over time (Brice, Citation2019, #3; Lauro Grotto et al., Citation2014, #15). Methods dealing with stochastic processes were used to describe the evolution of a system over time and identify patterns in the data. Dynamic complexity studies time-dependent fluctuations and changes in distributions and was used as a pragmatic alternative to other measures of time-series complexity that require larger numbers of data points (e.g., correlation dimension, Schiepek et al., Citation2017, #29). The chaoticity and stability of the system are usually measured by the local largest Lyapunov exponent. These methods are often used to transform raw data into time series of complexity/instability measures, enabling better identification of phase-space/order transitions. Typically, studies used a combination of methods rather than a single method. For more detailed information about the methods, see Supplement 7 in general.

Mathematical models based on differential equations

Sets of differential equations are used to describe time-dependent complex system behavior. A differential equation describes a relationship between dependent and independent (often time) variables and contains derivatives of dependent variables (Zill, Citation2012). Numerical methods allow us to conduct simulated experiments even without collecting empirical data. Solving a differential equation requires knowledge of the initial (value of a system at T = 0 at the initial state before time-evolution) and boundary conditions (if known, represent constraints for the solution of the equation).

In the present review, 16 studies specified a model based on differential equations. Of these, 13 studies used ordinary differential equations (ODEs; comprising unknown functions of a single independent variable, unknown function derivatives, and other functions of the independent variable), while three studies used partial differential equations (PDEs; comprising unknown functions of more than one independent variable and their partial derivatives to model multidimensional systems) (Tschacher & Haken, Citation2019, #37; Tschacher et al., Citation2015a, #38, Citation2015b, #39). Except for one study, all ODEs were first-order differential equations. All PDEs were second-order differential equations. Numerical methods are often applied to solve ODEs and PDEs because their analytical solution is usually unavailable. For example, Runge‒Kutta methods (Butcher, Citation1996) are used to estimate a numerical solution to ODEs by applying discretization of the temporal derivatives. None of the included studies in the present review provided an analytical solution for their set of equations.

The three most influential models employing differential equations included (1) the adjusted Gottman et al.’s (Citation2002) model of dyadic interaction (Liebovitch et al., Citation2011, #17) leading to stable attractors (Schiepek et al., Citation2016a, #27), see Supplement 8a; (2) Schiepek’s model of psychotherapy including problem intensity, success in therapy, motivation to change, emotions, and insight (see Supplement 8b); and (3) the Fokker-Planck equation (Tschacher et al., Citation2015b, #39), see Supplement 8c.

After 2019, more models outside of these three aforementioned branches of models emerged (see Supplement 8d for details). Brice (Citation2019, #3) focused on a system dynamics model with inputs from an agent-based model to describe the dynamic evolution of a system of depressed patients flowing from general practitioners’ care to visits of psychotherapists and psychiatrists and back in several feedback loops. Burger et al. (Citation2020, #5) created a computational model of eight differential equations of one hypothetical patient to create a formal idiographic theory, modeling the potential effects of two possible interventions on the vicious circle of discriminant stimuli, catastrophizing, panic, avoidance, and credibility of catastrophic interpretations. Jason et al. (Citation2020, #12) focused on describing dynamic interrelations between social support, abstinence self-efficacy, and stress using three differential equations. Paz et al. (Citation2021, #22) employed a combination of differential equations with mixed models to simultaneously model the first-order change in attractor patterns while allowing variations at multiple levels, which is important when using data with a nested structure. This approach can be further utilized, for instance, to directly predict therapeutic outcomes.

Emergence of chaotic patterns

Based on the included studies, psychotherapy seems to be a discipline where the emergence of chaos is possible. Even though dynamic systems modeled in the included studies were often stable and predictable (e.g., leading to fixed-point attractors), several studies investigated the emergence of chaotic patterns (n = 14). The emergence of deterministic chaos mainly in the unpredictable client-therapist interaction (Kowalik et al., Citation1997, #13; Schiepek et al., Citation1997, #31, Citation2016b, #30) was inferred from the positive Lyapunov exponents (Bob, Citation2007, #2; Schiepek & Strunk, Citation2010, #32), bifurcations creating chaotic states (Redington & Reidbord, Citation1992, #24; Sabelli et al., Citation1990, #26; Schiepek & Strunk, Citation2010, #32), visual inspection of phase-space portraits (Reidbord & Redington, Citation1992, #25), phase transitions, multistability, chaos-to-chaos phase transitions (Schiepek et al., Citation2017, #29; Schöller et al., Citation2018, #34), and the emergence of nonlinear patterns of global complexity with a low-dimensional chaos-like system or fractal dimension (Fuhriman & Burlingame, Citation1994, #7; Lauro Grotto et al., Citation2014, #15). Furthermore, 10 other studies suggested that chaos was potentially present but did not test it directly.

Visualization

The included studies used various methods to visualize the results. Visualizations were often provided for both raw data and for DST/CT phenomenon measurement trajectories. Types of visualizations used in the included studies comprised line graphs (e.g., n = 17 raw data, or n = 10 DST/CT data), phase-space portraits (n = 13), heatmaps and/or resonance diagrams (n = 7), cluster dendrograms and/or bifurcation diagrams (n = 4), recurrence plots (n = 3), and conceptual models. See Supplement 9 for more information about the visualization methods used.

Psychotherapeutic/Psychological Phenomena and Interventions

The studies focused on therapeutic outcome (n = 14), therapeutic process (n = 13), or a combination of process and outcome (n = 14). In summary, the psychotherapeutic phenomena suitable for DST/CT methods that were investigated in the included studies could be clustered into more general categories of (1) psychotherapeutic relationship (most investigated, e.g., alliance, dyadic reactions, transference), (2) group processes (e.g., member interactions), (3) physiological processes (e.g., heart rate variability), (4) therapeutic destabilization of pathological states, and (5) specific syndromes (e.g., depression, bipolar disorder) and ideographically approached phenomena (e.g., client cognitive–affective dynamics). A detailed description of extracted psychotherapeutic/psychological phenomena and interventions is provided in Supplements 10a and 10b.

Studies that Used Empirical Data

In this section, we focus in more detail on studies that used empirical data (i.e., we omit the nine simulation studies). Overall, nine studies used only an empirical example based on data from another study (Buckman et al., Citation2018, #4; Fisher et al., Citation2011, #6; Fuhriman & Burlingame, Citation1994, #7; Laramée et al., Citation2016, #14; Sabelli et al., Citation1990, #26; Schiepek & Strunk, Citation2010, #32; Schiepek et al., Citation2016a, #27; Schöller et al., Citation2018, #34; Tschacher et al., Citation2015b, #39) and not a proper cross-validation. In terms of the type of data and data granularity considered, we extracted the smallest unit of measurement for each variable in the included studies and divided them into micro, meso, and macro levels. At the micro level, 16 studies analyzed data with high time intensity focusing on the processes during therapeutic sessions (e.g., heart rate variability and verbal utterances). At the meso level, 9 studies analyzed data focusing on processes in patients’ daily life, typically using data from daily administered self-report questionnaires. At the macro level, four studies analyzed data focusing on the overall trajectories of the therapeutic change based on session-by-session or weekly measurements. Additionally, three studies combined meso and macro levels (e.g., using two different empirical examples with different data granularity). Overall, nine studies used dyadic data, while the remaining studies used individual data (typically based on patients’ perspectives). See Supplement 11 for more information.

Auxiliary Information

This review identified 23 studies that claimed to bring novelty to the field. The list of innovations demonstrates that the field of DST/CT application in psychotherapy research is still at its beginning—nearly half of the studies make novel contributions (see Supplement 12a for information about claimed novelty of the studies). The most often used software was MATLAB. However, R seems to be increasing in popularity in recent years and is slowly replacing the usage of MATLAB in this field. Open data or code are not truly common thus far in terms of open and reproducibility science. For more detailed information about software used and transparency see Supplement 12b. We also gathered several claimed limitations from the included studies, see Supplement 12c.

Discussion

This review narratively synthesized descriptive information from 41 psychotherapy-related studies applying methods from dynamic systems theory and complexity theory (DST/CT) frameworks within simulated or real time-series data. The present review aimed to provide information about the current state of adoption of these methods and their application in the field. Our review shows that these techniques have provided powerful tools for analyzing psychotherapy as a complex dynamic system and understanding the emergent behavior of these systems.

This review showed that the topic of dynamic modeling and complex systems in psychotherapy is currently understudied. We observed that the range of dynamic modeling methods used before 2019 was rather narrow, as only three models were developed. They utilized only one of the following specific models or their variations: (1) describing the therapeutic relationship as a set of influence functions between client-therapist dyad (based on an adapted marital dyads model, Gottman et al., Citation2002), (2) coevolution of key therapeutic variables and their control parameters, and (3) as a synchrony.

Recently, several other models and applications have emerged, driven by readily accessible computational resources and advanced easy-to-use modeling software. The most recent of the included studies focused mainly on combining multiple methods (Brice, Citation2019, #3; Paz et al., Citation2021, #22), constructing formal theories (Burger et al., Citation2020, #5), continuing with formalizing specific dynamics of selected variables inspired by the tradition of Liebovitch et al.’s (Citation2011, #17) approach and specifying their own model (Jason et al., Citation2020, #12). Other models, such as the model of couples therapy adding another layer to client-therapist interactions, remain to be operationalized and tested in the context of psychotherapy.

New researchers are encouraged across other countries to continue in this regard. Researchers collecting specific data or wanting to implement complexity in their analyses might want to use the shortcut provided by our review to quickly learn about what has been done, which research questions have been pursued, and what methods may be feasible for their specific data (see Supplement 3). Further studies should examine the use of other DST/CT methods, such as discrete event simulation, optimization algorithms, evolutionary algorithms, cellular automata, and swarm intelligence, within the context of psychotherapeutic interventions. For a scoping review of the use of machine learning in psychotherapy research, see Aafjes-van Doorn et al. (Citation2021).

The importance of computational accessibility is demonstrated by the fact that the majority of computational studies were performed on personal computers and not in supercomputer centers. The importance of the easy accessibility of the simulation capability is demonstrated by the frequent use of high-level languages such as MATLAB or R for simulations as opposed to low-level C or Fortran. Additionally, the authors of the included studies usually did not explicitly mention specific numeric methods to solve the equations, such as linear multistep methods or Runge‒Kutta methods (Butcher, Citation1996), but used readily available tools, e.g., MATLAB.

Authors usually simulated their data using differential equations, plotted the attractor dynamics, or computed characteristics of the dynamical systems, such as the largest Lyapunov exponents or fractal dimensions. However, even though psychotherapy researchers already assumed self-organization and the emergence of complex behavior of psychotherapeutic systems, many of the identified studies remained in the computational and theoretical framework without cross-validation using empirical data. A similar situation has been observed in the field of health research (Rusoja et al., Citation2018). Validating the assumptions, as well as tentative conclusions of the existing studies, using real-life data thus remains another challenge in the DST/CT field.

Additionally, a relatively large number of studies (approximately 50%) employed DST/CT methods only to provide transformation of raw data into complexity measures, further imputed in more linearized analytical methods to compare means or seek relationships between specific variables, usually in a multilevel mixed model framework. This is supported by the dynamic models in the individual studies where estimated attractors were often stable, leading to fixed point or limit cycle equilibriums. However, linearization is a generally used practice when dealing with nonlinear systems (cf. Cheng et al., Citation2010), the full potential of DST/CT methods thus often remains unemployed in a psychotherapeutic context. When used as a full-bodied method, DST/CT may offer an apparatus to create formal psychotherapeutic theories describing the evolution of state variables given the evolution of control parameters enabling transitions between qualitatively different system phases. Importantly, several studies have identified chaotic behavior in psychotherapy data with an important example of Kowalik et al. (Citation1997, #13), who also showcased chaos-to-chaos transitions. Additionally, while deterministic chaos has gained attention in the past, its prominence as a serious topic of study has waned in recent years, with a few exceptions. Empirical validation of chaotic dynamics in actual therapeutic time series remains limited mainly because of limited data resolution (i.e., number of measurement points) and the stochastic nature of measurement issues. However, future advancements in data collection methods may provide opportunities for a more nuanced exploration of chaotic dynamics in the therapeutic processes. Moreover, theoretical explorations and simulations of deterministic chaos in psychotherapy have already yielded valuable insights.

The focus on case studies and time-intensive collection of within-session data, both physiological and verbal, was expected given that the applied DST/CT methods usually require large numbers of measurement points. However, some methods, such as dynamic complexity measures that were already ideographically implemented in clinical practice, were adopted even for daily-based questionnaire data using dynamic complexity instead of raw time series to inform about sudden changes and critical instability (Olthof et al., Citation2020a). The DST/CT methods, therefore, also demonstrated the possibility of approaching psychotherapeutic phenomena from the ideographic perspective while remaining in the quantitative framework. In such cases, they allowed for the computation of simple complexity estimations based on fluctuation and distribution of time series even of the single item in as few as 20 observations (Schiepek & Strunk, Citation2010). Moreover, dynamic complexity was validated as a measure of critical instability and phase transitions using correlations with other similar measures, such as permutation entropy or Lyapunov exponents. This is a pragmatic tool accessible even when naturalistically less rich data are analyzed and could be easily employed by other psychotherapy researchers.

This review aimed to provide a description of the applied mathematical DST/CT methods and summarize specific models used in the studies. The review suggests that the application of DST/CT methods in psychotherapy research is still in its infancy, with many studies demonstrating novel contributions but seldom cross-validating the findings except by using empirical examples. Moreover, if we cluster the authors’ affiliations to institutions on the country level, two major clusters emerge: USA and German-speaking countries (i.e., Germany, Austria, and Switzerland). Furthermore, it seems that the majority of included studies are influenced by a small number of influential researchers who coauthored many of these studies. For instance, one of the researchers coauthored 14 of the included studies. However, the validation and calibration of these methods together with seeking good-practice guidelines is an important next step in the development of the field.

The present review has several limitations (see Supplement 13 for more details). The review does not include SCOPUS as a source of studies (because of an excessive number of search results) and the very recent developments in the field (e.g., dynamic latent-class structural equation modeling or Markov switching models, Flückiger et al., Citation2022) because the search was conducted in 2021 and because several relevant keywords might have been excluded for producing a large number of false positives (e.g., “synchrony”) or omitted unintentionally. For specific reviews targeting “synchrony in psychotherapy,” see Kleinbub (Citation2017) or Koole and Tschacher (Citation2016). Moreover, studies targeting a specific subset of complexity theory, e.g., network modeling, or studies targeting other than human adult populations were omitted arbitrarily, limiting the potential prevalence of other eligible models. Finally, the abstract screening was conducted only by a single reviewer, the methodological quality of the included studies was not assessed, and some variables were only subjectively evaluated (e.g., theoretical comprehensiveness of included studies). Future reviews should focus on novel models and approaches such as computational neuroscience.

In conclusion, to utilize the descriptive nature of the present review, characteristics of individual studies were dummy coded according to several criteria to provide a quick combination of identifiers that the readers might want to filter for when seeking the exact inspiration for their DST/CT research question. A relatively large number of studies provided not only a set of mathematical formulae but also comprehensive theoretical or conceptual explanations. This might clarify the fundamental principles of DST/CT and the intended usage in psychotherapy.

CRediT Authorship Statement

(1) Adam Klocek: funding acquisition, project administration, conceptualization, data curation, formal analysis, investigation, methodology, writing—original draft, writing—review & editing; (2) Jan Premus: conceptualization, formal analysis, methodology, supervision, writing—review & editing; (3) Tomáš Řiháček: conceptualization, supervision, writing—review & editing.

Supplemental material

Supplemental Material

Download MS Word (467.6 KB)

Supplemental Material

Download MS Excel (130.3 KB)

Disclosure Statement

No potential conflict of interest was reported by the author(s).

Supplemental Data

Supplemental data for this article can be accessed online at https://doi.org/10.1080/10503307.2023.2252169.

Additional information

Funding

This study was supported by Specific University Research Grant No. MUNI/A/1187/2021 provided by the Czech Ministry of Education, Youth, and Sports.

References

  • References marked with an asterisk indicate primary studies included in the review.
  • Aafjes-van Doorn, K., Kamsteeg, C., Bate, J., & Aafjes, M. (2021). A scoping review of machine learning in psychotherapy research. Psychotherapy Research, 31(1), 92–116. https://doi.org/10.1080/10503307.2020.1808729
  • *Baker, A. Z. (2018). Examining affectual interaction within the therapeutic relationship across three psychotherapeutic theoretical approaches. Dissertation Abstracts International: Section B: The Sciences and Engineering. ProQuest Information & Learning. https://fau.digital.flvc.org/islandora/object/fau%3A39818.
  • *Bob, P. (2007). Hypnotic abreaction releases chaotic patterns of electrodermal activity during dissociation. International Journal of Clinical and Experimental Hypnosis, 55(4), 435–456. https://doi.org/10.1080/00207140701506532
  • Boulding, K. E. (1956). General systems theory—the skeleton of science. Management Science, 2(3), 197–208. https://doi.org/10.1287/mnsc.2.3.197
  • *Brice, S. N. (2019). Modelling disease progression and treatment pathways for depression [Doctoral dissertation]. https://orca.cardiff.ac.uk/id/eprint/129861/1/Brice_S__N_Thesis_corrected.pdf
  • *Buckman, J. F., Vaschillo, E. G., Fonoberova, M., Mezić, I., & Bates, M. E. (2018). The translational value of psychophysiology methods and mechanisms: Multilevel, dynamic, personalized. Journal of Studies on Alcohol and Drugs, 79(2), 229–238. https://doi.org/10.15288/jsad.2018.79.229
  • *Burger, J., van der Veen, D. C., Robinaugh, D. J., Quax, R., Riese, H., Schoevers, R. A., & Epskamp, S. (2020). Bridging the gap between complexity science and clinical practice by formalizing idiographic theories: A computational model of functional analysis. BMC Medicine, 18(1), 1–18. https://doi.org/10.1186/s12916-020-01558-1
  • Butcher, J. C. (1996). A history of Runge-Kutta methods. Applied Numerical Mathematics, 20(3), 247–260. https://doi.org/10.1016/0168-9274(95)00108-5
  • Carey, G., Malbon, E., Carey, N., Joyce, A., Crammond, B., & Carey, A. (2015). Systems science and systems thinking for public health: A systematic review of the field. BMJ Open, 5(12), e009002. https://doi.org/10.1136/bmjopen-2015-009002
  • Cheng, D., Hu, X., & Shen, T. (2010). Linearization of nonlinear systems. In A. Astolfi & L. Marconi (Eds.), Analysis and design of nonlinear control systems (pp. 279–313). Springer. https://doi.org/10.1007/978-3-642-11550-9_10
  • *Fisher, A. J., Newman, M. G., & Molenaar, P. C. M. (2011). A quantitative method for the analysis of nomothetic relationships between idiographic structures: Dynamic patterns create attractor states for sustained posttreatment change. Journal of Consulting & Clinical Psychology, 79(4), 552–563. https://doi.org/10.1037/a0024069
  • Flückiger, C., Horvath, A. O., & Brandt, H. (2022). The evolution of patients’ concept of the alliance and its relation to outcome: A dynamic latent-class structural equation modeling approach. Journal of Counseling Psychology, 69(1), 51–62. https://doi.org/10.1037/cou0000555
  • *Fuhriman, A., & Burlingame, G. M. (1994). Measuring small group process. Small Group Research, 25(4), 502–519. https://doi.org/10.1177/1046496494254005
  • Gelo, O. C. G., & Salvatore, S. (2016). A dynamic systems approach to psychotherapy: A meta-theoretical framework for explaining psychotherapy change processes. Journal of Counseling Psychology, 63(4), 379–395. https://doi.org/10.1037/cou0000150
  • Gottman, J., Swanson, C., & Swanson, K. (2002). A general systems theory of marriage: Nonlinear difference equation modeling of marital interaction. Personality and Social Psychology Review, 6(4), 326–340. http://doi.org/10.1207/S15327957PSPR0604_07
  • Guastello, S. J., Koopmans, M., & Pincus, D. (2008). Chaos and complexity in psychology: The theory of nonlinear dynamical systems. Cambridge University Press.
  • *Gumz, A., Geyer, M., & Brähler, E. (2014). Psychodynamic therapy from the perspective of self-organization. A concept of change and a methodological approach for empirical examination. Clinical Psychology & Psychotherapy, 21(4), 299–310. https://doi.org/10.1002/cpp.1840
  • *Gumz, A., Kästner, D., Geyer, M., Wutzler, U., Villmann, T., & Brähler, E. (2010). Instability and discontinuous change in the experience of therapeutic interaction: An extended single-case study of psychodynamic therapy processes. Psychotherapy Research, 20(4), 398–412. https://doi.org/10.1080/10503300903551021
  • *Haken, H., & Tschacher, W. (2017). How to modify psychopathological states? Hypotheses based on complex systems theory. Nonlinear Dynamics, Psychology, and Life Sciences, 21(1), 19–34. http://www.exp.unibe.ch/research/papers/HakenTschacher-2017-free-version.pdf
  • *Helmich, M. A., Wichers, M., Olthof, M., Strunk, G., Aas, B., Aichhorn, W., Schiepek, G., & Snippe, E. (2020). Sudden gains in day-to-day change: Revealing nonlinear patterns of individual improvement in depression. Journal of Consulting and Clinical Psychology, 88(2), 119–127. https://doi.org/10.1037/ccp0000469
  • Heylighen, F. (2008). Complexity and self-organization. In M. J. Bates & M. N. Maack (Eds.), Encyclopedia of library and information sciences, 3 (pp. 1215–1224). CRC Press, Inc.
  • *Jason, L. A., Stevens, E. B., Light, J. M., & Doogan, N. J. (2020). An empirically based theory of the relationships among social embeddedness, economic viability, learned recovery skills and perceived quality of life in recovery homes. Alcoholism Treatment Quarterly, 38(1), 126–142. https://doi.org/10.1080/07347324.2019.1633977
  • Kleinbub, J. R. (2017). State of the art of interpersonal physiology in psychotherapy: A systematic review. Frontiers in Psychology, 8, 2053. https://doi.org/10.3389/fpsyg.2017.02053
  • Koole, S. L., & Tschacher, W. (2016). Synchrony in psychotherapy: A review and an integrative framework for the therapeutic alliance. Frontiers in Psychology, 7, 862. https://doi.org/10.3389/fpsyg.2016.00862
  • *Kowalik, Z. J., Schiepek, G., Kumpf, K., Roberts, L. E., & Elbert, T. (1997). Psychotherapy as a chaotic process II. The application of nonlinear analysis methods on quasi time series of the client–therapist interaction: A nonstationary approach. Psychotherapy Research, 7(3), 197–218. https://doi.org/10.1080/10503309712331331973
  • *Laramée, P., Millier, A., Brodtkorb, T.-H., Rahhali, N., Cristeau, O., Aballéa, S., Montgomery, S., Steeves, S., Toumi, M., & Rehm, J. (2016). A comparison of Markov and discrete-time microsimulation approaches: Simulating the avoidance of alcohol-attributable harmful events from reduction of alcohol consumption through treatment of alcohol dependence. Clinical Drug Investigation, 36(11), 945–956. https://doi.org/10.1007/s40261-016-0442-7
  • *Lauro Grotto, R., Guazzini, A., & Bagnoli, F. (2014). Metastable structures and size effects in small group dynamics. Frontiers in Psychology, 5. https://doi.org/10.3389/fpsyg.2014.00699
  • *Li, X., & Kivlighan, D. M. (2019). Examining therapy dynamics and session outcome using differential equations model and multilevel data disaggregation. Psychotherapy Research, 30(5), 604–621. https://doi.org/10.1080/10503307.2019.1649730
  • *Lichtenberg, J. W., & Knox, P. L. (1991). Order out of chaos: A structural analysis of group therapy. Journal of Counseling Psychology, 38(3), 279–288. https://doi.org/10.1037/0022-0167.38.3.279
  • *Liebovitch, L. S., Peluso, P. R., Norman, M. D., Su, J., & Gottman, J. M. (2011). Mathematical model of the dynamics of psychotherapy. Cognitive Neurodynamics, 5(3), 265–275. https://doi.org/10.1007/s11571-011-9157-x
  • McGill, E., Marks, D., Er, V., Penney, T., Petticrew, M., & Egan, M. (2020). Qualitative process evaluation from a complex systems perspective: A systematic review and framework for public health evaluators. PLoS Medicine, 17(11), e1003368. https://doi.org/10.1371/journal.pmed.1003368
  • Mellado, A., Tomicic, A., Martínez, C., & Krause, M. (2022). Resultados en psicoterapia individual de adultos desde la perspectiva de los Sistemas Dinámicos: una revisión sistemática. Terapia psicológica, 40(2), 231–256. https://doi.org/10.4067/S0718-48082022000200231
  • Moher, D., Shamseer, L., Clarke, M., Ghersi, D., Liberati, A., Petticrew, M., Shekelle, P., & Stewart, L. A. (2015). Preferred reporting items for systematic review and meta-analysis protocols (PRISMA-P) 2015 statement. Systematic Reviews, 4(1), 1–9. https://doi.org/10.1186/2046-4053-4-1
  • *Olthof, M., Hasselman, F., Strunk, G., Aas, B., Schiepek, G., & Lichtwarck-Aschoff, A. (2020a). Destabilization in self-ratings of the psychotherapeutic process is associated with better treatment outcome in patients with mood disorders. Psychotherapy Research, 30(4), 520–531. https://doi.org/10.1080/10503307.2019.1633484
  • *Olthof, M., Hasselman, F., Strunk, G., van Rooij, M., Aas, B., Helmich, M. A., Schiepek, G., & Lichtwarck-Aschoff, A. (2020b). Critical fluctuations as an early-warning signal for sudden gains and losses in patients receiving psychotherapy for mood disorders. Clinical Psychological Science, 8(1), 25–35. https://doi.org/10.1177/2167702619865969
  • *Orsucci, F. F., Musmeci, N., Aas, B., Schiepek, G., Reda, M. A., Canestri, L., Giuliani, A., & de Felice, G. (2016). Synchronization analysis of language and physiology in human dyads. Nonlinear Dynamics, Psychology, and Life Sciences, 20(2), 167–191.
  • *Paz, A., Rafaeli, E., Bar-Kalifa, E., Gilboa-Schectman, E., Gannot, S., Laufer-Goldshtein, B., Narayanan, S., Keshet, J., & Atzil-Slonim, D. (2021). Intrapersonal and interpersonal vocal affect dynamics during psychotherapy. Journal of Consulting & Clinical Psychology, 89(3), 227–239. https://doi.org/10.1037/ccp0000623
  • *Peluso, P. R., Liebovitch, L. S., Gottman, J. M., Norman, M. D., & Su, J. (2012). A mathematical model of psychotherapy: An investigation using dynamic non-linear equations to model the therapeutic relationship. Psychotherapy Research, 22(1), 40–55. https://doi.org/10.1080/10503307.2011.622314
  • *Redington, D. J., & Reidbord, S. P. (1992). Chaotic dynamics in autonomic nervous system activity of a patient during a psychotherapy session. Biological Psychiatry, 31(10), 993–1007. https://doi.org/10.1016/0006-3223(92)90093-F
  • *Reidbord, S. P., & Redington, D. J. (1992). Psychophysiological processes during insight-oriented therapy: Further investigation into nonlinear psychodynamics. Journal of Nervous and Mental Disease, 180(10), 649–657. https://doi.org/10.1097/00005053-199210000-00007
  • Rusoja, E., Haynie, D., Sievers, J., Mustafee, N., Nelson, F., Reynolds, M., Sarriot, E., Swanson, R. C., & Williams, B. (2018). Thinking about complexity in health: A systematic review of the key systems thinking and complexity ideas in health. Journal of Evaluation in Clinical Practice, 24(3), 600–606. https://doi.org/10.1111/jep.12856
  • Ryo, M., Aguilar-Trigueros, C. A., Pinek, L., Muller, L. A., & Rillig, M. C. (2019). Basic principles of temporal dynamics. Trends in Ecology & Evolution, 34(8), 723–733. https://doi.org/10.1016/j.tree.2019.03.007
  • *Sabelli, H. C., Carlson-Sabelli, L., & Javaid, J. I. (1990). The thermodynamics of bipolarity—a bifurcation model of bipolar illness and bipolar character and its psychotherapeutical applications. Psychiatry-Interpersonal and Biological Processes, 53(4), 346–368. https://doi.org/10.1080/00332747.1990.11024519
  • *Schiepek, G., Aas, B., & Viol, K. (2016b). The mathematics of psychotherapy: A nonlinear model of change dynamics. Nonlinear Dynamics, Psychology, and Life Sciences, 20(3), 369–399.
  • *Schiepek, G., Kowalik, Z. J., Schütz, A., Köhler, M., Richter, K., Strunk, G., Mühlnickel, W., & Elbert, T. (1997). Psychotherapy as a chaotic process I. Coding the client–therapist interaction by means of sequential plan analysis and the search for chaos: A stationary approach. Psychotherapy Research, 7(2), 173–194. https://doi.org/10.1080/10503309712331331953
  • *Schiepek, G., & Strunk, G. (2010). The identification of critical fluctuations and phase transitions in short term and coarse-grained time series—a method for the real-time monitoring of human change processes. Biological Cybernetics, 102(3), 197–207. https://doi.org/10.1007/s00422-009-0362-1
  • *Schiepek, G., Tominschek, I., Heinzel, S., Aigner, M., Dold, M., Unger, A., Lenz, G., Windischberger, C., Moser, E., Plöderl, M., Lutz, J., Meindl, T., Zaudig, M., Pogarell, O., & Karch, S. (2013). Discontinuous patterns of brain activation in the psychotherapy process of obsessive-compulsive disorder: Converging results from repeated fMRI and daily self-reports. PLoS One, 8(8), 1–13. https://doi.org/10.1371/journal.pone.0071863
  • *Schiepek, G. K., Stöger-Schmidinger, B., Aichhorn, W., Schöller, H., & Aas, B. (2016a). Systemic case formulation, individualized process monitoring, and state dynamics in a case of dissociative identity disorder. Frontiers in Psychology, 7. https://doi.org/10.3389/fpsyg.2016.01545
  • *Schiepek, G. K., Tominschek, I., & Heinzel, S. (2014). Self-organization in psychotherapy: Testing the synergetic model of change processes. Frontiers in Psychology, 5. https://doi.org/10.3389/fpsyg.2014.01089
  • *Schiepek, G. K., Viol, K., Aichhorn, W., Hütt, M.-T., Sungler, K., Pincus, D., & Schöller, H. J. (2017). Psychotherapy is chaotic—(Not only) in a computational world. Frontiers in Psychology, 8. https://doi.org/10.3389/fpsyg.2017.00379
  • *Schöller, H., Viol, K., Aichhorn, W., Hütt, M.-T., & Schiepek, G. (2018). Personality development in psychotherapy: A synergetic model of state-trait dynamics. Cognitive Neurodynamics, 12(5), 441–459. https://doi.org/10.1007/s11571-018-9488-y
  • *Sharman, N. J. (1998). Cognitive and emotional aspects of maladaptive interpersonal patterns: A dynamic systems approach. Dissertation Abstracts International: Section B: The Sciences and Engineering. ProQuest Information & Learning. https://hdl.handle.net/1807/10791
  • Sivakumar, B. (2004). Chaos theory in geophysics: Past, present and future. Chaos, Solitons & Fractals, 19(2), 441–462. https://doi.org/10.1016/S0960-0779(03)00055-9
  • *Stapelberg, N. J. C. (2016). Quantitative relationships between clinical measures of depression And heart rate variability As measured By linear and nonlinear methods [Doctoral dissertation]. https://doi.org/10.25904/1912/1692
  • Strathern, M., & McGlade, J. (2014). The social face of complexity science: A festschrift for professor peter M. Allen. Emergent Publications.
  • Strogatz, S. H. (2018). Nonlinear dynamics and chaos with student solutions manual: With applications to physics, biology, chemistry, and engineering. CRC press.
  • *Tschacher, W., & Haken, H. (2019). Causation and chance: Detection of deterministic and stochastic ingredients in psychotherapy processes. Psychotherapy Research, 30(8), 1075–1087. https://doi.org/10.1080/10503307.2019.1685139
  • *Tschacher, W., Haken, H., & Kyselo, M. (2015a). Empirical and structural approaches to the temporality of alliance in psychotherapy. Chaos & Complexity Letters, 9(2), 177–191.
  • *Tschacher, W., Haken, H., & Kyselo, M. (2015b). Alliance: A common factor of psychotherapy modeled by structural theory. Frontiers in Psychology, 6. https://doi.org/10.3389/fpsyg.2015.00421
  • *Tschacher, W., Scheier, C., & Grawe, K. (1998). Order and pattern formation in psychotherapy. Nonlinear Dynamics, Psychology, and Life Sciences, 2(3), 195–215. https://doi.org/10.1023/A:1022324018097
  • von Bertalanffy, L. (1968). General systems theory: Foundations, developments, and applications. Braziller.
  • *Walter, S., Schiepek, G. K., Schneider, S., Strunk, G., Kaimer, P., & Mergenthaler, E. (2010). The synchronization of plan activations and emotion-abstraction patterns in the psychotherapeutic process: A single-case study. Psychotherapy Research, 20(2), 214–223. https://doi.org/10.1080/10503300903277437
  • Zill, D. G. (2012). A first course in differential equations with modeling applications. Cengage Learning.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.