Carrying Von Bertalanffy’s baton: Systems, complexity and network science, Part 2

Title: Carrying Von Bertalanffy’s baton: Systems, complexity and network science, Part 2
Author: Dr. Burgert Senekal, University of the Free State.
Ensovoort, volume 42 (2021), number 6: 1

Abstract

Ludwig von Bertalanffy wrote extensively on General Systems Theory (GST) and defined and discussed the properties of systems. In the current article, Von Bertalanffy’s views on aspects such as emergence, the methodological implications of emergence, and self-organisation, are discussed in the context of contemporary complexity theory and network science. It is shown how Von Bertalanffy viewed these key aspects of systems and their study in similar ways as contemporary network science.
Keywords: Von Bertalanffy, General Systems Theory, systems science, complexity, network science

1. Introduction

Part 1 of this series of articles situated network science and systems theory within the context of the Information- and Connected Age and recent technological advances. The present article focuses on network science as an approach to complex systems.
Network science is related to several theoretical frameworks. First, however, it is an approach to complex systems, together with nonlinear dynamics and statistical physics (Barabási, 2016:16, Duijn, 2016:15, Amaral and Ottino, 2004:1655). Although network science has developed its own theoretical concepts (see e.g. Barabási, 2016, Caldarelli, 2013 and Newman 2010), the current and following articles focus on general principles that network science shares with systems theory. Furthermore, a distinction can be made between simple and even complicated systems on the one hand, and complex systems on the other, and the distinction is also discussed in the current series of articles. The current article is not intended to provide a complete overview of all aspects of systems theory, but rather to lay the foundation for a better understanding of interdependence and network science.
More specifically, the current article discusses the following characteristics of systems and networks: the concept of emergence, methodological implications of emergence, causality, and self-organization.

2. The relationship between part and whole: emergence in complex systems

Laing (1990:204) quotes one of his patients, Julie, who was diagnosed with schizophrenia, “I’m an in divide you all.” This statement illustrates a number of concepts that are important for the study of systems. First, there is the relationship between words (entities) that accomplishes something greater when combined in a sentence: when this sentence is read quickly, one can hear her state that she is an individual. The whole therefore represents something more than what is present in the individual words. Second, on a semantic level, Julie illustrates what happens to her when that whole is broken down into its constituent parts: her existence literally perishes when the word individual is broken down into syllables – she is “divided.” These are core concepts of systems theory that are discussed in the current article.

A network or system consists of entities and their links or relationships, always with the emphasis on relationships (Nøkleberg, 2014:43, Wyseur, 2011:25). Stegbauer (2017:18) argues, “It is not the combinations of properties of individuals that are central, nor their attitudes or subjectivity; network research focuses on the structure of the relations.”[1] These relationships result in interdependence, where the existence of entities is dependent on their relationships with other entities (Page, 2011:17, Bar-Yam, 1997:12). Plsek (2001:309) also notes the importance of relationships when defining a system as, “the coming together of parts, interconnections, and purpose” (see also Kilcullen, 2010:193).

As with other system properties, this emphasis on relationships is clearly articulated by Von Bertalanffy (1968:55; 1972:417), who defines a system as, “a set of elements standing in interrelations.” More exact and comprehensive is however Strauss’s (1985:1-2) description of a system from the perspective of GST,

According to general systems theory, it must be considered that each system consists of interconnected parts which in their interrelations are mutually dependent on one another and which precisely – thanks to this reciprocal involvement – constitute the dynamic nature of each true totality.[2]

In aggregates, on the other hand, there is no significant interaction between entities. Viljoen (1986:4) provides the example of a heap of sand and a diamond to illustrate the difference between a system and an aggregate: while the molecules in a diamond are arranged in a specific pattern, the same is not the case with a heap of sand. Systems are about “order and coherence,” and “To be able to understand the system, one must not only know the elements, but also the relationships between them”[3] (Senekal, 1987:24, 25, see also Ehlers, 1989:188). Because the relationships within a system represent such an important facet of a system, the operation of a system can only be understood by investigating its structure (Nistor, Pickl and Zsifkovits, 2015:10).

Interdependence is a key concept that distinguishes a system from a collection, and therefore also an important concept in a complex versus a simple system. Although it is often stated that complex systems consist of a large number of components, the distinction between a simple and a complex system does not lie in the number of components from which it is composed (Amaral and Ottino, 2004:1654): since complexity is the result of interactions between entities within the system, it is insufficient to classify a system as complex merely on the basis of the number of components of which it is composed. Glattfelder (2010:175) mentions for example that a human being consists of about 25 000 genes and a grain of rice consists of 50 000 genes, but the former is obviously more complex than the latter. In other words, it is the interactions between elements that lead to the distinction between complex and simple systems.

There is no clear definition of a complex system (Duijn, 2016:15, Mobus and Kalton, 2015:169, Kwapień and Drożdż, 2012:119), and Amaral and Ottino (2004:1653) speculate that if one were to ask 10 scientists to formulate a definition, one would probably end up with 11 definitions. Several attempts have been made to define a complex system, and in the examples below it can be seen that interdependence is an important component of these efforts:

  • “A complex system is a system formed out of many components whose behaviour is emergent, that is, the behaviour of the system cannot be simply inferred from the behaviour of its components” (Bar-Yam, 1997:10).
  • “A complex system is a system with a large number of elements, building blocks or agents, capable of interacting with each other and with their environment. […] The common characteristic of all complex systems is that they display organization without any external organizing principle being applied. The whole is much more than the sum of its parts” (Amaral and Ottino, 2004:148).
  • “Complexity examines how components of a system, through their dynamic interactions, “spontaneously” develop collective properties or patterns that are not implicit in, or at least not in the same way implicit in, individual components”[4] (Urry, 2004:23).
  • “it is a system composed of many interacting parts, such that the collective behaviour of those parts together is more than the sum of their individual behaviours. The collective behaviours are sometimes also called ‘emergent’ behaviours, and a complex system can thus be said to be a system of interacting parts that displays emergent behaviour” (Newman, 2011:1).
  • “a complex system is a system built from a large number of nonlinearly interacting constituents, which exhibits collective behaviour and, due to an exchange of energy or information with the environment, can easily modify its internal structure and patterns of activity” (Kwapień and Drożdż, 2012:118).

An important term implied in these definitions (and included explicitly in the definitions of Bar-Yam and Newman), is emergence.[5] Emergence refers to Aristotle’s well-known statement that the whole is more than the sum of the parts (Glattfelder, 2013:2, Page, 2011:7, Von Bertalanffy, 1972:407, 1968:18). Ropohl (2005:27) describes emergence as when, “the properties of wholeness cannot be explained solely from the properties of individual parts, but only from the special way in which the parts interact”[6] (see also Kwapień and Drożdż, 2012:118, Wyseur, 2011:24, and DeLaurentis, 2007:364 for similar definitions).

Von Bertalanffy also considered emergence as a key feature of systems. He (1956:33) defines emergence as follows, “Each higher level presents new features that surpass those of the lower levels” (see also Von Bertalanffy, 1928:64).

Emergence is found in all systems. A written text, for example, is composed of letters, words and sentences, but can elicit feelings of happiness, sadness or anger from a reader. Language itself is a complex system (Kwapień and Drożdż, 2012) which on many levels consists of the interactions between entities. Senekal and Geldenhuys (2016) note that the letters ‘o’, ‘e’, ‘s’ and ‘p’ can be combined in such a way in Afrikaans that they express surprise (“oeps,” English “oops”), a medium of communication (“e-pos,” English “email”), a literary genre (“epos,” English “epic”), and vulgarity (“poes,” English “cunt”) – it is, among other things, the way in which the letters are combined and thus how the elements are connected that leads to meaning (for an overview of language as a complex system, see Kwapień and Drożdż, 2012, and Senekal and Geldenhuys, 2016).

Chemical compounds are frequently used to illustrate the concept of emergence. Christakis and Fowler (2010:26) state, for example, that the taste of a cake is quite different from the sum of the cake’s ingredients. Kilcullen (2010:195) also maintains that the taste of sugar (C12H22O11) originates at the molecular level and that a study of the atoms from which sugar is composed does not give an indication of the taste of their combination. Mobus and Kalton (2015:505) in turn note that Sodium (Na) and Chlorine (Cl) are on their own toxic, but the right combination, Sodium Chloride (NaCl) (table salt), is an essential part of the human diet.[7] It is the relationships that are crucial in systems; in the words of Wilden (1980:215), “entities do not create relationships so much as relationships create entities” (original emphasis).

One of the clearest expositions of emergence can be found in Page (2011:217), who frames the concept in the form of Equation (1).[8]

In other words, the function of x plus y is greater than the function of x plus the function of y – the whole is more than the sum total of the parts. Steyn (1984:8) illuminates further,
This organized whole is not traceable to the individual elements of which it is composed, and on the whole, there is a holistic perspective in this approach: The organization of components in systemic relationships gives the whole characteristics that do not only differ from the components taken in isolation, but are often also not present at all in the components alone; something that is a consequence of the organizational effect.[9]

The example of Sodium Chloride above illustrates, for example, Steyn’s argument: emergence means that the properties of the whole are not present in the elements themselves. Note, however, that emergence is present in any system, whether simple or complex, and is therefore not a distinctive feature of complex systems.
Given the importance of emergence in a complex system, any view of a complex network must include this characteristic of complex systems. Lawson, Ferris, Cropley and Cook (2006:9) define a network as follows,
A network is formed when a number (between two and infinity) of distinct entities that may be similar or dissimilar (nodes, elements, components, people, military formations, software instructions) are connected and interact such that new properties or behaviours emerge that are beyond the capabilities of any of the entities acting alone. These emergent properties cannot be predicted using reductionist consideration of the distinct entities. They are of interest because of the functions they perform and the purposes they serve, while the distinct and dissimilar entities included within a particular network boundary are those that are understood to be most significant in determining the emergent properties.
This definition of a complex network is the most comprehensive definition I have encountered to date and forms the backdrop of all conceptualisations of networks discussed in the current series of articles.

3. Reductionism versus holism: methodological implications for the study of systems

Since the properties of the whole are not present in the properties of the elements, emergence has important methodological implications for the study of systems. Mobus and Kalton (2015:11), Plsek (2001:311), De Beer (1997:81) and Von Bertalanffy (1950:134) contend that the paradigm of science over the past few centuries has been reductionism, in other words that a better understanding of the parts would lead to a better understanding of the world. This mechanistic Newtonian worldview suggested that the whole could be fully understood and predicted if science could comprehend the parts in totality. Wyseur (2011:24) refers to Julius Caesar’s saying, “Divide et impera” (divide and conquer). In other words: divide (analyse) and understand. He (2011:11) clarifies,

The mathematical laws led to a world view in which the universe was seen as a great mechanism. Mankind proved able, on the basis of mathematical language and empirically sound experiments, to derive absolute laws from nature in order to make predictions about the future. If one could derive all natural laws and know the positions of all particles present in the cosmos, it was possible to fully calculate and thus predict all movements and changes in time. This casts an absolutely deterministic view of the workings of the universe. The universe in its entirety was seen as one big clock. The ultimate goal of science was to unravel all natural laws in order to maximize the prediction possibilities.[10]
Planck’s quantum theory (as later refined by Heisenberg and Schrödinger) – together with Einstein’s theory of relativity – ushered in a new perspective within science, which called the Newtonian perspective into question (Wyseur, 2011:12, De Beer, 1997:78, Marais, 1992:20). Einstein’s theory of relativity showed that the position and time of the observer plays a role in observations (De Beer, 1997:82-83), while quantum physics, where Heisenberg identified the uncertainty principle (Mobus and Kalton, 2015:273, De Beer, 1997:85), also raised questions about the measurability of objects and further contributed to a questioning of science’s ability to undertake absolutely objective observations.
The most important scientific paradigm shift for the purposes of the current article, however, was the realization that the properties of the whole cannot be predicted by a better understanding of the parts, in other words that analysis will bring only limited knowledge of the world. Von Bertalanffy (1968:18-19) argues,

The system problem is essentially the problem of the limitations of analytical procedures in science. […] ‘Analytical procedure’ means that an entity investigated be resolved into, and hence can be constituted or reconstituted from, the parts put together, these procedures being understood both in their material and conceptual sense. […] Application of the analytical procedure depends on two conditions. The first is that interactions between ‘parts’ be nonexistent or weak enough to be neglected for certain research purposes. Only under this condition, can the parts be ‘worked out,’ actually, logically, and mathematically, and then be ‘put together.’ The second condition is that the relations describing the behaviour of parts be linear; only then is the condition of summativity given, i.e., an equation describing the behaviour of the total is of the same form as the equations describing the behaviour of the parts; partial processes can be superimposed to obtain the total process, etc. (see also Von Bertalanffy, 1968:36-37).

In line with Von Bertalanffy’s argument, Bar-Yam (1997:11) and others writing about complexity, e.g. Smaling (2013:91), Koithan, Bell, Niemeyer and Pincus (2012:10), Macklem (2008:1844), De Beer (1997:86), and Anderson (1972:393), argue that emerging properties cannot be studied by reductionist methods; entities must be studied within the complex web of relationships within which they exist. Anderson (1972:393) contends for example,
The behaviour of large and complex aggregates of elementary particles, it turns out, is not to be understood in terms of a simple extrapolation of the properties of a few particles. Instead, at each level of complexity entirely new properties appear, and the understanding of the new behaviours requires research which I think is as fundamental in its nature as any other.

It is therefore insufficient to try to achieve a better understanding of the human brain, for example, by simply studying individual neurons (Telesford, Simpson, Burdette, Hayasaka and Laurienti, 2011:295), or to try to understand why traffic jams occur by examining only the individual driver (Christakis and Fowler, 2010:25, Kumpula, 2008:2). One of the examples of a complex system that is most commonly mentioned in the literature on complex systems is a living organism, and Kresh (2006:24) maintains that emergence here also points out the shortcomings of reductionism, since the collective behaviour of living organisms cannot be reconstructed from the components of the system. Glückler (2017:23) also reasons in terms of human beings, “People act in dependence on situational and structural opportunities and not solely on the basis of ‘oversocialized’ internalization of given norms or ‘under-socialized’ compliance with logic such as rationality.”[11]

A more precise example. In the Afrikaans film industry from 1994-2014, as studied in Senekal (2015), there are 88 023 collaborative relationships between 1 866 actors, with an average path length of 2,35 between actors in this network. For the entire film industry, there are 805 103 collaborative relationships between 6 274 people (including all role players). There are therefore 814,65% more collaborative relationships for the whole industry (the whole) than for the actors (a component of the film industry), but the average path is not 814,65% longer for the whole industry, but rather shorter: 2,17. The average path length in the actor network thus provides no indication of what the average path length will be for the entire network, which illustrates that some facets of complex systems are not reducible to the study of components.

Because of emergence, systems theory then places the emphasis on studying the relationships that constitute the system. Ropohl (2005:28) summarises the systems-theoretical approach, “Systems thinking prefers systematization over elementalization, holistic models over atomistic models, multidimensionality over one-dimensionality, integration over differentiation, synthesis over analysis.”[12]

According to Cong and Liu (2014:599), network science is ideally suited to take emergence into account when studying complex systems; network science “makes it possible to probe into the complexity of real-world systems in their entirety and thus constitutes one, if not the only, solution to the challenge of reassembling complex systems and capturing their holistic properties.” Network science enables the scientist to consider each entity and each relationship in a system, as well as to study the individual entity within this complex web of relationships.

However, network science still suffers from an important limitation: all relationships that affect an entity cannot be practically included in a study. Part of the problem is data: finding all relevant data points, and getting them in a format where they can be included in the study, is currently not possible. In other words, network science seeks to consider as many connections as possible, but never includes all relationships that affect the component. The important issue here is however that systems theory and network science are not under the impression that parts can be understood in isolation, even though it is currently not yet feasible to take all relationships into account in every study.

4. Causality in complex systems

Emergence also affects our understanding of causality. In a simple system, if one understands the workings of the parts, predictions can be made with a great deal of certainty. For example, if one understands the parts of a motor vehicle and how they function together, it can be predicted whether a car will function as expected. Because linear causality applies to the simple or complicated system such as a motor vehicle, the mechanic’s job is relatively simple: find the part that is not functioning as it should, and replace it. When all parts are in working condition, the motor vehicle will also be in working condition. Von Bertalanffy (1968:45) explains,

The only goal of science appeared to be analytical, i.e., the splitting up of reality into ever smaller units and the isolation of individual causal trains. Thus, physical reality was split up into mass points or atoms, the living organism into cells, behaviour into reflexes, perception into punctual sensations, etc. Correspondingly, causality was essentially one-way: one sun attracts one planet in Newtonian mechanics, one gene in the fertilized ovum produces such and such inherited character, one sort of bacterium produces this or that disease, mental elements are lined up, like the beads in a string of pearls, by the law of association.

The same degree of predictability does not apply to complex systems. Prigogine (1997:4) – one of the key theorists of unpredictability and non-linear processes – writes, “Classical science emphasized order and stability; now, in contrast, we see fluctuations, instability, multiple choices, and limited predictability at all levels of observation.” In complex systems, causality is not a single or even an accumulation of causes (Byrne and Callaghan, 2014:190, Smaling, 2013:90, Weideman, 2009:67). In simple systems, a cause can be identified with certainty, for example that a pen has fallen due to gravity, or that ice has melted due to a temperature above 0°C and air pressure falling within a normal spectrum of surface air pressure (an accumulation of causes). In complex systems, however, in the words of Wilden[13] (1980:39), “causes cause causes to cause causes” (original emphasis). Nicolis (1995:1-2) explains,

A striking difference between linear and nonlinear laws is whether the property of superposition holds or breaks down. In a linear system the ultimate effect of the combined action of two different causes is merely the super-position of the effects of each cause taken individually. But in a nonlinear system adding two elementary actions to one another can induce dramatic new effects reflecting the onset of cooperativity between the constituent elements. This can give rise to unexpected structures and events whose properties can be quite different from those of underlying elementary laws, in the form of abrupt transitions, a multiplicity of states, pattern formation, or an irregularly markedly unpredictable evolution of space and time referred to as deterministic chaos.

A clear example of unpredictability in a complex system can be found in the fall of the Soviet Union. Papp, Alberts and Tuyahov (1997:30-31) explain that although reforms in the 1980s were aimed at decentralizing economic decision-making and leading to improved production, it rather created confusion and economic uncertainty in the Soviet Union and reduced Soviet production. Reforms were to encourage popular support for communism by involving more people in the political decision-making process, but rather led to more Soviet citizens questioning and eventually rejecting the system. There are many reasons why the Soviet Union collapsed, including their weak economy, technological disadvantage compared to the West, the war in Afghanistan and these reforms, but the important issue to keep in mind here is that it is the way in which various factors are combined that ultimately led to this unpredictable outcome. Interactions are the key to unpredictability in complex systems, as Eusgeld, Nan and Dietz (2011:681) contend, “a complex system can never be fully knowable not only due to rapid changes in the system state (high dynamic) and nonlinear behaviour, but also interconnections within the system (interdependencies).”

Because communities are also complex systems, unpredictability sometimes hinders efforts to improve communities’ conditions. One example of such an unexpected result is the well-meaning efforts of Western film stars to distribute mosquito nets in Africa, which undermine local netmakers’ businesses and thus increase people’s vulnerability to malaria in the long term (Kilcullen, 2013:243). In South Africa, the African National Congress (ANC) banned alcohol sales in March 2020 to combat alcohol-related crimes such as domestic violence during the Covid-19 epidemic (Harding, 2020), but less than two months later, President Ramaphosa had to admit that violence against women increased during the lockdown (Chothia, 2020).

Simple linear causality is deeply embedded in human thinking. One such example is the cause of shooting incidents at high schools in the US, which is a phenomenon that is frequently attributed to teenagers playing violent video games. This alleged causal link is criticised by Watts (2011:116-117), as it ignores the vast majority of adolescents who never harm anyone, even though they also play these video games. Similarly, the high availability of firearms in this country is also frequently blamed for these incidents. One could offer the same critique of this claim: if firearms are the cause, why are there so few incidents? There were 583 shooting incidents at American schools between 2013 and 2019 (Everytown, 2019) in a country with around 300 000 000 firearms (33-34% of adults own firearms, Agresti and Smith, 2010), which raises the question: what about the other firearms? And why are there no such incidents in South Africa, where there are also large quantities of firearms (between 3 700 000 and 7 200 000) (GunPolicy, 2020) that are privately owned? If the firearms are the cause, why does one American adolescent take his father’s firearm and shoot his fellow students, while the other one does not, and why does the South African teenager not take his father’s firearm and shoot his fellow students?

The problem with such alleged simple causal relationships is that a human being is a complex system where such simplistic causal relationships are not valid. People who advocate to ban video games or firearms in the US to prevent shooting incidents, try to replace the faulty part – like the mechanic – in the hope that the whole system will function flawlessly thereafter. But a human being (a complex system) is not a motor vehicle (a simple system). For one American adolescent, his decision to perpetrate a massacre can be helped by the fact that he can easily obtain a firearm, and he may have grown up with violent video games, but his neighbour may also easily obtain a firearm and play the same video games, but he does not harm anyone. The superficial similarities are not as important as the many differences: the two adolescents have different genetic compositions, are exposed to different influences, come from different parental homes, have different diets and many more. It is the combination of all factors in an individual’s social, biological and psychological spheres that leads to such incidents, not a single one.

Moreover, by identifying the wrong cause due to an outdated Newtonian, mechanistic worldview, the problem cannot be addressed because it cannot be correctly identified. What is needed is a complexity perspective. Prigogine (1984:203) contends, “We are trained to think in terms of linear causality, but we need new ‘tools of thought’.”

5. The web without a spider: self-organization in complex systems

Self-organization refers to the phenomenon where elements organize without any outside control (Duijn, 2016:15, Wyseur, 2011:35). Examples include the formation of patterns in a snowflake, as well as the formation of crystals, diamonds, dunes and chemical compounds. These examples represent simple systems, but the phenomenon is of course also found in complex systems. Koithan, Bell, Niemeyer and Pincus (2012:11) define self-organisation as follows,

A nonlinear distributed process throughout complex systems that is dependent on energy and information networks occurring in far-from-equilibrium states and yields order and emergent or arising behaviour befitting the environment or circumstance. Self-organization is complexification.
Von Bertalanffy (1968:98) proposed that organisms are capable of self-organisation, but extended his view to include all open systems,
Self-differentiating systems that evolve toward higher complexity (decreasing entropy) are, for thermodynamic reasons, possible only as open systems – e.g., systems importing matter containing free energy to an amount overcompensating the increase in entropy due to irreversible processes within the system […]. However, we cannot say that “this change comes from some outside agent, an input”; the differentiation within a developing embryo and organism is due to its internal laws of organization, and the input […] makes it only possible energetically.

To illustrate self-organisation, Cilliers (1998:88-90) provides the example of a school of fish, where the condition of the fish will depend on a large number of factors, including the availability of food, the temperature of the water, the amount of oxygen and light, the time of year, and so on. As these factors vary, the size of the school of fish will adapt itself optimally to the prevailing conditions, despite the fact that each individual fish can only look after its own interests. The system of the school of fish organizes itself as a whole to ensure the best relationship between the system and the environment. There is no agent who decides for the school, and also each individual fish does not understand the complexity of the situation.

A well-known human example of self-organization is the Mexican Wave at sports events, which is also mentioned by Christakis and Fowler (2010:25) and studied in detail by Farkas, Helbing and Vicsek (2002). A group of people stand up and throw their arms in the air, the group closest to them does the same, and the one closest to the next group does the same, and finally the ‘wave’ moves around the sports field with remarkable regularity. Again, there is no central control present; no single person coordinates the wave.

According to Christakis and Fowler (2010), self-organization is responsible for smoking cessation, spreading suicide over social networks, as well as happiness and depression, and also – their most famous finding – obesity (Christakis and Fowler, 2007). Kilcullen (2013:80-86) also discusses the tactics of Somali fighters as swarm tactics, where they appear coordinated in the absence of a central governing body. In his view, the American Task Force Ranger confronted such a “self-organizing swarm” (2013:73) in Mogadishu in October 1993 (this battle is familiar through Riddley Scott’s film, Black Hawk Down).

A related phenomenon is that of synchronization,[14] as discussed by Strogatz (2004), Boccaletti, Latora, Moreno, Chavez and Hwanga (2006:237-251), Barrat, Barthélémy and Vespignani (2008:136-159), and Csermely (2006:81). Synchronisation refers to “a process wherein many systems (either equivalent or non-equivalent) adjust a given property of their motion due to a suitable coupling configuration, or to an external forcing” (Boccaletti, Latora, Moreno, Chavez, and Hwanga (2006:237). Christian Huygens’s (1893[1665]) well-known 17th century discovery of synchronization between two pendulums is an example of synchronization between simple systems, but a variety of studies of synchronization between complex systems have already been undertaken. Through visual and acoustic interactions, fireflies flash in sync (Buck, 1938), crickets cricket in harmony (Walker, 1969), audiences clap their hands in sync (Néda, Ravasz, Vicsek, Brechet, and Barabási (2000), and women living together synchronize menstrual cycles (McClintock, 1971). On a much larger scale, the outbreak of syphilis was synchronized from Houston to New York between 1960 and 1993 (Grassly, Fraser and Garnett (2005).
Schmidt (1997:125) argues that cultural systems are also self-organising, “[M]odern literary systems are self-organising insofar as all decisions concerning literariness and literary values are made ‘inside’ the social system literature, i.e. in literary communications” (see also De Berg, 1997:145-147). There are gatekeepers, but the gatekeepers themselves are also part of the system. Roberts (1973:87) also writes that all identification of specific objects (books and writing traditions) as literature is carried out by the literary community itself. No central governing body coordinates the functioning of the literary system.

The same holds for works of art, where the intrinsic characteristics of the work of art do not determine the ultimate reputation of the work. Fraiberger, Sinatra, Resch, Riedl and Barabási (2018:825) explain,

Quality in art is elusive; art appeals to individual senses, pleasures, feelings, and emotions. Recognition depends on variables external to the work itself, like its attribution, the artist’s body of work, the display venue, and the work’s relationship to art history as a whole. Recognition and value are shaped by a network of experts, curators, collectors, and art historians whose judgments act as gatekeepers for museums, galleries, and auction houses.
Since self-organization is “the key concept in the science of complexity”(Plsek, 2001:313), any valid understanding of a complex network must also consider self-organization. This is how Barabási (2003:221) considers complex networks, “No central node sits in the middle of the spider web, controlling and monitoring every link and node. There is no single node whose removal could break the web. A scale-free network is a web without a spider.”

6. Conclusion

This article discussed some of the features of systems from a GST, complexity and network science perspective. Whether the issue is emergence or the methodological issues thereof, causality, or self-organisation, in each case, network science shares the views articulated by Von Bertalanffy. In terms of these features, then, it can be argued that network science continues many of Von Bertalanffy’s insights, albeit in a new guise.
The following article discusses further features associated with network science that correspond to Von Bertalanffy’s views.

Bibliography

Agresti, J. D. and Smith, R. K. 2010. Just Facts. http://www.justfacts.com/guncontrol.asp#f8 [Accessed 5 April 2016].
Amaral, L. and Ottino, J. 2004a. Complex networks. Augmenting the framework for the study of complex systems. European Physical Journal, 38:147–162.
Amaral, L. and Ottino, J. 2004b. Complex systems and networks: challenges and opportunities for chemical and biological engineers. Chemical Engineering Science, 59:1653 – 1666.
Anderson, P.W. 1972. More is different. Science, 177(4047):393-396.
Bar-Yam, Y. 1997. Dynamics of complex systems. Colorado: Westview Press.
Barabási, A.-L. 2003. Linked. London: Plume.
Barabási, A.-L. 2016. Network science. Cambridge: Cambridge University Pres.
Barrat, A., Barthélémy, M. and Vespignani, A. 2008. Dynamical processes on complex networks. Cambridge: Cambridge University Press.
Boccaletti, S. et al. 2006. Complex networks: Structure and dynamics. Physics Reports, 424:175–308.
Buck, J. 1938. Synchronous rhythmic flashing of fireflies. Quarterly Review of Biology, 13:301–314.
Byrne, D. and Callaghan, G. 2014. Complexity Theory and the Social Sciences. The state of the art. Oxon: Routledge.
Caldarelli, G. 2013. Scale free networks. Complex webs in nature and technology. Oxford: Oxford University Press.
Chothia, A. 2020. ‘Men in our country have declared war on women’ – Cyril Ramaphosa. https://www.thesouthafrican.com/news/gender-based-violence-gbv-increase-lockdown-cyril-ramaphosa/ [Accessed 19 May 2020].
Christakis, N. A. and Fowler, J. H. 2007. The spread of obesity in a large social network over 32 years. New England Journal of Medicine, 357(4):370-379.
Christakis, N. and Fowler, J. 2010. Connected. London: Harper.
Cilliers, P. 1998. Complexity and Postmodernism: Understanding Complex Systems. London: Routledge.
Cong, J. and Liu, H. 2014. Approaching human language with complex networks. Physics of Life Reviews, 11:598–618.
Csermely, P. 2006. Weak links: Stabilizers of complex systems from proteins to social networks. Heidelberg: Springer.
De Beer, S. J. 1997. Holisme en ontwikkeling. Unpublished MA-dissertation: Randse Afrikaanse Universiteit.
De Berg, H. 1997. Communication as a challenge to systems theory. Canadian Review of Comparative Literature, March: 141-151.
Duijn, P. 2016. Detecting and Disrupting Criminal Networks. A Data Driven Approach. Unpublished PhD-thesis: University of Amsterdam.
Ehlers, C. 1989. Die burgermag en die gesin. Suid-Afrikaanse Tydskrif vir Sosiologie, 20(3):187-194.
Eusgeld, I., Nan, C. and Dietz, S. 2011. “System-of-systems” approach for interdependent critical infrastructures. Reliability Engineering and System Safety, 96(6):679-686.
Everytown. 2019. Gunfire on School Grounds in the United States. https://everytownresearch.org/gunfire-in-school/2866/#2866 [Accessed 19 May 2020].
Farkas, I., Helbing, D. and Vicsek, T. 2002. Mexican waves in an excitable medium. Nature, 419:131–132.
Fraiberger, S. P. et al. 2018. Quantifying reputation and success in art. Science, 362:825–829.
Glattfelder, J. B. 2010. Ownership networks and corporate control: mapping economic power in a globalized world. Unpublished PhD-thesis: ETH Zurich.
Glattfelder, J. B. 2013. Decoding Complexity: Uncovering Patterns in Economic Networks. Heidelberg: Springer.
Glückler, J. 2017. Relationales Denken diesseits und jenseits der Netzwerkforschung. Soziologie, 46(1):22-27.
Grassly, N., Fraser, C. and Garnett, G. 2005. Host immunity and synchronized epidemics of syphilis across the United States. Nature, 433:417–421.
GunPolicy, 2020. South Africa — Gun Facts, Figures and the Law. https://www.gunpolicy.org/firearms/region/south-africa [Accessed 19 May 2020].
Harding, A. 2020. South Africa coronavirus lockdown: Is the alcohol ban working?. https://www.bbc.com/news/world-africa-52358268 [Accessed 19 May 2020].
Huygens, C. 1893[1665]. Letter to R. Moray (February 27, 1665). In: M. Nijhoff, ed. Euvres Completes de Christian Huygens. The Hague: Société Hollandaise des Sciences:246–249.
Kilcullen, D. 2010. Counterinsurgency. London: C. Hurst and Co.
Kilcullen, D. 2013. Out of the mountains. The coming age of the urban guerrilla. London: C. Hurst and Co.
Koithan, M., Bell, I. R., Niemeyer, K. and Pincus, D. 2012. A complex systems science perspective for whole systems of complementary and alternative medicine research. Forschende Komplementärmedizin/Research in Complementary Medicine, 19(Supplement 1):7-14.
Kresh, J. Y. 2006. Integrative systems view of life: Perspectives from general systems thinking. In: T. S. Deisboeck and J. Y. Kresh, eds. Complex Systems Science in Biomedicine. Heidelberg: Springer:3-29.
Kumpula, J. 2008. Community structures in complex networks: detection and modeling. Unpublished PhD dissertation: Helsinki University of Technology.
Kwapień, J. and Drożdż, S. 2012. Physical approach to complex systems. Physics Reports, 515(3):115-226.
Laing, R. 1990. The divided self. London: Penguin.
Lawson, E., Ferris, T., Cropley, D. and Cook, S. 2006. Development of a Foundation for Military Network Science. s.l.:Systems Engineering and Evaluation Centre (SEEC), University of South Australia.
Macklem, P.T. 2008. Emergent phenomena and the secrets of life. Journal of Applied Physiology, 104:1844–1846.
Marais, M. 1992. Die persoon in die sisteem: Kritiek op gesinsterapie. Unpublished MA-dissertation: Randse Afrikaanse Universiteit.
McClintock, M. 1971. Menstrual synchrony and suppression. Nature, 229:244–245.
Mobus, G. E. and Kalton, M. C. 2015. Principles of systems science. Heidelberg: Springer.
Néda, Z. et al. 2000. The sound of many hands clapping. Nature, 403:849–850.
Newman, M. E. J. 2010. Networks. Oxford: Oxford University Press.
Newman, M. E. J. 2011. Complex systems: A survey. arXiv preprint:1-10.
Nicolis, G. 1995. Introduction to Nonlinear Science. Cambridge: Cambridge University Press.
Nistor, M. S., Pickl, S. W. and Zsifkovits, M. 2015. Visual Analytics of Complex Networks: A Review from the Computational Perspective. International Conference on Modeling, Simulation and Visualization Methods:10-15.
Nøkleberg, M. 2014. Sammen for sikkerhet? En casestudie av sikkerhetsforvaltningen i Bergen. Unpublished MA-dissertation: Universiteit Bergen.
Page, S. E. 2011. Diversity and complexity. Princeton: Princeton University Press.
Papp, D. S., Alberts, D. S. and Tuyahov, A. 1997. Historical Impacts of Information Technologies: An Overview. In: D. S. Alberts and D. S. Papp, eds. The information age: an anthology on its impact and consequences. s.l.:CCRP Publication Series:13-35.
Plsek, P. 2001. Redesigning health care with insights from the science of complex adaptive systems. In: C. o. Q. o. H. C. i. America, ed. Crossing the quality chasm: A new health system for the 21st century. Washington: National Academy Press:309-322.
Prigogine, I. 1984. Order out of chaos: Man’s new dialogue with nature. New York: Bantam Books.
Prigogine, I. 1997. The end of certainty. Time, chaos, and the new laws of nature. New York: The Free Press.
Roberts, T. J. 1973. The Network of Literary Identification: A Sociological Preface. New Literary History, 5(1):67-90.
Ropohl, G. 2005. Allgemeine Systemtheorie als transdisziplinäre Integrationsmethode. Technikfolgenabschätzung – Theorie und Praxis, 2(14):24-31.
Schmidt, S. J. 1997. A systems-orientated approach to literary studies. Canadian Review of Comparative Literature, March: 119-136.
Senekal, B. A. and Geldenhuys, C. 2016. Afrikaans as ‘n komplekse netwerk: Die woord ko-voorkoms netwerke van woorde in André P. Brink se Donkermaan in Afrikaans, Nederlands en Engels. Suid-Afrikaanse Tydskrif vir Natuurwetenskappe en Tegnologie.
Senekal, B. A. 2015. ʼn Veelvlakkige netwerkontleding van die Afrikaanse filmbedryf (1994-2014). LitNet Akademies (Geesteswetenskappe), 12(2).
Senekal, J. H. 1987. Literatuuropvattings: ‘wese’ en ‘waarhede’ van ‘n nuwe literêre teorie. Bloemfontein: University of the Free State.
Smaling, A. 2013. Complexiteitsdenken en kwalitatief onderzoek. Kwalon, 18(2):89-98.
Stegbauer, C. 2017. Interdisziplinäre Netzwerkforschung. Soziologie, 36(1):17-22.
Steyn, A. F. 1984. Buckley: die moderne oop-sisteembenadering. Suid-Afrikaanse Tydskrif vir Sosiologie, 15(1):5-16.
Strauss, D. 1985. Sisteemteorie en die sosiologie. Suid-Afrikaanse Tydskrif vir Sosiologie, 16(1):1-8.
Strogatz, S. 2004. Sync. The emerging science of spontaneous order. London: Penguin.
Telesford, Q. K. et al. 2011. The Brain as a Complex System: Using Network Science as a Tool for Understanding the Brain. Brain Connectivity, 1(4):295-308.
Urry, J. 2004. Globale complexiteit en het autosysteem. Tijdschrift voor sociologie, 25(1):21-35.
Viljoen, H. 1986. Die Suid-Afrikaanse romansisteem. ‘n Vergelykende studie. Unpublished PhD-thesis: University of the Northwest.
Von Bertalanffy, L. 1928. Kritische Theorie der Formbildung. Berlin: Borntraeger.
Von Bertalanffy, L. 1950. An Outline of General System Theory. The British Journal for the Philosophy of Science, 1(2):134-165.
Von Bertalanffy, L. 1956. A Biologist Looks at Human Nature. The Scientific Monthly, 82(1):33-41.
Von Bertalanffy, L. 1968. General systems theory: Foundations, development, applications. New York: George Braziller.
Von Bertalanffy, L. 1972. The History and Status of General Systems Theory. The Academy of Management Journal, 15(4):407-426.
Walker, T. 1969. Acoustic synchrony: Two mechanisms in the snowy tree cricket. Science, 166:891–894.
Watts, D. J. 2011. Everything is obvious. Once you know the answer. London: Atlantic.
Weideman, A. 2009. Uncharted territory: A complex systems approach as an emerging paradigm in applied linguistics. Per Linguam, 25(1):61-75.
Wilden, A. 1980. System and structure: essays in communication and exchange. New York: Tavistock.
Wyseur, S. 2011. De evolutie naar non-lineaire complexiteit in digitale mediacultuur; een exploratieve literatuurstudie. Unpublished MA-dissertation: University of Gent.

Endnotes

[1] Own translation from the original German, “Zentral sind dabei nicht die Kombinationen von Eigenschaften einzelner Personen und auch nicht deren Einstellungen oder deren Subjektivität; in den Mittelpunkt stellt die Netzwerkforschung die Struktur der Relationen.”
[2] Own translation from the original Afrikaans, “Wel moet volgens die algemene sisteemteorie verreken word dat elke sisteem uit onderling ineengeweefde onderdele bestaan wat in hul interrelasies wederkerig van mekaar afhanklik is en wat juis danksy hierdie wederkerige betrokkenheid die dinamiese aard van elke egte totaliteit konstitueer.”
[3] Own translation from the original Afrikaans, “Om die sisteem te kan verstaan, moet ʼn mens nie net die elemente ken nie, maar ook die relasies tussen hulle.”
[4] Own translation from the original Dutsch, “Complexiteit onderzoekt hoe componenten van een systeem doorheen hun dynamische interactie ‘spontaan’ collectieve eigenschappen of patronen ontwikkelen die niet impliciet zijn aan, of op zijn minst niet op dezelfde manier impliciet zijn aan, de individuele componenten.”
[5] Emergence is by no means a new concept in science. In addition to, for example, Aristotle, John Stuart Mill (1843:III(6)1) argued, “To whatever degree we might imagine our knowledge of the properties of the several ingredients of a living body to be extended and perfected, it is certain that no mere summing up of the separate actions of those elements will ever amount to the action of the living body itself.”
[6] Own translation from the original German, “die Eigenschaften der Ganzheit nicht allein aus den Eigenschaften einzelner Teile, sondern nur aus der besonderen Art des Zusammenwirkens der Teile erklärt werden können.”
[7] Strauss (1985:3) argues that it is problematic to claim that a system consists of parts, and uses the example of sodium chloride to illustrate his argument. He writes, “Wanneer soutkorrels verdeel word, word telkens kleiner stukkies sout gekry – met as ondergrens ʼn enkele NaCl molekuul. Sodra hierdie kleinste (sout-gewys gesproke ondeelbare) eenheid verdeel word, is die sout tot niet en sit ons bloot met ʼn Natrium atoom en ʼn Chloor atoom! Dit toon onmiskenbaar dat Natrium en Chloor nie gesien kan word as integrale (egte) dele van tafelsout nie, hoeseer tafelsout ookal daaruit opgebou is. [When salt grains are divided, smaller pieces of salt are obtained each time – with a single NaCl molecule as the lower limit. Once this smallest (salt-wise indivisible) unit is divided, the salt is destroyed and we are left with a Sodium atom and a Chlorine atom! This unmistakably shows that Sodium and Chlorine cannot be seen as integral (real) parts of table salt, no matter how much table salt consist of it.]” Strauss’s argument, however, is similar to that presented here: the relationships between elements lead to something quite different from what is present in the properties of the individual elements.
[8] Page uses the symbol ≥, but when he explains the concept, he calls the function “greater than.” Using ≥ is probably a printing error.
[9] Own translation from the original Afrikaans, “Hierdie georganiseerde geheel is nie tot die individuele elemente waaruit dit bestaan, herleibaar nie, en in geheel gesien, is daar ʼn holistiese perspektief in hierdie benadering: Die organisasie van komponente in sistemiese verhoudings gee aan die geheel kenmerke wat nie net verskil van die komponente in isolasie geneem nie, maar dikwels ook glad nie in die komponente alleen geneem, gevind word nie; iets wat ʼn uitvloeisel van die organisasie-effek is.” See also Von Bertalanffy (1952:148; 1950:148) and Bar-Yam (1997:10) for similar statements.
[10] Own translation from the original Dutch, “De mathematische wetmatigheden leidden tot een wereldbeeld waarbij het universum als een groot mechanisme werd gezien. Het menselijk wezen bleek in staat om, op basis van wiskundige taal en empirisch verantwoorde experimenten, absolute wetten af te leiden van de natuur teneinde voorspellingen te kunnen maken over de toekomst. Indien men alle natuurlijke wetmatigheden zou kunnen afleiden en men de posities van alle aanwezige deeltjes in de kosmos zou weten, bestond de mogelijkheid om alle bewegingen en veranderingen in tijd volledig te berekenen en dus te voorspellen. Dit werpt een absoluut deterministische visie op de werking van het heelal. Het heelal in zijn totaliteit werd gezien als één groot uurwerk. Het ultieme doel van de wetenschap bestond erin alle natuurlijke wetmatigheden te ontrafelen om zo de voorspellingsmogelijkheden te maximaliseren.”
[11] Own translation from the original German, “Menschen handeln in Abhängigkeit situativer und struktureller Gelegenheiten und nicht allein aufgrund ‘übersozialisierter’ Verinnerlichung vorgegebener Normen oder ‘untersozialisierter’ Befolgung von Logiken wie zum Beispiel der Rationalität.”
[12] Own translation from the original German, “Das Systemdenken präferiert die Systematisierung gegenüber der Elementarisierung, holistische Modelle gegenüber atomistischen Modellen, die Mehrdimensionalität gegenüber der Eindimensionalität, die Integration gegenüber der Differenzierung, die Synthese gegenüber der Analyse.”
[13] Wilden does not write within the framework of complexity theory, but within the theory of open systems, as does Von Bertalanffy (who proposes the same argument (1968:45)), but as in the case of the latter, his understanding of an open system is highly similar to a complex system. One also finds a comparable similarity between the theory of open systems and that of complex systems in Steyn (1984).
[14] Carl Jung referred to the comparable concept of synchronicity, but as Csermely (2006:84) writes, the majority of Jung’s examples are “difficult to accept […] in the current state of scientific knowledge.”