Realism and Psychological Science [1st ed.] 9783030451424, 9783030451431

The book provides an argument why realism is a viable metatheoretical framework for psychological science. By looking at

460 97 2MB

English Pages VIII, 151 [157] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Realism and Psychological Science [1st ed.]
 9783030451424, 9783030451431

Table of contents :
Front Matter ....Pages i-viii
Introduction (David J. F. Maree)....Pages 1-12
The Methodological Division: Quantitative and Qualitative Methods (David J. F. Maree)....Pages 13-42
The Applicative Split: The Science-Practitioner Model of Training and Practice (David J. F. Maree)....Pages 43-53
The Metatheoretical Opposition: Positivism and Constructionism (David J. F. Maree)....Pages 55-92
Realism in Psychological Science (David J. F. Maree)....Pages 93-136
The Realist Image of Science (David J. F. Maree)....Pages 137-151

Citation preview

David J. F. Maree

Realism and Psychological Science

Realism and Psychological Science

David J. F. Maree

Realism and Psychological Science

David J. F. Maree Department of Psychology University of Pretoria Pretoria, South Africa

ISBN 978-3-030-45142-4 ISBN 978-3-030-45143-1 https://doi.org/10.1007/978-3-030-45143-1

(eBook)

© Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG. The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

To Leo David

Contents

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1 11

2

The Methodological Division: Quantitative and Qualitative Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Clarifying the Paradigm Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . Origin of the Qualitative Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . Constructing the Qualitative Approach . . . . . . . . . . . . . . . . . . . . . . . . Mixed Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

13 13 15 21 24 35 38 40

The Applicative Split: The Science-Practitioner Model of Training and Practice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Boulder Model and Beyond . . . . . . . . . . . . . . . . . . . . . . . . . . . . Struggle for Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Theory and Practice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Empirical Evidence and Positivism . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

43 43 43 45 47 49 50

The Metatheoretical Opposition: Positivism and Constructionism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Positivism, Empiricism and Its Ontology . . . . . . . . . . . . . . . . . . . . . . Classical Empiricism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Logical Positivism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Kuhn’s Postpositivism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Social Constructionism and Its Roots in Idealism . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55 55 56 57 58 65 67 67

3

4

vii

viii

5

6

Contents

Modernist Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Constructionist Image of Science . . . . . . . . . . . . . . . . . . . . . . Human Nature Militating Against Positivism . . . . . . . . . . . . . . . . . Critical Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69 73 80 85 90

Realism in Psychological Science . . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Scientific Realism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Entity Realism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structural Realism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Semirealism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Implications of Scientific Realism for Social Science and Psychology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Minimal Scientific Realism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mind-Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Unobservables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Commonsensibles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Critical Realism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Transitive Dimension, Epistemic Access and the Epistemic Fallacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Intransitive Dimension of Science . . . . . . . . . . . . . . . . . . . . . . The Nature of Intransitive Reality: The Transcendental Argument . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Nature of Social Reality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Process of Science . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Problem with Closure in the Natural and Social Sciences . . . . . . . . . Lessons Learned . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Situational Realism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nuovo Realismo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . “Real” Science . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

93 93 96 97 97 98

The Realist Image of Science . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Science as Criticism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Mythic Image of Science . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Quantification and Measurement in Psychology and Social Sciences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Doing Realist Science in Psychology . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

101 102 103 105 105 107 107 108 109 110 112 115 118 119 120 121 126 132 132

. . .

137 137 139

. . .

142 147 150

Chapter 1

Introduction

The argument of this book is that the view(s) of science that has been applied to psychology from early in the previous century up to now does not do justice to either psychology or science. Various voices in the past and present were raised against the application of the positivist template to psychological science. In the process this template became the mirror against which suggestions and innovations for new approaches towards studying psychology were reflected. Sometimes this reflection mutated into a straw man but it was still effectively used to forge new approaches, metatheories and methods. The aim of this book is not to provide a comprehensive historical overview of the process of utilising images of science although broad references to some events and developments will made. Others already provided wonderful and comprehensive interpretations of the development of images of science (Danziger, 1994). I should say something about my use of the term “image” because it will frequently be woven into discussions. When I speak of an image I mainly have the mythic image of science in mind. This is the template I spoke about above, mutating into various straw men. Rather than utilising terms such as template, straw man (it is probably not expedient to be too gender sensitive with a pejorative term involving straw), view, construction or approach, I will be using mythic image to indicate, on the one hand, that this image is the constructed picture people have of what science is. On the other hand, it is mythic and we know that myths usually harbour some truths but due to a long history of interpretation and re-interpretation, it can colour our perception of things severely. However, we do not know that the image is mythic: we take it for the truth. Fortunately, we can uncover the myth by pointing to its lessons and how it was utilised in stories constructed about science. It almost functions as those urban legends and myths: at some stage one may regard the rumour of the claw against the roof of the car as true, but as soon as it becomes clear that it is a myth one’s perception of the situation changes. There is a lesson in this particular myth; it harbours some truths (stay out of dark alleys!) but it can be used by anybody to manipulate people in believing mythical aspects of a phenomenon. By means of its power as myth it functions as a very © Springer Nature Switzerland AG 2020 D. J. F. Maree, Realism and Psychological Science, https://doi.org/10.1007/978-3-030-45143-1_1

1

2

1 Introduction

effective marketing ploy. If you believe the myth you will be more likely to buy the newest alarm system. Postmodernists are very adept in identifying marketing ploys: they call it rhetorical devices. It is necessary to be deconstructive with some assumptions and beliefs. When scientists, researchers, governing bodies and the public stubbornly hold on to dogmatic beliefs and practices to their own or others detriment, one should initiate a process of change. Not that this is always easy. As can be seen in an era of #metoo, violence against women and children and religious intolerance, it is not easy to change perceptions and practices. Although I do believe that we live in an age where societies tend to be more sensitive to oppressive practices on a global scale and they are able to voice their concerns much better and more publicly than in the past. Despite this freedom and the fruits of a number of intense struggles by among others political-philosophically grounded movements (such as various liberation movements, feminism, ecological movements and the like), myths, fake news and oppression abound. Applied to the topic at hand, the current image of science prevalent in psychology can be called mythic. The truth of the image is that it at some stage encapsulated some of the problems scientists, philosophers and psychologists had with positivistic metatheory. From the work of Logical Positivists some core tenets were advocated as characteristic of what it means to be scientific. These tenets formed the template for positivism in science and had an enormous influence on psychosocial disciplines. Even a cursory glance at journal articles published since the 1900s will indicate a steep increase in empirical studies in psychology over a period of 50–100 years (Danziger, 1994). The extreme proliferation of empirical studies probably cemented the positivist template in the minds of researchers and its critics. However, I do not think that all empirical studies can unqualified be characterised as positivist. Proponents in the domain of psychosocial research methods realised that the positivist template does not apply to them and much of the work they do cannot easily be fit into positivism despite its empirical nature. This is where the template became murky and the scientific image turned into a myth. The positivist template and ever strengthening mythic image of science served as a target for a strong postmodernist movement against modernist ideas of progression, individualism and science. It is not our task to explore postmodernism here but one of the central offshoots will concern us in this book, namely, social constructionism. A number of central ideas of social constructionism taken over from various postmodern directions had a fundamental influence on how we view science or at least what is called the received or traditional view. This is the view I call the mythic image of science. Social constructionism and its cousins vehemently opposed the received view of science in order to rehabilitate or regain the important dimension of meaningful human existence. It is no wonder that continental philosophy and its various emphases on meaning of being served as feeding ground for postmodernism. Phenomenology, hermeneutics, critical theory and poststructuralism played their part in challenging received science and its methods. Unfortunately, the target of attacks was more often than not the mythic image of science. Especially in the arena of methods ammunition were amply available from the sources mentioned above to render the image of science crippled and scorned.

1 Introduction

3

The fallout of the mythic image of science is serious. Beyond the fact that it is responsible for misunderstanding and miscommunication in psychology as science, it is also responsible for maintaining a number of dichotomies in psychology. These dichotomies or contrasts can be found on various levels of analyses, namely, applicative, methodological and metatheoretical. These dichotomies form constant themes in the practise, training, research and theorising of psychology. They pop up at conferences, publications, teaching and in students’ dissertations. While we continue to battle these dichotomies psychology will struggle to move forward as a science. It is an important science, one focusing on the well-being of people and societies and one that is becoming all the more important as we move into the twenty-first century. I am not claiming we can dissolve the dichotomies and polarities by unmasking the mythic image of science. However, by understanding what science is, and what the mythic image of science involves, I think we should be able to make some progress in psychological science. Obviously, this is not a new story but one discussed in various guises for as long as psychology was on the table as a science. As the various debates across decades have shown, agreement does not come easy and villains were identified across the various divides. For instance, in the famous Loughborough debate between relativists and realists, one reaction of the realist to the relativists’ claim that nothing is real but constructed was to hit the table and shout that the table is real, the stone you hit your toe against is real, death and suffering are real and so on. I am taking a different tack. It is easy to pick a fight and point out to a social constructionist such as Kenneth Gergen that because even a social constructionist takes illness seriously it must be real (and not constructed) or causal thinking is part of the constructionist and positivist’s conceptual tools (so suck it up, reality is causally structured). No, I want to pose the argument differently, namely, that we are explicitly or implicitly harbouring a mythic image of science which feeds various oppositions in our field, and if we succeed in unmasking this image, a different view of what science is might help us move psychology as science along. Let me put my cards on the table: I believe science is a realist endeavour but before the reader shouts, “another positivist!” and throws the book in the trash, bear with me for a few moments. Science is a realist enterprise but one that can accommodate a range of epistemological stances. I would like to make an argument that psychological science boils down to a critical enterprise which is not only epistemological or methodological but probably ontological. Thus, I would like to base my claim on a proper metatheoretical view for psychological science. At its core science is critical, a process of claim and counter-claim and even a constructionist can agree with this claim. A constructionist or for that matter a relativist in the proper postmodernist sense is in the business of deconstructing, challenging claims, beliefs, narratives and constructions, yes, and even realities. This essential critical movement is what I regard as science (but I motivate it as fundamental to realism). It is important to understand what the view of researchers and practitioners is about psychological science. The most obvious position living in the mind of the public is probably the view of psychology as a helping science or an avenue to selfimprovement. The latter view differs from what students of psychology believe psychology aims to do, namely, understanding human behaviour and mental

4

1 Introduction

functioning with an eye to applicative processes. Our concern in this book is the professional, academic, researcher and practitioner who moved through postgraduate studies in psychological science, knows that psychology is a science and whose work is embedded in the psychological principles taught in academic institutions. By saying that the professional knows that psychology is a science of sorts does not mean that there is one agreed-upon view of what this science involves and how it is practiced. I will, however, be advocating the fact that most of us have a particular idea of what science involves and that this view informs the mythic image of science prevalent in psychology (and probably in related social sciences). Thus, knowing that psychology is supposed to be a science does not mean that we are correctly conceptualising what this science involves! Psychology as a discipline and science have a relatively short history and some still regard psychology as an immature science, if at all. Although the roots of psychological thought are as old as recorded history ranging from the Greek philosophers to the present, psychology as a discipline’s origin usually is attached to Wundt’s establishment of a psychological laboratory in 1879. This event in itself is significant for the self-image of psychology. Psychology was largely viewed as a scientific enterprise from its inception. What counts as science was modelled on the natural and medical sciences of the day. In the laboratory of Wundt experiments were done involving careful measurements and controlled conditions. Psychology as an experimental science on par with other sciences infused laboratories and departments established throughout America. The story is relatively well known. In the early 1900s the difficulty of utilising introspection as a measurable and controllable method to investigate mental phenomena led to a shift of emphasis to observable and clearly measurable behaviour. In undergraduate textbooks Wundt is quoted as the father of modern experimental psychology and in the same breath the rejection of his introspectionism is touted as a hallmark of problems with early structuralism in experimental psychology (Coon & Mitterer, 2010; Kalat, 2008; Plotnik & Kouyoumdjian, 2011; Sternberg, Sternberg, & Mio, 2012, p. 8). What Danziger (1980a, p. 244) called the Wundt-myth is perpetuated in textbooks and is a good example of how popular perceptions influence so-called historical facts. According to Danziger (1980a, p. 248), of the 180 studies published between 1883 and 1903, only four studies referred to introspection associated with structuralism later on. The majority of studies utilised behavioural measurements. Clearly Wundt did not employ introspection as some textbooks would have it. His experiments focused on behavioural measurements such as perceiving sensations and reaction time measurements (Danziger, 1980a, p. 248). Wundt distinguished between two types of introspection, namely, Selbsbeobachtung (self-observation) and Innere Wahrnehmung (inner perception). The first Wundt regarded as unreliable whilst the second, as fundamental to psychology (Danziger, 1980a). The idea was to enable the scientist to get data in a similar way as in natural science, namely, with observing events taking place and reporting on them. The problem with self-observation as an approximation of objective and external observation is that it cannot report on events as they happen. It would always be a mediated and interpreted reflection. Retrospection was

1 Introduction

5

proposed as a way to overcome the problem of introspection but memory of the event created distortion so it was no real solution. Wundt rejected self-observation but tried to find a way that immediate experiences can be reported by means of one’s inner perception of events. Understandably, a very circumscribed range of events could be reported in this way by making inner perception accessible to external measurement (Danziger, 1994, p. 35). The latter were thus limited to experiments of sensations. The solution to the difficulties of introspection was a focus on measuring behaviour. Here one can quote the work of Wundt or someone similar. Wundt is maybe not the best example because as Danziger (1980b) argued, Wundt had a principled different view of psychological science than his peers and those that came after him. Wundt viewed the psychological as dealing with the mental and the methodological access to the latter is made possible by physiological measurement and experimentation. However, contrary to positivist dogma, data for Wundt was empirical and not loose standing and accumulated facts. Empirical data had to cohere to answer theoretical questions about the psychological, an attitude that was surprisingly postpositivist. However, his contemporaries largely ignored his methodological contributions in favour of positivist approaches to psychological science. Wundt had a very specific view on the psychological experiment. It is a precise and careful manipulation of conditions in order to have a desired psychological effect. The experiment was not merely designed to find correlations between events but to find underlying explanatory causal mechanisms. Although he spoke of psychological or psychic cause as opposed to causality in the natural domain, his aim was to understand and uncover those causal processes underlying psychological events (Danziger, 1980b). The development of behaviourism with its negation of the mental gained prominence in the first half of the twentieth century. In the 50s and 60s cognitivism developed as a reaction to the wholesale externalisation of behaviourism bringing the focus to the mind and mental operations. Cognitivism under which diverse fields such as cognitive psychology and computer sciences can be grouped, became consolidated as cognitive science. Advances in biology and neurosciences added further impetus to cognitive science and since the 1990s neuroscience became the catch phrase of the early twenty-first century. The prominence of neuroscience, neuropsychology and cognitive science currently dominates much of the empirical endeavours of psychology to such an extent that numerous departments remove the term psychology from their names. Experimental work, controlled trails and precise measurement are essential to its image of science. However, looking at the development of professional psychological associations in the USA the apparent straightforward line from Wundt to current neurosciences was not so straightforward. The development of psychology in the USA corresponds with the development of the APA and APS. The Wundtian laboratory was central to the image of science of psychologists but slowly since the early 1900s a large practitioner base developed in reaction to the exclusive experimentalist focus. Practitioners also developed in other areas of psychology such as occupational, education and counselling psychology. The balance towards practitioners became so extreme that it was suggested that APA include practitioner in its name. The

6

1 Introduction

criticism against APA was the loss of a scientific and research focus (which is not really true) (Ardila, 2007), and a group of research-minded psychologists initiated the establishment of the APS. However, it should be apparent by mere inspection that both associations have applicative and scientific foci. The pendulum is constantly moving between practitioner and science issues. In the following chapters the issue of the Boulder model of scientist-practitioner will be discussed. The reaction back to a scientific orientation is of course not uncontested. Early in the growth of cognitivism, which incidentally did much for psychology reclaiming the mental, reactions developed against the view of science, research and what developed into a psychological image of humanity. Studies in artificial intelligence, linguistics, computational psychology, cognition, memory, perception and so on multiplied and flourished but along the way their view of the mind was restricted to computational processes residing in the brain. Various reactions developed. An emphasis on humanistic approaches was called for. Psychology should capture the humane and human in its work, something cognitivism struggled to do. Rogers and others turned towards the subject and the importance of the individual to change and function properly in the world. Non-directive techniques along with the primacy of the individual were emphasised. Psychology following the medical model emphasised training practitioners especially clinicians in diagnosing and treating psychological deficits. The medical model based on empirical research strengthened the empirical scientific image in psychology in addition to establishing the diagnostic and symptomatic approach to treatment. This model, related but independent from psychiatry, culminated in the various DSM diagnostic manuals and played an integral role in the training of clinical psychologists. The image of science arising from this can be summarised as one focusing on external and measurable behaviour. How does this square with common sense, namely, our intuition that I am feeling things, I have a sense of identity and I am the one that has to stop my bad habits and so on? Social constructionism was a reaction to cognitivism, individualism and the deficit model. Interestingly, positive psychology developed in the footsteps of humanistic approaches and social cognitive approaches also as reaction to the deficit medical model and behaviourism’s one-dimensional treatment of human beings. Currently positive psychology is developing on its own course and subscribes to the empirical scientific image of science. On the other hand, social constructionism reacted much more vehemently to the traditional empirical, practitioner and cognitivist approaches to psychology. Shotter (1993), for instance, who started his career within cognitive psychology, rejected the latter’s view of the self and subsequently developed his social constructionist views (Lock & Strong, 2010). Gergen (1982) who similarly started off in empirical science and statistics is one of the leading figures in social constructionism in psychology. Constructionists think that the only way to understand people is linguistically and that the full grasping of the other (and the self) can only happen in interaction or in relation to other people. Something similar to what happens in a conversation or a game takes place. In the act of talking to another person, mutual meanings are constructed, misunderstood, clarified, co-constructed and so on. This

1 Introduction

7

dynamic interplay between people in conversation is regarded to be so powerful that a narrative approach to therapy is advocated strongly by constructionists. How does this differ from what Freud and other psychodynamic practitioners did? This is not the place to analyse psychodynamic theory but suffice to say that over and above processes such as assessing people with different instruments and changing behaviour and mental processes chemically, psychology’s fundamental access to other people is linguistic, i.e., by talking and questioning. The constructionist and interpretivist view of meaning construction, according to them, explains why talk therapy works. Understanding and changing people are effected linguistically, but how does this assist us in understanding and explaining psychological phenomena such as consciousness, memory and personality (I am being facetious, fully realising the antiessentialism of constructionists)? This question touches upon a number of aspects of our image of science. Some would say that understanding personality or intelligence can be sufficiently done by measuring these phenomena by assessment instruments. Others would advocate experimental studies and so on. But what can we learn about memory, for instance, by engaging in a conversation? The processes and reactions playing out in American psychology are also reflected in the development of psychology and psychological science in South Africa. The reaction against scientific psychology or the particular image of science in South Africa was largely based on socio-political reasons. Science, experiment and measurement were viewed as positivistic and South African psychology were preoccupied with eradicating measurement from its vocabulary in order to push a qualitative agenda based on the belief that it is a necessary corrective to positivist views of science in psychology (Stevens, 2003, p. 201; Terre Blanche, Durrheim, & Painter, 2006). Psychology for the past 100 years in South Africa was dominated by a white male minority, which within the apartheid ideology, did not help the case of psychology at all (Nicholas, 1990; Stevens, 2003, p. 190). Before the end of apartheid, and for 20 years after its demise, a qualitative based psychology and methodology based on a largely constructionist and critical view of psychology were advocated by liberal and mainly black academic institutions (Stevens, 2003, pp. 191, 203; Zietkiewicz & Long, 1999). The historically white institutions were to some extent slow on the uptake but critical and constructionist psychology won the day eventually. Understandably, because of racial and gender discrimination, the coarse neglect of living conditions, health and mental health of Black communities, and South Africa’s poor human rights record in general, critical psychology and its local and international proponents were recruited for the purpose of deconstructing traditional psychology and opening the field to transformation (De la Rey & Ipser, 2004; De la Rey & Kottler, 1999; Duncan, Stevens, & Canham, 2014; Hickson & Kriegler, 1991; Ratele, 2003). The endeavour was relatively successful and in the process professional psychology is still being transformed. However, this critical resistance and transformation came at a price, scientifically speaking. Building on the insights of traditional qualitative and critical researchers, qualitative psychology was promoted on grounds of the belief that quantitative psychology and methods are positivistic, and that

8

1 Introduction

positivism is fundamentally not able to account for the rich meaningful experience of social groups, and of course, politically and psychologically suppressed minorities/ majorities (Hickson & Kriegler, 1991; Stevens, 2003). In addition, psychological measurement is seen as instrumental in maintaining racial divides by consistently showing Black groups as mentally and psychologically inferior to whites (Louw & Foster, 1992). The very necessary critical work has been done but in the process psychology students in the country became functionally illiterate with regard to quantitative methods. It is, consequently, of utmost importance to clarify and unravel the underlying philosophical and methodological muddles, so that psychology as science can get back on track. The current climate in professional psychology in South Africa is almost similar to that of the USA 20–30 years ago. Professional and practitioner training dominates postgraduate education and research and scientific foci struggle to get recognition despite many universities seeing themselves as research institutions. The practitioner-scientist imbalance is nowhere more apparent than in the struggle of research psychologists (unique to the South African system) to be recognised as professional and registered psychologists. The strong emphasis universities have on research is at odds with practitioner training and it will be interesting to see how the practitioner-scientist division of labour will crystallise in the university and private training college developments in South Africa. This book was conceived in this tension and although it is determined by unique socio-historical forces in South Africa there are surprising commonalities with the development of psychology in other countries. Mackay and Petocz (2011, pp. 18–21) suggest that the current major themes in psychological science and research can be divided in three broad directions, namely, a cognitive/neuroscientific experimental grouping, a unreflective metatheoretical group and the opposition to the mainstream consisting of mainly postmodern and constructionist thinkers. (a) The traditional cognitive-based theories (Fodor, Minsky and others) fuel the cognitive/neuroscientific experimental group. It is closely allied to neuroscience, its epistemological position is largely representationalist and it views the brain as a computer. (b) The second group is in some sense allied with the first. The unreflective metatheoretical group is the majority group (Mackay & Petocz, 2011) and consists of those that support the hope of detailed discoveries made by the first group. They work in diverse areas such as personality, social psychology, applied psychology, emotion and cognition but they are theoretically unreflective and in a sense naive. Mackay and Petocz (2011, p. 19) hint that the statements about what psychological research involves found in many methods textbooks express the sentiment of the second metatheoretically unreflective group. Research is done with the rather naive assumption that psychologists study behaviour and the processes underlying it. This conglomeration of theories and studies assume physiological and mental processes underlie behaviour, that it is possible to measure these processes and to utilise

1 Introduction

9

various statistical methods to establish relationships between constructs. There are literary thousands of constructs posited, operationalised and measured. This grouping aligns itself with the so-called cognitive revolution but differs from the first meta-position by being critically unreflective. It is followed by the majority of psychologists, can be characterised as eclectic and is caught up in the false belief that psychology as a science is progressing (Mackay & Petocz, 2011, p. 19). (c) The third group consists of those opposed to the mainstream. It is heterogenous and consists of social constructionists and other postmodern approaches mentioned above. Importantly, as Mackay and Petocz (2011, p. 20) say, this group is united in its rejection of the mainstream view, usually typified as positivist. The group also rejects personality and social psychological theories as well as the experimental-computational grouping. The group advocates the constructionist nature of knowledge, rejects traditional empirical psychology and makes a clear distinction between methods for the natural and social/human sciences (Mackay & Petocz, 2011, p. 21). Underlying all three streams are particular images of science and from the discussion about the origin of science and experimentation above, the outlook for realism as a metatheory for psychology seems bleak. However, the story does not end here. Thousands of institutions train psychology students and they have been trained under the umbrella of a particular idea of what psychological science is, namely, the scientist-practitioner model. The emphasis of this model is that a practitioner should first of all be a scientist; although harbouring a lofty and valuable ideal, it is a model that probably did psychological science more damage than good. In order to get to the discussion of realism as a metatheory for psychological science let me briefly provide an overview of the discussions in the book in the order they appear in the chapters. When referring to the polarities in psychology, I will use the various terms such as contrast, dichotomy or split interchangeably. The polarities appear on three levels which in order from concrete to conceptual are the applicative, methodological and metatheoretical. However, the discussion starts with the methodological, followed by the applicative then the metatheoretical. The reason for this order is that the methodological probably houses the most pervasive dichotomy in the psychosocial sciences, namely, between quantitative and qualitative methods. It provides an appropriate introduction to defining the mythic image of science and allows us to identify one of the important reasons the scientist-practitioner applicative split is maintained. Finally, the image of science will be examined in the two metatheories, positivism and constructionism, that clash in the previous polarities. On the methodological level the quantitative-qualitative divide perpetuates the impression that to be scientific is intimately associated with measurement and quantification. It is on the methodological level that researchers eschew the positivist template and maintain the importance of human existence as meaningful. Because numbers cannot encapsulate meaning and meaningful experience, the only sensible way to study people is qualitatively. The arguments motivating qualitative methods range from sophisticated to merely repeating the mythic image of science.

10

1 Introduction

Unfortunately, these arguments get reiterated without much critical thought and find their way into research methods textbooks. Thus, even the student utilising quantitative methodology characterises her approach as positivist without even batting an eyelid. The applicative level refers to the practice of psychology as a profession. On the applicative level the scientist-practitioner split informs the mythic image of psychological science. Students are trained as psychologists and get exposed to practical training in therapy and counselling but also in the science of establishing, devising and testing therapeutic interventions. However, in reality psychologists as practitioners pay little attention to matters of science. The category of scientist-practitioner was created in the 1940s in Boulder, Colorado to encourage an integrative mindset for training psychologists. This integrative principle became a split and forms one of the dichotomies psychology as a science and profession still struggles with. The final level where the mythic image of science maintains a dichotomy is that of the metatheoretical. This level refers to the mainly philosophical discussion and theorising about what science studies and how it can know anything about what it studies. Our interest is psychological science or psychology as science with implications for social sciences where the subject of psychology features. Metatheoretical issues thus have something to say about what there is, namely, what reality consists of or ontology. In addition, it considers questions about how knowledge about this reality is possible and what the nature of this knowledge is; it therefore investigates epistemological questions. In the philosophical context these metatheoretical issues would be considered under the topic of a theory of science or philosophy of science. Examples of metatheories are realism, relativism, social constructionism, idealism and so on. Some metatheories focus on both ontological and epistemological issues while others emphasise the one or the other. Social constructionism, for example, holds that knowledge of reality is constructed linguistically and socially to such an extent that ontological questions recede into the background or is impacted by epistemological assumptions. The latter is the case with strong versions of social constructionism which maintain that reality as such is constructed by minds. Different communities experience different realities. Weak social constructionism acknowledges different interpretations of a relatively similar reality, i.e., different communities experience the same reality differently. Realism, in contrast to constructionism, holds ontology to be primary and epistemological questions secondary. The implication is that reality is essentially mind-independent: what there is cannot be determined by the mind. However, various realisms have different answers about how this reality is known. We will encounter some of these realist versions later on. Positivism as a metatheory has implications for reality and has much to say about knowing reality. Some things can be known if they in some way can be brought back or related to observable events or phenomena. The observable part refers to positivism’s empiricism. Thus, one of the names for the brand of positivism found in the early part of the previous century was Logical Empiricism. It was also known as positivism or Logical Positivism, the label which will be used here. For the Logical Positivist, knowledge or facts are anchored in empirical reality. Their ontology is

References

11

thus an implicit empirical realism, a term and view that should not be confused with the type of realism discussed later on in the book. The proposal put forward in this book, epistemologically and ontologically speaking, is based on realism. We would like to identify the mythic image of science and move towards a proper understanding of what science is. This means avoiding the mistakes and incorrect assumptions of both social constructionism and positivism. It would be important to examine whether realism has any commonality with either social constructionism or positivism and how realism differs from both. In order to develop an intelligible metatheory for psychological science it is crucial to tease realism and positivism apart because many misconceptions about both are lodged in the mythic image of science.

References Ardila, R. (2007). The nature of psychology: The great dilemmas. American Psychologist, 62(8), 906–912. https://doi.org/10.1037/0003-066x.62.8.906. Coon, D., & Mitterer, J. O. (2010). Introduction to psychology: Gateways to mind and behavior (12th ed.). Belmont, CA: Cengage Learning. Danziger, K. (1980a). The history of introspection reconsidered. Journal of the History of the Behavioral Sciences, 16, 241–262. Danziger, K. (1980b). Wundt’s psychological experiment in the light of his philosophy of science. Psychological Research, 42(1–2), 109–122. https://doi.org/10.1007/bf00308696. Danziger, K. (1994). Constructing the subject: Historical origins of psychological research. Cambridge: Cambridge University Press. De la Rey, C., & Ipser, J. (2004). The call for relevance: South African psychology ten years into democracy. South African Journal of Psychology, 34(4), 544–552. Retrieved from http:// repository.up.ac.za/handle/2263/14269. De la Rey, C., & Kottler, A. (1999). Societal transformation: gender, feminism and psychology in South Africa. Feminism and Psychology, 9(2), 119–126. Duncan, N., Stevens, G., & Canham, H. (2014). Living through the legacy: The Apartheid Archive Project and the possibilities for psychosocial transformation. South African Journal of Psychology, 44(3), 282–291. Gergen, K. J. (1982). Toward transformation in social knowledge. New York: Springer. Hickson, J., & Kriegler, S. (1991). The mission and role of psychology in a traumatised and changing society: The case of South Africa. International Journal of Psychology, 26(6), 783–793. Kalat, J. W. (2008). Introduction to psychology (8th ed.). Belmont, CA: Thomson. Lock, A., & Strong, T. (2010). Social constructionism: Sources and stirrings in theory and practice. Cambridge: Cambridge University Press. Louw, J., & Foster, D. (1992). Intergroup relations and South African social psychology: Historical ties. Canadian Psychology/Psychologie Canadienne, 33(3), 651. Mackay, N., & Petocz, A. (2011). Realism and the state of theory in psychology. In N. Mackay & A. Petocz (Eds.), Realism and psychology: Collected essays (pp. 17–51). Boston: Brill. Nicholas, L. J. (1990). The response of South African professional psychology associations to apartheid. Journal of the History of the Behavioral Sciences, 26(1), 58–63. Plotnik, R., & Kouyoumdjian, H. (2011). Introduction to psychology (9th ed.). Belmont, CA: Cengage Learning.

12

1 Introduction

Ratele, K. (2003). We black men. International Journal of Intercultural Relations, 27(2), 237–249. https://doi.org/10.1016/s0147-1767(02)00094-9. Shotter, J. (1993). Conversational realities: Constructing life through language. London: Sage. Sternberg, R. J., Sternberg, K., & Mio, J. S. (2012). Cognitive psychology. Belmont, CA: Wadsworth/Cengage Learning. Stevens, G. (2003). Academic representations of race and racism in psychology: Knowledge production, historical context and dialectics in transitional South Africa. International Journal of International Relations, 27, 189–207. Terre Blanche, M. J., Durrheim, K., & Painter, D. (2006). Research in practice: Applied methods for the social sciences (2nd ed.). Cape Town: UCT Press. Zietkiewicz, E., & Long, C. (1999). Speaking for a change: Strategies for transgressive praxis. Feminism and Psychology, 9(2), 142–151.

Chapter 2

The Methodological Division: Quantitative and Qualitative Methods

Introduction The quantitative-qualitative divide is one of the most prominent divisions in social science and especially psychology. Of course, the emphasis placed on the one or the other depends geographically. In South Africa, the use of qualitative approaches overshadows quantitative approaches in some areas. One reason for the qualitative emphasis is its close association with critical psychological approaches utilised in race and gender projects. A good example would be that of the Apartheid Archive project which involved inviting ordinary citizens to share their experience of racism during the Apartheid years in South Africa (Duncan, Stevens, & Canham, 2014). Another example would be the demise of psychometric test development in education and psychology in South Africa due to critical opposition to discriminatory practices in utilising psychometric tests in schools and industry (Ferreira, Maree, & Stanz, 2016). The implication is that the use of various methods either enabled discriminatory practices or facilitated critical exploration and dismantling of structures and processes used for suppression of people. This very brief sketch of the socio-political situated emphasis on qualitative methods is an indication of the utility of qualitative methods on the one hand but also the rhetorical strategy employed to strengthen the case for qualitative methods. This chapter explores how the quantitative-qualitative divide contributes to establishing the mythic image of science. More to the point, one may assume a reciprocal relationship between the two aspects. In a sense, the mythic image of science is informed by the polarity but also maintains it. For example, the image lets us believe that the hallmark of science is the ability to measure constructs. Thus, the reaction against the received view of science is to eschew measurement as not applicable in the psychosocial domain. On the other hand, those working with measurement is confronted with a realisation that an activity such as cross-cultural assessment poses difficulties for comparing cultures and groups because (a) of the classical measurement models used and (b) realising that real group differences © Springer Nature Switzerland AG 2020 D. J. F. Maree, Realism and Psychological Science, https://doi.org/10.1007/978-3-030-45143-1_2

13

14

2 The Methodological Division: Quantitative and Qualitative Methods

cannot be negated. In the classical test theory (CTT) model test characteristics depend on sample characteristics thus tests are not transportable across groups. Thus, measurement methods are forced to refine and seek solutions. The movement from classical test theory to item response theory is a good example of this tendency. The result is the impression that refined and different measurement models and techniques would provide valid answers to scientific investigation thus also reinforcing the quantitative-qualitative polarity. Almost as soon as qualitative methods became popular advocates for mixed methods started canvassing the combination of quantitative and qualitative methods. Reasons for combining methods are numerous and we will explore some of them, but one may assume that mixed methods are viewed as useful to a number of researchers realising that neither measurement nor interpretative analysis is sufficient on their own to understand psychosociological phenomena. The popularity of mixed methods is not indicative of the picture of science becoming clear. It is true though that measurement played and is still playing a strong role in what is regarded as science but also in what is executed as research in practice, but I think the mythic image of science inadvertently bolsters measurement as a fundamental characteristic of science rather than a methodological avenue for understanding. Constructionism and related postmodern variants feed off the mythic image of science and justifies its position in social science against this image. In this chapter we would like to investigate how the establishment of the quantitative-qualitative divide along with the attempt to bridge this divide both contributed to forming the mythical image of science. In the following section we look at the concept of paradigm because it plays a large role in setting the context for both advocates of qualitative and mixed methods. I will briefly discuss the development of the characteristics of the various paradigms found in the work of Denzin and Lincoln (2018b). This work was deliberately chosen because of its wide influence and longevity as a standard reference book for qualitative researchers. Finally, we will briefly discuss the strategy of mixed methods. In order to prepare the reader for what is coming the non-mythic corrective can be stated bluntly: measurement is not a characteristic of science but one of the ways of understanding and explaining but always appropriate to its subject matter. Quantification, inferential statistics and statistical data manipulation provide epistemic access to certain but not all constructs and phenomena. It is important to determine the limits of epistemic access of methods and their subject matter. It is more effective to utilise a hammer for knocking nails into a wall than to remove a splinter from a finger. However, a hammer might be utilised for more than just hammering in nails (like removing cement from a wall). Thus, interpreting phenomena is not contrary to the scientific enterprise, both natural and social. The constructionist fundamental principle of questioning all claims and assumptions is essentially scientific. These correctives are metatheoretically justified by realism which is discussed later on.

Clarifying the Paradigm Concept

15

Clarifying the Paradigm Concept The postpositivist debate in the 60/70s focused on a number of issues pertinent to natural science. The story is reasonably familiar, namely, Kuhn (1996) published his book about paradigm shifts in 1969, focusing on how scientific traditions might supersede one another. Issues such as the relationship between theory and empirical data, the growth of scientific knowledge, the demarcation of science from non-science, the underdetermination of observation by theory, truth of scientific theories and the like were debated (Chalmers, 2013). The postpositivism of Feyerabend, Kuhn, lead to contributions from Rorty which on its turn informed qualitative and constructionist theorising. The paradigm concept provided a fitting vehicle for investigating the history of science to describe shifts in frameworks and of course incompatibility of frameworks. However, the concept of paradigm as a core construct of scholarly thought has many meanings and refer to a number of contexts. Thus, what different researchers would like to express with the term paradigm might not necessarily have the same meaning (Morgan, 2007). One can propose a number of ways of defining a paradigm but I would like to focus on a suggestion of two sets of conceptions. The first refers to a set of concepts indicating the scope of conceptual frameworks. Generally speaking, a paradigm refers to a conceptual framework similar to but broader in scope than a theoretical model, or theory. Thus, a model is a restricted conceptual framework offering a partial description of a phenomenon or parts of the phenomenon. A theory is a broader and coherent conceptual framework that explains the phenomenon whilst a paradigm is a collection of theories similar in theme, usually focused on a topic of investigation, but still different explanatory frameworks. An example of the three conceptual frameworks might be the following. Maslow’s (1943) familiar pyramid model of hierarchical human needs is an example of a restricted but preliminary conceptual framework, a proto-theory if you will. His (1954) broader theory serves as an explanatory framework for the phenomenon of motivation within which the hierarchy of needs are embedded (Kenrick, Griskevicius, Neuberg, & Schaller, 2010). The collection of different but thematic similar theories would serve as a humanistic paradigm. One can, for instance, distinguish this humanistic paradigm from others such as the behaviourist or psychodynamic paradigms. This view of a paradigm is hierarchical and nested. The model is the most basic and specific conceptual framework and the paradigm the most general. Within a paradigm the theory is imbedded, followed by the model. However, this view of paradigm might not be what Kuhn had in mind. Morgan (2007) extracted four senses of the term paradigm that are also hierarchical and nested. From the specific to the most general they are (a) the paradigm as a model or exemplar, (b) paradigms as shared beliefs among members of a specialty area, (c) paradigm as an epistemological stance, and (d) paradigm as worldview. Let us start with the broadest framework. A paradigm as a worldview is the broadest framework involving how we think about the world including values and beliefs. A worldview in this sense is much

16

2 The Methodological Division: Quantitative and Qualitative Methods

broader than a conceptual framework in the sense described above although one may argue that those aspects forming a worldview, namely, one’s beliefs, morals, values and so on inform the conceptual framework one develops about the world. One can thus distinguish between two meanings of conceptual framework. The first set of frameworks discussed above refer to the content of a discipline or area of investigation sans a-theoretical issues. The use of paradigm as worldview as specified by Morgan (2007), fuses the contexts of discovery and justification. Thus, theoretical and a-theoretical contexts are combined in forming the worldview. The paradigm as epistemological stance refers to philosophical systems but circumscribed in terms of scope and application, such as constructivism or realism. It still relates to a worldview but is much narrower because it focuses on the nature of research or doing science. The nature of knowledge within a particular epistemological approach is the main focus. Deeper issues may be implied such as ontology, and it is usually these ontological issues that may lead to fundamental paradigm incompatibility in the social sciences. When we discuss realism and constructivism in this book, the epistemological, methodological and ontological are made explicit, therefore, our decision to call these onto-epistemological stances metatheories. Morgan (2007) says that this use of paradigm as epistemological stance reflects how most social science research discussions use the term. This is also how Guba and Lincoln (1994, 2005) and Lincoln, Lynham, and Guba (2018) use the term. The third use of paradigm refers to a more specific level of shared beliefs of smaller groups of specialists within specific areas of speciality and the participants’ agreement on what to study and how especially within the context of natural science. Morgan (2007, p. 53) points out that this was Kuhn’s preferred use of paradigm. This sense is not popular in the social sciences but when it was appropriated it referred to disciplines (such as nursing) which is broader than Kuhn’s smaller specialist groups and communities sharing beliefs and practices. Kuhn’s discussion of paradigm incommensurability and paradigms superseding each other applies to this restricted sense. A popular example that, to the mind of the non-natural scientist, is associated with a paradigm incommensurability and supersession, is the Newtonian and Einsteinian views of time and space (to the social scientist’s ear their influence on physics and worldview sounds admittedly very general and far reaching and certainly not like the ideas of a small specialist group). Finally, paradigms as exemplars or model examples refer to how research is done in various fields (Kuhn, 1996, p. 187). This sense was used by Kuhn in a limited way but an increasing number of social scientists tend to provide exemplars of how research is done by means of providing method templates (Morgan, 2007). A good example is Tabachnick, Fidell, and Ullman’s (2019) discussion of full examples of case studies utilising particular multivariate analysis techniques. The four meanings of paradigms range from very specific and restricted (models) to the most general view of reality or the world (worldview). Kuhn’s aimed at the second level of specificity whilst paradigm as an epistemological stance is mostly used in the social sciences. This third level is the same as metatheory as is used in this book but for this chapter, paradigm2 (metatheory) will be used. The sense of paradigm in the list of conceptual frameworks discussed initially fits well with

Clarifying the Paradigm Concept

17

Fig. 2.1 The intersection between two sets of paradigm classification schemes

Kuhn’s shared beliefs about knowledge and processes of a specialist group with slight expansion in the context of social sciences. Thus, humanism, psychodynamic theory and behaviourism could be appropriate for level two although Kuhn might have had specialist groups in mind within, for instance, humanism or psychodynamic theory. This sense of paradigm will be called paradigm1. Figure 2.1 summarises the way the two sets of understanding paradigms intersect. Paradigms1 refer to fields of speciality or knowledge as the result of scientific work. The latter consists of the interplay between theory and empirical methods. The metatheoretical then, considers paradigms1 or themes of theories from the theory of science vantage point looking at the underlying methodological, epistemological and ontological matters. This means that one may consider, for instance, humanism on a metatheoretical level. Metatheories that would be important are those that Guba and Lincoln (1994, 2005) and Lincoln et al. (2018) utilise as paradigms2, namely, positivism, postpositivism, interpretivism and others such as constructionism. The characteristics of a paradigm1 Kuhn specified are, among others, its tendency to be regarded as the normal way science on a particular topic is done. The different phases in the development and demise of a paradigm, or as (Kuhn, 1996, p. 182) called it, a disciplinary matrix, are normal science which involves puzzle solving. At some stage the participants and critics of the paradigm find them unable to solve issues, called anomalies, and a new way of thinking about the field and its methods is required. A paradigm overthrow means that the previous paradigm in terms of its theoretical explanations of phenomena is incompatible with the newer paradigm. This is why paradigms are incommensurable which may imply incompatibility but not necessarily so. Thus, in Kuhn’s description paradigm overthrow and succession are characteristic of the progress of science. In the social sciences and, for instance, psychology we do not experience similar paradigm successions on the level of subject matter very frequently although there are some examples of paradigms that are of historical interest. Titchener’s structuralism, behaviourism, psychophysics, classical

18

2 The Methodological Division: Quantitative and Qualitative Methods

psychoanalytic, information processing theory in cognitive psychology and so on may qualify as superseded paradigms. Neuroscience and its impact on psychology can also be regarded as a tentative new framework for studying psychology and human behaviour especially on the level of popular psychology. Various forms of positive psychology in the humanist paradigm are in favour in contrast to deficit models. Although there are thus examples of superseded subject matter paradigms1, others remain simultaneously active as valid research programmes within psychology. We cannot speak of one dominant paradigm1 in psychology. Even the neuroscientific perspective does not dominate all areas of psychology; subject matter pluralism seems to be accepted within psychology. What I have characterised as subject matter pluralism, others call different schools of thought and characterise them as not yet on the level of a paradigm (Weimer & Palermo, 1973). In this sense, psychology is a pre-paradigmatic science (Hergenhahn & Henley, 2014). However, the fact that reasonably coherent research programmes, as mentioned above, were superseded by others on the level of subject matter, makes it appropriate to characterise them as paradigms1 (Weimer & Palermo, 1973). Thus, at any moment in psychology’s history there are a number of viable paradigms guiding researching programmes. Psychology can therefore be viewed as a multi-paradigmatic discipline, which is not what Kuhn had in mind, but it is a description that probably fits the state of the psychosocial sciences better (Hergenhahn & Henley, 2014). However, in contrast to subject matter or multi-level paradigms, Gross (2017) insists on identifying a singular paradigm, namely, modernism with its associated view of science that is in the process of becoming superseded by the new paradigm in psychology. The modernist characteristics of the old paradigm associated with psychology are the following: essentialism, individualism, universalism and decontextualisation. Since the enlightenment science has served as a deep correction for religious authoritarianism. The resulting scientism in psychology, namely, incorporating natural science as the model of doing science into its own self-image, is closely related to the four issues mentioned above. For instance, essentialism means that psychology regards certain properties as constitutional, thus, making a person what she is (Gross, 2017, p. 68). This essentialism is bolstered by individualism or the belief that agency resides fully within the individual. In addition, because one individual, like an atom, can be substituted for another, understanding one is generalisable and people function the same despite their contexts, cultures and history (decontextualisation) (Gross, 2017). In psychology, these modernist principles and the scientistic principles reciprocally support each other. People are seen mechanistically and as atoms whose patterns of behaviour establish linear cause and effect and allows prediction of future behaviour. Psychological science requires the exercise of particular methods in order to arrive at certain and true knowledge. This “methodological imperative” is also known as a methodolatry stance: The reason for this methodolatry has more to do with the way Psychologists wish to present themselves than with an open-minded evaluation of its practical utility: it offers the credentials of being ‘serious’ and ‘scholarly’, since empirical investigation can only be conducted by ‘experts’, and so confers exclusivity (Gross, 2017, p. 70).

Clarifying the Paradigm Concept

19

Gross (2017) points to one reason for insisting on sticking to methods, namely, a conscious or unconscious motive that has got nothing to do with science but everything with subjective and psychological needs that can express themselves in a tendency to foster a sense of exclusivity and self-importance. It might be expressed as a will to power or the need to validate oneself as the expert, or maybe involve some other form of scholarly posturing. These personal issues form part of one’s constitution as a human being and as a scientist. Idiosyncrasies, bias, values and even political agendas form part of our background knowledge or horizon of understanding. What is at stake in a modernist paradigm are not these subjective issues but the epistemic consequences of viewing methods and numbers as recipes for truth. In contrast, the “new paradigm” incorporates concepts from social constructionism, hermeneutics, feminist theory and discursive psychology, positions that can be characterised as critical psychology (Gross, 2017). What is thus characterised as old and new paradigms are not so much subject matter or particular theories of human functioning but ways of practicing science. The old paradigm engenders experimental, empirical and quantitative methods whilst the new paradigm focuses on acknowledging humanity in all its subjectivity and meaningfulness. The one paradigm is built upon epistemological assumptions of the possibility of objectivism, methods leading to certain knowledge that applies universally, whilst the new paradigm sees knowledge and truth (if it exists) as relative, local and historical. These issues lie on a metatheoretical level but also on the worldview level so we can certainly ask whether we are justified to talk about paradigm conflicts at this level of analysis? To recap: the levels of analyses that we have identified from the second set of paradigm usage discussed above, are exemplar/model, paradigm1 (subject matter level), paradigm2 (metatheory), and worldview. Gross (2017) contrasts paradigms at the paradigm2 level because of the epistemological and methodological issues he compares. Morgan (2007) follows a similar strategy. Positivism was superseded by the qualitative paradigm which Morgan (2007) calls the metaphysical paradigm. The reason for grouping the different epistemological stances in the qualitative camp under metaphysical is because of their top-down argument for a different ontology, epistemology and methodology than that of positivism. Thus, the motivation for choosing any of the qualitative approaches falls within the paradigm2 level of analysis but explicitly goes beyond the epistemological stance Kuhn assigned to this level. The inclusion of ontology makes it a metaphysical paradigm. Morgan (2007) then argues that the pragmatic paradigm can supersede the metaphysical paradigm, an issue we will discuss below. However, Morgan (2007, p. 56) points out that the metaphysical paradigm discussion focused on the nature of research, thus on methodology. Methodology is not the same as methods. The latter refers to the actual use of particular methods while methodology refers to the justification of the use of those methods. Ontology (what reality is) and epistemology (how we know that reality) provide the justification. The contrast, discussion and choice between different approaches thus lie on the level of paradigm2 (metatheory) and not on the level of subject matter or paradigm1.

20

2 The Methodological Division: Quantitative and Qualitative Methods

The emphasis on the nature of research or the methodological aspect of paradigm2 can be clearly seen with talk of quantitative and qualitative paradigms (Collins, 2015; Michell, 1997; Onwuegbuzie & Collins, 2007). However, is this not putting the cart before the horse? It would not make sense to place the quantitative and qualitative paradigms on the same level as humanism, psychodynamic theory or behaviourism. Different groups of theory (or specialist areas) in a humanistic paradigm, for example, might employ qualitative and quantitative (or mixedmethods) approaches. Thus, what is it about the quantitative/qualitative divide that makes it so compelling that it even manages to eclipse subject matter? Guba and Lincoln (1994) point out that the terms quantitative-qualitative are sometimes used as if they are paradigms. Talk of quantitative-qualitative “paradigms” probably arose from repeated emphasis on quantification as the main characteristic of positivism and that a proper and mature science fully utilises mathematical methods: The “received view” of science (positivism, transformed over the course of this century into postpositivism; see below) focuses on efforts to verify (positivism) or falsify (postpositivism) a priori hypotheses, most usefully stated as mathematical (quantitative) propositions or propositions that can be easily converted into precise mathematical formulas expressing functional relationships. Formulaic precision has enormous utility when the aim of science is the prediction and control of natural phenomena . . . In recent years, however, strong counter pressures against quantification have emerged (Guba & Lincoln, 1994, p. 106).

The question is what determines method in any discipline. Would it be sensible to say that humanism, psychodynamic theory or behaviourism determine how phenomena should be studied? The latter is tricky because its subject matter was intimately intertwined with quantitative methods from its inception. Let us take less obvious examples, for instance, health psychology, humanism, psychodynamic theory or positive psychology; it is not immediately obvious which method should be appropriate when looking at health issues in psychology. A topic on its own do not indicate which method is appropriate to investigate its phenomena. However, one would be hard pressed to find a quantitative study within feminist psychology or critical race studies in psychology. Nevertheless, quantities such as the distribution of gender within the sciences play a significant critical role in bolstering feminist perspectives. The motivation for the type of method lies in the nature of the data that a discipline and its subdisciplines work with and is justified on a methodological level, thus, in a particular metatheory applicable to a paradigm1; methodological issues are the outcome of epistemological and ontological considerations. I am therefore suggesting that talk of qualitative and qualitative paradigms is inappropriate. Methodological approaches are considered in a metatheory and one should refer to these as approaches (to data analysis). Guba and Lincoln (2005) made a distinction in their earlier 1994 work between four paradigms2, namely, positivism, postpositivism, critical theory and constructivism, and since 2005 added the participatory paradigm. The five paradigms2 are distinguished mainly on the ontological, epistemological and methodological levels, but Guba and Lincoln (2005) included further distinctions such as the nature of

Origin of the Qualitative Approach

21

knowledge, its accumulation, quality criteria, values, ethics, the inquirer posture and training. The quantitative and experimental methods are mainly restricted to positivism and postpositivism whilst the remainder utilise various discursive and qualitative methods. Positivism is characterised by a naive realism whilst postpositivism by “critical” realism. The role of the subject or community in co-constructing reality in various degrees are characteristic of the qualitative paradigms. On the epistemological level knowledge is regarded as true or false by positivists, probably true by postpositivists but transactional by the qualitative approach proponents. The influence of Guba and Lincoln’s (2005) characterisation of the qualitative paradigms may be enormous given that it was cited 32,411 times according to Google-Scholar across the different editions of their publication. The use of the concept paradigm facilitated conceptualisation of the qualitative epistemology and ontology as a coherent system. According to Morgan’s (2007) analysis, the success of the metaphysical paradigm (i.e., the four qualitative paradigms) relied to a large extent on the bolstering of its ontological assumptions and establishing an epistemological conceptual framework from which differences between research paradigms can be pointed out. The metaphysical paradigm includes the three areas of ontology, epistemology and methodology. This top-down approach of the metaphysical paradigm framed and impelled the discussion of social research methods to focus on epistemological and ontological issues rather than technicist ones. The debates justified qualitative research approaches but also permeated much of the discussion of research methodology in social science. Principally, the prevailing attitude towards methods is that of paradigmatic differences and incompatibility, meaning one cannot easily mix methods because the ontological, epistemological and methodological assumptions differ fundamentally. Morgan’s (2007) interpretation is that the metaphysical paradigm’s focus on ontological issues, i.e., matters of truth and reality, effectively cut off communication between the various paradigms, thus bolstering the idea of incommensurability between paradigms (Biesta, 2010). The approach in the next section is to look at the way Lincoln et al. (2018) described the methodological approaches and how they contrasted each in order to make a case for the qualitative approach. For the moment I will use “metaphysical paradigm2” to refer to the different qualitative paradigms as a set (Morgan, 2007). The different paradigms under quantitative and qualitative refer to paradigm2 and this is consistent with how Lincoln et al. (2018) use the terms and probably the most familiar to the reader (Guba & Lincoln, 1994, p. 105).

Origin of the Qualitative Approach Qualitative methods (QM) are largely a movement of the social sciences. Sociology, geography, psychology, cultural studies, humanities and the arts are disciplines where various qualitative approaches are utilised (Chamberlain & Murray, 2008).

22

2 The Methodological Division: Quantitative and Qualitative Methods

The development and use of QM in psychology in a sense parallels that of social science and usually the same texts are utilised. Some of the QM proponents specialised in psychology and the interpretative turn in psychology coincided largely with the use of QM (Willig & Stainton-Rogers, 2008b). Qualitative methods were used long before it became fashionable to call them qualitative methods. Qualitative methods did not appear out of the blue but was an approach, although not by this name, that was available to researchers from the early 1900s or even earlier according to Johnson and Gray (2010) (Willig & Stainton-Rogers, 2008a). However, it is generally agreed that qualitative methods as a systematic approach towards doing social science gained some prominence in the 1960s which escalated during the 1970 and 80s and continued to 2000 (Morgan, 2007; Schwandt, 2000). Denzin and Lincoln (2018b) included a chapter in their textbook since the late 90s sketching the development of qualitative methods over the years. This sketch was updated in the latest version of their narrative. Denzin and Lincoln (2005a) indicated 8–11 so-called historical moments in the development of qualitative methods (Denzin & Lincoln, 2018b). These phases of development show that the metaphysical paradigm is not a single and monolithic approach to studying psychosocial reality. The metaphysical paradigm is complex, varying in epistemologies and supports diverse methods (Schwandt, 2000). Numerous journals and related publications saw the light and it can be demonstrated that journal articles employing qualitative methods number in their thousands. Despite the growing popularity both proponents and opposition acknowledge that qualitative methods and approaches can still be regarded as a minority movement in comparison to quantitative and positivist methods (Mackay & Petocz, 2011, p. 20). Researchers invoking the qual-quant polarity can easily make the mistake of presenting both sides as relatively unitary in approach and method. Quantitative is characterised as positivistic, implying a single and unified approach to social science as opposed to qualitative which focuses on meaning and thus appropriate for psychosocial phenomena. The mythic image of science is shaped by these oversimplified generalisations and it is important to look at the complexities of both sides. According to Denzin and Lincoln (2005a) the development of the qualitative approach can be divided into several historical “moments,” namely, traditional (1900–1950), modernist (1950–1970); blurred genres (1970–1986); the crisis of representation (1986–1990); the postmodern (1990–1995); post-experimental inquiry (1995–2000); the methodologically contested present (2000–2004) and the eight comprises what Denzin and Lincoln (2005a, p. 3) call the fractured future post2005. In the most recent edition of the influential “The Sage handbook of qualitative research” these historical moments were revised in the light of past debates and developments (Denzin & Lincoln, 2018b). The paradigm wars (1980–1985) was inserted and after 2004, paradigm proliferation (2005–2010) was added. A posthumanist present (2010–2015) followed, as well as an “uncertain, utopian future, where inquiry finds its voice in the public arena” (2016) (Denzin & Lincoln, 2018a, p. 42). It must be noted that this narrative of moments apply largely to the development of qualitative methods in Northern America (Denzin & Lincoln, 2018b).

Origin of the Qualitative Approach

23

The traditional period was characterised by the dominance of positivism and by the struggle of alternatively minded qualitative proponents challenging positivist epistemology. The modernist era coincided with publications of landmark postpositivist works such as those by Kuhn (1962), Rorty (1979) and Feyerabend (1993). In this time a number of philosophical and epistemological approaches were incorporated in the repertoire of the qualitative paradigm. These include hermeneutical, structuralism, semiotics, phenomenology, cultural studies and feminism (Denzin & Lincoln, 2005a, p. 3). The blurred genre phase saw the researcher, in Denzin and Lincoln’s (2005a) terms, as a bricoleur, utilising many genres and disciplines. The crisis of representation moment indicates the movement between social sciences and humanities. Simply put, social scientists drew upon the interpretation of texts, an activity mainly located in humanities, to invigorate their methods and critical potential, whilst the humanities drew upon social theories to bolster their interpretation of local cultures. The postmodern moment became increasingly experimental and deconstructive searching for alternative evaluative criteria (Denzin & Lincoln, 2005a). Denzin and Lincoln’s (2005b, p. x) agenda for QM for the twenty-first century was clearly spelled out: they want to encourage the use of QM as a form of “radical democratic practice.” The post-2005 era Denzin and Lincoln (2005a) characterised as a critical era focusing on incisive socio-political, mainly liberal values, discussions playing out in social science disciplines and humanities. Topics of marginalisation, race, gender and alternative sexualities were critically discussed. The 80s and 90s saw key qualitative works published and set the qualitative paradigm on its course (Denzin & Lincoln, 1994), whilst some works were already published in the 70s along with new journals exploring qualitative issues (Bryman, 2008). Despite the enormous growth and adoption of qualitative methods, Denzin and Lincoln (2018a, p. 42) paint a picture of perpetual struggle for qualitative methods in the light of recent developments in academic and research contexts, namely, an environment of organisational auditing, funding based on the demonstration of quantitative efficacy and qualitative students and scholars regarded as “soft” or quasi scientists (Denzin & Lincoln, 2018a, p. 40). Instead of a decrease of reliance on quantification the world sees an intensification: The QUAN/QUAL divide is blurring; perhaps it is time to give up the war . . . Radical feminists are using biostatistics and pursuing biosocial studies. Poststructuralists and posthumanists are interrogating the underlying assumptions and practices that operate in the era of big data, digital technologies, the data sciences, software analytics, and the diverse practices of numeracy . . . (Denzin & Lincoln, 2018a, p. 32).

24

2 The Methodological Division: Quantitative and Qualitative Methods

Constructing the Qualitative Approach The precursor to Denzin and Lincoln (2018b) respected and authoritative handbook is the work of Lincoln and Guba (1985) called “Naturalistic inquiry.” In the subsequent editions of the “Handbook of qualitative methods” the term naturalistic was eventually replaced with “qualitative.” Naturalistic in this context meant an investigative approach of human experience as opposed to so-called non-natural approaches which would be the typical experimental and quantitative methods. As discussed above, dissatisfaction with these methods lead to a naturalistic approach or one amenable to the character of human nature and culture. In the following paragraphs we will look at what Lincoln and Guba (1985) said about the paradigms and then proceed to the discussion found within Denzin and Lincoln (2005c, 2018b). Lincoln and Guba (1985) made a broad comparison between the positivist and naturalistic paradigms. Naturalistic approaches included a number of qualitative perspectives whilst the largely quantitative methods were seen as positivism. Although it is understandable that Lincoln and Guba (1985) wanted a unitary target to aim at and against which they can formulate the characteristics of the naturalistic paradigm, the first step in generalising positivism can be seen here. In Table 2.1, Lincoln and Guba (1985, p. 37) summarised the difference between the two paradigms in terms of ontology and epistemology. Guba and Lincoln (1994) can be credited for making the now ubiquitous distinction between ontology, epistemology and methodology part of the philosophical characteristics of a paradigm whilst Lincoln and Guba (1985) initially indicated general characteristics now recognisable as fitting into the tripartite portrayal of a paradigm. Two of the statements apply to ontology namely, the nature of reality and the possibility of causal linkages. Lincoln and Guba (1985) regard the positivist view of reality as single, tangible and fragmentable. It is not readily apparent what a single reality refers to, but the context implies space-time uniformity which in the end Table 2.1 Contrasting positivist and naturalist axioms (Lincoln & Guba, 1985, p. 37) Axioms about The nature of reality The relationship of knower to the known The possibility of generalization The possibility of causal linkages The role of values

Positivist paradigm Reality is single, tangible, and fragmentable. Knower and known are independent, a dualism.

Naturalist paradigm Realities are multiple, constructed, and holistic. Knower and known are interactive, inseparable.

Time- and context-free generalizations (nomothetic statements) are possible. There are real causes, temporally precedent to or simultaneous with their effects.

Only time- and context-bound working hypotheses (idiographic statements) are possible. All entities are in a state of mutual simultaneous shaping, so that it is impossible to distinguish causes from effects. Inquiry is value-bound.

Inquiry is value-free.

Constructing the Qualitative Approach

25

enables derivation of regularities or laws. The naturalist multiple realities do not allow for control and prediction (Lincoln & Guba, 1985, p. 37). The possibility for causal linkages refers to reality as fragmentable and linkable, i.e., the presentation of readily apparent categories of being and events and the possibility of identifying regularities between these. Nature can be carved at the bones which means for the positivist constructs present themselves as they can be found nature. The principle also refers to the ability of identifying causes. The naturalist contrast is that no linear and directional causality can be identified. The two aspects stipulated under epistemology are the relationship between the knower and known and the possibility of generalisation. Generalisation refers to the ability of establishing principles or laws applicable across space-time. Laws are thus a-historical and not locally bound. The contrast between nomothetic and ideographic applies here. For the naturalist only time and context-bound claims are possible. The individual case cannot be made universally applicable (Lincoln & Guba, 1985, p. 38). The knower-known dualism is crucial to the positivist. This means that the knower cannot and should not influence the object. Usually this is what is known as the positivist emphasis on objectivity, namely, an uninvolved scientist. The naturalist version would like to emphasise the reciprocal influence of the subject and object. The effect of the knower-known dualism leads to the last characteristic which is not an epistemological question but an implication of the particular positivist ontology and epistemology: the requirement of objectivity understood as the non-influence of the scientist implies that enquiry is value free. In contrast Lincoln and Guba (1985, p. 38) emphasise multiple ways in which enquiry are influenced by values, namely, by the choices the enquirer makes in terms of particular problem, the paradigm one works in, contextual values and the theory utilised in the enquiry. Value-ladeness expresses the opposite sentiment of objectivity in this context of establishing the basic contours of the quantitative-qualitative image of science. The characteristics Lincoln and Guba (1985) indicated in their “Naturalist inquiry” were repeated to some extent and modified in subsequent volumes of Denzin and Lincoln (1994, 2005c, 2018b). Initially Guba and Lincoln (1994) indicated the differences between positivism, postpositivism, critical theory and constructivism (Lincoln et al., 2018, p. 110). Thus, where Lincoln and Guba (1985, p. 37) started with a dual comparison, two paradigms became a quartet, namely, two for each methodological approach (Guba & Lincoln, 1994). These four paradigms were compared across the familiar rubrics (or axioms) of ontology, epistemology and methodology. In addition, Denzin and Lincoln (2018b, p. 110) compared the following issues: inquiry aim, nature of knowledge, knowledge accumulation, goodness or quality criteria, values ethics, voice, and training. We focus on the first three which they group under metaphysical assumptions, namely ontology, methodology and epistemology. From Table 2.2 it can be seen that positivism is associated with naive realism. The ontological assumption of postpositivism is called critical realism. The difference between the two is that in the former reality is believed to be fully known but in the latter only probabilistically. For positivism “An apprehendable reality is assumed to exist, driven by immutable natural laws and mechanisms. Knowledge of the “way

26

2 The Methodological Division: Quantitative and Qualitative Methods

Table 2.2 Basic beliefs or metaphysical assumptions for different paradigms (Guba & Lincoln, 1994, p. 109) Item Ontology

Positivism Naive realism— “real” reality but apprehendible

Postpositivism Critical realism— “real” reality but only imperfectly and probabilistically apprehendible

Epistemology

Dualist/objectivist; findings true

Methodology

Experimental/ manipulative; verification of hypotheses; chiefly quantitative methods

Modified dualist/ objectivist; critical tradition/community; findings probably true Modified experimental/manipulative; critical multiplism; falsification of hypotheses; may include qualitative methods

Critical theory Historical realism—virtual reality shaped by social, political, cultural, economic, ethnic, and gender values; crystallized over time Transactional/subjectivist; valuemediated findings

Constructivism Relativism— local and specific constructed and co-constructed realities

Dialogic/ dialectical

Hermeneutical/ dialectical

Transactional/ subjectivist; created findings

things are” is conventionally summarized in the form or time- and context-free generalizations, some of which take the form or cause-effect laws” (Guba & Lincoln, 1994, p. 109). The “way things are” as implied by the realism for both positivism and postpositivism can be understood with its contrast, namely, the historical realism and ontological relativism of critical theory and constructivism. Historical realism implies a view of “historically situated structures that are, in the absence of insight, as limiting and confining as if they were real” (Guba & Lincoln, 1994, p. 111). The reality in a sense is still out there, i.e., subject-independent but because people cannot see them for what they are, they appear, for instance, as oppressive. For the constructivist reality is no longer something out there or subject-independent. As Guba and Lincoln (1994, p. 111) explain: “constructivism’s relativism, which assumes multiple, apprehendable, and sometimes conflicting social realities that are the products of human intellects, but that may change as their constructors become more informed and sophisticated.” A similar movement from subject to reality is present as with historical realism. How reality appears depends on the human mind but clearly reality is a product of this human mind. Turning to epistemology Guba and Lincoln (1994, p. 109) distinguish between a dualist and transactional epistemology. Both positivisms are dualist thus knower and known are separate and do not influence each other except that the positivist obtains certain knowledge and the postpositivist fallible knowledge. For the remaining two postmodern paradigms knowledge is constructed transactionally, between knower and known or as they call them: investigator and respondents. Knowledge is also

Constructing the Qualitative Approach

27

value-mediated for both, thus, influenced and formed by values and interests and not objective or a-historical. For positivism methodology implies verification but for postpositivism, falsification. Critical theory methods are dialogic and dialectical while that of constructivism hermeneutic/dialectic. Both aim at the “reconstruction of previously held constructions” (Guba & Lincoln, 1994, p. 112). In the latest edition of Denzin and Lincoln (2018b), Table 2.2 above is greatly expanded by Lincoln et al. (2018) after Guba and Lincoln (2005) added a third paradigm to the post-modern ones, namely, the participatory paradigm. Heron and Reason (1997) proposed a participatory paradigm and argued that there are aspects which both the critical perspective and the constructivist paradigm cannot account for. Guba and Lincoln (2005) took the proposals to heart and added the participatory paradigm to their list of postmodern ones as a third paradigm that opposes the positivist/postpositivist paradigms (Table 2.3). According to Guba and Lincoln (2005, p. 195) a participative view of reality entails a “subjective-objective reality, cocreated by mind and given cosmos.” In addition, the practical rubrics are somewhat modified and amended. Some of the issues worth mentioning is that the aim of inquiry for the positivist/postpositivist paradigms is explanation, which is seen as prediction and control. Positivism regards those verified facts as true and immutable laws while postpositivism regards the same for non falsified hypotheses. Both paradigms regard knowledge as accumulable in terms of building blocks. In addition both regard quality criteria as the conventional internal and external validity as well as reliability and objectivity. Ironically, the quality criteria for constructivism are trustworthiness and authenticity which sounds like mere synonyms for validity and reliability (Lincoln et al., 2018, p. 110). The positivism/postpositivism image developed from Lincoln and Guba’s (1985) initial axioms to that of Guba and Lincoln (1994, p. 112) where it stayed relatively the same up to now (Lincoln et al., 2018). The characteristic of a single and tangible reality is reiterated in subsequent views. Tangible made way for objective and external and presumably the positivist image of reality is that of something separate from people. This concrete and separate domain seems not to be problematical when dealt with by natural science. Only when social reality is regarded in a similar manner, i.e., as external to people, unchangeable and objective, it becomes a problem for the social scientist. The springboard motivation is the objectivist view of reality, natural or otherwise, that cannot be dealt with positivistically in any of the postmodern paradigms. Critical theory see reality as essentially historically conditioned, thus, infested with human values, political and cultural interests. Constructivism goes even beyond the historical situatedness of humans. People co-create their realities, histories, stories and the like within a social community. The social co-creation means that there are communities that experience their realities uniquely. The local nature of social realities implies epistemological relativism in contrast to the search for certainty of the positivisms. Lincoln and Guba (1985) epistemological principle of the subject-object split expresses the typical positivist requirement of a disinterested observer (Lincoln et al., 2018, p. 112). The emphasis is on interests not influencing the objects of

Item Ontology

Positivism Belief in a single identifiable reality. There is a single truth that can be measured and studied. The purpose of research is to predict and control nature.

Postpositivism Recognize that nature can never fully be understood. There is a single reality, but we may not be able to fully understand what it is or how to get to it because of the hidden variables and a lack of absolutes in nature.

Critical theory Human nature operates in a world that is based on a struggle for power. This leads to interactions of privilege and oppression that can be based on race or ethnicity, socioeconomic class, gender, mental or physical abilities, or sexual preference

Constructivism Relativist: Realities exist in the form of multiple mental constructions, socially and experientially based, local and specific, dependent for their form and content on the persons who hold them. Relativism: local and specific constructed and co-constructed realities. “Our individual personal reality–the way we think life is and the part we are to play in it–is selfcreated. We put together our own personal reality.” Multiple realities exist and are dependent on the individual. “Metaphysics that embraces relativity”. “We practice inquiries that make sense to the public and to those we study.” Assumes that reality as we know it is constructed intersubjectively through

Participative Participative reality: subjective-objective reality, co-created by mind and the surrounding cosmos. Freedom from objectivity with a new understanding of relation between self and other. Socially constructed: similar to constructive, but do not assume that rationality is a means to better knowledge. Subjective-objective reality: Knowers can only be knowers when known by other knowers. Worldview based on participation and participative realities.

Table 2.3 Expanded characterisation: basic beliefs or metaphysical assumptions for different paradigms (Lincoln et al., 2018, p. 114)

28 2 The Methodological Division: Quantitative and Qualitative Methods

Epistemology

Belief in total objectivity. There is no reason to interact with who or what researchers study. Researchers should value only the scientific rigor and not its impact on society or research subjects.

Assume we can only approximate nature. Research and the statistics it produces provide a way to make a decision using incomplete data. Interaction with research subjects should be kept to a minimum. The validity of research comes from peers (the research community), not from the subjects being studied. Research is driven by the study of social structures, freedom and oppression, and power and control. Researchers believe that the knowledge that is produced can change existing oppressive structures and remove oppression through empowerment.

the meanings and understandings developed socially and experientially. To me this means that we construct knowledge through our lived experiences and through our interactions with other members of society. As such, as researchers, we must participate in the research process with our subjects to ensure we are producing knowledge that is reflective of their reality. Subjectivist: Inquirer and inquired into are fused into a single entity. Findings are literally the creation of the process of interaction between the two. Transactional/subjectivist: co-created findings. The philosophical belief that people construct their own understanding of reality; we construct meaning based on our interactions with our (continued)

Holistic: “Replaces traditional relation between ‘truth’ and ‘interpretation’ in which the idea of truth antedates the idea of interpretation.” Critical subjectivity in participatory transaction with cosmos; extended epistemology of experiential, propositional, and practical knowing; co-created findings. Critical subjectivity: Understanding how we

Constructing the Qualitative Approach 29

Item

Positivism

Table 2.3 (continued)

Postpositivism

Critical theory

Participative know what we know and the knowledge’s consumating relations. Four ways of knowing: (1) experiential, (2) presentational, (3) propositional, and (4) practical.

Constructivism surroundings. “Social reality is a construction based upon the actor’s frame of reference within the setting.” Findings are due to the interaction between the researcher and the subject “We cannot know the real without recognizing our own role as knowers.” “Simultaneously empirical, intersubjective, and process-oriented.” “We are studying ourselves studying ourselves and others.” Assumes that we cannot separate ourselves from what we know. The investigator and the object of investigation are linked such that who we are and how we understand the world is a central part of how we understand ourselves, others, and the world. This means we are shaped by our lived experiences,

30 2 The Methodological Division: Quantitative and Qualitative Methods

Methodology

Belief in the scientific method. Value a “gold standard” for making decisions. Grounded in the conventional hard sciences. Belief in the falsification principle (results and findings are true until disproved). Value data produced by studies that can be replicated.

Researchers should attempt to approximate reality. Use of statistics is important to visually interpret our findings. Belief in the scientific method. Research is the effort to create new knowledge, seek scientific discovery. There is an attempt to ask more questions than positivists because of the unknown variables involved in research. There is a unifying method. Dialogic/dialectical (Guba & Lincoln, 2005) Search for participatory research, which empowers the oppressed and supports social transformation and revolution.

and these will always come out in the knowledge we generate as researchers and in the data generated by our subjects. Hermeneutic, dialectic: Individual constructions are elicited and refined hermeneutically, and compared and contrasted dialectically, with aim of generating one or a few constructions on which there is substantial consensus. Hermeneutical; dialectical. Hermeneutical discussion. Hermeneutics (interpretation, i.e., recognition and explanation of metaphors) and comparing and contrasting dialectics (resolving disagreements through rational discussion). “Everyday consciousness of reality and its chameleon like quality pervade politics, the media, and literature.” (continued)

Political participation in collaborative action inquiry, primacy of the practical; use of language grounded in shared experiential context. Use deconstruction as a tool for questioning prevailing representations of learners and learning in the adult education literature; this discredits the false binaries that structure a communication and challenges the assertions of what is to be included or excluded as normal, right, or good. Experiential knowing is through face-to-face learning, learning new knowledge through the application of the knowledge. Democratization and

Constructing the Qualitative Approach 31

Item

Positivism

Table 2.3 (continued)

Postpositivism

Critical theory

Constructivism

co-creation of both content and method. Engage together in democratic dialogue as co-researchers and as co-subjects.

Participative

32 2 The Methodological Division: Quantitative and Qualitative Methods

Constructing the Qualitative Approach

33

study; certainly, the scientist is interested in what she is doing but those interests, motivations, background knowledge and values cannot be allowed to contaminate observed facts. The subject-object dualism is required to establish objectivity, i.e., to ensure a lack of influence of the knower’s bias and values on what is observed that can be established as fact and knowledge. This dualism is also required to ensure the truth of facts that can contribute to the certain foundation for scientific knowledge. Certain and true knowledge can apply a-spatiotemporarily. This sharp and fundamental contrast between subject and object is driven home in version 3’s positivist epistemological characteristic (see Table 2.3): “Belief in total objectivity. There is no reason to interact with who or what researchers study” (Lincoln et al., 2018, p. 115). This framework or image of positivist science cannot be applied to social science and Lincoln and Guba (1985) contrast with post-modern approaches drives this point home. No subject-object split is possible in the social domain. The object is always coloured by the subject’s interests. It is even more than just looking with tinted lenses to objects. Social objects or the known is constructed by the subject or knower. Some postmodernists might view reality as tinted, others might view reality as wholly constructed; it is a matter of degree and the point is the influence of the human mind on perceiving reality. If we acknowledge reality is tinted or even slightly constructed it is a small step to ontological relativism, namely, there are as many realities and truths as there are people and communities. This ontological relativism grounds epistemological relativism. If there are only local communities and local histories it is very difficult to establish a-spatiotemporal truths. This ontological and epistemological grounding should have been sufficient to distinguish positivist and postmodern approaches to science. However, Lincoln and Guba (1985, p. 37) went further and axiomatised the issue of causality for positivism as well. They speak of “real causes” emphasising the fact that positivists believe that there are things among the whole spectrum of observable reality that can be called a cause. As Hume’s empiricism showed, we cannot observe a cause, merely a sequence of events. If positivists are empiricists they certainly do not believe in cause as Lincoln and Guba (1985, p. 37) indicated in their table. However, the reason they emphasise real causes can be found in its postmodern contrast, namely, there are no causes but things influence each other simultaneously in the world of the social. Thus, it must be the result of the dissolution of the subject-object dualism in the postmodern paradigm: if realities or subject-object co-create each other there is no neat one-to-one cause from subject to object or object to subject but a complex process of co-creation. However, in the process Lincoln and Guba (1985) confused cause with single and one-directional cause. Denying directionality and emphasising complexity does not mean that no influence is taking place. Even “simultaneous shaping” is a form of things determining each other (Lincoln & Guba, 1985, p. 37). Therefore, in their picture of the positivist/postmodernist image of science, both acknowledge cause but more accurately, positivists think they can separate cause and effect whilst it is not possible to the postmodernist. The one characteristic we have identified distinguishing the paradigms proceeds from the relationship between the knower and the known or the knowing mind versus objective reality. Positivism needs ontological dualism to establish

34

2 The Methodological Division: Quantitative and Qualitative Methods

epistemological objectivity, truth, certainty, and a-spatiotemporality. This dualism is reiterated in Guba and Lincoln’s (1994, p. 109) table of paradigm characteristics and axioms (Table 2.1). The characteristic is the nature of knowledge which depends on being. For the positivist it is infallible knowledge and for the postmodernist it is relative knowledge. The idea of causality was introduced in Table 2.1 above, but it does not appear explicitly in Table 2.2 above. Instead, the issue of testing hypotheses by means of quantitative methods is introduced by Guba and Lincoln (1994, p. 109). This is sharply contrasted with the focus of understanding meaning in postmodern paradigms. The focus on understanding meaning is better understood when contrasted with the quantitative nature of positivist methods rather than its focus on linear causality as in the first table. One may argue that the choice of a contrastive axiom required either quantification or causality depending on how one views both aspects. For example, quantification implies reduction to numbers which rarely can express the meaningful nature of social and human phenomena. Or, the focus on causality can be typically viewed in a behaviourist manner, namely stimulus and response mirrors cause and effect and this particular reduction also strips human actions of meaning. Despite these examples it is more important to point out that a particular understanding of cause-effect and quantification infiltrated the contrast between the paradigms so that we have these two characteristics informing the image of science. From the expanded Table 2.3 of Lincoln et al. (2018) we can see that the concept of quantification is now essential to identifying positivism by its inclusion on the ontological level: “There is a single truth that can be measured . . .” (Lincoln et al., 2018, p. 114). A cursory glance at the definitions of the postmodern paradigms in Table 2.3 shows that quantification is not needed for its contrastive power. Lincoln et al. (2018, p. 114) are more interested in distinguishing the three paradigms and their different conceptualisations for reality. The ontological characteristics of the three postmodern paradigms focus on power, creation of multiple realities by the mind(s) and the co-creation of the real by subject-objects (more on the latter later). No specific need to contrast meaning construction with quantification can be found in the three postmodern paradigms. Quantification as an ontological characteristic for positivism is now a given, accepted as part of the reality of the positivists and integral to its scientific image. It is thus not even necessary to make quantification explicit under the methodological characteristics as in the previous version of the table (Table 2.2). One may assume that scientific method of the hard sciences naturally implies quantification. The purpose of quantification is to give accurate descriptions of reality and to enable the identification and formulation of laws that can enable prediction and control. Up to this point we spoke about positivism and postpositivism as marching under the same positivist flag, and in a sense this is how Guba and Lincoln (1994) and Lincoln et al. (2018) treat the two positivisms. The only discernible difference is the way each regard facts and knowledge. Whereas positivism ensures infallible, true and certain knowledge, postpositivism embraces probability. Facts might be true. The fact that we cannot, like the positivists, guarantee the truth of knowledge, does not mean that we have to abandon the scientific enterprise. Therefore,

Mixed Methods

35

everything applies as is in the positivist case except that knowledge is fallible. The search for causality, to enable prediction and control equally applies to postpositivism except that we now talk about knowledge as probably true and do not claim absolute truth. One may assume that the change of heart of postpositivists with regard to certainty of facts is the replacement of verification with falsification as methods (see Table 2.2). Of course, this is not really true because the proposal of falsification as method was Popper’s (1963/2002) criticism on the logical faulty verificationist method of the positivists. Many debates later falsification was also shown to be inadequate as an empiricist method (Chalmers, 2013). A final remark about Guba and Lincoln’s (1994) image of postpositivism. In Table 2.2 it can be seen that they call the postpositivist reality “critical realist.” I do not think it refers to Bhaskar’s critical realism as discussed later in this book, but “critical” according to Guba and Lincoln (1994) has a bearing on the fallible and probabilistic nature of knowledge despite regarding reality as real. It is thus critical in the sense of not being infallible. On its own, typifying the realism of postpostivism as critical realist is misleading and wrong.

Mixed Methods The force of the disjunction between qualitative and quantitative methods is such that in the psychosocial sciences one of the first distinctions made about researchers is whether they work qualitatively or quantitatively. However, this position of exclusivity is making place for researchers using both. Currently, some researchers talk about both mixed methods research (MMR) and multimethod mixed methods research (MMMR). Although there is some debate about how to distinguish mixed from multi method, the principle is reasonably simple. MMR refers to mixing qualitative and quantitative in various designs, emphasis, and combinations (Johnson, Onwuegbuzie, & Turner, 2016). Multimethod refers to using more than one method, not necessarily a mix between quantitative-qualitative, to investigate a problem (Anguera, Blanco-Villaseñor, Losada, Sánchez-Algarra, & Onwuegbuzie, 2018). Multimethod approaches is based on Campbell and colleagues’ so-called postpositivist triangulation of methods but fits easily in the structure and intention of mixed research methods (Mark, 2015, p. 24). How does one motivate exclusivity or a mixed approach? From the discussion of previous section it was seen that qualitative proponents explicitly motivate their stance against quantitative approaches usually identified as forms of positivism. The paradigms2 or metatheoretical frameworks are regarded as not compatible. This is known as the incompatibility thesis and the ontological and epistemological contrasts between the two approaches precludes any form of combining methods. Similar views based on the incompatibility of paradigms can be found from researchers of both sides. Incompatibility is usually based on the requirements of Kuhn’s paradigm discussed above, namely that incommensurability is one requirement for one paradigm to be overthrown by another. The argument for either

36

2 The Methodological Division: Quantitative and Qualitative Methods

approach is based on metatheoretical considerations thus incompatibility between, for instance, a relativist vs. an empiricist ontology and epistemology. Thus, incompatibility lies on the paradigm2 level and should not spill over to the methods level: again speaking of quantitative-qualitative paradigm wars or incompatibility places the emphasis on the wrong level of analysis. We will return to this issue. Johnson and Onwuegbuzie (2004) make a distinction between purists in the qualitative and quantitative camps. Quantitative purists view science as based on positivists principles, namely, there is no difference between natural and social phenomena and they should be treated the same; the subject (the observer) is separate from the object or the observed; objectivity is defined as time and context free and finally “real causes of social scientific outcomes can be determined reliably and validly” (Johnson & Onwuegbuzie, 2004, p. 14). Finally, the aim of positivist research is to establish social or psychological laws. The qualitative purists, under which constructivists and interpretivists are grouped, maintain the primacy of “constructivism, idealism, relativism, humanism, hermeneutics and, sometimes, postmodernism” (Johnson & Onwuegbuzie, 2004, p. 14). They reject timeless and context-free generalisations. Values play a fundamental role in research; cause and effect cannot be distinguished properly; conceptually the scientist works inductively, knower and known cannot be separated and reality is multiply constructed. Despite the incompatibility thesis, researchers in real life use mixed methods. As is the case with qualitative methods, Maxwell (2016) says that mixed methods was very much a familiar practice before the 1950s although not by that name (Bryman, 2008). He proceeds to give a few examples of pre-1950 projects that clearly demonstrated the integrative and triangulation practices of both natural and social scientists. Currently, researchers mostly acknowledge that they are using MMR but in the light of the incompatibility thesis how do they justify combining methods? Given the fact that method combination did happen and happens in practice, there are mainly two strategies of justification, namely, following a broad eclecticism or quoting pragmatic criteria and elevating the discussion to the metatheoretical level, thus invoking a pragmatist paradigm2 (Morgan, 2007; Tashakkori & Teddlie, 2010a). Denzin and Lincoln (2018a, p. 35) call this third force “a soft, apolitical paradigm” that allows pragmatist proponents to quote the principle of “what works.” As is the case with quantitative methods and qualitative methods no systematic historical overview of MMR has been done although sporadically (Hanson, Creswell, Clark, Petska, & Creswell, 2005; Johnson & Gray, 2010; Johnson et al., 2016). A historical overview and analysis would probably be able to provide an interesting story about the rise of the various methodological approaches which do not correspond perfectly with what we find in textbooks. Maxwell (2016) points out that most textbooks and qualitative texts say that mixed methods started in the 1950s up to the 1980s. What is well-known is that the rise of mixed methods as a distinct approach came to prominence since the 1980s, similar to the rise of qualitative methods (Denzin & Lincoln, 2018a). The focus on the rise of MMR is a “distinct and self-conscious strategy.” Maxwell (2016, p. 12) aimed at advocating the deliberate combining of quantitative and qualitative approaches:

Mixed Methods

37

In developing and promoting mixed methods research, it is more persuasive to present this as a new and exciting development, a “third paradigm” for social research, than to acknowledge that people have been doing mixed methods research for centuries, and far more broadly than most mixed methods work recognizes. It’s also more advantageous to position yourselves at the center of this movement, as the people who are developing the methods and standards for this approach, than to accept that many other researchers are systematically combining qualitative and quantitative approaches with little or no input from the self-defined mixed methods community (Maxwell, 2016, p. 13).

Thus, as in the case of the rise of the qualitative or metaphysical paradigm, the deliberate advertising of mixed methods as a viable alternative to either quantitative or qualitative is based on a construction of a research narrative against, for instance, the inadequacies of either quantitative and qualitative approaches on their own. Maxwell (2016, p. 13) quotes Platt (1996) that text books tend to propagate “origin myths” in order “to legitimate contemporary preferences.” Although certainly not true in all cases, these origin myths probably also feed the image of science we regard as the standard or received view. Contemporary MMR is advocated by a few key authors and a number of publications and journals saw the light as was the case with qualitative methods (Creswell, 2010). Many subsequent works rely on these authors for current views on MMR (Creswell & Clark, 2011; Johnson & Onwuegbuzie, 2004; Mertens, 1998; Onwuegbuzie & Collins, 2007; Tashakkori & Teddlie, 2010b; Teddlie & Tashakkori, 2009). Most MMR justify its use based on either ignoring the paradigm incompatibility or by taking a pragmatic view of methods and combining of methods (Bryman, 2008). The avenues of paradigm incompatibility and pragmatism are based on the positivist-constructionist polarity. Rarely, if ever, is realism quoted as an alternative to positivism and postpositivism and usually lumped with either of these. An exception in the advocates of MMR is Maxwell (2004) who from a realist vantage point argues for the integration of quantitative and qualitative methods. Morgan (2007) pointed out that amongst a list of so-called paradigms within which social science can be practiced, pragmatism was consistently ignored, while positivism, critical theory, interpretivism, and participatory were maintained in Guba and Lincoln’s (2005) list. Furthermore, Morgan (2007) says that both the metaphysical paradigm and pragmatism hold epistemological issues to be important, a matter that needs to be retained by pragmatism. However, the emphasis on ontology of the metaphysical paradigm is unnecessarily “too narrow” (Morgan, 2007, p. 68). What is the scientific image of MMR/MMMR? The image of science these approaches have correspond with that of qualitative methods. At least Mertens and Tarsilla (2015) regard postpositivism and not positivism as the appropriate representative for qualitative approaches. Its ontology regards reality as single and measurable; its epistemology incorporates objectivity defined as distance and neutrality; methodology consists of experimental and quasi experimental analysis methods, mostly quantitative. What is interesting is the inclusion of axiology, an issue qualitative researchers raised, and thus taken on board by MMR proponents. In this instance, Mertens and Tarsilla (2015, p. 433) point out that the postpositivist approach upholds ethical principles of respecting privacy and confidentiality,

38

2 The Methodological Division: Quantitative and Qualitative Methods

beneficence and non-maleficence, and it treats people fairly especially when utilising random selection when sampling. The last principle is based on the issue of individuals having an equal chance to be compared to others to be included in the sample and that population characteristics are distributed equally throughout the sample. Beyond including axiological considerations as part of its metatheoretical structure, this version of postpositivism is close to the qualitative image of positivism delineated above. Johnson and Gray (2010) laments the fact that qualitative proponents remain stuck in the long superseded positivist framework and hold that most quantitative research are postpositivist, which moved beyond positivism in the following matters: facts are (a) theory-, and (b) value-laden, (c) knowledge is fallible and provisional, (d) theory is underdetermined by facts, and (e) postpositivists accept that some parts of reality are socially constructed.

Conclusion The mythic image of science is formed by certain beliefs about what science is and these beliefs we find in discussion of methods in publications utilised to train researchers. Thus, students get used to the particular ideas we have of science and that methods can be divided into qualitative and quantitative approaches. One of the unfortunate strategies of advocating a particular method is by arguing with reference to counter points. Of course, this critical strategy is one that we have as scientists: someone states a so-called fact, and we may criticise it. However, as we know critical strategies can become rhetorical and be used in service of advocacy. Not that advocacy is bad, but due to the lack of counter arguments qualitative methods informed the image of science we work with, especially in the social science and psychology where the image of science was befuddled from the start. Psychosocial science in their development had been subjected to ridicule and judgements of quasi or proto science but never a mature science. The positivist framework established at the beginning of the previous century gave a number of psychosocial scientists a vehicle to climb on while trying to act as respectable scientists. The role of quantification, experimentalism and behaviourism is wellknown in psychology. The image of science that we in the psychosocial sciences have today is the result of a large contingent of scientists critical of the reductive nature of quantitative methods trying to salvage the integrity of the human subject. It is with some irony that this process coincided with post-modern deconstruction of the same subject. The qualitative methods advocates successfully canvassed their paradigm within the arena of the psychosocial sciences. From the brief discussion above we can discern a few strategies, namely, the way the term paradigm was recruited for advocating qualitative methods, the construction of the positivist image of science, equating quantification with positivism, and utilising these issues to firmly establish and entrench the quantitative-qualitative dichotomy.

Conclusion

39

Let us summarise the image of science we sit with. It is positivistic which means that quantification is required to measure and analyse an objective and single reality. Epistemologically, knowledge is established empirically, the scientist must be objective or distant from the subject matter as to not contaminate knowledge with subjective and personal issues such as values. The main method for knowledge generation which includes the ability to establish laws is deductive. Positivism and postpositivism only differ on the certainty of knowledge and facts. For positivism facts are certain but for postpositivism they are probable. The effect of this image is that as soon as any scientist or researcher uses measurement, it is assumed that she subscribes to this scientific image (Biesta, 2010). To acknowledge the relativist nature of constructions about persons or to utilise interpretation to find meaning in the people and phenomena, means that one does not believe in measuring things, which is of course not true! In the development of postpositivism, among others, due to Kuhn’s (1996) work, the discussion of paradigms provided sufficient conceptual material for some qualitative proponents to advocate for the establishment of qualitative methods. As we have seen, paradigm should be used for a particular level of conceptual frameworks, namely, those referring to subject matter, which in psychology is difficult due to a number of co-existing explanatory frameworks about human behaviour and functioning. Paradigm or rather paradigm2 or a metatheory incorporates epistemological and by implication ontological levels of discussion. As Guba framed it, it includes methodological and axiological considerations as well. My sense is that these issues should be grouped under a metatheoretical level of discussion or analysis. However, the recruitment of the term paradigm to refer to metatheoretical issues succeeded in demarcating different views of reality to such an extent that purists could invoke the incompatibility thesis for using one or the other methodological approach. The incompatibility thesis underlies a number of discussions in the quantitativequalitative debate and is responsible for maintaining the dichotomy. Thus, the talk of paradigm differences bolstered the uniqueness of the qualitative approach whilst maintaining its counterpoint to quantitative methods. It is important to realise that whereas the initial reaction was against positivism as evidenced in early work, the equality between positivism and measurement were systematically established so that talk of the one implied the other. The fact that there is scarcely any difference between positivism and postpositivism in Guba’s schema supports the view that in the qualitative proponent’s mind, they are essentially of the same sort, and both easily function as counterpoint for justifying the metaphysical paradigm. At this stage we need to set the record straight. While we used the term metaphysical paradigm to distinguish the metatheories of the qualitative set of theories/ paradigms2, and talked about the quantitative paradigm2 to refer to positivism and postpositivism, we need to slowly rid ourselves of these terms. Using, as Gorard (2010) called them, q-words merely re-affirms the quantitative-qualitative split. In addition, we need to get rid of using the term paradigm when referring to metatheories when discussing epistemological and ontological issues. The q-words cannot be equated to paradigms (Gorard, 2010; Greene & Hall, 2010). We should

40

2 The Methodological Division: Quantitative and Qualitative Methods

also understand that methods and methodology are two different issues, the latter referring to justifying methods based on metatheoretical considerations. The question is whether we need to remain on the metatheoretical level of analysis or discard it as some authors suggest, when speaking about methods. Methods describe the nature of data, and data cannot be determined. To what extent is it feasible to keep justifying our versions of reality ontologically and maintain dichotomies? Do we only have a choice between say empirical realism as in the case of positivism and constructionism with its relativist ontology? The suggestion of using a third paradigm, namely, pragmatism, by mixed methods proponents perpetuates the discussion on a metatheoretical level of analysis (Morgan, 2007).

References Anguera, M. T., Blanco-Villaseñor, A., Losada, J. L., Sánchez-Algarra, P., & Onwuegbuzie, A. J. (2018). Revisiting the difference between mixed methods and multimethods: Is it all in the name? Quality and Quantity, 52(6), 2757–2770. https://doi.org/10.1007/s11135-018-0700-2. Biesta, G. (2010). Pragmatism and the philosophical foundations of mixed methods research. In C. Teddlie & A. Tashakkori (Eds.), Sage handbook of mixed methods in social and behavioral research (2nd ed., pp. 95–118). Thousand Oaks, CA: Sage. Bryman, A. (2008). The end of the paradigm wars? In P. Alasuutari, L. Bickman, & J. Brannen (Eds.), The Sage handbook of social research methods (pp. 12–25). London: Sage. Chalmers, A. F. (2013). What is this thing called science? (4th ed.). St Lucia, QLD: University of Queensland Press. Chamberlain, K., & Murray, M. (2008). Introduction. In C. Willig & W. Stainton-Rogers (Eds.), The Sage handbook of qualitative research in psychology (pp. 390–406). London: Sage. Collins, K. M. T. (2015). Validity in multimethod and mixed research. In S. Hesse-Biber & R. B. Johnson (Eds.), The Oxford handbook of multimethod and mixed methods research inquiry (pp. 240–256). Oxford: Oxford University Press. Creswell, J. W. (2010). Mapping the developing landscape of mixed methods research. In A. Tashakkori & C. Teddlie (Eds.), Sage handbook of mixed methods in social and behavioral research (2nd ed., pp. 45–68). Thousand Oaks, CA: Sage. Creswell, J. W., & Clark, V. L. P. (2011). Designing and conducting mixed methods research. Thousand Oaks, CA: Sage. Denzin, N. K., & Lincoln, Y. S. (1994). Handbook of qualitative research (1st ed.). Thousand Oaks, CA: Sage. Denzin, N. K., & Lincoln, Y. S. (2005a). Introduction: The discipline and practice of qualitative research. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (3rd ed., pp. 1–32). Thousand Oaks, CA: Sage. Denzin, N. K., & Lincoln, Y. S. (2005b). Preface. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (3rd ed., pp. ix–xix). Thousand Oaks, CA: Sage. Denzin, N. K., & Lincoln, Y. S. (2005c). The Sage handbook of qualitative research (3rd ed.). Thousand Oaks, CA: Sage. Denzin, N. K., & Lincoln, Y. S. (2018a). Introduction: The discipline and practice of qualitative research. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (5th ed., pp. 29–71). Thousand Oaks, CA: Sage. Denzin, N. K., & Lincoln, Y. S. (Eds.). (2018b). The Sage handbook of qualitative research (5th ed.). Thousand Oaks, CA: Sage.

References

41

Duncan, N., Stevens, G., & Canham, H. (2014). Living through the legacy: The Apartheid Archive Project and the possibilities for psychosocial transformation. South African Journal of Psychology, 44(3), 282–291. Ferreira, R., Maree, D. J. F., & Stanz, K. (2016). Contextualising psychological service provision in South Africa. In R. Ferreira (Ed.), Psychological assessment: Thinking innovatively in contexts of diversity (pp. 3–19). Pretoria: Juta. Feyerabend, P. K. (1993). Against method (3rd ed.). London: Verso. Gorard, S. (2010). Research design, as independent of methods. In C. Teddlie & A. Tashakkori (Eds.), Sage handbook of mixed methods in social and behavioral research (2nd ed., pp. 237–252). Thousand Oaks, CA: Sage. Greene, J. C., & Hall, J. N. (2010). Dialectics and pragmatism: Being of consequence. In C. Teddlie & A. Tashakkori (Eds.), Sage handbook of mixed methods in social and behavioral research (2nd ed., pp. 119–144). Thousand Oaks, CA: Sage. Gross, R. D. (2017). Psychology in historical context: Theories and debates. New York: Routledge. Guba, E. G., & Lincoln, Y. S. (1994). Competing paradigms in qualitative research. In N. K. Denzin & Y. S. Lincoln (Eds.), Handbook of qualitative research (1st ed., pp. 105–117). Thousand Oaks, CA: Sage. Guba, E. G., & Lincoln, Y. S. (2005). Paradigmatic controversies, contradictions, and emerging confluences. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (3rd ed., pp. 191–215). Thousand Oaks, CA: Sage. Hanson, W. E., Creswell, J. W., Clark, V. L. P., Petska, K. S., & Creswell, J. D. (2005). Mixed methods research designs in counseling psychology. Journal of Counseling Psychology, 52(2), 224–235. https://doi.org/10.1037/0022-0167.52.2.224. Hergenhahn, B. R., & Henley, T. B. (2014). An introduction to the history of psychology (7th ed.). Belmont, CA: Wadsworth Cengage Learning. Heron, J., & Reason, P. (1997). A participatory inquiry paradigm. Qualitative Inquiry, 3(3), 274–294. https://doi.org/10.1177/107780049700300302. Johnson, R. B., & Gray, R. (2010). A history of philosophical and theoretical issues for mixed methods research. In A. Tashakkori & C. Teddlie (Eds.), Sage handbook of mixed methods in social and behavioral research (2nd ed., pp. 69–94). Thousand Oaks, CA: Sage. Johnson, R. B., & Onwuegbuzie, A. (2004). Mixed methods research: A research paradigm whose time has come. Educational Researcher, 33(7), 14–26. https://doi.org/10.3102/ 0013189X033007014. Johnson, R. B., Onwuegbuzie, A. J., & Turner, L. A. (2016). Toward a definition of mixed methods research. Journal of Mixed Methods Research, 1(2), 112–133. https://doi.org/10.1177/ 1558689806298224. Kenrick, D. T., Griskevicius, V., Neuberg, S. L., & Schaller, M. (2010). Renovating the pyramid of needs: Contemporary extensions built upon ancient foundations. Perspectives on Psychological Science, 5(3), 292–314. https://doi.org/10.1177/1745691610369469. Kuhn, T. S. (1962). The structure of scientific revolutions. Chicago: University of Chicago Press. Kuhn, T. S. (1996). The structure of scientific revolutions (3rd ed.). Chicago: University of Chicago Press. Lincoln, Y. S., & Guba, E. G. (1985). Naturalistic inquiry. Beverly Hills, CA: Sage. Lincoln, Y. S., Lynham, S. A., & Guba, E. G. (2018). Paradigmatic controversies, contradictions, and emerging confluences, revisited. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (5th ed., pp. 108–150). Thousand Oaks, CA: Sage. Mackay, N., & Petocz, A. (2011). Realism and the state of theory in psychology. In N. Mackay & A. Petocz (Eds.), Realism and psychology: Collected essays (pp. 17–51). Boston: Brill. Mark, M. M. (2015). Mixed and multimethods in predominantly quantitative studies, especially experiments and quasi-experiments. In S. Hesse-Biber & R. B. Johnson (Eds.), The Oxford handbook of multimethod and mixed methods research inquiry (pp. 21–41). Oxford: Oxford University Press. Maslow, A. H. (1943). A theory of human motivation. Psychological Review, 50(4), 370–396.

42

2 The Methodological Division: Quantitative and Qualitative Methods

Maxwell, J. A. (2004). Using qualitative methods for causal explanation. Field Methods, 16(3), 243–264. https://doi.org/10.1177/1525822X04266831. Maxwell, J. A. (2016). Expanding the history and range of mixed methods research. Journal of Mixed Methods Research, 10(1), 12–27. https://doi.org/10.1177/1558689815571132. Mertens, D. M. (1998). Research methods in education and psychology: Integrating diversity with quantitative and qualitative approaches. Thousand Oaks, CA: Sage. Mertens, D. M., & Tarsilla, M. (2015). Mixed methods evaluation. In S. Hesse-Biber & R. B. Johnson (Eds.), The Oxford handbook of multimethod and mixed methods research inquiry (pp. 426–446). Oxford: Oxford University Press. Michell, J. (1997). Quantitative science and the definition of measurement in psychology. British Journal of Psychology, 88, 355–385. Morgan, D. L. (2007). Paradigms lost and pragmatism regained. Journal of Mixed Methods Research, 1(1), 48–76. https://doi.org/10.1177/2345678906292462. Onwuegbuzie, A. J., & Collins, K. M. T. (2007). A typology of mixed methods sampling designs in social science research. The Qualitative Report, 12(2), 281–316. Platt, J. (1996). A history of sociological research methods in America: 1920–1960. Cambridge: Cambridge University Press. Popper, K. R. (1963/2002). Conjectures and refutations: The growth of scientific knowledge. London: Routledge. Rorty, R. (1979). Philosophy and the mirror of nature. Princeton, NJ: Princeton University Press. Schwandt, T. A. (2000). Three epistemological stances for qualitative inquiry: Interpretivism, hermeneutics, and social constructionism. In N. K. Denzin & Y. S. Lincoln (Eds.), Handbook of qualitative research (2nd ed., pp. 189–213). Thousand Oaks, CA: Sage. Tabachnick, B. G., Fidell, L. S., & Ullman, J. B. (2019). Using multivariate statistics (7th ed.). Boston: Pearson. Tashakkori, A., & Teddlie, C. (2010a). Overview of contemporary issues in mixed methods research. In C. Teddlie & A. Tashakkori (Eds.), Sage handbook of mixed methods in social and behavioral research (2nd ed., pp. 1–42). Thousand Oaks, CA: Sage. Tashakkori, A., & Teddlie, C. (Eds.). (2010b). Sage handbook of mixed methods in social and behavioral research (2nd ed.). Thousand Oaks, CA: Sage. Teddlie, C., & Tashakkori, A. (2009). Foundations of mixed methods research: Integrating quantitative and qualitative approaches in the social and behavioral sciences. Los Angeles: Sage. Weimer, W. B., & Palermo, D. S. (1973). Paradigms and normal science in psychology. Science Studies, 3(3), 211–244. https://doi.org/10.1177/030631277300300301. Willig, C., & Stainton-Rogers, W. (2008a). Introduction. In C. Willig & W. Stainton-Rogers (Eds.), The Sage handbook of qualitative research in psychology (pp. 1–12). London: Sage. Willig, C., & Stainton-Rogers, W. (Eds.). (2008b). The Sage handbook of qualitative research in psychology. London: Sage.

Chapter 3

The Applicative Split: The Science-Practitioner Model of Training and Practice

Introduction The second polarity following closely on the heels of the methodological one is the scientist-practitioner (SP) split. The SP model was developed to encourage practitioners to embrace the scientific basis of their profession. The model’s development, which serves as training model for numerous institutions locally and abroad, encouraged practitioners to base therapeutic interventions on evidence but also encourage psychological researchers to consider the applicability or translational value of their research. Unfortunately, the split seems large as ever and various investigations point to practitioners and other professionals having an incorrect image of what science is. Even here the mythic image looms large. It paints science as inextricably bound to positivist practices involving objectivity, measurement and causality.

The Boulder Model and Beyond The SP model is universally recognised as one of the important training models for psychologists (Malott, 2018b). Many universities in the USA, Canada, Britain and even South Africa subscribe to the SP model or at least variants of it, such as the practitioner-scholar, reflective-practitioner, science-based practitioner and clinicalscientist (Corrie & Callahan, 2000; Hunsley, 2007; Malott, 2018a; Peluso, Carleton, & Asmundson, 2010; Pillay, Ahmed, & Bawa, 2013; Pillay & Johnston, 2011; Skourteli & Apostolopoulou, 2015; Stoltenberg et al., 2000). All of these emphasise the need for psychologists to be competent in the use and interpretation of scientific methods. The history of the training of psychologists is almost as long or short as the history of psychology itself. The turn of the century and especially the conclusion of the two world wars had a major stimulating effect on psychology and the need for therapeutic training (Baker & Benjamin Jr., 2000). The development of psychology © Springer Nature Switzerland AG 2020 D. J. F. Maree, Realism and Psychological Science, https://doi.org/10.1007/978-3-030-45143-1_3

43

44

3 The Applicative Split: The Science-Practitioner Model of Training and Practice

as science commenced with the well-known history of Wundt and others and influenced what developed as the science of psychology. The major attention given to physics and subsequent positivist philosophy of science also influenced the image of psychological science. The development of professional psychology in Britain went a different route than, for instance, the USA. Lead by Eysenck in Britain, emphasis for the development of clinical psychology was on empiricist methods and based on the medical and thus positivist model. Apparently Eysenck was not very fond of psychoanalysis and regarded therapy and practice as interfering with science and its worthwhile quoting him as in Corrie and Callahan (2000, p. 415): We must be careful not to let social need interfere with scientific requirements ... Science must follow its course according to more germane arguments than the possibly erroneous conceptions of social need (Eysenck, 1949, p. 173).

The impetus to train practitioners within the scientist-practitioner (SP) model came from decisions made at Boulder, Colorado in 1949, consequently, known as the Boulder model (Baker & Benjamin Jr., 2000; Jones & Mehr, 2007; Petersen, 2007; Raimy, 1950). The SP model was based on the assumptions that research knowledge and skills play an important role in practice, finding evidence for proven practices would enhance practitioner impact and research-based practice can address social issues (Jones & Mehr, 2007). In addition, the relationship between science and practice was seen as reciprocal (Jones & Mehr, 2007, p. 766). In some way research and science-mindedness should enhance practice, but practice could also be utilised to promote research. These activities was initially conceptualised as parallel and separate, which allowed for some interpretative possibilities (Jones & Mehr, 2007, p. 768). Seeing science and practice as separate also encouraged numerous views of how they cannot be integrated. Speaking from experience with training practitioners, Malott (2018a), for instance, argues that the scientist and practitioner are two different species—one can at most train an excellent science-based practitioner because those opting for practice do so because of personal interest and an inclination to help (Horn et al., 2007, p. 810; Huber, 2007): Alongside, it seems that the practitioner aspect of counselling psychologists’ identity is underpinned by an inherent motivation to invest in helping others rather than an interest to engage in what they perceive as sterile, clinically-irrelevant research (Skourteli & Apostolopoulou, 2015, p. 17).

However, the initial purpose of the Boulder model was to define the role of the practitioner in close relation to that of the scientist (Baker & Benjamin Jr., 2000). Jones and Mehr (2007) indicate three roles for the practitioner, namely, (a) that of researcher which means that a practitioner should foremost generate knowledge and contribute to the field of psychology; (b) of consumer, with the ability to find, read, assess and apply research especially in her practice, and (c) lastly, the role of empirical evaluator indicating her responsibility to approach therapy and programmes with a mind to evaluate their efficacy and in this way contribute to the profession of psychology. Almost 70 years after Boulder the ideal of the SP model still informs the aims and assumptions of training as psychologist. Overholser (2015, p. 221) acknowledges its

Struggle for Integration

45

importance as an ideal for training but also the difficulty in actualising this ideal in practice: “Unfortunately, it is difficult to truly integrate the science and practice of psychology.” The difficulty of integrating the roles in training clinical, counselling, educational and IO psychologists is frequently mentioned (Chwalisz, 2003; Hagstrom, Fry, Cramblet, & Tanner, 2007; Horn et al., 2007; Huber, 2007; Lane & Corrie, 2007, p. 2; Myers, 2007; Ridley & Laird, 2015; Rourke, 1995; Rowe, 2018; Stoner & Green, 1992; Stricker, 2002; Wakefield & Kirk, 1996), but on occasion there are examples where integration seems to take place or of voices encouraging integration (Kison, Moorer, & Villarosa, 2015; Spengler & Lee, 2017; Stoltenberg & Pace, 2007; van der Watt, 2016). These examples lift out specific training experiences such as (a) having research orientated role models as supervisors on the postgraduate level (Stoltenberg & Pace, 2007; Williams, Sayegh, & Sherer, 2018), (b) broadening the definition of research and science to include logistics and processes such as reading and applying research, applying findings in one’s own practice (Corrie & Callahan, 2000, p. 420), (c) utilise empirical problemsolving approaches (Huber, 2007); cultivate a scientific attitude, i.e., critical thinking skills and scientific skepticism (Kison et al., 2015; Pilecki & McKay, 2013, p. 543), and (d) understand theory of psychotherapy and mental functioning (Pilecki & McKay, 2013). Jones and Mehr (2007) summarised the role of the practitioner as follows: A scientist-practitioner is someone who applies critical thought to practice, uses proven treatments, evaluates treatment programs and procedures, and applies techniques and practices based on supportive literature. As such, one who embodies the role of a scientistpractitioner neatly integrates science and practice to best serve clients in a psychological realm (Jones & Mehr, 2007, p. 770).

Struggle for Integration Although training at many institutions implicitly or explicitly accepts the tenets of the SP model, the ideals of Boulder did not exactly realise as envisaged (Chang, Lee, & Hargreaves, 2008). A number of papers argue for a stronger emphasis on the SP model and in a number of special issues the matter was seriously investigated for clinical and counselling training (Asay, Lambert, Gregersen, & Goates, 2002; Barnette, 2006; Lampropoulos, Goldfried, et al., 2002; Lampropoulos & Spengler, 2002; Lampropoulos, Spengler, Dixon, & Nicholas, 2002; Lueger, 2002; Mellott, 2007; Stone, 2006; Stricker, 2002; Vespia, 2006; Vespia & Sauer, 2006; Vespia, Sauer, & Lyddon, 2006). The consensus is that the scientist part of the model is neglected and recent surveys revealed interesting attitudes of practitioners towards the model (Lilienfeld, Ritschel, Lynn, Cautin, & Latzman, 2013). As with the SP model, many practitioners are reluctant to embrace even the evidence-based approach, probably for similar reasons, while others regard the evidence based practice (EBP) approach as a way to honour the SP model given certain provisions (Hunsley, 2007).

46

3 The Applicative Split: The Science-Practitioner Model of Training and Practice

Some of the issues raised against adopting the EBP approach are, firstly, a lack of adequate emphasis on scientifically assessed therapies at graduate level, and practitioners’ entrenched beliefs in the efficacy of their own approaches. The trust in one’s own intuition and only accepting the efficacy of an approach by direct involvement can be called naive realism. Coupled with erroneous beliefs about scientific evidence, one can understand practitioners’ reluctance to drop “tried-and-trusted” therapies in favour of unfamiliar but evidence based ones (Lilienfeld, Ritschel, Lynn, Cautin, & Latzman, 2013). Secondly, one’s interest and skills influence to what extent one endorses the SP model. The natural polarisation of scientist and practitioner is readily apparent in those that teach, students and even within the body of practitioners as well. Some people are just not that interested in research and are naturally inclined towards practice (Chang et al., 2008; Maddux & Riso, 2007). Thirdly, the role of the supervisor in facilitating this process of exposure to science and integration of science and practice is crucial (Jones & Mehr, 2007). However, long-term career goals play an important role: academics, and thus trainers, focusing on research, loose the crucial exposure to a practice over time. The inclinations of trainers tend to rub off on students, and the onus is on trainers to set the example for proper science-practice integration (Overholser, 2007). The converse is also true—trainers focusing on skills training often do not have time or interest to publish. Lastly, the most important finding was that practitioners view the science model they need to subscribe to as largely positivist, thus, limited in utility and applicability. Despite the Boulder ideal, integrating science and practice will never take place if practitioners and scientists alike regard the scientific paradigm underlying the SP model as outmoded and inapplicable. For counselling psychology, the tension between science and practice . . . essentially reflects the philosophical and epistemological incongruence between the medical model’s adherence to positivist values, commitment to objectivity and evidencebased practice and counselling psychology’s emphasis on meaning, subjectivity, intuitive clinical practice in real-life situations, mutually constructed realities and multiplicity of perceptions (Skourteli & Apostolopoulou, 2015, p. 17).

Thus, psychologists seem to face a choice between an epistemological approach reflecting their succourative orientation and one that is objective, disinterested and disengaged (Corrie & Callahan, 2000, p. 416): “Psychology’s current commitment to positivist explanation, scientific knowledge characterized by law-governed causal processes, is at the core of the scientist-practitioner split” (Chwalisz, 2003, p. 497) and . . . Efforts to establish psychology as a scientific discipline underlie the alignment with a logical positivist philosophy of science . . . and its assumption that an objective reality exists that can be observed by researchers. However, psychological phenomena are far too complex . . . to be adequately captured using positivist scientific methods, and psychology has inadvertently limited its knowledge by aligning with positivist science (Chwalisz, 2003, p. 499).

Theory and Practice

47

The notion of evidence should thus be broadened to include non-positivist knowledge such as those gathered by qualitative analysis. Single case study evidence must also be allowed because this is where the practitioner can contribute to psychology. Meta-analytic studies may also contribute as an alternative source for evidence because as Chwalisz (2003, p. 504) says, over-reliance on null-hypotheses testing “hindered the progress” of psychological science. This claim is not clearly explained but one might assume that its hindrance is associated with positivistic approaches. It is telling that Chwalisz (2003) points to single case studies and metaanalysis as additional evidence beyond positivistic evidence although one would have thought these experimental and quantitative approaches would fall straight forwardly in the (post)positivist context. The slight confusion about what counts as positivist continues with Chwalisz’s (2003) discussion of what counts as additional support for psychological phenomena: “Theory provides the foundation on which positivist scientific knowledge is built, yet the role of theory in professional psychology decision making has not been adequately recognized within current scientific structures” (Chwalisz, 2003, p. 505). The second part of the claim is certainly true: theory as such has been grossly underestimated within psychological science although the paradigms1 are not lacking theory. Humanism and specifically positive psychology (note, not positivist psychology!), or CBT approaches do have theories about human functioning and treatment (Pilecki & McKay, 2013). Thus, the claim that theory is the foundation of positivist psychological knowledge is certainly false—the claim speaks of a misunderstanding of what theory is and how it functions as our discussion on paradigms and theories has shown. Theory and evidence cannot be separated however one defines the methodological or metatheoretical approach. The positivists denied theoretical entities, i.e., unobservable entities that had no chance of ever being observed, but we are referring to the explanatory framework helping us understand psychological phenomena.

Theory and Practice Continuing the discussion about theory’s role in psychology Chwalisz (2003, p. 505) says “Systematic approaches to evaluating theoretical contributions, beyond positivist hypothesis testing, should be developed”. This confuses the matter slightly especially with regard to understanding theory and the role of theory. If Chwalisz (2003) meant “empirical contributions” the claim would make perfect sense. A broader definition of science, i.e., one not restricted to a narrow positivist approach, would guide us in what counts as evidence for supporting a theory, something more than statements dependent on a 0.05 exceedance probability level. In order to reject a null-hypothesis the knowledge claim needs to be formulated in a certain manner and the study designed in a certain manner to support this eventual statistical decision. Not all problems can be investigated in this manner but some certainly can. To briefly return to the matter of theory. Pilecki and McKay (2013) call for a careful consideration of the relationship between practice and theory. This

48

3 The Applicative Split: The Science-Practitioner Model of Training and Practice

discussion certainly relates to the SP debate because theory forms an integral part of the process of science, an issue seemingly not considered seriously in discussions of EBP. The focus in the latter is mostly on empirical evidence that a particular therapy works in certain conditions, sometimes neglecting how its practical success informs the theory about mechanisms and behaviour. Theory, as the major form an explanatory framework takes on at some stage of its life, has to be tested or evidence accrued for its ability to explain behaviour and behaviour change in the case of therapy. Cognittive behavioural therapy (CBT), for instance, is based on particular theoretical principles which can be assessed. The same applies to its principles of treatment: these can be assessed for efficacy. It may well happen that therapists stumble upon on certain techniques that work without knowing why and how they fit into their theoretical framework. Pilecki and McKay (2013, p. 543) make the following worrying statement: “In the present state of the field, it is unclear to what degree such theoretical foundations are incorporated into training. . . .” Along with the lack of integrative discussions of theory in EBP discussions, a lack of familiarity with theory in practitioner training indicates a serious misunderstanding of the role of theory in the scientific process. Among the advantages of theoretical training Pilecki and McKay (2013) indicate the following: (a) theoretical knowledge assists in delivering better therapy. Not understanding comorbid conditions and, for instance, CBT’s applicative range, can seriously hamper the efficacy of therapy (Pilecki & McKay, 2013, p. 544). Theoretical knowledge further assists in (b) therapeutic identification, (c) adaptation, and (d) identifying progress markers for patients. These advantages are practical matters related to the implications of a theory and can be derived from the basic raison d’être of theory, namely, serving as a conceptual explanatory framework for a range of phenomena and processes. Thus, what Pilecki and McKay (2013) identifies as advantages still do not cut to the central purpose and role of theory. A similar disjunction between theory and practice is apparent for the disadvantages of theory, which according to Pilecki and McKay (2013) are issues such as time and financial constraints to increasing theoretical aspects in practitioner training, i.e., it might become too expensive to train students in theory. In addition, some therapies might not require or do not have solid theoretical underpinnings and by increasing theoretical components students might be dissuaded to qualify as practitioners because it is too difficult. These disadvantages reveal a narrow understanding of theory. As implied above, some therapies might work similar to adding eggs to flour to make dough—it just works and one need not know why. The end result might well be a satisfactory cake. However, this instrumentalist view of theory and its relationship to empirical reality is unscientific, at least in the concept of science we are aiming for in this book.

Empirical Evidence and Positivism

49

Empirical Evidence and Positivism Within the framework of defining evidence in a broader manner, Chwalisz (2003) argues for training psychologists not as scientist-practitioners but as evidence based practitioners. However, this can only be done if evidence is defined broader than quantitatively established facts. Although some tolerance for numbers can be seen in Chwalisz’s (2003) proposal, Denzin and Lincoln (2018) lament the rise of evidence based strategies. They call it strategic positivism which utilises quasi-statistics: researchers employ frequencies and percentages to make work “palatable to positivist colleagues,” presumably whom include funders (Denzin & Lincoln, 2018, p. 30). Thus, evidence based practice is merely another incarnation of positivism and science’s obsession with quantity. The rise of the notion of evidence based practice (EBP) since the early 1990s, originating in the medical environment but adopted by the helping social sciences, is another articulation of the SP model (Lilienfeld, Ritschel, Lynn, Cautin, & Latzman, 2013). The EBP model is based on three assumptions, namely, what evidence is best and available for treatments that work, utilising clinical expertise to make decisions about treatment and the values and interest of clients influencing treatment choice (Lilienfeld, Ritschel, Lynn, Cautin, & Latzman, 2013, pp. 885–886; Wilson, Armoutliev, Yakunina, & Werth, 2009). EBP should be distinguished from empirically supported therapies (EST) which refers to only one of the assumptions of both the SP model and the EBP approach (Lilienfeld, Ritschel, Lynn, Brown, et al., 2013, p. 388; Wilson et al., 2009, pp. 403–404). EBP commences with the patient and the psychologist has to ask what research evidence there is to best serve the needs of the client while EST is an already confirmed effective treatment by means of randomised clinical trials (Levant & Hasan, 2008, p. 659). EBP is more than the specific treatment and involves additional processes among others assessment, clinical expertise and case formulations including client preferences (Hunsley, 2007). Levant and Hasan (2008) defended the guidelines determined by the APA in 1995 about EBP for psychology. The reasons are clear: EBP originated in the medical environment and if psychologists do not define EBP for themselves, someone else will. The crucial aspect is to determine what counts as evidence for the efficacy of psychological treatments. The APA report on EBP were among others praised but also criticised for its assumption that empirical and apparent quantitative evidence are preferred above qualitative evidence. In this way EBP’s association with empiricist epistemology is strengthened and the qualitative approach is regarded as a “second class methodology” (Levant & Hasan, 2008, p. 660). Related to narrow epistemic considerations are concerns that professional psychology can become dominated by special interest groups, dehumanising of psychological services and inadequate research underlying EBP (Hunsley, 2007, p. 33). Thus, the evidence for empirically supported therapies and evidenced based practices is regarded as positivist especially where statistics is involved to decide the efficacy of outcomes (Chwalisz, 2003). Some studies call for an expansion of

50

3 The Applicative Split: The Science-Practitioner Model of Training and Practice

what is viewed as evidence, namely, including evidence from single case and qualitative studies and from less controlled situations of observation and interpretation. There is no need to redefine what counts as evidence. Such a plea merely exposes one’s ignorance of what counts as science; I argue for a revision of what we understand as science and in the process the myth of positivistic science needs to be exposed for what it is. The quotes from Chwalisz (2003) above, should not, in the light of my remarks, be taken at face value. I agree with the point Chwalisz (2003) wants to make, namely, the image of science both practitioners and scientists have is that of the mythic beast. However, the science they criticise is not positivist, but realist; the approach they advocate, namely qualitative, is also realist and also scientific. Some opinions were voiced about another possibility of why the SP model failed to integrate training and be adopted despite it acting as regulative ideal (Huber, 2007; Lane & Corrie, 2007). It is closely related to our argument above, namely, that the constant association of the science part of the SP model with positivism and quantity precludes its wholesale adoption and integration. It is worthwhile to quote Corrie and Callahan (2000, p. 413) who state that in the UK, EBP was becoming a requirement given the legal changes in the National Health Service (NHS) which necessitated the development of an adequate framework: The scientist–practitioner model is one such framework and despite its contentious history, it is argued that the difficulties traditionally associated with its implementation are largely owing to an outdated view of scientific activity which relies on a positivist philosophy of science (Corrie & Callahan, 2000, p. 413).

The difference between their observation and those we have discussed above blaming positivist science for the woes of the SP model, is the crucial phrase “outdated view of scientific activity.” The outdated view is precisely the mythic image we are extracting from the different polarities, discussed in this book. Corrie and Callahan (2000, p. 421) accurately point out that the assumptions of so-called neutral observational methods of positivism and “...that reality could be ‘discovered’ independently of the observer’s cultural and social context has long been recognized as problematic.” In fact, the postpositivists, namely, Kuhn, Feyerabend and even Popper, repeatedly offer alternative conceptions of science which, as Corrie and Callahan (2000, p. 421) say, had little impact on the SP-model debate. However, it is not only the SP-model debate that is operating in a bubble outside mainstream philosophy of science developments: the important and influential quantitativequalitative and related paradigm debates also seem to share the bubble with other dichotomies in psychology.

References Asay, T. P., Lambert, M. J., Gregersen, A. T., & Goates, M. K. (2002). Using patient-focused research in evaluating treatment outcome in private practice. Journal of Clinical Psychology, 58 (10), 1213–1225.

References

51

Baker, D. B., & Benjamin, L. T., Jr. (2000). The affirmation of the scientist-practitioner: A look back at Boulder. American Psychologist, 55(2), 241–247. https://doi.org/10.1037//0003-066X. 55.2.241. Barnette, V. (2006). A scholarly work commitment in practice. Counselling Psychology Quarterly, 19(3), 253–263. https://doi.org/10.1080/09515070600959334. Chang, K., Lee, I.-L., & Hargreaves, T. A. (2008). Scientist versus practitioner–An abridged metaanalysis of the changing role of psychologists. Counselling Psychology Quarterly, 21(3), 267–291. https://doi.org/10.1080/09515070802479859. Chwalisz, K. (2003). Evidence-based practice: A framework for twenty-first-century scientistpractitioner training. The Counseling Psychologist, 31(5), 497–528. https://doi.org/10.1177/ 0011000003256347. Corrie, S., & Callahan, M. M. (2000). A review of the scientist-practitioner model: Reflections on its potential contribution to counselling psychology within the context of current health care trends. British Journal of Medical Psychology, 73, 413–427. Denzin, N. K., & Lincoln, Y. S. (2018). Introduction: The discipline and practice of qualitative research. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (5th ed., pp. 29–71). Thousand Oaks, CA: Sage. Eysenck, H. J. (1949). Training in clinical psychology: An English point of view. American Psychologist, 4(6), 173–176. Hagstrom, R. P., Fry, M. K., Cramblet, L. D., & Tanner, K. (2007). Educational psychologists as scientist-practitioners an expansion of the meaning of a scientist-practitioner. American Behavioral Scientist, 50(6), 797–807. Horn, R. A., Troyer, J. A., Hall, E. J., Mellott, R. N., Coté, L. S., & Marquis, J. D. (2007). The scientist-practitioner model a rose by any other name is still a rose. American Behavioral Scientist, 50(6), 808–819. Huber, D. R. (2007). Is the scientist-practitioner model viable for school psychology practice? The American Behavioural Scientist, 50(6), 778–788. https://doi.org/10.1177/0002764206296456. Hunsley, J. (2007). Training psychologists for evidence-based practice. Canadian Psychology, 48 (1), 32–42. Retrieved from https://psycnet.apa.org/record/2007-04963-006. Jones, J. L., & Mehr, S. L. (2007). Foundations and assumptions of the scientist-practitioner model. American Behavioral Scientist, 50(6), 766–771. https://doi.org/10.1177/0002764206296454. Kison, S. D., Moorer, K. D., & Villarosa, M. C. (2015). The integration of science and practice: Unique perspectives from counseling psychology students. Counselling Psychology Quarterly, 28(3), 345–359. https://doi.org/10.1080/09515070.2015.1060193. Lampropoulos, G. K., Goldfried, M. R., Castonguay, L. G., Lambert, M. J., Stiles, W. B., & Nestoros, J. N. (2002). What kind of research can we realistically expect from the practitioner? Journal of Clinical Psychology, 58(10), 1241–1264. https://doi.org/10.1002/jclp. Lampropoulos, G. K., & Spengler, P. M. (2002). Introduction: Reprioritizing the role of science in a realistic version of the scientist-practitioner model. Journal of Clinical Psychology, 58(10), 1195–1197. Lampropoulos, G. K., Spengler, P. M., Dixon, D. N., & Nicholas, D. R. (2002). How psychotherapy integration can complement the scientist-practitioner model. Journal of Clinical Psychology, 58(10), 1227–1240. Lane, D. A., & Corrie, S. (2007). The modern scientist-practitioner: A guide to practice in psychology. London: Routledge. Levant, R. F., & Hasan, N. T. (2008). Evidence-based practice in psychology. Professional Psychology: Research and Practice, 39(6), 658–662. https://doi.org/10.1037/0735-7028.39.6. 658. Lilienfeld, S. O., Ritschel, L. A., Lynn, S. J., Brown, A. P., Cautin, R. L., & Latzman, R. D. (2013). The research-practice gap: Bridging the schism between eating disorder researchers and practitioners. International Journal of Eating Disorders, 46(5), 386–394. https://doi.org/10.1002/ eat.22090.

52

3 The Applicative Split: The Science-Practitioner Model of Training and Practice

Lilienfeld, S. O., Ritschel, L. A., Lynn, S. J., Cautin, R. L., & Latzman, R. D. (2013). Why many clinical psychologists are resistant to evidence-based practice: root causes and constructive remedies. Clinical Psychological Review, 33(7), 883–900. https://doi.org/10.1016/j.cpr.2012. 09.008. Lueger, R. J. (2002). Practice-informed research and research-informed psychotherapy. Journal of Clinical Psychology, 58(10), 1265–1276. Maddux, R. E., & Riso, L. P. (2007). Promoting the scientist–practitioner mindset in clinical training. Journal of Contemporary Psychotherapy, 37(4), 213–220. https://doi.org/10.1007/ s10879-007-9056-y. Malott, R. W. (2018a). A model for training science-based practitioners in behavior analysis. Behavior Analysis in Practice, 11, 196–203. https://doi.org/10.1007/s40617-018-0230-3. Malott, R. W. (2018b). A science-based practitioner model. Education and Treatment of Children, 41(3), 371–384. https://doi.org/10.1353/etc.2018.0020. Mellott, R. N. (2007). The scientist-practitioner training model for professional psychology. American Behavioral Scientist, 50(6), 755–757. Myers, D. (2007). Implication of the scientist-practitioner model in counseling psychology training and practice. The American Behavioural Scientist, 50(6), 789–796. https://doi.org/10.1177/ 0002764206296457. Overholser, J. C. (2007). The Boulder model in academia: Struggling to integrate the science and practice of psychology. Journal of Contemporary Psychotherapy, 37(4), 205–211. https://doi. org/10.1007/s10879-007-9055-z. Overholser, J. C. (2015). Training the scientist–practitioner in the twenty-first century: A risk– benefit analysis. Counselling Psychology Quarterly, 28(3), 220–234. https://doi.org/10.1080/ 09515070.2015.1052779. Peluso, D. L., Carleton, N., & Asmundson, G. J. (2010). Clinical psychology graduate students’ perceptions of their scientific and practical training: A Canadian perspective. Canadian Psychology, 51(2), 133–139. Retrieved from https://psycnet.apa.org/record/2010-12180-008. Petersen, C. A. (2007). A historical look at psychology and the scientist-practitioner model. American Behavioral Scientist, 50(6), 758–765. https://doi.org/10.1177/0002764206296453. Pilecki, B., & McKay, D. (2013). The theory-practice gap in cognitive-behavior therapy. Behavior Therapy, 44(4), 541–547. https://doi.org/10.1016/j.beth.2013.03.004. Pillay, A. L., Ahmed, R., & Bawa, U. (2013). Clinical psychology training in South Africa: A call to action. South African Journal of Psychology, 43(1), 46–58. https://doi.org/10.1177/ 0081246312474411. Pillay, A. L., & Johnston, E. R. (2011). Intern clinical psychologists’s experiences of their training and internship placements. South African Journal of Psychology, 41(1), 74–82. https://doi.org/ 10.1177/008124631104100108. Raimy, V. (1950). Training in clinical psychology. New York: Prentice-Hall. Ridley, C., & Laird, V. (2015). The scientist–practitioner model in counseling psychology programs: A survey of training directors. Counselling Psychology Quarterly, 28(3), 235–263. https://doi.org/10.1080/09515070.2015.1047440. Rourke, B. P. (1995). The science of practice and the practice of science: The scientist-practitioner model in clinical neuropsychology. Canadian Psychology, 36(4), 259–277. Retrieved from https://search-proquest-com.uplib.idm.oclc.org/saveasdownloadprogress/ 20D9BFB4D4214F18PQ/false?accountid¼14717. Rowe, P. M. (2018). Work experience, the scientist-practitioner model, and cooperative education. Canadian Psychology/Psychologie Canadienne, 59(2), 144–150. https://doi.org/10.1037/ cap0000148. Skourteli, M. C., & Apostolopoulou, A. (2015). An interpretative phenomenological analysis into counselling psychologists’ relationship with research: Motives, facilitators and barriers – A contextual perspective. Counselling Psychology Review, 30(4), 16–33. Spengler, P. M., & Lee, N. A. (2017). A funny thing happened when my scientist self and my practitioner self became an integrated scientist-practitioner: A tale of two couple therapists

References

53

transformed. Counselling Psychology Quarterly, 30(3), 323–341. https://doi.org/10.1080/ 09515070.2017.1305948. Stoltenberg, C. D., & Pace, T. M. (2007). The scientist-practitioner model: Now more than ever. Journal of Contemporary Psychotherapy, 37(4), 195–203. https://doi.org/10.1007/s10879-0079054-0. Stoltenberg, C. D., Pace, T. M., Kashubeck-West, S., Biever, J. L., Patterson, T., & Welch, I. D. (2000). Training models in counseling psychology: Scientist-practitioner versus practitionerscholar. The Counseling Psychologist, 28(5), 622–640. https://doi.org/10.1177/ 0011000000285002. Stone, G. L. (2006). The scientist-practitioner in context. Counselling Psychology Quarterly, 19(3), 305–312. https://doi.org/10.1080/09515070600960522. Stoner, G., & Green, S. K. (1992). Reconsidering the scientist-practitioner model for school psychology practice. School Psychology Review, 21(1), 155–166. Retrieved from https:// psycnet.apa.org/record/1992-29394-001. Stricker, G. (2002). What is a scientist-practitioner anyway? Journal of Clinical Psychology, 58 (10), 1277–1283. van der Watt, R. (2016). Strengthening doctoral supervision in a Doctor of Psychology (DPsych) specialisation in child and adolescent psychology. Journal of Psychology in Africa, 26(1), 84–91. https://doi.org/10.1080/14330237.2016.1149332. Vespia, K. M. (2006). Integrating professional identities: Counselling psychologist, scientistpractitioner and undergraduate educator. Counselling Psychology Quarterly, 19(3), 265–280. https://doi.org/10.1080/09515070600960555. Vespia, K. M., & Sauer, E. M. (2006). Defining characteristic or unrealistic ideal: Historical and contemporary perspectives on scientist-practitioner training in counselling psychology. Counselling Psychology Quarterly, 19(3), 229–251. https://doi.org/10.1080/ 09515070600960449. Vespia, K. M., Sauer, E. M., & Lyddon, W. J. (2006). Counselling psychologists as scientistpractitioners: Finding unity in diversity. Counselling Psychology Quarterly, 19, 223–227. https://doi.org/10.1080/09515070600960506. Wakefield, J. C., & Kirk, S. A. (1996). Unscientific thinking about scientific practice: Evaluating the scientist-practitioner model. Social Work Research, 20(2), 83–95. Retrieved from https:// academic.oup.com/swr/article-abstract/20/2/83/1631932. Williams, M. E., Sayegh, C. S., & Sherer, S. (2018). Promoting scholarly training in a clinical psychology postdoctoral fellowship. Training and Education in Professional Psychology, 12 (2), 90–95. https://doi.org/10.1037/tep0000185. Wilson, J. L., Armoutliev, E., Yakunina, E., & Werth, J. L. (2009). Practicing psychologists’ reflections on evidence-based practice in psychology. Professional Psychology: Research and Practice, 40(4), 403–409. https://doi.org/10.1037/a0016247.

Chapter 4

The Metatheoretical Opposition: Positivism and Constructionism

Introduction Let us summarise what we have up till now. We started off with a look at the mythical image created by various role players which, may I say, seem ideological or political, because the particular formulations of what science is serves particular purposes in the various approaches. The mythic image of science at this stage is the belief that science is largely modelled on what the natural sciences do; it involves quantification to a large extent; it uses experimentation and it regards the individual as primary target. When it, ironically, hides the individual within aggregate and large sample data, then all the characteristics we hold dear of the subject disappears within an average “man.” Science on human beings is largely aimed at explaining, namely, finding causes that are based on a regularity conception of causation. Apparently, if we can say that an increase in smoking causes cancer this is enough to understand the phenomenon of lung cancer. Or to take another example: mothers drinking coffee while pregnant tend to have overweight children (Papadopoulou et al., 2018). In order for the psychological scientist to do his/her research and experiments, they should do so objectively, thus, not letting their own bias influence the study because facts and findings should be generalisable, i.e., timeless and objective. This image of science has been labelled positivist, post-positivist or even realist. So, which is it? Within the polarity of the quantitative and qualitative debates about methods, we saw that proponents of the qualitative paradigm can have one of two approaches towards the quantitative paradigm. The first entails paradigm incompatibility, thus, the study of human phenomena cannot in principle be done from within the quantitative paradigm. Positivist science cannot account for understanding the meaning of what it means to be human—meaning cannot be captured by numbers and even the only way one can access human experience is through language. The second approach is based on the ability to utilise both approaches to study people. Thus, triangulation means having different perspectives on similar phenomena and © Springer Nature Switzerland AG 2020 D. J. F. Maree, Realism and Psychological Science, https://doi.org/10.1007/978-3-030-45143-1_4

55

56

4 The Metatheoretical Opposition: Positivism and Constructionism

the paradigm compatibility thesis usually spills over into a pragmatic stance and mixed methods approach. Here we also observe some mythmaking about methods. Some staunch incompatibilists regard mixed methods as a mere postpositivist capitulation. The scientist-practitioner polarity contributes to the mythic image of science from another perspective. The intention of advocating the SP model was praiseworthy but unfortunately boomeranged against its originators. Practice or skills training is seen as de-valuing scientific approaches and research. Thus, even advocating evidence based practices seems to fail because they are based on science, but a specific view of science, namely, positivist science! The failure to embrace the SP model at least by practitioners can be traced to the view of science they have been trained in and exposed to, namely, quant/positivist science which in principle cannot be applied to human experience. Now we have to ask what lies at the root of the mythic image of science. The third polarity lies on the metatheoretical level and as we have seen in the discussion of the paradigms, it is a very convincing and powerful level influencing subsequent practices and beliefs. The marketing machine of the qualitative proponents worked immensely well and all because of the, rightfully so, emphasis on the metatheoretical level. It is on this level where we need to do battle and expose the foundations of the mythic image of science. We need to examine what both positivism and constructionism involve on a metatheoretical level, which implies looking at their ontologies and epistemologies. In the process of identifying the mythic image of science and pointing to a realist one it is crucial to understand what positivism is, on the one hand, but on the other hand the same goes for the classical leftist reaction, namely, constructionism. When looking at positivism as a metatheory then it becomes clear that it is a very restricted empiricist theory so it is necessary to describe its characteristics briefly and how it relates to empiricism. Similarly, for constructionism as a metatheory: looking at its epistemological and ontological implications then it becomes clear that its roots lie in idealism, i.e., making the real dependent on the mind. It is important to distinguish constructionism’s relativistic implications from its methodological approach, an issue as said above will become crucial for our understanding of science: a critical stance is crucial for science, but one needs to tease this from the relativism underlying the idealism of constructionism.

Positivism, Empiricism and Its Ontology Positivism and its variants from Comte to the Logical Positivists (LP) have some common principles, among others, maintaining the importance of empirical experience in establishing facts or knowledge. Despite some commonalities the views of Logical Positivists differed on a number of points and between each other. Although the tendency is to present it as such, positivism is not an united front and has many

Positivism, Empiricism and Its Ontology

57

variants (Vincent, 2017). Carnap differed in his views from other Logical Positivists despite them known as the Vienna Circle. The Logical Positivists initiated a linguistic turn and in their delineation of language as formal and material (Carnap) or structure and content (Schlick) they tried to distinguish between the subjective and observational aspects of what constitutes knowledge of reality. They attempted to solve the problem of the subjectivity of meaning by externalising it: language needs to be reduced to empirical, i.e., observable, simple units and only then these expressions of fact are meaningful. Their empiricism extended to causality denying things such as unobservable mechanisms.

Classical Empiricism Locke, Berkeley and Hume, or the British empiricists, provided their answers to how knowledge of reality is established. According to them knowledge can be found in the ideas in the mind but which have their origin in sense perception. Chalmers (2013, p. 3) calls this a psychological justification of empirical knowledge, while the positivists eliminated the psychological or mind part—they base knowledge on empirical observation. Locke rejected innate knowledge and held that all knowledge can be reduced to experience, whether external experience caused by external sensations (such as perceiving), or internal sensations (such as feeling). Ideas (or concepts) play a role in both thought (introspective experience) and sensation (external experience). Locke’s version of empiricism wants to deny the innateness of some of the conceptual material and knowledge the rationalists assumed to be part of the mind’s constitution, but he wants to show that thought and its operations can be linked to external sensations by means of ideas (or concepts and precepts) (Winkler, 2010, p. 47). Everything required for the mind to understand is thus given by experience. According to Winkler (2010, p. 46) Kant clarified Locke’s vague usage of idea by referring to ideas as representations. Hume viewed ideas as the less clear counterpart of sensations or the direct impressions of sensational experience. Importantly Kant then distinguished between three types of representations, namely, (a) intuitions (Anschauungen) or a representation with a direct relation to its object, (b) concepts (Begriffe) or mediate presentations of objects (Winkler, 2010, p. 46) and (c) ideas (Ideen) that are Platonic concepts beyond experience. Hume made a distinction between clear and distinct sensations and ideas which entail concepts about the sensations. Knowledge need to be traced to its simple sensations. An important contribution of Hume was his denial that concepts such as necessity and universality can be empirically grounded. We will encounter the realist interpretation and rejection of Hume’s concept of necessity and causality later on. Suffice to say that Hume (1962) denied the reality of something like a cause, because it cannot be observed: all that one can observe are two events following each other. If billiard ball A bumps into billiard ball B, we normally say A caused B to move (or something to this effect). However, Hume denies that there is something like a cause: all that we see are two events taking place

58

4 The Metatheoretical Opposition: Positivism and Constructionism

in close proximity spatially and in time. This constant conjunction of events is what our minds associate as a cause: thus, causality can be written off as a psychological habit or association (Hume, 1962, p. 48). The “constant conjunction” concept of cause is called a regularity conception of causation and will play a role when discussing realism and why an empiricist conception of science cannot be adequate.

Logical Positivism LP is not a unified doctrine of ideas about philosophy and science (Creath, 2017). It can be called a movement that was visible in the 1920–30s especially in Europe and then after WW2 in the USA in the 1940–50s (Creath, 2017). It also cannot be centralised around one figure but had many representative voices and was located in different European centres, such as Vienna or Berlin. The prominent voices differed on some important matters from each other (Mormann, 2007). Currently a re-evaluation of positivism is underway (Richardson & Uebel, 2007). Stadler (2007, p. 33) notes that research findings “. . . render obsolete all previous characterisations of the Vienna Circle as a homogenous, (anti-) philosophical school.” As we have seen in previous chapters the image of science usually called the received view is wholly infested with ideas about positivism. It is this, quasireceived view we call the mythic image of science. A reconsideration of positivism and specifically LP does not signal—to the qualitative scientist’s dismay—a resurgence of positivist and quantitative ideas of science. It merely means that we can start separating the chaff from the corn, or dismantle the positivist straw man. One of the commentators in Richardson and Uebel (2007) observed that LP can be distinguished from the received view of science that originated in the 1950s (Stadler, 2007, p. 35). Much of the negative associations of positivism can be ascribed to this received view which Reisch (2007) ascribes to Kuhn’s (1962) publication. New research showed that LP cannot be reduced to the two principles they are widely known for, namely verification and value-neutrality. A broader scientific conception was advocated by proponents of the LP but the popularity of these two principles as describing LP can be ascribed to Ayer (1936/1971) (Stadler, 2007, p. 27). The LPs fully supported methodological pluralism, and were accommodating in their debates of diverse views. As Reisch (2007, pp. 58–59) observed: For Kuhn, logical empiricism was a philosophy of science concerned mainly with logic and using logic to understand science. But only two decades before, logical empiricism was engaged with progressive, modern trends in science, social life, education, architecture, and design. Originally logical empiricism was in the business of social enlightenment.

Different names are used for positivism, namely, the Vienna Circle, Logical Positivism or Logical Empiricism or just positivism. The broader term, Logical Empiricism (LE), is preferred by proponents of LE and Logical Positivism (the term I shall use) (LP), is used also albeit more generically. Godfrey-Smith (2003, p. 19) restricts the term LP to pre-WW2 whilst LE is used for moderate versions post

Positivism, Empiricism and Its Ontology

59

war. It is difficult to make clear distinctions and Vienna Circle refers to a subgroup of LPs (Creath, 2017). The latter was established by a few philosophers pre-WWI and developed while LP was in a non-public phase (Stadler, 2007, 2015). Its public phase was announced with the publication of their manifesto called “Wissenschaftliche Weltauffassung. Der Wiener Kreis” (The Scientific WorldConception. The Vienna Circle) (Stadler, 2007). This manifesto aimed at distinguishing itself from the world view of continental philosophy and emphasised “principles of this-worldliness, practical relevance, and interdisciplinarity” (Stadler, 2007, p. 14). An important issue for the LPs was taking up the challenge of the new phase in physics and science in the light of the work of Ernst Mach, Max Planck, Albert Einstein and others, i.e., countering the strong metaphysical inclinations of the science community and the consumers of these philosophies and religion (Stadler, 2007, pp. 16, 30). Some of the LPs were involved in the Viennese adult education initiative pre-WW2 focusing on popularising scientific philosophy reacting against the rising authoritarianism in Austrian and German Universities (Stadler, 2003, p. xvi, 2015, pp. xx–xxi). LP developed with the aim to be open to the community and education of the public. Its liberal socio-political approach by no means found acceptance in the academic circles the LPs moved. Some struggled to get appointments and given the WW2 government in Germany almost all emigrated to mainly the UK and the USA (Stadler, 2007, p. 29). LP developed through phases. The first was a private phase (1907–1911) where interested parties met informally and discussed the issues that were to become part of their principles. The public phase, announced by above mentioned manifesto and a conference in 1929, saw the establishment of the Vienna Circle in Prague, mainly, constituted by Moritz Schlick, Rudolf Carnap, Hans Hahn, Otto Neurath and Herbert Feigl (Stadler, 2007). Members of the circle were Gustav Bergmann, Rudolf Carnap, Herbert Feigl, Philipp Frank, Kurt Gödel, Hans Hahn, Viktor Kraft, Karl Menger, Marcel Natkin, Otto Neurath, Olga Hahn-Neurath, Theodor Radakovic, Moritz Schlick, and Friedrich Waismann. Stadler (2007) argued for a core group of 19 persons that regularly attended the Circle’s meetings and identified 16 visitors. Wittgenstein and Popper never attended but were associated with some members (Stadler, 2007). It was from this stage that the group was known as the Vienna Circle. According to Stadler (2015, pp. 13–14) basic positivist principles could already be discerned before WW1, namely, (a) their acceptance of an empirical approach with operationalism, instrumentalist and pragmatism well integrated, (b) the aim of the unity on science, i.e., scientific monism, (c) their preference for methodological and linguistic nominalism, (d) a critical orientation towards metaphysics, (e) conventionalism combined with a holistic theory of science, and (f) a formal range of logical tools and methodological phenomenalism. In the following I will give a brief description of these issues as an indication of some of the core aspects of LP. However, it must be pointed out as we have seen trying to pinpoint the characteristics of LP is a difficult task because of the differences between LPs. Stockman (1983) found some agreement between authors he surveyed, especially on the unity of science and empiricism but not on other characteristics. However, if

60

4 The Metatheoretical Opposition: Positivism and Constructionism

most authors agree on the LP’s empiricism and science monism then one needs to have a careful look how they interpret each characteristic because these may also differ (Stockman, 1983, p. 6). For instance, Stadler (2007, p. 25) points out that even the principle of scientific monism, generally agreed to be a key aspect of LP, was not generally accepted amongst LPs for intermittent periods. Empiricism as LPs most salient characteristic is based on early empiricist dogma that all that the mind has to work with are the sensations obtained from the empirical world. Their version of empiricism can be called phenomenalism (Bunge, 2017): we do not have direct access to the world but to structured and organised collections of sensations (Godfrey-Smith, 2003). Kołakowski (1968, p. 3) puts it nicely: there is no difference between essence and phenomenon, i.e., what you see is what you get (or what there is). The importance of their empiricism is also illustrated by the unification of science principle or scientific monism (Stadler, 2007, p. 16). As physics was the ultimate illustration of the LP approach to reality, coupled with their drive to establish philosophy on scientific footing (Stadler, 2007, p. 19) along with establishing a world conception based on science (Stadler, 2007, p. 14), they expected other sciences to comply with the strict empiricist and non-metaphysical principles of their system. Closely related to phenomenalism is the LPs’ nominalism. Nominalism is the denial of the existence of universals and/or abstract objects (Rodriguez-Pereyra, 2015). Universals and abstract objects are not the same and one might be a nominalist about universals or abstract objects or both. This may come as a surprise to those holding the mythic image of science but LPs deny the universality of objects (Kołakowski, 1968, p. 7)! Universality is a characteristic of language, thus, the way we talk about things seems to indicate universal properties but again, all we have is the experience of empirical reality, and talk of anything else uses merely names for things (Heller, 2011, p. 8). Science also uses language, sometimes referring to ideal things, such as triangles which comply to strict requirements (such as inside angles adding up to 180 ), but this usage is to order, calculate and systematise. There is no ideal, existing and real triangle, only those we encounter in experience: “According to nominalism, in other words, every abstract science is a method of ordering, a quantitative recording of experiences, and has no independent cognitive function in the sense that, via its abstractions, it opens access to empirically inaccessible domains of reality” (Kołakowski, 1968, pp. 6–7). It is important to know what the LPs reacted against and why they formulated some of their basic theses. Kant made a distinction between a priory and a posteriori knowledge or statements. The first remains true irrespective of experience while the latter depends on experience (Scruton, 2001). His interest was on a priori statements but not analytic a priori claims which are tautologically true (such as bachelors are unmarried). It was the synthetic a priory statements that needs to be justified, i.e., those claims that are true irrespective of experience but qualifies as substantial knowledge (such as bachelors are arrogant) (Scruton, 2001). He defended the existence of synthetic a priori knowledge, i.e., knowledge not derived from empirical experience but knowledge that fundamentally applies to the empirical world (Friedman, 1998). This means that certain a priori principles are required to make sense of

Positivism, Empiricism and Its Ontology

61

the empirical world and these principles can be found in the mind of the knower. An example would be the principle or category of cause that allows us to perceive causal relations when observing the empirical world. Kant rejected metaphysical knowledge, namely, knowledge not originating in empirical experience. Kant regarded the mathematical physics of Newton with its reliance on Euclidean geometry as an example of synthetic a priori knowledge (Friedman, 1998; GodfreySmith, 2003, p. 26). However, the development of non-Euclidian geometry provided challenges to staunch neo-Kantians that defended the centrality of Euclidean geometry for Kant’s system (Friedman, 1998). However, others aimed to widen the applicability of the synthetic a priori to include non-Euclidian geometry. This was especially the case with the LPs who had to adapt how one can account for the synthetic a priori in the light of developments in mathematics and physics that differed from Kant’s Newtonian view. The development of Einstein’s relativism made further demands on the LPs philosophical beliefs because it is not compatible with the Kantian a priori. They had to accept a form of conventionalism initially adopted by Poincaré (Friedman, 1998). This conventionalism holds that, for instance, the Euclidian axioms are not based on empirical fact neither on synthetic a priori but convention, a convenient choice based on practical matters but one could just as well have chosen a non-Euclidian system (Rescorla, 2019). The LPs thus applied this line of reasoning to the choices made in other areas of science and physics. Rescorla (2019) summarises their conventionalism clearly using Carnap’s view of language as an example: Beginning with Logical Syntax of Language (1937/2002), Carnap developed a particularly thoroughgoing version of linguistic conventionalism. Carnap invites us to propose various linguistic frameworks for scientific inquiry. Which framework we choose determines fundamental aspects of our logic, mathematics, and ontology. For instance, we might choose a framework yielding either classical or intuitionistic logic; we might choose a framework quantifying over numbers or one that eschews numbers; we might choose a framework that takes sense data as primitive or one that takes physical objects as primitive. Questions about logic, mathematics, and ontology make no sense outside a linguistic framework, since only by choosing a framework do we settle upon the ground rules through which we can rationally assess such questions. There is no theoretical basis for deciding between any two linguistic frameworks. It is just a matter for conventional stipulation based upon pragmatic factors like convenience.

The LPs’ conventionalism allowed them to not base mathematics and logic on empirical experience, neither on metaphysical intuitions, but regard them as analytic: “For logical positivism, mathematical propositions do not describe the world; they merely record our conventional decision to use symbols in a particular way” (Godfrey-Smith, 2003, p. 26). Thus, the experiential claims (empirical experience or synthetic statements) about the world are verbalised by means of something such as a mathematical language without the language system describing the real world. Mathematics and logic used for these purposes are wholly analytic and a priori, thus devoid of factual content but knowable without reference to experience. The LPs had a holistic view of empirical testing in science. Hypotheses can thus not be rejected on their own but only when consider within the whole theory for research. Although the LPs defended this principle which is a remarkable

62

4 The Metatheoretical Opposition: Positivism and Constructionism

post-positivist one, Godfrey-Smith (2003, p. 32) points out that their focus on verifiability of specific protocol sentences sounded very much individualistic so that the import of their holism was overlooked: hypotheses tend to get tested one at a time. However, despite their own holism Quine (1951, p. 38) levelled substantial criticism against LP: theories as a whole are put to empirical test, not isolated assumptions. What is now known as the Duhem-Quine thesis entails that no single observation or test can falsify a theory (Chalmers, 2013, p. 83). It is always possible to generate post hoc reasons why a theory should be retained especially in the light of its background knowledge and beliefs, auxiliary hypotheses, measurements and so on (Chernoff, 2012, p. 231). The principle is that a theory is always underdetermined by the data, which means it is possible for more than one theory to fit observed data. The first implication is that knowledge is fallible, i.e., it is always possible to change a theory in the light of available evidence, and the second implication is that single hypotheses cannot be falsified or verified conclusively because they form part of a whole body of theory and related aspects. According to Chernoff (2012, p. 233) Quine’s underdetermination is more radical than that of Duhem (1991/1954). Whereas adjustments in the light of anomalous observations can be made to the theory and its relations within for instance a paradigm, Quine (1951, p. 38) holds that adjustments can be made anywhere, even in the laws of logic, because “our statements about the external world face the tribunal of sense experience not individually but only as a corporate body.” Thus, clearly, the LPs’ version of holism did not follow its implications to the end because the Duhem-Quine versions of holism point to the impossibility of testing any hypothesis conclusively. This does not bode well for the LP hypothetico-deductive methodology for scientific testing (Stadler, 2007, p. 21). The methodological phenomenalism and logical approaches are clearly evident in how the LPs rejected metaphysics. As we have seen above the empirical stance was largely a phenomenalism, i.e., all that we have are impressions or sensations of our environment (Kołakowski, 1968, p. 4). By combining logical atomism with this phenomenalism they manage to reduce statements to protocol sentences that have to be supported by sensations or empirical observations. Logical atomism refers to statements or claims about phenomena that can be reduced to their constituent sentences or statements (Stockman, 1983, p. 20). These statements should correspond to atomic facts of the world and are then known as protocol statements. Combining atomic sentences into molecular ones with logical connectors allowed one to evaluate the truth of these molecular statements. Different categories of molecular statements can be discerned. (a) As we have seen analytic a priori statements refer to mathematics and logic whose truth is not dependent on experience. They are true by virtue of being tautologies like all bachelors are unmarried, or contradictions of each other. (b) Other molecular statements depended on the truth of atomistic statements in so far as they correspond with empirical facts. Lastly, (c) there are statements whose truth cannot be determined at all. These are metaphysical claims that are not contradictions, tautologies or can not be supported by an empirical fact. Stockman (1983, p. 21) says that the combination of logical atomism and phenomenalism (or sensationalism) formed the verifiability criterion of

Positivism, Empiricism and Its Ontology

63

meaningfulness. Thus, the principle of verifiability indicates states that “the meaning of a proposition is its method of verification” (Stockman, 1983, p. 21). A metaphysical statement cannot be verified and is thus not meaningful. Implicit in the discussion about the various characteristics of LP is the question of how they viewed the relationship between theory and empirical reality. Clearly, its conventionalism showed that there is a language to be used when describing observable reality and it may also be the case that so-called conceptual, abstract or non-observable “entities” can figure in a theory. Theories can be viewed as syntactical, semantic or pragmatic (Mormann, 2007, p. 137). The syntactical kind of theory consists of a system of axiomatised sentences, while the semantic type views theories as a collection of models, usually non-linguistic and/or mathematical. LPs have a syntactical view of theories but this is challenged by the semantic view, and both are challenged by the pragmatic view that typifies theories as consisting of skills, practices, sentences, models and problems. Although we characterise the LP view as syntactical, usually only one LP’s view is taken to illustrate the view because others vary too much in their views. Again, as stated above there is no standard LP view, only views of different proponents and Mormann (2007, p. 138) makes a case that to some extent all three types of theories were represented by LPs at different stages of their development. The LP view of a theory as a self-contained axiomatic and mathematical system is not strictly what they had in mind. Although admiring the exigencies of a syntactical and axiomatic system, their fundamental empiricism required the theory to have some connection with empirical experience. An oft quoted example of an LP view of the relationship between empirical reality and theory is that of Herbert Feigl in Fig. 4.1. The theory can be described in terms of a two component language (L), namely, a vocabulary (LO) describing the empirical level, and a vocabulary (LT) describing the theoretical level, consisting of postulates and concepts (Mormann, 2007, p. 141). The two levels need to be translated into each other by means of correspondence rules and the experiential level provided meaning to the upper theoretical level. Feigl’s model is one of the solutions of the relationship between theory and empirical reality. For instance, taking our cue from Duhem (1991/1954) mentioned above, holism means that there is no one-to-one correspondence between a theoretical statement (consisting of a symbolic or mathematical statement) and an empirical statement expressed in natural language: “Rather, there is a many-many correspondence: to every symbolic fact there corresponds an infinity of experimental facts and vice versa” (Mormann, 2007, p. 142). However, the LPs strived to make a connection between theoretical statements or systems and empirical statements. Devising an axiomatic system, for instance, for mathematics, is obvious given its terms are internally defined and truth and falsity of statements can logically or deductively be inferred. The challenge is to have a similar system for empirical language, hence their emphasis on deductively related protocol statements as we have indicated above and to relate the empirical and symbolic systems (Mormann, 2007, p. 141). The LP criterion of meaning, namely, relating a theoretical statement to an observation statement was one way to conceptualise the relationship between theory and empirical reality but the LPs each had different answers to the problem

64

4 The Metatheoretical Opposition: Positivism and Constructionism

Fig. 4.1 Feigl’s LP presentation of theory (Mormann, 2007, p. 141)

but the mainly syntactical approach to theory proved to be unfruitful and insolvable (Mormann, 2007, p. 141). The development of LP, its two decade development pre-WW2 and its sedimented image formulated by Kuhn (1962) and others in the 1950s and 60s as the received view must be seen against the political developments of the first part of the twentieth century. The movement of LP proponents beyond the borders of Europe when they emigrated carried with them socialist ideas of education, society and government which was intimately bound with their concept of the unity of science and the importance of science as worldview contra metaphysical and religious ideas about the world. In a revealing analysis of the socio-political history in the USA, Reisch (2007) points out that the unity of science ideal and its socialist associations were severely repressed by the conservative and anti-communist sentiment in the USA. The opposition to the LPs ideas were thus such that it was easier to espouse a sanitised and non-threatening LP focusing on logic and dropping the unity of science principle. It is ironic that the liberal, progressive and socially involved LPs eventually became the symbol of authoritarianism, value-free science and cold, objective distance between scientist and objects/subjects of study. One might venture the guess that within reasonable bounds today’s post-modernists and various brands of anti-establishment neo-liberalists would have pleased the original LPs. A post war sanitised version of science formed the received view, a view that functioned as a springboard for the postpositivists (Mormann, 2007).

Positivism, Empiricism and Its Ontology

65

However, one should be careful to ascribe too much to socio-political context especially about what the LPs said about science. For one, they, or some of them, believed in a strict unity of science initiative. The unity of science meant empirical approaches for those disciplines who regarded themselves as sciences and in particular that no metaphysical and non-verifiable claims can be tolerated. Thus, they had an affinity over and above their mathematical and physics theorising for experimental psychology, especially of the behaviouristic variety (Hardcastle, 2007). Anything clinical, therapeutic or non-experimental could not qualify as scientific: The logical empiricists’ insistence that psychology be approached as a natural science, subject to the general principles of logic, should squelch any impression that the use of psychological concepts by the logical empiricists inclined them toward psychologism, that is, the view that logical entities like propositions, sentences, meanings, or the like were ultimately psychological entities, or further that logical laws can be explained by reference to psychological laws (Hardcastle, 2007, p. 231).

One is reminded of Popper’s (1963/2002) insistence that psychoanalysis cannot qualify as scientific because its theories cannot be falsified. Popper used psychological theories as an example of pseudoscience where the claims of a theory are justified when confirmed by observed instances for whomever seeks such confirming instances. It is much harder to find instances that contradicts a theory or claim, hence his proposal of falsification. The requirement of scientificity for the LPs was then empirical or observational grounds for any knowledge claim. Its rejection of large parts of humanities, social sciences and psychology because of these disciplines’ noncompliance to experimental criteria was too strict and can rightly be rejected. Although the programme of LP is long gone, it must be acknowledged that LP ideas and debates had a major impact on science theoretical and philosophical thinking since the 1950s (Richardson, 2007). When one looks at the LPs socialist and liberal undertones of their a-metaphysical and strict empirical stance its fundamental critical orientation towards authority of various sorts makes sense (Stadler, 2007, 2015). Unfortunately, LP inspired hegemonic versions of itself bolstering the conservative and ruling scientific empire of western empirical science.

Kuhn’s Postpositivism Historians and other commentators generally agree that LP was overthrown or largely superseded by one single work, namely, that of Kuhn (1962) (Devlin & Bokulich, 2015; Godfrey-Smith, 2003; Kindi & Arabatzis, 2012; Pandit, 1991; Stadler, 2015; Toulmin, 2001). His work introduced postpositivism in the philosophy of science. In essence his work made the following claims about the development of science (Chalmers, 2013). One should look to actual history of science namely what scientists did and how science is practiced to develop a theory of science. Philosophy of science should not be an ideal theory against which practice is measured. Thus, in reality science is practiced very differently from, for instance, the

66

4 The Metatheoretical Opposition: Positivism and Constructionism

guidelines of LPs. Science goes through phases of normalcy and revolutions. The normal phase of science practice is characterised largely by puzzle solving activities thrown up by empirical data diverging from theories. It is unlikely that single empirical experiments would overthrow a theory and if such discrepancies arise it leads to adaptations to theories to accommodate those findings. In a subdued sense rational but also non-scientific factors play a role in theory formulation and testing. While the LPs formulated strict criteria for the distinction between metaphysics and empirical science, Kuhn (1962) says that social, metaphysical and other non-rational factors play a role in scientific progress (Chalmers, 2013, p. 102; Richardson, 2007, p. 350). Popper (1963/2002), for instance, allowed proto-scientific theories to influence scientific theorising. Initially, a clear distinction between extra-rational factors and validation of knowledge was made by means of the context of discovery and the context of justification. The context of discovery was reserved for the psychological sphere that leads to scientific discoveries but the context of justification is the sphere where theories are justified by logic and empirical data. However, with Kuhn, in a sense, the context of discovery widened to include social and other non-rational factors. Importantly, the LPs held the context of justification as the only valid sphere where scientific theories can be validated. Kuhn (1962) along with others increasingly argued for non-rational factors influencing the progress of science. In fact, scientific revolutions, viewed from outside, seem irrational. When one paradigm is replaced by another, the reasons for doing so do not seem rational but are based on many different socio-historical factors. Revolutions take place en-masse, relatively quickly and manage to convert adherents to the new views. For Kuhn, incommensurability was one of the hallmarks of paradigm shifts: usually, they are incompatible. Richardson (2007, p. 349) says that Kuhn’s broad principles listed above countered LP theory of science comprehensively and was responsible for the diminishing influence of LP’s ideas. However, Richardson (2007) makes us aware of an interesting puzzle. On the one hand, Kuhn’s work was responsible for the demise of LP: almost everybody quotes Kuhn’s work as the introduction of postpositivism. On the other hand, some early and recent work on the LPs showed that Kuhn’s ideas was largely in line with what many LPs believed about science and its progression (Richardson, 2007). For example, Kuhn’s “Structure of Scientific Revolutions” were published in “The Encyclopaedia for Unified Science” in 1962 and received a favourable review from some LPs (Reisch, 2007, p. 59; Stadler, 2007, 2015, p. xxiv). Richards and Daston (2016, p. 1), however, note that Kuhn regretted his publication commitment made in 1953 to publish in the Encyclopaedia. The concept of a-rational factors playing a role in scientific work was, as said above, even accepted by LPs as the distinction between the context of discovery and justification shows. Kuhn saw himself as not forging a fundamental break with the science theoretical tradition that developed between the 1930s and 50s. However, reviews that appeared of his 1962 work clearly acknowledge the new direction his work introduced in epistemology of science theory (Richards & Daston, 2016, p. 6). The time was ripe in the 1960s for new directions. Some theorists that proposed alternative views specifically non-empiricist views were Popper, Polanyi (1958) and

Social Constructionism and Its Roots in Idealism

67

Hanson (1958) (Richardson, 2007). Expanding on and criticising Kuhn’s ideas were philosophers such as Lakatos and Feyerabend (Chalmers, 2013). It should be apparent by now that there is a difference between theoretical work done on science in different spheres. The involvement and interest of the LPs in psychological experimental science definitely had a permeable effect on psychology and social science, thus philosophical work in the natural science impacted social science. For instance, as we have seen postpositivism from the psychological side is seen as a largely positivist continuation whilst simultaneously not taking cognisance of serious criticism of relativist and constructionism on the relativist side. The history of development is complex and it is not my place to examine the out-ofstep development of psychology and philosophy but to make the reader aware of the problematic image of science that developed in psychological circles. So how can we summarise the tenets of postpositivism without effecting another reduction and oversimplification. The discussions of LPs and PPs should alert us to the fact that glib summaries, although understandable, are not always correct and do more harm than good. As the good constructionists taught us, history and context should be primary!

Social Constructionism and Its Roots in Idealism Introduction Constructionism, or versions thereof, arose as a reaction against positivism and foundationalism in science at the beginning of the previous century (Hacking, 1999; Laudan, 1996). The search for a scientific method that provides true knowledge and the search for an assured foundation for the edifice of scientific knowledge were overthrown by the work of, among others, Kuhn, Rorty and Feyerabend (Hacking, 2002; Laudan, 1996). As we have seen, there is no method that can lead to truth or certain knowledge and the basis of empiricism and positivism, namely, that observation guarantees true and certain knowledge, was abolished by these postpositivist theorists when they showed that observation is always theory-laden (Parker, 1998). Observation takes place through the lens of theories, background knowledge, assumptions and perspectives and it was subsequently realised that scientific work was done in social contexts; if no universal criteria for truth, method and knowledge existed, these were socially determined and constructed. The shift from universal to local criteria emphasised the relative nature of science (Kvale, 1992). However, we need not dissolve science into a quicksand pool of relativism, because, as Sayer (1998, p. 122) says, theory-laden does not imply “theorydetermined.” Kenneth Gergen is a prominent advocate of social constructionism and he reacted consistently against the empiricist and positivist image of science which he regards as the received view. Situated within the postpositivist linguistic turn, social constructionism emphasises the central role of language in constructing reality for

68

4 The Metatheoretical Opposition: Positivism and Constructionism

communities of people. Social constructionism emphasises the constitutive nature of social communities in forming the identities of individuals. Individuality and subjectivity are denied to a large extent locating the individual within society and language. Gergen emphasises the discursive nature of reality and calls for an alternative way of doing psychological science in contrast to the scientistic version, namely positivist, that emphasises measurement, control, and experimentation. The social nature of the psychological requires an alternative mode of investigating on par with how that particular reality is constructed, namely, through language and discursively. Epistemic access to psychological reality is gained though understanding meaning and discourse. Gergen has a peculiar view of natural reality—he does not deny the importance of getting the mechanics right for getting to the moon or trying to save a person’s life though medial science, but calls the realities they work with alternative discourses—engineers at NASA need to call a spade a spade otherwise their attempts to launch rockets will fail: “When my children were physically ill I sought medical attention as opposed to a spiritual advisor. And I will readily talk about my pain and pleasure, seek annual medical check-ups and donate funds to famine relief and clothing to the Purple Heart” (Gergen, 2001, p. 42). Gergen denies subjective idealism, the position critics of social constructionists accuse them of. However, everything that there is can be described otherwise through language. Gergen (1999, p. 47) says that we can easily describe a world where gravity does not play a role: our discursive reality can be constructed differently from the one that we are used to. This is the point of trying to understand cultural differences and in a sense this outcome for social constructionism is relatively trivial. We may even concede that what people regard as real, such as a flat world or a rising sun, are real to them. However, marrying nonessentialism with discursive processes stimulates the slide to serious relativism, subjective idealism and solipsism. Gergen also distinguishes between a natural and psychosocial world where different rules apply, namely, determinate causality vs. freedom. He holds the usual objections against the received view of science, namely, predictability, objectivity, and the quest for laws. These issues cannot be applied to psychosocial phenomena because human nature is indeterminate and based on the generation of meaning. It is useful to learn from his insights on the social development of science, the criticism of the primacy of the individual or knowledge atomism and even the hegemonic authority of the scientist. Kenneth Gergen (1934–) was responsible for bolstering social constructionism in psychology, especially after his first controversial publication in 1973 called “Social psychology as history” (Colman, 2008). As will be seen in the discussion of Bhaskar’s theory of science, a version of constructionism is also popular in Critical Realism and some psychological constructionists and realists frequently cross boundaries (and swords). One strategy to examine the value of, for instance, constructionist postulates is to identify modernist assumptions and then see how postmodernist critiques attack these assumptions. Social constructionism (SC) as advocated by Gergen, identifies some of these assumptions and then provides an

Social Constructionism and Its Roots in Idealism

69

image of science. I examine the assumptions Gergen holds as important, as well as discuss what he regards as characteristic of scientific knowledge. According to Gergen (1999, p. 29), modernist assumptions lie in a “sense of self” as knowing, rational and autonomous, and in a view of objective knowledge, reason and moral foundations. Gergen indicates the limitations of these beliefs that boil down to the problematic relationship between language and reality. The assumption of my argument is that SC forms a crucial and core component of scientific thinking, and in fact as will become clear later on, this principle forms the core component of realism as well. The component I am referring to is the fundamental capacity to engage in criticism. However, it is not so easy to distil this component from SC discourse. Gergen’s view of traditional or westernised critique will be briefly examined along with what he regards as criticism SC proper.

Modernist Assumptions Modernist assumptions about science and humanity lie at the root of various epistemological and ontological conundrums according to the social constructionist. We will briefly look at the primacy of the individual, knowledge and then the role of language and reality. (a) Primacy of the individual One of the most important aspects that probably formed and influenced modernism and postmodernism in a fundamental way was the development of an awareness of the primacy and importance of the subject or individual. The development of the primacy of the self and the focus on the subject’s ability to think for itself, observe, evaluate and choose actions developed in reaction to the authority of medieval rulers (Gergen, 1999, p. 7). The modernist development in the enlightenment was made possible by the development of the primacy of the individual. The focus on the individual and her capacity to think and act resulted in a particular view of the self that is still prevalent. The most important consequence was on the power the subject had to discover her environment. The individual figured prominently in knowing his world in contrast to the authority of the church and state dictating beliefs and behaviour. Thus, a conception of rationality developed which formed the impetus behind enlightenment and modernism. According to Gergen (1999, p. 8) a distinction for various reasons was formed between within and without, namely, the reality of the mind and reality of an outside world. The result was a dualist ontology. The material world is deterministic and involves cause and effect but the world of the mind is governed by free-will. The certainty that the predictability of the natural domain brought, required its preservation. Therefore, because the law-bounded material reality cannot be sacrificed to indeterminism implied by free-will, the two domains

70

4 The Metatheoretical Opposition: Positivism and Constructionism

needed to be separated. Gergen is very aware of the implications of either a materialist or idealist monism: according to the idealist . . . the presumption that there is a material world is something we generate with our minds. Although idealist views remain alive in certain quarters, they are generally rejected. Few wish to accept the solipsism invited by this view, that is, the assumption that we each live in totally private worlds, and that even the belief that other persons exist is nothing more than a private fantasy” (Gergen, 1999, pp. 8–9).

Materialism likewise has unacceptable consequences by reducing the mind to mere cause and effect or machine-like behaviour subjected to laws. (b) Knowledge The focus on the subject has consequences for epistemology: how do we as subjective individuals get to know the reality or world out there? One way is to observe and describe nature or reality. This is the classical empiricist way. Mind was seen as the mirror of nature; we describe what we see as accurately as possible and this constitutes knowledge. Gergen (1999, p. 14) says “Knowledge begins, it is said, with careful observation.” He then shows that observation cannot be the correct starting point because each specialist or scientist views a phenomenon from her own perspective, patterns and categories. For example, the botanist and the physicist will describe a piece of wood from their own respective frameworks. Consequently, observation cannot lead to an accurate mirroring of nature. In order to obtain similar results from observation, the framework from which observation was made needs to be replicated. The framework is not only the immediate context but also research and related traditions. The categories used by specialists cannot be simply read off from nature. These categories are seated within traditions: . . . we must suppose that everything we have learned about our world and ourselves— that gravity holds us to the earth, people cannot fly like birds, cancer kills, or that punishment deters bad behaviour—could be otherwise. There is nothing about what there is that demands these particular accounts; we could use our language to construct alternative worlds in which there is no gravity or cancer, or in which persons and birds are equivalent, and punishment adored (Gergen, 1999, p. 47).

One problem associated with knowledge gathered objectively is the value neutrality of the exercise. According to Gergen the problem of value neutrality is also imbedded in the modernist paradigm. In the modernist context science requires value neutrality as a corollary of objectivity. Knowledge cannot be infested with subjective and thus untrue perspectives. However, according to (Gergen, 1999), so-called scientific interpretations and valuations are not valuenatural but interest-laden. This case does not only apply to social science, humanities, politics, etc. but also to natural science. (Gergen, 1999, p. 47) quite correctly quotes an example of how women’s bodies are referred to in medical texts. All interpretation is value-laden and the balance of power requires vested interests to be protected in the expression of science and culture.

Social Constructionism and Its Roots in Idealism

71

(c) Language and reality The distinctive link between the individual and reality can be disrupted by shifting to relational aspects of a community knowing and practicing science, but this leaves the medium for linking thought and reality, namely, language, intact. The realisation that language might be the source of the modernist quandary, namely, laying claim to objective truth and knowledge, led to the semiotic and deconstructionist critique of language (Gergen, 1999, p. 29). The question is how language manages to bridge the divide between world and subjective knowledge. The modernist view of language leads to what Gergen calls the picture theory of language (or the correspondence theory of language), i.e., words describe the world as it is. It provides us with pictures of the world. However, there is a fundamental divide between reality and language or as he states it “The terms by which we understand our world and our self are neither required nor demanded by ‘what there is’” (Gergen, 1999, p. 47). Photographs and verbal descriptions do represent, describe or picture what there is—it is open to many interpretations—in fact, there is no way to distinguish better or worse descriptions “. . . in principle (though not in practice)” (Gergen, 1999, p. 47). The divide between language and reality is motivated by postmodern developments about language. Gergen (1999, p. 24) refers to de Saussure’s (1959, pp. 67, 73) semiotics and two basic ideas, namely, that the relationship between signifier and signified is arbitrary and that sign systems are closed, i.e., the rules determining relationships and meaning are indigenous to the system. The implication is that in language as a sign system meaning is assigned to words entirely arbitrarily. There is nothing in the world that dictates what a specific word should mean, accordingly, what a word means is arbitrary. Also, words find their meaning within the relationship between other words without the need to refer to the world. Not only is the meaning of words internally determined, syntax and grammar also function as internal rules to make sense of what we are saying. If this is true then meaning has no link with the outside world according to Gergen (1999, p. 26). Deconstruction drove the next nail into the coffin of the correspondence theory of language. Referring to Derrida, Gergen (1999, pp. 26–27) points out two aspects in deconstruction theory that undermines the correspondence theory. First, efforts to find meaning depend on “a massive suppression of meaning” and second, no rational argument is immune to critical scrutiny. According to Derrida’s (1997, 2005) theory, meaning of words depend on binaries, i.e., the meaning that one wants to bring into focus always depends on the second binary, namely, its difference. Meaning is a matter of presence and absence. For example, against the meaning of white is absence or the differentiation from non-white (or black). However, the absences as the second pole of the binary are usually suppressed, but as Gergen says, without these, the current meaning cannot exist. The suppression of the meaning of the binary takes place to focus on the actual meaning of words and concepts. In this sense language involves meaning suppression.

72

4 The Metatheoretical Opposition: Positivism and Constructionism

The second aspect involves the vacuity of any rational foundation. Rational argument is regarded as the basis for making scientific and moral decisions but according to Derrida this confidence in rationality is unfounded. Given his method of difference and deferring, the meaning of any word is in principle ambiguous and not decidable. This fragility underlies rational argumentation— clarifying meaning involves moving within the encapsulated system of language where meaning is constantly shifting (Gergen, 1999, pp. 28–29). Modernist assumptions are thus based on the privileged position of the individual to find truth through the medium of language that we can call epistemological atomism. Epistemological atomism or the primacy of the knowing individual influenced western traditions fundamentally (Gergen, 1999). Cultural imperialism was one of the effects due to truths regarded as universal. Western traditions were seen as superior to other traditions and that led to various socio-political and religious catastrophes. The noble cause of the enlightenment, namely, establishing democracy on all levels of society by giving power to the individual lead to a new hegemony: the constitution of a new class of experts. This hegemony of knowledge established a special domain of knowledge and to enter that domain one has to be initiated: “As scientific communities have grown strong, so they have developed specialized vocabularies, methodologies, modes of analysis and practices of reason. Thus, we confront the emergence of a new ‘knowledge class,’ groups who claim superiority of voice over all others” (Gergen, 1999, p. 18). The emphasis on the individual leads further to the neglect of social relations. Importantly, the important role groups, society and communities play in knowledge production and morality is not acknowledged. Gergen (1999, p. 18) blames the emphasis on individual rights for the neglect of our obligation towards the social: communities need to be sustained for the collective good. Relations in a modernist society are merely instrumental: with the new individualism one is “. . . invited into an instrumental posture toward others . . .” (Gergen, 1999, p. 1847). This is a relationship of power constituted by the distinction between the self and other: what is not the self is the “other.” The other is thus a means to provide the self with fulfilment. Simply put, the other can only mean something as long as the self benefits from the relationship. This “instrumental posture” is carried through to other spheres such as nature: it involves an orientation of exploitation rather than sustainability of which the last is rooted in fulfilling our obligations towards the community. Gergen also explicates the role of language in terms of Wittgenstein’s language game metaphor and life forms. What are the implications of Wittgenstein’s metaphor for social constructionism? According to Gergen (1999, p. 48) “Our modes of description, explanation and/or representations are derived from relationship.” This means that we establish meaning in language only within a relationship with somebody else, and deeper with culture and history/tradition. The solipsism of a self-contained meaning generating language system is broken since language is a game acted out by players. The

Social Constructionism and Its Roots in Idealism

73

individual is thus denied the privileged position of establishing or originating meaning. Maintaining traditions, culture, institutions, and history requires, first of all, relationships and the language emerging from those relationships. Maintaining a ritual, for example, requires relatively enduring relationships and the ability to maintain the ritual’s intelligibility or meaning: “As we describe, explain or otherwise represent, so do we fashion our future” (Gergen, 1999, p. 48). Thus, meaning is co-created and this creation or construction of meaning maintains morality, values or as Wittgenstein put it, life forms. According to Gergen, the past does not determine the future—things that was maintained in the past can easily go to waste if there are no meaning-making relationships to sustain it; traditions and culture need to be continuously reinterpreted to make them relevant: “The same is true of our intimate connections in daily life, in our families, and circles of friendship; we must continuously reconstruct their nature (for example, ‘who we are to each other’) in order to keep them alive” (Gergen, 1999, p. 49). He continues “At the same time, constructionism offers a bold invitation to transform social life, to build new futures” (Gergen, 1999, p. 49). What Gergen calls “generative discourses” focus on not only making the past relevant but on generating new discourse that simultaneously criticises the existing forms of life and that opens up new possibilities. In this way constructionism focuses on the future. Usually when we consider certain alternative issues it is always from within a particular tradition and it is difficult to consider alternatives outside of this tradition as possibilities: For constructionists such considerations lead to a celebration of reflexivity, that is, the attempt to place one’s premises into question, to suspend the “obvious,” to listen to alternative framings of reality, and to grapple with the comparative outcomes of multiple standpoints. For the constructionist this means an unrelenting concern with the blinding potential of the “taken-for-granted.” If we are to build together toward a more viable future then we must be prepared to doubt everything we have accepted as real, true, right, necessary or essential. This kind of critical reflection is not necessarily a prelude to rejecting our major traditions. It is simply to recognize them as traditions—historically and culturally situated; it is to recognize the legitimacy of other traditions within their own terms. And it is to invite the kind of dialogue that might lead to common ground (Gergen, 1999, p. 50).

The Constructionist Image of Science The social constructionist motivation for its view almost invariably takes place against the ills of positivism as the standard view of science. As we have seen, on the one hand, taking positivism as starting point is valid in a sense because social science and specifically empirical sociology and psychology went through a phase during the first half of the twentieth century that largely reflected the orientation of empiricist science. Thus, it was quite valid at some stage to break free from the

74

4 The Metatheoretical Opposition: Positivism and Constructionism

hegemony of empiricist/positivist science, to criticise certain assumptions and develop an orientation that can now be called postmodern and constructionist (Alasuutari, 2010; Michell, 2004, 2011a; Morgan, 2007). However, to continue this line of defence for qualitative/postmodern/constructionist approaches on the other hand, is currently untenable (Cruickshank, 2012; Martin, 2003; Michell, 2003, 2011b; Vinden, 1999). According to Gergen (1982, pp. 7–9) positivist-empirical science has been characterised by a few assumptions, namely, that the major task of science is to construct general laws, the laws should be consistent with empirical fact, and scientific knowledge is cumulative. Although Gergen acknowledges the changed assumptions of Logical Positivism over time and that it was superseded by postpositivism, the basic tenets of what science should do and be remain the same. The following tenets form the received view. (a) General laws should indicate the relationship between observed events enable the scientist to explain why those events occurred. The scientist ought to be able to predict future events with these laws in hand. Gergen’s (1982, pp. 44–57) view of the positivist implications for psychosocial science includes three important principles, namely, predictability of behaviour, universality of behaviour and reduction of behaviour. The predictability of behaviour refers to the quest of finding stable patterns of behaviour and laws of behaviour. However, if the empiricist grants that human behaviour is variable, then to study it scientifically its variability must be minimised. An excellent illustration of this is humans’ ability to innovate when becoming the objects of scientific scrutiny: when a subject is aware of the scientist intentions, she may react accordingly. This is the so-called reactivity problem (Shadish, Cook, & Campbell, 2001). Ironically, the reactivity problem as a threat to scientific validity in the social sciences merely confirms that human behaviour might be very predictable in certain contexts. However, Gergen regards the fact that some behaviours might seem similar or universal as incidental. The universality of behaviour refers to similar behaviour across groups (multicultural), places and times. Any resemblance of actions across, for instance, cultures is ephemeral, because it might be coincidental, but, more importantly, . . . with proper sensitivity one could as well generate evidence for the universality of sitting in the shade, gazing at the moon, and hopping on one foot. The mere location of occasional likenesses does not serve as an adequate basis for establishing universality (Gergen, 1982, p. 50).

Reduction of human behaviour is not possible. It refers to the argument that human behaviour can ultimately be reduced to physiological processes and the body is amenable to physical analysis as a physical system. Gergen’s (1982) objection is that knowledge of the physical system cannot provide the content of behaviour but depends largely on input and processing which cannot be controlled or predicted. The system is so complex and surrounded by such a

Social Constructionism and Its Roots in Idealism

75

complex environment (its history, everyday experience and impinging factors) that control of all variables and responses is not possible. (b) Laws are based on systematic observation of occurring events. Despite Popper’s inversion of the logical positivist verification principle, falsification does not rule out that theory, at some stage, must be tested against the facts (Maree, 1990, pp. 201–203). (c) Gergen (1982) bases the cumulative view of scientific knowledge on what the hypothetico-deductive method implies. According to him it requires that hypotheses are formulated about states of affairs, subsequently the hypotheses are empirical tested and if the verification is successful, the probability of the truth or validity of the statements under examination is increased. By eliminating statements and theories in this way, those that are truthful may be retained and progress ensured. (d) Empirical science must be dispassionate or objective. According to Gergen (1999, p. 91) for the empiricist “the aim of good empirical research is to reflect the world as it is.” The empiricist has access to direct experience of reality and should not have any bias, feelings, or political motives towards her science. Objectivity (or rather the impression of objectivity), depends on rhetoric, namely, using distancing language, language that establishes authority and unemotional, i.e., distant or detached language. Objectivity as a way of talk is a functional necessity—it depends on the particular scientific community. Gergen says something interesting, namely, that persons would not be able to walk on the moon had the space agents not talked in a particular way because the language of science establishes trust and communicative ability: “They are ‘calling a spade a spade’ in terms of the community’s standards, and as a result humans can walk on the moon” (Gergen, 1999, p. 76). He continues “The rhetoric of the real may be essential to effective community functioning; problems result primarily when the internal realities are treated as universal or ‘really real’” (Gergen, 1999, p. 76). Against the belief that empirical science must be dispassionate, the constructionist responds that the scientist does research for a reason: “It is disingenuous to cloak these investments in the language of neutrality” (Gergen, 1999, p. 91). Gergen (1999, p. 47) creates the impression that the so-called dispassionate nature of science negates the actual interests scientists have but I think he is confusing objectivity with objectivism. One should indicate why objectivity is a necessary prerequisite for doing science, at least from a realist perspective. In fact, the constructionist demand for critical enquiry establishes objective distance. The way Gergen (1999, p. 91) poses the issue, namely, that social constructionism grounds passionate relations between people and the scientist can thus not be distancing and dispassionate, confuses two types of relations. Two further aspects are indicated, namely, that we would not like to live with scientists that believe the best way to get results is control and manipulation, and results gained from controlled studies only serve those in power. I am not denying that there is research that keeps institutions, companies and even governments in power but stating that science as a social enterprise is

76

4 The Metatheoretical Opposition: Positivism and Constructionism

authoritarian is going too far, although the claim might apply to institutionalised forms of science. As Gergen acknowledges, any research can be interpreted in more than one way! I would like to know how constructionist research refrains from supporting power structures—is it in some way exempt from being recruited for socio-political, unethical or economic purposes? Ironically, the fact that research can be framed and contextualised by whomever is applying research to inappropriate situations, emphasises the requirement for objectivity! Taking this argument from the other side: if a researcher doing research for a tobacco company with well-controlled research found that smoking does not cause cancer, fully and shamelessly exposed her bias, namely, that the aim was to support the tobacco company’s sale of tobacco, then the researcher actually complied to the constructionist requirement of not remaining dispassionate. (e) The empiricist must have control over conditions and uses the model of cause and effect (Gergen, 1999, p. 91). In social sciences behaviour can be deterministically divided into cause and effect. The initial conditions can be specified and the eventual behaviour predicted. Against this issue of the control of conditions, the constructionist holds that cause and effect is a social construction. It is read into observations. The division of world into events, precise segments (as Gergen puts it) or discrete actions “. . . is again an a priori commitment, a fore-structure of understanding as opposed to a derivative as it is” (Gergen, 1999, p. 91). This statement is very contentious: in the first place what does he view as “a derivative of reality as it is”—how does the constructionist know reality as it is? The constructionist purports to be antirealist and reality cannot be known without construction or interpretation: the human mind always enters into any attempt to know “reality.” The second aspect that Gergen emphasises is the implication of the cause-effect model, namely, that human agency is thereby negated. “People are by implication reduced to robots . . .” (Gergen, 1999, p. 92). He follows with the following: “It is not that constructionism holds agency to be true, but there are very good reasons for nurturing this vulnerable tradition” (Gergen, 1999, p. 47). This view of cause and effect and determinism is false: the realist does not hold a regularity concept of causation but a deep one as we will see in the next chapter. The fact that underlying mechanism or other events, agents or processes cause changes in things does not imply determinism. Linear, singular and one-directional causality are simplifications based on the mechanistic and atomistic view of billiard balls and levers that cannot be applied to all of reality (Bhaskar, 1978). (f) The scientist uses quantification as a precise language (Gergen, 1999, p. 92). The empiricist knows that language is not precise and the aim of science is precision. The assumption is that numbers are neutral (thus complying to objectivity). The scientist can statistically analyse observations when they are converted to numbers. In this way cause-effect can be determined and, consequently, predictions can be made.

Social Constructionism and Its Roots in Idealism

77

Against the quantitative imperative the constructionist holds that number is merely a different way of representing. Numbers reduce and eliminate the personal and subjective. Moreover, statistics is an expert language only for the initiated—it functions as a silencing device! This is a false view of quantification. As one of the languages and methods (for example, statistical analysis) of science it provides epistemic access to certain parts of reality, not all. (g) For the empiricist the aim is to find the one true answer from among many opinions (Gergen, 1999, p. 92). Against the search of the scientist for the one truthful answer the constructionist says that there are multiple interpretations possible: how can we choose only one true voice? The others are thereby silenced! The realist denies that there is one absolute truthful answer because someone else might make a different claim. This opens up the way for criticism and debate. (h) Truth and practice must be separated (Gergen, 1999, p. 93). For the empiricist the aim is to provide ahistorical and widely applicable laws that can be applied in all circumstances. Against the tenet that truth and practice must be separated the constructionist says that theories cannot be proven true or false by data. The scientist’s view of what data is comes from the theoretical standpoint in any case and this theoretical standpoint is formed by the community the scientist stands in—Gergen thus typifies the search for universal laws/theories as “cultural imperialism” (Gergen, 1999, p. 93). This issue relates to how we view truth and our ability to state truth with certainty (Bernstein, 1983). This point is based on a false image of science: fallibility is essentially part of realism as is a critical view of universality. Underlying the tenets of the received view is an assumption of the stability of the empirical world, i.e., that regular events or patterns of events occur. The assumption of the scientist is that there is a stability to be found and discovered, even in human behaviour. Gergen (1982) maintains that one of the fundamental differences between natural and social science is the extent of stability. While the former enjoys relative stability, the latter far less so, and patterns of behaviour are not reliable, replicable or enduring. Hence, human behaviour is relatively free of stimulus-conditions as opposed to more simple life forms. Social and psychological phenomena are influenced by contextual and temporal factors—in fact any semblance to universal behaviour or invariable behaviour is coincidental and not really universal, atemporal or invariable because human psychosocial behaviour is essentially contextual, variable and temporal: . . . most would probably agree that the vast share of human activity is not genetically programmed, that human conceptual capacities enable people to generate many new and different understandings of the world, and that prevailing investments in such values as freedom, unpredictability, and uniqueness may function as sources for counternormative behaviour (Gergen, 1982, p. 43).

Thus, the quest for laws or stable patterns of human behaviour is not based on sure footing and can never yield adequate knowledge. Human behaviour is capable of creative and new behaviours and not easily moulded into patterns.

78

4 The Metatheoretical Opposition: Positivism and Constructionism

In terms of freedom of behaviour versus patterned events the subject matter of social science differs fundamentally from that of the natural science. Secondly, because humans are capable of interpreting their reality symbolically, the infinite variation possible in this process works against the ability to predict behaviour: “The scientist’s capacity to predict is precipitously dependent on the conceptual proclivities of the population under study” (Gergen, 1982, p. 17). In this instance, the population under study is creative, able to find novel solutions and behaviours and fundamentally unpredictable. Gergen (1982) finds three aspects within human nature that militate against patterning and predictability. The first is a person’s quest for freedom—the more one tries to shackle human beings the stronger the move towards freedom becomes. Similarly, Gergen finds a strive for uniqueness and a need to differentiate oneself from others as fundamental in some cultures. In the last instance, he acknowledges that a society requires some form of predictability to function properly, but where there is social pressure to conform and behave predictable, the individual will be the first to try and be unpredictable. Why can the researcher not be separated from her object of study as in positivistic concepts of science? Gergen says that this typical view of science requires a gap between scientist and phenomenon and points out that, for instance, the astronomer’s theory does not change the path of the planets and position of the stars (Gergen, 1982, p. 22). However, as he correctly points out, there are at least two aspects involved in doing science that involve a reciprocal influence between observer and phenomenon. First of all, the theory-laden nature of perception, and secondly, the value-laden nature of human conduct influences the way we do science. Reality does not present itself to the observer in ready-made categories of understanding. This implies that despite the so-called position of non-influence of the scientist, her categories of understanding divides reality in slices dictated by the theory, history and culture/community of particular groupings of scientists. In this sense then there is no objectivity because reality is viewed through whatever lens the scientist holds up to her phenomena. Furthermore, Gergen (1982) shows that this lens has an important impact on the scientific community but also the social community within which her science is practised. Our concepts have implications for the way we act. Thus, the way the scientist segments reality and the explanations she provides invariably colours her own research programme but also the behaviour of the consumer of scientific knowledge. The second aspect is the value-laden nature of human conduct. The scientist is required to remain dispassionate, i.e., not to allow emotions, interests and values to enter the practice of science. However, as Gergen (1982) again correctly argues, this is impossible: (a) the segmentation of reality often harbours interests, values and intentions not always conscious at the time but influenced by the socio-cultural climate—some distinction can lead to oppression of minorities, others to systematic blindness to relevant issues. Gergen uses the example of social theories of conflict that has been developed that were influenced by

Social Constructionism and Its Roots in Idealism

79

scientists’ fear of nuclear warfare. (b) Terms developed and used in a research programme can subtly conceal bias, such as a term in a theory of aggression that spurns the aggressor (Gergen, 1982, p. 29). (c) Along with segmentation and terminology, explanation also can include bias, for instance, individualism perpetuated in social theories. Finally, (d) the process of how science is conducted can also harbour interests and bias. Gergen uses the interesting example of traditional psychology that entertains a Humean conception of causality and a positivist view of observation. According to Gergen (1982, p. 31) a Humean/positivist view of science aims at “. . . locat(ing) reliable patterns of contingency among observable events.” Consequently, psychology has to find links between stimuli, which is what the erstwhile behaviourism did. The implication is that if persons are stimulus-response organisms then one can control behaviour by controlling the stimuli conditions. Psychological research based on these assumptions encourage a view of persons as “manipulable automaton(s)” (Gergen, 1982, p. 32). According to Gergen constructionist criticism of the received view of science is aimed at re-establishing the lost democracy that was the noble aim of enlightenment in the first place: “Removing the mantle of scientific authority and fostering democratic participation has been a chief aim of constructionist inquiry” (Gergen, 1999, p. 52). The exclusive domain of the high-priests of science needs to be broken open in order to allow the populace in; democratic participation in science is the aim: “The point of this discussion is not to undermine scientific efforts, but to remove their authority and to place them into the orbit of everyday scrutiny” (Gergen, 1999, p. 51). Gergen agrees with the general belief of social constructionism, namely, that scientific knowledge is socially constructed which means that what is regarded as facts, are generally agreed upon within a particular community of scientists. However, according to Gergen (1999, p. 55), the outcome of social determination can be too extreme as the statement “Scientific knowledge is nothing but social convention” shows. He refers to Latour and Woolgar’s (1979) exposition of how scientific fact gets established: scientists conscript others to defend their so-called factual findings and ward of negative criticism. Enough support in the social network of scientists establishes a fact as true. Scientific fact thus emerges from conscripting support in a network of interacting influences. The problem with this emergentist theory is that each time one confronts a factual claim, one is referred to other sources much like opening a series of Russian dolls: the fact cannot be isolated and identified but consists of this web of “interlocking arrangement of assumptions, equipments, writings, and so on-in effect, an entire tradition or form of life” (Gergen, 1999, p. 57). In the end the view of the scientist as the “new high-priest” is strengthened by means of keeping the uninitiated away from this privileged knowledge. What do the ignorant populace then do with a scientific fact such as smoking causes cancer? If Gergen’s purpose is to devolve epistemic authority to the populace, what do they do when confronted with such a claim? Gergen (1999, p. 57) hovers between saying that this fact stays a mere construction and that one should

80

4 The Metatheoretical Opposition: Positivism and Constructionism

accept it as the truth—he actually says one should pay attention to a fact like this and adjust our lives accordingly because such a proposition has “functional value” in the light of the value of “what we call ‘life’” (implying that life is also a cultural construction). Gergen (1999) said that constructionists do not deny the usefulness of empiricist studies, only that they need to tone down the claim of truth and then “. . . there would be far less resistance to such work” (pp. 93–94). This statement is surprising in the light of the fact that literature is replete with references from constructionist and postmodernists whom deny the empirical scientist’s right of existence. The issue of fallibility, truth and universality of laws will be touched upon later in this study as I have mentioned above. Realists are constantly battling the positivist straw man fabricated by constructionists validating the latter’s views. The second way empirical research can be modified to become more acceptable is to remove its universal truth claim—truth is local and empirical research must be locally relevant and applicable. Gergen also claims that empirical data can be used as illustrations much as the journalist uses photographs: to “inject life into an idea in a way that helps us appreciate its significance and plausibility” (Gergen, 1999, p. 94). I would like to see the NASA scientist trying to market an idea for failsafe tiles for the Challenger in this way—to inject life! Imagine the empirical scientist working in this way and what disastrous consequences a failed rhetoric would have. Empirical research is ok if its predictions can be applied to social, moral and political issues. Gergen claims that much empirical research is trivial and restricted to pushing buttons and completing questionnaires.

Human Nature Militating Against Positivism Before we come to the main constructionist thrust I would like to recruit for science, we need to discuss the human characteristics that exempts psychosocial science from positivist analysis. The major focus of Gergen’s (1982) criticism of positivistempiricist sciences is based on the changing nature of human beings, the interpretive nature of human reality and the linguistic nature of humanity. (a) Inconstancy of human nature In countering the behavioural scientists’ search for laws, durability and tendencies in human behaviour, Gergen (1982) motivates its impossibility at length by stressing the inconstancy of human behaviour. It is forever changing, free, no two acts are ever the same and human beings have the unlimited capacity to choose, create and change at will. How on earth can one through the aeons of human history find semblances of similar behaviour that can point to certain repetitive streaks in humans? According to Gergen the putative perceived similarities are merely that, namely, hastily drawn conclusions based on coincidental observation. To take an extreme example, if one observes cooperative behaviour between ants and at some stage observe similar behaviour in groups of

Social Constructionism and Its Roots in Idealism

81

humans, does this warrant the inference of a general principle governing groups of organism? Probably not, and the social researcher should indeed heed Gergen’s (1982, p. 50) warning that coincidental observed similarities between groups, theories and cultures (like shade-sitting behaviour) do not constitute laws or tendencies. Granted, Gergen (1982) restricts his examples carefully to (a) the search for universal patterns of behaviour based on so-called similarities between theories and cultures; (b) the construction of models of behaviour (which implies that the ability to construct such a model does not mean the patterns or tendencies are real—recall the ant example above); (c) correlating a large number of variables across groups, populations and cultures where the existence of correlations often are interpreted as indicating stable patterns of cross-group behaviour and characteristics. However, given the empirical scientist’s investigation of universal behaviour, the abstract notion of universal behaviour should correspond to observable instances of such behaviour. There should be correspondence rules connecting the abstract concept and the observable— without such correspondence rules, the concept of universals is without empirical content. Yet, if human activity is undergoing continuous change, if cultures develop novel patterns of action over time, then systems of correspondence rules are constantly threated by obsolescence (Gergen, 1982, p. 55).

However, I think that he overstates the case of human inconstancy, which was less pejoratively termed as changing nature. If no “theme” existed in human behaviour, speech and actions, then it would be impossible to negotiate and maintain relationships. If the wife cannot at least guess the mood of her husband (or the other way around) given certain typical actions, then she walks into a conversation blindly with the knowledge that no background information informing her interpretive stance towards her husband could assist her in communicating. Each conversation will start anew, each meeting between so-called friends and groups would be fresh—there would be no history except unpredictability, nothing to at least build upon. Nations would have to renegotiate each contract anew and in this world of chaos each day would be new, left with nothing durable or any promise of something better in future. My contention is that total inconstancy erodes the very heart of constructionism, namely, the ability to construct relationships. Gergen (1982, pp. 11–12) calls for a radical division between the phenomena of natural and social science, even though he acknowledges that social sciences cannot be devoid of any stability. The division is based on the more or less stability of their phenomena in the natural and social sciences. According to Gergen (1982, p. 13) the continued failure of the human and social sciences to provide any form of law or invariable principle for human behaviour ought to prompt us to re-examine the nature of phenomena under study. Social and psychological phenomena are qualitatively different than natural science phenomena and the fundamental difference lies in their suitability to be described with laws or

82

4 The Metatheoretical Opposition: Positivism and Constructionism

principles at certain levels of abstraction: social and psychological phenomena cannot be described thusly; they are fundamentally unstable and variant. (b) Interpretative nature of human reality Human reality as described and patterned through the eyes of human beings is equivocal. According to Gergen (1982) there are only multiple interpretations of behavioural acts. These interpretations depend on context or history. Context is determined by the stream of historical events within which a particular target of interpretation is embedded in. However, its embeddedness is also circumscribed by the sociocultural context. Although Gergen stipulates the contextual factors quite well, he is careful not to say that context limits interpretation but instead that it opens up limitless possibilities. More specifically, an interpretation can constantly be revised in the light of past and present events. Interpretation also varies between cultures. Indeed, culture, history and context determine how we react and feel: “. . . the assumption that we privately think, feel, desire, intend, and the like is not demanded by ‘what there is’ but is essentially optional” (Gergen, 1999, p. 117). This means that we are socially, culturally or systemically constructed or how ever one would like to put it. He calls it “the ideology of the self-contained individual” (Gergen, 1999, p. 118). This leads to a gulf between individuals: I cannot know the real you. From the individualistic perspective relationships are artificial whilst the individual still stays the primary entity. “When the self is the essential atom of society, we find invitations to isolation, distrust, narcissism, and competition . . .” (Gergen, 1999, p. 122). How we think about ourselves is culturally and socially determined and language or narratives play a fundamental role in forming these beliefs about ourselves. Gergen (1999, pp. 66–68) shows that metaphor is used to construct images of the self. For example, metaphoric speech is used to speak about the mind, thought and emotions. In the case of the mind and subjective experience a metaphor of inner/outer prevails which reminds one of a container metaphor of the mind. Gergen (1999, p. 67) refers to examples such as “what’s on my mind” and “in my thoughts.” What is inside can be distinguished from what is outside. The self and what the self believes about the real (the world) and the good (values) are created socially (Gergen, 1999, p. 81). Ontology and ethics are not so much dependent on what there is but on what is constructed and only later become “solidified” in order to seem like something objective and durable. A self is constructed within a relationship with others: as Gergen (1999, p. 82) says, who a person is and the nature of her actions are established in interaction with others. For example, parents provide a baby with a gender and the actions towards the baby establishes what the child eventually believes about herself. Furthermore, a person’s talk about herself establishes her belief of who she is in terms of a single agent. Relationships can only be established with shared meanings that in turn are established by sharing a language. The shared meanings refer to a shared reality. Shared reality establishes shared ethics or as Gergen puts it, patterns of interaction provide a certain structure to a relationship in terms of actions and

Social Constructionism and Its Roots in Idealism

83

communication. Breaking the patterns of actions and communications is disruptive and therein lies good or bad (Gergen, 1999, p. 81). The construction of the self takes place in relationships contra the modernist view of a self-contained, rational individual (Lock & Strong, 2010, pp. 300–303). One of the problematic views of the concept of individual is the division between inner and outer. The construction of the senses as the receptors of outside impulses is problematic because the person will remain selfencapsulated. Gergen points out that the way we perceive are socially constructed within a locally bound community. Wine tasting is an example. However, this shifts the problem because constructing what counts as good wine, tastes, smells, and paintings does not answer the question how the person registers these sensations (Gergen, 1999, pp. 102–104). The point is emphasised with Gergen’s (1999, pp. 104–105) discussion of pain. According to him there cannot be a universal phenomenon beyond construction because pain throughout history has been viewed differently—from Christian views that pain purges the sinner to masochistic invitation of pain by the boxer. However, despite the various social constructions of pain, there remains a phenomenon that elicits a certain response in human beings and, for instance, in cultures where pain must be stoically endured, this very expression gives the nature of pain away: “endured” implies an initial avoidant reaction that must be revised. Gergen should also explain the fear and anxiety elicited when a child is repeatedly and severely beaten—of course, the child might decide to endure and survive no matter what and in this way grows stronger and callous but it does not negate the hurtful nature of pain. Emotions are unduly substantialised in western psychology. Gergen (1999, pp. 107–111) tries to illustrate the futility of this exercise by providing a number of cultural examples where emotions are given labels that denotes feelings for which we do not have words. Consequently, there is cultural variation in what is experienced and how it is named. What the westernised person regards as primary emotions is not necessarily reflected in other cultures. Even history shows that names for emotions do not even apply today—does this mean that that emotion ceased to exist (an example is “curiosity”)? Does this mean that experiences, feelings and emotions come into being and that they differ when the context changes? Or does it merely mean that language describing those variations of emotions capable by human beings differs? It might well be the case that a strange culture has words for a phenomenon or emotion experienced by a community, but that does not mean that it cannot be experienced or understood by an outsider. I come back to the point of different emotions existing for different groups of people. The way the issue is posed presumes the reality of emotions experienced in the sense that an individual or group of people experience. Does this imply substantiality of the emotion (namely, that it is a thing)? I think from all the human states ranging from consciousness to emotions one could agree that emotions are the least likely to be understood as a substance.

84

4 The Metatheoretical Opposition: Positivism and Constructionism

One useful answer that Gergen gave that opens up different avenues for research is that in opposition to emotional-realism, emotions should be viewed as a way of acting: . . . so it is with the entire range of emotional performances; they are ways of being, historically provided ways of life. We need not measure them, wait for them to move us, or be concerned that we feel more that is “natural”. Rather, like forms of dress or moves in a game or a dance, they are vocabularies of action, and life will be filled or emptied according to how we press then into action (Gergen, 1999, pp. 111–112).

(c) Establishing the relational nature of self A fully relation view of being is necessary in order to establish the social constructionist view of persons. Thus, al vestiges of the individual as retaining control or the remaining the originator of acts: an expression such as “I think” should constitute relational being and not individual being (Gergen, 1999, p. 131). Gergen (1999, pp. 131–132) motivates the primacy of relation to the individual self from the performative nature of language—when one says something like “I love you,” it is not the expression of a private and subjective thought but the expression of an act that influences a relationship. However, the performative quality of language is not sufficient to establish the primacy of relational being since one merely acknowledges that spoken language is somehow embedded in relationships. Thus, examples of verbalised thought that presumes a dyadic relationship falls short of illustrating the relocation of being from single individuals to things that can only exist in multiplicity. The next step for Gergen is to acknowledge the embeddedness of language and actions within a cultural history. As Gergen (1999, p. 47) says, acts and words make sense by carrying a tradition of meaning with it: “. . . when I perform I am carrying a history of relationships, manifesting them, expressing them. They inhabit my every motion” (Gergen, 1999, p. 133). Thus, relationships, i.e., those that formed the meaning of words and acts, inform current words and acts (performances). In a converse manner performances are intended for an audience, thus, an utterance such as “I think” has a recipient in mind. The question is whether the formative and informative nature of performances is sufficient to establish them as relational. According to Gergen (1999, p. 47), they are: We now find that one’s performances are essentially constituents of relationship; they are inhabited not only by a history of relationships but as well by the relationships into which they are directed. By making these two theoretical moves, treating psychological discourse as performative and embedding performances within relationships, we are now positioned to see the entire vocabulary of the mind as constituted by and within relationship (Gergen, 1999, p. 133).

The dissolution of the individual originating mind is now final: “. . . there is no independent territory called “mind” that demands attention. There is action, and action is constituted within and gains its intelligibility through relationship” (Gergen, 1999, p. 133). The jump from subjective to social actions is as follows: “Thus, we may be doing something privately—which we might want to call

Social Constructionism and Its Roots in Idealism

85

reasoning, pondering, or feeling—but from the present standpoint these are essentially public actions carried out in private” (Gergen, 1999, p. 134). The examples he uses to emphasise the formative function, for example, memory is seen as a relational phenomenon. It is “not an individual act but a collective one” (Gergen, 1999, p. 134). A simple example is a child trying to recount a movie and getting some facts right and others wrong. Luckily, her mother is present to correct “facts.”

Critical Resistance One consequence of a strong version of SC is a thoroughgoing relativism and usually the clash between constructionists and realists and/or empiricists concerns the extent to which an external reality is real. As we have seen the mythic image of science includes this idea of an external and unitary reality, usually implying a concrete, observable and thus empirical reality. But this is not the reality a realist refers to. The nature of the realist concept of reality is discussed in the next chapter but suffice to say is that the realist’s reality is not confined to what can be observed. However, the constructionist, or rather the strong version of constructionism, makes reality wholly dependent on the mind. If we take Gergen’s claim about gravity at face-value then it means that gravity does not exist except as a construct in the minds of some human beings and scientists. In principle it means that the gravity that we think exists only exists because people formed a conception about it. However, is this what Gergen has in mind? We saw in different quotations above that he acknowledges that NASA sent people to the moon by using a common language, calling a spade a spade and not a lollipop. Does this mean that some parts of reality for the constructionist is a bit more solid or out there then other parts? It seems as if certain facts of life are a bit more real than others, and with “a bit more real” I mean somewhat independent of the human mind. See what he says about illness and death (although we already know that he took his kids to a modern medical professional): If illness is simply illness, then we might consign our bodies to the care of the medical professions and be done with it. Yet, as Frank demonstrates, there are other stories to be told under such conditions. In particular, he points to the advantages of a ‘quest narrative’, in which the individual uses the occasion of the illness to embark on a quest—a movement eventuating in a deeper understanding of self and society. One may see the possibility for a change of character through suffering, an opportunity to speak for those who suffer, to bear testimony and thereby contribute to the greater wisdom of the community. In the same way, if material death is simply material death, and that’s the end of it, what an impoverished world of meaning we inhabit! If death can also be a rejoining with lost loved ones, a place at the right hand of God, a movement in an endless cycle of becoming or an ultimate tranquillity, our world is enriched: cultural life is more inhabitable (Gergen, 2001, p. 42).

Gergen would like to retain the constructionist link between illness/death and the mind. Humans cannot but assign meaning to experience. Illness and death do happen but always within a fundamental network of meaning. However, his is not a strong version of constructionism when certain mind-independent realities such as illness

86

4 The Metatheoretical Opposition: Positivism and Constructionism

and death is involved. By granting NASA their success story, taking kids to a medical practitioner, mourning the death of a loved one within a religious or other cultural context, do not make these realities any more dependent for their existence on minds. This is a weak version of constructionism a realist can live with and even embrace as we will see critical realism did in the next chapter. In this weaker version of constructionism we need not be too concerned about relativising ontology, i.e., making reality dependent on mind, but we can embrace a relativistic epistemology. This relativistic epistemology to my mind is crucial to the scientific enterprise. I do not think Gergen would agree with this interpretation but let us take his point further. In a statement that almost sounds realist Gergen (2001, p. 42) says: “We are prompted to explore alternative understandings of ‘what is the case’, and to locate meanings that enable us to go on in more adequate ways.” The reader will see in the next chapter’s discussion about realism that making claims about the way things are is an important realist strategy. By making a claim about something I firstly assume that it is sufficiently real for me but also to others to enable them to refute my claim or make counter claims. However, Gergen puts “what is the case” in quotation marks well knowing this is what a realist would say, meaning that someone would claim mind-independence for a particular state of affairs. This is where social constructionism features at its best. It can and does contest any claim: The constructionist resistance continues: such terms as ‘real,’ ‘true,’ ‘rational’ and ‘objective’ possess deadly potentials. To be sure, they are enormously useful within particular communities for affirming traditions, facilitating mutual trust, ensuring forms of coordination and generating collective enthusiasm. However, they are fraught with danger when participants in such communities extend what is local to the plane of the universal—real for all people, transcendentally true, fundamentally rational, indisputably objective (Gergen, 2001, p. 42).

The power of language is such that it can make certain things true for a number of communities. I would like to venture the claim that as soon as beliefs resist attempts to dismantle it, it is already mind-independent in a sense. This will also be discussed in the next chapter but the force, for example, a timetable exerts on its followers, demonstrates its realness. What Gergen says above is that certain concepts such as truth, objectivity and universality can exert a very real force especially when we start transgressing local boundaries. A good example would be the psychologist doing empirical research and who needs to identify variables for her study. Gergen (1982, pp. 34–41) showed that the choice of constructs to be examined can be totally arbitrary. Identification of independent, dependent and intervening/moderating variables and successful testing thereof does not necessarily lead to valid conclusions. When identifying dependent, independent and intervening variables, many more may be added by means of careful thought based on previous research and conceptual analysis. Although not explicitly stated, the ability to generate a number of possible variables impacting on the problem at hand either as independent and confounding variables should caution the researcher of committing, what I shall call, the entity fallacy. The mere fact that others and one self are able to think of putative constructs does not imply that such constructs, processes and relations exist—the ability to think of unicorns does not guarantee their substantiality as

Social Constructionism and Its Roots in Idealism

87

entities. Gergen’s (1999, p. 13) fallacy of misplaced concreteness is related to what I have in mind as the entity fallacy. In his case, the fact that we use words such as thinking or intent leads to words such as thought and intention. It is a small jump to attribute concreteness to these words, i.e., “I had a thought” implies there is a concrete object in the head. This fallacy focuses on the way we use language and as such alerts us to carefully examine the way we speak about constructs. It also warns against a too glib adoption of folk psychology into our psychological theories. The entity fallacy, on the other hand, warns against confusing thought and concreteness. Gergen (1982, p. 38) reports the first problem as entity frequency. He assumes the constructs exist as entities in a particular community. Given that it does exist it is prone to the problem of entity frequency, namely, its relevance depends on the frequency of its use in a community. Entities change in their relevance over time. For example, self-image might be important at certain stages in a culture’s history. Gergen’s point is that these entities are not universally relevant and prone to change depending on the local context and not so much whether they actually exist or not. The relationship between dependent and independent, independent and intervening, and intervening and dependent variables not only change over time but are also subject to brief and rapid changes. The point is that one cannot assume the universal applicability of the relationship per se due to their local nature. For example, attitudes towards sexuality changed dramatically over time in certain western countries: a gender-attitude relationship with other variables would look different over 50 years in the United States. Thus, one the one hand, from Gergen’s SC we have an epistemological relativism that might fit the realist image of science, and on the other hand, we do have an ontological claim: speaking, thinking and in general minds, create things that are real. The nature of these real things is explored in the next chapter but we know that communities may regard certain things as real such as timetables, rituals, traditions, religious beliefs and so on. These things might exert real power over people but we need to realise that these things might have been otherwise. Along with the epistemological and ontological insights we would like to harness SC’s critical thrust of unmasking what communities believe is real. Let us look at a rather long section from Gergen (2001) to probe the nature of this particular skill of unmasking. The clash between social constructionism and its alternatives can be dealt with in two ways in the academic context, namely, by “waging the war” in what Gergen (2001, p. 13) calls protected enclaves. Thus, by isolating each other and dealing with the arguments internally and locally, in these academic circles protected by like-minded colleagues, it is not necessary to confront the other but by means of this “self-satisfying and self-affirming option” dismiss the standpoint of the enemy. This has been the way of dealing with opposing voices in academia all too often. However, (Gergen, 2001, p. 14) continues, “There is a second option animated by the existing animus: argumentative confrontation.” Again as scholars we are fully prepared for this ritual, and in certain respects our mode of conduct contributes to the society at large. We demonstrate the potential for replacing armed combat with an exchange of words, extending the resources for peaceful resolution of

88

4 The Metatheoretical Opposition: Positivism and Constructionism conflict. Thus, we can pursue rational debate, drawing the battle lines more sharply and clearly in hopes that, over time, the superior paradigm will win. We avoid the torpor of solidification and move toward a superior level of comprehension—or so it is supposed. I have severe doubts about the potential of traditional argumentation to yield compelling solutions in such cases . . . However, there is one major flaw in the process of argumentation specific to the present case: both realism and constructionism are lodged within differing presumptions about the nature of knowledge, reason and value. Resultingly the positions are inherently incommensurable. They cannot properly be compared within the terms of either position, because the very presumptions from which each argues automatically foreclose on the alternative intelligibility (Gergen, 2001, p. 42).

Gergen finds argumentation rooted in human beings’ tendency to extend the local beyond its boundaries and conflict is thus natural (Gergen, 2001, p. 42). In other words, as soon as local traditions are universalised and objectified it serves to suppress and destroy other local traditions and cultures. Of course, this colonial tendency is based on the arch binary informing every universalising tendency: good against evil. Someone always believes he or she is on the right side of the binary and this belief, along with power interests, lead to war, i.e., to annihilate the opposition. Thus, war underlies argumentation and especially scholarly debate: “Failure to participate in this drama of good-over-evil is to miss out on one of the most acute pleasures of scholarly life” (Gergen, 2001, p. 42). Gergen calls this behaviour selfgratifying and self-satisfying and serves the fundamental constitution of differences between opposing groups and opinions. The opponent’s view has to be obliterated in order for the protagonist to stay in power along with the so-called correct or true paradigm. Gergen locates argumentation in the self-serving pleasure of dominating over others. Thus, argumentation cannot lead to agreement and especially not in academia and science: Gergen (2001, p. 42) even points to Kuhn’s statement that paradigmatic strife can usually only be settled by “the expiration of the opposition”! Gergen takes a dim view of the possibility of argumentation to facilitate understanding and dissolving of differences between what he calls competing paradigms and in this case realism and constructionism. This despite using argumentation to criticise realist, empiricist and positivist views. Gergen (2001, p. 14) calls these two ways of dealing with opposition “passive and active modes of antagonism.” However, one should not be misled by his deconstruction of traditional forms of argumentation and critique precisely because the structure of what he views as crucial constructionist deconstruction remains the same. Even though he sets up the argument against argumentation in a specific manner, it is clear that he does not want to rid SC of the basic structure of argumentation. In the one context of largely enemies declaring their own truths and blind devotion to their truths, it is understandable that we should frame the mode of confrontation as modes of antagonism. Using her fundamental skill of critical resistance, the constructionist points to the ontological power of language at work when framing argumentation as academic antagonism in whatever form. The core of SC is essentially (sic) one of radical criticism or the unmasking of difference which I contend is the hallmark of realist science. In the long quotation above Gergen (2001, p. 42) says: “We demonstrate the potential for replacing armed combat with an exchange of words, extending the resources for peaceful resolution of conflict.” It points to a similar sentiment Popper

Social Constructionism and Its Roots in Idealism

89

(1966, p. 25) expressed years ago: “Our schema allows for the development of erroreliminating controls (warning organs like the eye; feed-back mechanisms); that is, controls which can eliminate errors without killing the organism; and it makes it possible, ultimately, for our hypotheses to die in our stead.” Gergen expressed the handling of differences as antagonistic argumentation rather than killing each other literally, whilst Popper regarded handling difference properly as necessary for survival, i.e., the same structure of managing difference found in survival can be found in the practicing of science, namely, conjecture and refutation. All organisms encounter problems, they find tentative solutions, test some of them and if they survive go on to the next round of conjecture and refutation. People (and scientists) have the fortunate ability to let their hypotheses die when disconfirmed instead of themselves. Gergen does have a proposal on how to deal with the differences between people (in this instance, between realists and constructionists). First, it must be acknowledged that opposing camps (again, between realists and constructionists although our interest is in the critical structure of constructionist resistance) both use language to pose their opinions, i.e., they utilise discourse and this fact enables one to find a discursive solution (Gergen, 2001, p. 15). Gergen (2001, p. 42) mentions three avenues that he regards as not so natural as the usual conflictual one: (a) shifting the discursive register, (b) separating expression from personal essence, and (c) exploring polyvocal potentials. With the first he means that one should shift into a meta-mode where one considers both the opposition’s and one’s own standpoint—very similar to standing in the other’s shoes (Gergen, 2001, p. 20). In a situation of seemingly irreconcilable conflict such as in a marriage, it takes a major effort to see your own standpoint through the eyes of your partner’s. Of course, one needs not agree with one’s partner’s standpoint but looking back empathetically, through eyes of the other towards oneself enables real understanding to take place. The moment one can say “I can fully grasp why the other person says and acts as she does,” common ground is established and a softening of one’s own resolute standpoint takes place. It is this willingness to empathetically climb in the skin of the opposition that melts irreconcilability. Secondly, to separate expression of opinions from the person expressing them means exactly what Popper intended with hypothesis dying instead of people: do not shoot the messenger, avoid ad hominem attacks and realise that behind the argument and opinions expressed there is a human being living in multiple contexts. The person vehemently defending positivist values is also a grandfather and a loving spouse. This is what Gergen (2001, p. 21) means by exploring polyvocal potentials. By imagining my opponent within multiple contexts allows me also to not kill the messenger. “One important outcome” according to Gergen (2001, p. 21) “of separating words from persons is that we reduce tendencies toward the psychological essentializing of evil.” In addition, exploring polyvocality is a good principle but we are also reminded that some of the most notorious criminals in history were also loving family members. I wholeheartedly support Gergen’s proposals for conflict resolution. However, the basic structure of difference between things and people remain. That there is

90

4 The Metatheoretical Opposition: Positivism and Constructionism

difference and because people for whatever reason will continue to differ, oppose and absolutise their own truths, it is important that constructionism retains its fundamental critical resistance. However, as the next chapter shows, difference, and hence our ability to make claims and counter claims, i.e., criticise, is fundamental to a realist epistemology and ontology.

References Alasuutari, P. (2010). The rise and relevance of qualitative research. International Journal of Social Research Methodology, 13(2), 139–155. https://doi.org/10.1080/13645570902966056. Ayer, A. (1936/1971). Language, truth and logic. Middlesex: Penguin. Bernstein, R. J. (1983). Beyond objectivism and relativism: Science, hermeneutics, and praxis. Oxford: Blackwell. Bhaskar, R. (1978). A realist theory of science (2nd ed.). London: Verso. Bunge, A. M. (2017). Doing science: In the light of philosophy. New Jersey: World Scientific. Chalmers, A. F. (2013). What is this thing called science? (4th ed.). St Lucia, QLD: University of Queensland Press. Chernoff, F. (2012). The impact of Duhemian principles on social science testing and progress. In H. Kincaid (Ed.), The Oxford handbook of philosophy of social science (pp. 229–258). Oxford: Oxford University Press. Colman, A. (2008). Social constructionist psychology. A dictionary of psychology. Retrieved from http://0-www.oxfordreference.com.innopac.up.ac.za/view/10.1093/acref/9780199534067.001. 0001/acref-9780199534067-e-9399. Creath, R. (2017). Logical empiricism. The Stanford encyclopedia of philosophy. Retrieved from https://plato.stanford.edu/archives/fall2017/entries/logical-empiricism/ Cruickshank, J. (2012). Positioning positivism, critical realism and social constructionism in the health sciences: A philosophical orientation. Nursing Inquiry, 19(1), 71–82. https://doi.org/10. 1111/j.1440-1800.2011.00558.x. de Saussure, F. (1959). Course in general linguistics. New York: Philosophical Library. Derrida, J. (1997). Of grammatology (G. C. Spivak, Trans. Corrected ed.). Baltimore: Johns Hopkins University Press. Derrida, J. (2005). Writing and difference (A. Bass, Trans.). London: Routledge. Devlin, W. J., & Bokulich, A. (Eds.). (2015). Kuhn’s structure of scientific revolutions – 50 years on (Vol. 311). Cham: Springer. Duhem, P. (1991/1954). The aim and structure of physical theory. New Jersey: Princeton University Press. Friedman, M. (1998). Logical positivism. Routledge encyclopedia of philosophy. Retrieved from http://0-www.rep.routledge.com.innopac.up.ac.za/article/Q061SECT3Zahavi Gergen, K. J. (1982). Toward transformation in social knowledge. New York: Springer. Gergen, K. J. (1999). An invitation to social construction. London: Sage. Gergen, K. J. (2001). Social construction in context. London: Sage. Godfrey-Smith, P. (2003). Theory and reality: An introduction to the philosophy of science. Chicago: University of Chicago Press. Hacking, I. (1999). The social construction of what? Cambridge, MA: Harvard University Press. Hacking, I. (2002). Historical ontology. Cambridge: Harvard University Press. Hanson, N. R. (1958). Patterns of discovery: An inquiry into the conceptual foundations of science. Cambridge: Cambridge University Press. Hardcastle, G. L. (2007). Logical empiricism and the philosophy of psychology. In A. Richardson & T. Uebel (Eds.), The Cambridge companion to logical empiricism (pp. 228–249). Cambridge: Cambridge University Press.

References

91

Heller, M. (2011). Philosophy in science: An historical introduction. Heidelberg: Springer. Hume, D. (1962). A treatise of human nature, book I of the understanding. In D. G. C. MacNabb (Ed.), A treatise of human nature (pp. 39–331). Glasgow: Fontana. Kindi, V., & Arabatzis, T. (Eds.). (2012). Kuhn’s “The structure of scientific revolutions” revisited. New York: Routledge. Kołakowski, L. (1968). The alienation of reason: A history of positivist thought. Garden City, NY: Anchor. Kuhn, T. S. (1962). The structure of scientific revolutions. Chicago: University of Chicago Press. Kvale, S. (1992). Psychology and postmodernism. London: Sage. Latour, B., & Woolgar, S. (1979). Laboratory life: The social construction of scientific facts. Beverly Hills: Sage. Laudan, L. (1996). Beyond positivism and relativism: Theory, method and evidence. Boulder, CO: Westview Press. Lock, A., & Strong, T. (2010). Social constructionism: Sources and stirrings in theory and practice. Cambridge: Cambridge University Press. Maree, D. J. F. (1990). Kritiese rasionalisme en teologie. (Unpublished doctoral thesis), University of Pretoria. Martin, J. (2003). Positivism, quantification and the phenomena of psychology. Theory and Psychology, 13(1), 33–38. https://doi.org/10.1177/0959354303013001759. Michell, J. (2003). Pragmatism, positivism and the quantitative imperative. Theory and Psychology, 13(1), 45–52. https://doi.org/10.1177/0959354303013001761. Michell, J. (2004). The place of qualitative research in psychology. Qualitative Research in Psychology, 1, 307–319. https://doi.org/10.1191/1478088704qp020oa. Michell, J. (2011a). The place of qualitative research in psychology. In N. Mackay & A. Petocz (Eds.), Realism and psychology: Collected essays (pp. 678–698). Boston: Brill. Michell, J. (2011b). Qualitative research meets the ghost of Pythagoras. Theory and Psychology, 21 (2), 241–259. https://doi.org/10.1177/0959354310391351. Morgan, D. L. (2007). Paradigms lost and pragmatism regained. Journal of Mixed Methods Research, 1(1), 48–76. https://doi.org/10.1177/2345678906292462. Mormann, T. (2007). The structure of scientific theories in logical empiricism. In A. Richardson & T. Uebel (Eds.), The Cambridge companion to logical empiricism (pp. 136–162). Cambridge: Cambridge University Press. Pandit, G. L. (1991). Methodological variance: Essays in epistemological ontology and the methodology of science. Dordrecht: Dordrecht Kluwer Academic Publishers. Papadopoulou, E., Botton, J., Brantsæter, A.-L., Haugen, M., Alexander, J., Meltzer, H. M., et al. (2018). Maternal caffeine intake during pregnancy and childhood growth and overweight: Results from a large Norwegian prospective observational cohort study. BMJ Open, 8(3), e018895. https://doi.org/10.1136/bmjopen-2017-018895. Parker, I. (Ed.). (1998). Social constructionism, discourse and realism. London: Sage. Polanyi, M. (1958). Personal knowledge: Towards a post-critical philosophy. Chicago: University of Chicago Press. Popper, K. R. (1963/2002). Conjectures and refutations: The growth of scientific knowledge. London: Routledge. Popper, K. R. (1966). Of clouds and clocks: an approach to the problem of rationality and the freedom of man. St. Louis: Washington University. Quine, W. V. (1951). Two dogmas of empiricism. The Philosophical Review, 60(1), 20–43. https:// doi.org/10.2307/2181906. Reisch, G. A. (2007). From “the life of the present” to the “icy slopes of logic”: Logical Empiricism, the Unity of Science Movement, and the Cold War. In A. Richardson & T. Uebel (Eds.), The Cambridge companion to logical empiricism (pp. 58–87). Cambridge: Cambridge University Press. Rescorla, M. (2019). Convention. The Stanford encyclopedia of philosophy. Retrieved from https:// plato.stanford.edu/entries/convention/

92

4 The Metatheoretical Opposition: Positivism and Constructionism

Richards, R. J., & Daston, L. (Eds.). (2016). Kuhn’s “Structure of scientific revolutions” at fifty: Reflections on a science classic. Chicago: University of Chicago Press. Richardson, A. (2007). “That sort of everyday image of logical positivism” Thomas Kuhn and the decline of logical empiricist philosophy of science. In A. Richardson & T. Uebel (Eds.), The Cambridge companion to logical empiricism (pp. 346–369). Cambridge: Cambridge University Press. Richardson, A., & Uebel, T. (Eds.). (2007). The Cambridge companion to logical empiricism. Cambridge: Cambridge University Press. Rodriguez-Pereyra, G. (2015). Nominalism in metaphysics. The Stanford encyclopedia of philosophy. Retrieved from https://plato.stanford.edu/entries/nominalism-metaphysics/ Sayer, A. (1998). Abstraction. In M. Archer, R. Bhaskar, A. Collier, T. Lawson, & A. Norrie (Eds.), Critical realism: Essential readings (pp. 120–143). London: Routledge. Scruton, R. (2001). Kant: A very short introduction. Oxford: Oxford. Shadish, W. R., Cook, T. D., & Campbell, D. T. (2001). Experimental and quasi-experimental designs for generalized causal inference. Boston: Houghton Mifflin. Stadler, F. (2003). What is the Vienna Circle? Some methodological and historiographical answers. In F. Stadler (Ed.), The Vienna Circle and logical empiricism: Re-evaluation and future perspectives (pp. xi–xxiii). New York: Kluwer. Stadler, F. (2007). The Vienna Circle: Context, profile, and development. In A. Richardson & T. Uebel (Eds.), The Cambridge companion to logical empiricism (pp. 13–40). Cambridge: Cambridge University Press. Stadler, F. (Ed.). (2015). The Vienna Circle: Studies in the origins, development, and influence of logical empiricism. Cham: Springer. Stockman, N. (1983). Antipositivist theories of the sciences critical rationalism, critical theory and scientific realism. Dordrecht: Springer. Toulmin, S. E. (2001). Return to reason. Cambridge, MA: Harvard University Press. Vincent, G. (2017). Comte and the positivist vision. In L. C. McIntyre & A. Rosenberg (Eds.), The Routledge companion to philosophy of social science (pp. 7–17). New York, NY: Routledge. Vinden, P. G. (1999). Gathering up the fragments after positivism: Can Ratner make us whole again? Culture and Psychology, 5(2), 223–238. Winkler, K. P. (2010). Kant, the empiricists, and the enterprise of deduction. In P. Guyer (Ed.), The Cambridge companion to Kant’s Critique of pure reason (pp. 41–72). Cambridge: Cambridge University Press.

Chapter 5

Realism in Psychological Science

Introduction In the previous chapters we saw that the mythic image of science is constructed on various levels within particular dualisms. Along the way some of the problematic elements were identified such as what positivism entails, what the essence of constructionism is and that we should distinguish between epistemic access and ontology, namely, what things are. We have also seen that for the social scientist and especially the psychological scientist the danger of creating constructs is especially easy. It is a short step from thinking to concretising. Any hope of clear guidelines of what is true and real is, as we have seen, very elusive and we as humans sit with the conundrum of talking about stuff with other people, an activity that can both free us from narrow strictures or lock us into fundamentalist beliefs. The next step in our argument is to explore how realism could provide us with a conceptual framework within which to view psychological science. The choice for realism is not very apparent and as we have seen in the discussions above realism and empiricism including positivism are frequently confused. Many SCs aim at positivism but call it realism although one can understand their intended target, namely foundationalist empiricism. That said, it must also be acknowledged that the prevalence of the paradigm wars does not make it easy for us to distinguish between different conceptual frameworks. In addition, the fact that realist and empiricist science both work with the empirical, can confuse those sceptical about so-called empirical methods merely because both realists and positivists work with empirical data. One may call this confusion by association. The situation gets worse. Because of their intentional disassociation from empirical natural science methods, postmodern researchers forget that they also have only empirical data to work with. Their inferential movement beyond the given, i.e., empirical data, align them with the realists. Both speculate beyond the given and do not restrict their conclusions to the empirical.

© Springer Nature Switzerland AG 2020 D. J. F. Maree, Realism and Psychological Science, https://doi.org/10.1007/978-3-030-45143-1_5

93

94

5 Realism in Psychological Science

Roy Bhaskar’s critical realism is widely regarded as a viable avenue between positivism and constructionism. He contrasts his own transcendental realism with the transcendental idealism of Kant and in a helpful way manages to focus on the limitations of positivism. The latter restricts the real to levels of what is observable. His explanation of Hume’s constant conjunctions of events also emphasises the difference between positivist and realist conceptions of causality. Scientists of both natural and social varieties aim for causal explanation by identifying causal mechanisms. Both domains of reality are open systems but only the natural domain can be closed by means of experiment according to Bhaskar. Not so for the social and psychological domains: experiment is largely excluded from their investigations and epistemic access is granted by means of language. Thus, Bhaskar’s attempt to naturalise social science ends in failure by ending up with two ontologies. The shortcoming of Bhaskar’s metatheory is contrasted with a discussion on varieties of realism ranging from scientific realism, Chakravartty’s semi-realism, Mäki’s social economic realism, and the lesser known situational realism of John Anderson. Finally, the new realism of Ferraris is discussed. The limitations of the solutions that the previous realisms from Bhaskar to Mäki provide can be overcome by the strengths of situational and new realism. It becomes clear that a realism that addresses ontological and epistemological issues of both natural and social domains of reality requires a naturalist orientation. This means first of all that one needs to work with a homogenous or an egalitarian ontology that assumes that the qualitatively different domains such as the social and the natural come from one reality. The fundamental nature of reality, namely, its ability to posit itself or resist underlies even our human tendency to make claims and counter claims about this reality. Reality’s resistance and our ability to question underlie what we know as science. It is common sense made systematically and coherently critical. Any claim about reality can be countered: this provides the epistemic access to reality which in science becomes a systematic and purposeful process. To clarify how realism differs from the “received” view of science, namely, its empiricist variants, a brief overview of scientific realism is provided. It is the most prominent version of realism in the natural sciences with some applications in social and economic sciences. A version of scientific realism, namely, Anjan Chakravartty’s (2007) semirealism is discussed as a position responding on issues structural and entity realism while accounting for anti-realist arguments. Semirealism explains the reality of relata and relations, and holds that first-order structures are real. Explanatory mechanisms reside in the properties of relata, thus relata cause other things to happen. The ontology of semirealism firmly maintains the mind-independence of things that exist and also acknowledges the realness of unobservables provided they comply with two criteria, namely, the ability to detect them and to manipulate them to do other things (Hacking, 1983). Causality is thus a fundamental principle underlying the reality of unobservable things and their structures. Semirealism also accounts for the stability of core elements in a theory when theories change. However, scientific realism is not usually applied to psychosocial sciences and to date only in a limited sense. Uskali Mäki (2002, 2005a, 2011, 2012a, 2012b) is one

Introduction

95

scientific realist that accepts the tenets of realism and applies it to economic theory. He acknowledges the difference between the realities being investigated in social science, economic theory and natural science. In order to provide a minimal application for realism in social science he proposes relaxing some of the principles of realism such as mind-independence and unobservables. However, he also proposes a particular view of modelling phenomena and systems that might be profitable for explanation outside of the natural sciences. Importantly, his view of modelling is based on the assumption of causality underlying the socio-economic sphere. A lesser known variation of realism originating from John Anderson (1962b) working in Australia in the 50s and 60s is discussed as it provides epistemological and ontological principles pertinent to the natural-psychosocial division of reality as well as allowing an ontological grounding of knowledge and science avoiding some of the pitfalls of positivism and constructionism. Anderson’s situational realism avoids empiricist reduction of facts to atomistic simple units, and regards what happens in reality as complex and causality as not something singular and linear. Situational realism avoids idealism at all costs, regards reality ontologically as one and roots all knowledge, science and argumentation in the ontological claim that x is y (in the previous chapter we were introduced to the realist claim of “what there is”). The implication of situational realism’s avoidance of idealism is rather radical: the Kantian categories are ontological and what we know, perceive or understand is reality per se and not mediated by representations. Thus, understanding of how things work is based on the principle of non-constitutive relations. Subject and object remain distinct and independent along with the relation but never can the one or the other along with a relation constitute either subject or object. The implication for constructionism is clear: the mind cannot in any way contribute to constituting reality. In a sense, Anderson’s principle of non-constitutive relations is paralleled by Kant’s possibility of synthetic a priori judgments. Knowledge of the world that is worth something can only be obtained synthetically: if a predicate was somehow contained in the subject it would have been analytic and thus trivial. Finally, we look at the new realism of Maurizio Ferraris (2014, 2015a, 2015b). He developed new realism or positive realism that forms part of a newer continental version of realism we will not cover here, called speculative realism. A group of continental philosophers independently, and in some instances, in collaboration constituted the interesting embracement of realism rising from continental philosophy—of course, one should be alert to the fact how ironic this is because continental perspectives underlie the paradigms2 we met in a previous chapter, namely, critical theory, constructivism (of which social constructionism forms part) and the participative that are especially not known for their realist advocacy. Ferraris’ version can easily fit into the speculative realist group but his is a relatively clear and very useful formulation of realism that overlaps with critical realism. However, his views on language and ontology makes his view very attractive and compatible with some of the tenets of Anderson and in combination can allow one to ground a homogenous ontology and in particular bolster the view of science as fundamentally critical. The moment we realise that science is critical then it becomes the substratum of any method one uses to access domains of reality.

96

5 Realism in Psychological Science

Scientific Realism Chakravartty’s (2007, p. 4) definition of scientific realism (SR) can suffice for now: “Scientific realism, to a rough, first approximation, is the view that scientific theories correctly describe the nature of a mind-independent world.” Stated more generally, realism refers to one’s belief that something is real, independent of one’s mind (Chakravartty, 2007, p. 8). Scientific realism is the view that progress and success in science can only be explained by a miracle if the stuff science works with is not real (Putnam, 1975, p. 73). The no-miracles argument as explanation for the success of science is countered by three crucial arguments and Chakravartty (2007) pointed out that this debate generated much of current SR literature. The arguments are (a) the inference to the best explanation or the use of abductive inference, (b) the underdetermination of theory by data and (c) pessimistic induction (PI) (Chakravartty, 2007, p. 5; Ladyman, 2014). Abductive inference or inference to the best explanation states that one should prefer the theory that best explains the available data. As every fan of Sherlock Holmes knows, a number of theories and speculations may fit the data. Like Sherlock, one can possibly devise another theory as soon as one theory is refuted by new evidence. This is as much the case for common sense, the detective and the scientist. Abductive reasoning is a good strategy for interpreting data but it cannot function as a valid argument for realism and the antirealist can easily dismiss it with counter arguments. The same goes for the principle of the underdetermination of theory by data, usually voiced by the Duhem-Quine thesis. In fact, the principle of underdetermination of theory by data is one reason why abduction fails as a realist defence. One version of the Duhem-Quine thesis states that a hypothesis can never be tested in isolation and it is possible to make adjustments to salvage the hypothesis (Psillos, 1999). Thus, it is impossible to decide between two empirically equivalent theories except on criteria other than truth. Pessimistic induction refers to the view that theories in the past were shown to be false because they were replaced by newer theories. The fact that past theories are false implies that current theories will also be regarded as false in future (Chakravartty, 2007, p. 7; Frigg & Votsis, 2011, p. 241). The issue then for the theorist is to determine whether there are stable elements surviving theories in the process of discarding old ones for newer theories. Chakravartty (2007) notes that two variants of scientific realism, structural realism and entity realism, have certain limitations for the scientist and he proposes his own variation of SR, namely, semirealism. He characterises both approaches as selective scepticism (Chakravartty, 2007, p. 30).

Scientific Realism

97

Entity Realism Entity realism (EnR) holds that the unobservable entities described in theories are real especially when it is possible to manipulate them and let them interfere with other things (Chakravartty, 2007, p. 30). However, EnR’s selective scepticism entails not believing everything that is postulated about entities. Consequently, the reality of theories is under suspicion (Chakravartty, 2007, p. 30; Frigg & Votsis, 2011, p. 259). Given the two criteria of manipulation and interference, a belief in the reality of entities is warranted but not necessarily in the reality of theories in which they occur. The two causal criteria might ensure the endurance of some entities through different versions of theories. However, entities do not operate in isolation but requires the information encapsulated in theories to make them intelligible (Chakravartty, 2007, p. 31). More specifically, the properties and relations attached to the entities might be differently understood over time, hence, playing into the pessimistic induction threat.

Structural Realism In order to counter objections that theory change produces for realism, structural realist (StR) theories say something about the structure of observable and unobservable reality, but not about the unobservable entities within those structures (Ladyman, 1998, pp. 409–410). The relations are foregrounded at the expense of the relata, and the structure can be mathematically described (French & Ladyman, 2003). What is transferred between successive and different theories, is the structure and not necessarily the individuals. StR thus involves scepticism about the nature of the relata or entities but provides support for the no-miracles argument in terms of enduring structures (Chakravartty, 2007, p. 33). Chakravartty (2007) provides a simple model of structure as point of departure: a picture on the wall consists of a frame, wire, glass and so on related in different ways, namely with glue, spatially and so on. The entities or relata are the things such as glass, frame and wire. For StR the relations are important and the relata almost functions as placeholders. A distinction must be made between epistemic and ontic or ontological StR (Ainsworth, 2010): (a) Epistemic StR is the view that we can have knowledge of structures but not of relata and (b) ontic StR denies the existence of relata altogether (Chakravartty, 2007, p. 32; Frigg & Votsis, 2011, p. 260; Ladyman, 1998). Although set-theoretic assumptions dominate discussions in SR, the structure character of reality does not imply that reality is essentially mathematical (Ainsworth, 2010, p. 50; French & Ladyman, 2003). An understanding of what SR means by structure is required. In its simplest form, a structure consists of relata and relations. In addition, a distinction between abstract and concrete structures can be made. In principle, an abstract structure can have

98

5 Realism in Psychological Science

many concrete instantiations. Chakravartty (2007, p. 40) says “A concrete structure consists in a relation between first-order properties of things in the world.” So what are first-order properties? Things such as processes, objects and events have firstorder properties such as weight, colour, charge, volume and so on (Chakravartty, 2007, p. 40). An abstract structure is typified as higher-order properties of relations. Thus, the relation of, for instance, “being smaller than” and having the properties of “being smaller than” can be described as having a higher order property of x. As soon as relations can be described by x’s and y’s, they constitute an abstract structure. However, Chakravartty (2007) proposes a variant of realism, namely, semirealism, that avoids the problems of epistemic StR—the latter views abstract structures as the answer to the problems of PI. However, Chakravartty (2007, pp. 40–41) thinks that it is possible for one abstract structure to describe two different concrete structures, which makes the abstract structure explanatory redundant. One should, therefore, as a realist focus on concrete structures: Concrete structures are relations between first-order properties of things, so to know them is to know something qualitative about relations, not merely their higher-order properties (Chakravartty, 2007, p. 41); Concrete structures are identified with specific relations between first-order properties of particulars, and first-order properties are what make up the natures of things (Chakravartty, 2007, p. 42).

Semirealism Semirealism, consequently, avoids the scepticism of epistemic StR by maintaining that one can say and know something about relations and relata. An essential point for semirealism (SeR) is its view of how the burden of work is shifted to the properties of individuals (particulars or relata). Because the first-order properties of particulars are causally efficacious, these particulars cause events and processes to happen. Knowledge of the intrinsic nature of particulars is possible through knowing its properties. Because the properties are responsible for relations as well, knowledge of the relations implies epistemic access to particulars’ intrinsic natures. This is the first point of SeR; the second is that properties “confer dispositions” on things and these dispositions are their tendency to act and behave in certain ways (Chakravartty, 2007, p. 42). Another way of putting this is, as Chakravartty (2007, p. 43) says, structure and nature come together, similar but still distinguishable. The next issue for SeR is how to regard real structure and particulars without SeR merely being a conflation of entity and epistemic structural realism. Chakravartty (2007) distinguishes between detection properties and auxiliary properties. Detection properties are those that are stable, carried over in different theories and which are minimally required for the theory to be about something. Detection properties are, of course, properties of things that constitute relations to other things with properties. Chakravartty (2007) provides the following guideline: “The recipe of the minimal interpretation is austere, and straightforward: commit only to structures

Scientific Realism

99

with which one has forged some significant causal contact, and understand the natures of detection properties in terms of dispositions for relations to other properties” (p. 54). (a) A brief summary of semirealism Semirealism (SeR) goes further than entity realism (ER) and epistemic structural realism (StR). SeR entails knowledge of both properties and relations as stated above in the sense of both providing epistemic access to the internal nature of particulars. It will be remembered that entity realism grounds our belief in entities in the extent to which they have effect on other things and to the extent we can manipulate them to do things. However, knowledge of relations is excluded from entity realism: By emphasizing causation, ER captures the common and deeply held realist intuition that the greater the extent to which one seems able to interact with something—at best, manipulating it so as to bring about desired outcomes—the greater the warrant for one’s belief in it (Chakravartty, 2007, p. 58).

Epistemic structural realism says that, despite theories that change over time, there are certain mathematical descriptions of relations that remain unchanged (Chakravartty, 2007, p. 59; Ladyman & Ross, 2007, p. 94). These core descriptions probably show that there is something true in these theories. However, philosophy of science showed that structure is not always defined at a sufficiently detailed level as to be helpful to clearly indicate what counts as structure for realists. Thus, both ER and StR address what they believe should be the target for realists in order to counter one of the problems of realism, namely, pessimistic induction but each lands in different problems. Semirealism solves the problems of both versions of realism by (a) having a clear understanding of what is studied in science, namely, the properties of things, and (b) their relations which form the structure of reality. In order to account for core and stable truths of theories, the relations are described with mathematical formulae. Structure is viewed as concrete and not abstract. The main aim of semirealism is to retain core content of theories and distinguish this from auxiliary content. In this way SeR can account for the historical development of theory. Because the causal work is done by properties of particulars and their relations, Chakravartty (2007, p. 64) distinguishes between causal properties and auxiliary properties and, of course, the latter cannot be involved in identifying an object in successive theories. (b) Causality According to Chakravartty (2007, p. 90) “. . . the epistemic engine that drives realist beliefs regarding certain unobservable properties, structures, and particulars is causation.” A more eloquent expression of the centrality of causation for the realist can rarely be found! However, this story comes with a twist. Unlike things, properties and events, the existence of the very unobservable “cause” cannot be empirically determined. This was, of course, Hume’s issue with causality. Chakravartty (2007, pp. 90–91) says that there are metaphysical concepts that the realist holds to be true, that cannot be empirically checked,

100

5 Realism in Psychological Science

but are required for the foundation of realism. Thus, natural kinds and causation are metaphysical concepts the realist presupposes for a realist description of reality. Furthermore, necessity is an aspect of causality required to enable distinction between contingent and necessary patterns of events (Chakravartty, 2007, p. 93). The type of causal necessity we are after is de re necessity instead of de dicto, i.e. objective necessity, rather than subjective or conceptual. The issue of causality is complex and has a long and chequered history (Cartwright, 1998; De Pierris & Friedman, 2013; Falcon, 2015; Groff, 2008; Porpora, 2008; Schaffer, 2014; White, 2013). Chakravartty (2007, pp. 96–102) discusses objections to the idea of causal necessity, one of which I will briefly refer to because it applies to a central concept of Bhaskar’s theory. The demand for a causal mechanism in science is necessitated by the so-called inability to present a coherent account of de re necessity. Causal effects can be described as thing A having an effect on thing B. If A is imagined as unchanging then it is very difficult to think that it can bring about a change in B. In fact, even if A changes, the issue of how it might effect a change in B still requires explanation. The postulation of a mechanism, which can easily turn into a mere black box, is required to act as go-between between A and B (Chakravartty, 2007, p. 101). Taking ordinary life instances of causation as exemplars, we might be misled to think that cause-effect holds between events. For example, the door opened because I pushed it or the wind blew it open; my teeth rot because I do not brush them and so on. The agent (thing A in the discussion above) causes something to happen expressed as an event. According to Chakravartty (2007, p. 109), referring to events as relata of causal relations is a rough approximation of what is actually happening: there are much more detail waiting to be unpacked in order to give a proper realist account of causality. Chakravartty (2007, p. 105) points out that events-based talk cannot work as an ontological explanation for causal mechanisms. Events can be demarcated in any way one pleases. A more sensible way is to locate causal mechanisms within the properties and their relations in concrete structures (Chakravartty, 2007, p. 107). A simple example by Chakravartty (2007), in his own words, is the property of mass, which allows a body that has this property, to accelerate when a force is applied to it, and he concludes that “Causal phenomena are produced by the ways in which particulars with properties are disposed to act in concert with others . . .” (p. 108). Events appear to cause other events but the actual causal agents are the properties of particulars predisposing them to act in certain ways. Chakravartty (2007) moves away from a simple linear concept of causality— although he explains it linearly: “Events that are changes generally overlap multitudes of changes in the properties of the objects concerned” (p. 109). The point that he wants to make though, is not about the complexity of interacting things, but about the continuous nature of interacting things that is summarised by the phrase causality as process (Chakravartty, 2007, p. 111). He is convinced that the process account of causality can counter many objections against the realist notion of causal mechanisms. However, one should concede that the demand for a causal mechanism might lie beyond the grasp of the realist:

Scientific Realism

101

precisely how change is brought about at the moment of interaction might not be explicable (Chakravartty, 2007, p. 112). Chakravartty (2007, p. 111) acknowledges that both Hume and the empiricist have a point, we cannot observe a mechanism, but the realist assumes that this unobservable “moment” or link is real in contrast to the empiricist’s sceptical conclusion.

Implications of Scientific Realism for Social Science and Psychology One should understand that semirealism is devised as a theory explaining the natural domain of reality. On its own it does not purport to say anything of the social or psychological domains of reality. When looking at psychosocial reality, how applicable is both entity and structural realism? Both, like semirealism, aim to answer pessimistic induction to show which elements in a theoretical description remain constant when theories evolve. These are difficult issues to address with local and psychosocially based explanations. The question is to what extent psychosocial theories try to identify causal mechanisms. How important is it to predict human behaviour or the behaviour of social groups? In other words, what is the nature of understanding when working with psychosocial phenomena? We have touched upon the nomothetic and ideographic aims of the two spheres of sciences and hopefully we are realising that these are oversimplifications of both natural and human/social sciences’ aims. If one of the main aims of science is to uncover causal mechanisms, does this apply to psychosocial science as well? The debates between quantitative and qualitative methods made it clear that a mathematical reduction of human behaviour and mental life is difficult if impossible. A possibility is to regard causal mechanism discovery as essential to science but require the method of its discovery, description and analysis to be epistemically appropriate. The second issue is the nature of the relationship between mind and reality. Scientific realism is clear in its requirement of mind-independence: truth and realness are not conferred by the mind but by the ability to affect things causally and letting those things affect other things causally. Observability is not a requirement for realness in the case of natural science which is good news for the psychosocial scientist because many things that we assume and work with cannot be observed. In a sense then it is easier for the psychologist to be a realist about the constructs she assumes to operate to produce behaviour, and in a similar sense a SC can also side with the realist. However, mind-independence for psychosocial reality is another matter entirely because many of the phenomena we depend on to describe psychosocial reality, and many phenomena constituting psychosocial reality, are mind dependent! A familiar realist thought experiment is to imagine a world without human beings and then ask what would remain. In our case as a human race, the remainder of the universe! Star

102

5 Realism in Psychological Science

formations, supernovae implosions and rocks forming and disintegrating will go on rather infinitely without human knowledge, intervention or appreciation. What would not remain without human beings? Culture, cultural artefacts such as language and motor vehicles, nor iPads or stone age tools. Definitely not timetables and bus schedules. No love or social constructionists; no anxiety disorders or bad dreams; no documents, no mathematics nor racism or human consciousness and so on. Now, does mind dependence mean that these things are not real? Furthermore, if we as a civilisation die and very little trace except a WiFi receiver here and there remain, does it preclude the rise of humanity 2.0 who similarly can discover mathematics and experience anxiety? No, it does not. Cultural things such as language may be lost forever, but they can exist as real in future, has done so in the past and certainly exerted a very concrete and real effect on its originators, participants and immediate environment.

Minimal Scientific Realism Mäki (2002, 2005b, 2011) can be counted as a representative of SR but one not concerned exclusively with natural science. He made a number of appeals for a softening of scientific realist requirements in order to accommodate other sciences not “on par” with physics as an enterprise for whom SR seems a tight fit. “On par” should be understood loosely because Mäki wants to avoid the exclusivist notion that physics is the model example of what a mature and successful science should be. However, it should be acknowledged that other sciences, such as his own field of economics, also qualify as sciences and also can claim realist status. As we have seen above, SR defends itself against anti-realism on grounds of the no-miracles argument: the success of science cannot be explained by a miracle. Thus, the first main assumption of SR is the success of science and in this case physics. The second assumption of SR is the mind-independence of natural reality and the third is the existence of unobservables. It should be readily apparent that psychology in our case, and economics in Mäki’s case, also work with unobservables but probably qualitatively different than those physics work with. The second major difference for Mäki is the lack of mind-independence in both the economy and in economics. In psychology, and as we have seen above in social psychology, a great deal of things we study are mind dependent: remove persons and psychosocial phenomena disappear. Mäki concedes that economics has not made such spectacular progress as physics and the same goes for psychology which is frequently viewed as an immature science or sometimes even a pseudo-science. Although a field such as economics had not had the predictive and technological success of physics, Mäki is still convinced of its viability as a science. Not sharing the miracle of physics does not imply that realism cannot be applied to economics. As a minimum, realism as applied to a science, means that its theories represent reality in some way and does so truthfully. Realist theories are thus not instrumentalist devices: they say

Minimal Scientific Realism

103

something true of reality and if we want to retain the realist slant of a theory, some adjustments need to be made to escape the rigorous requirements of SR. One can keep the philosophy constant and argue that the field does not comply to the ideal or one can adapt the philosophy based on what is found in the field. Mäki (2011, p. 4) prefers the latter. In the case of SR the predictive and technological success of physics determined the tenets of its realism. For a discipline to retain a realist claim for its theories—if SR does not apply—it might not be expedient to keep the ideal philosophy of science constant. The avenue Mäki (2005b) proposes is to look at the requirements of realism, then see whether it can be sufficiently relaxed to include a specific field as realist and scientific. Mäki (2005b) proposes a relaxation of the ideal of realism and allowing the requirements of a localised science inform realism. His (2005b, p. 246) solution of talking about science-independence and theory-independence takes into account that what is at stake with scientific realism are scientific phenomena. A version of minimal scientific realism (MSR) is then developed, with flexible criteria, to allow for local variations of various scientific disciplines (Mäki, 2011). In addition to the basic tenets of SR already indicated above, namely, (a) mind-independence, (b) unobservables, (c) that theories are justifiably regarded as more or less true, the following should also be relinquished: (d) that the epistemic performance of a theory within a discipline will always be successful; (e) that science progresses and one can invariably point to its successes so that the main task of a philosophy of science is to explain why this is not a miracle (Mäki, 2011, p. 5). Mäki (2011, p. 5) shows how these requirements can be relaxed to include varieties of local sciences that have varying degrees of success and are in various stages of development. He illustrates how MSR grounds economics and I would like to point to a few issues important for our movement towards a realism for psychological science.

Mind-Independence Mind-independence as a criterion does not hold for social science and psychology. Mäki (2005b, p. 243) made the accurate observation that without minds there will be no social reality, thus, social reality is mind- and representation-dependent. The basic tenet for realism cannot thus hold for psychosocial reality. Furthermore, “Social objects characteristically do not exist representation-independently either . . .” Mäki (2005b, p. 243). Social institutions, such as marriage, exist as partially socially constituted institutions, namely, it is partially by agreement, some documentary evidence exists and human behaviour maintains it as a phenomenon. Trade is represented to some extent by the physical representation of money and the actions executed by people when trading. One should understand Mäki as referring to the representations and symbols held in people’s minds as well. Thus, partial and some full representations are concept-dependent. The problem, consequently, is how one can maintain realism if some of the things we live with are mind-dependent? Maybe

104

5 Realism in Psychological Science

one should regard mind-independence differently when speaking about the reality of things not depending on the mind as in the case of natural science. Natural objects are science-independent, i.e., they exist prior to and separately from science. Mind-independence works fine as a realist criterion for some sciences such as physics but most of our psychosocial world will (partially) disappear without human minds. Thus, rather than retaining mind-independence as a criterion for realism, one should acknowledge the things we regard as real in our world and adapt realism to this view. Mäki’s aim is a broader applicability of realism by letting local and sometimes undeveloped sciences into the realist fray. The question is what it is to be a realist without the criterion of mind-independence. For starters when we claim something to be real or true we believe in some minimalist sense in its existence. Of course, such a belief may be expressed emphatically and justified epistemically such as “I believe electrons exist because you can see their traces in a cloud chamber” or “what else could explain electricity?” However, according to Mäki (2005b) strong beliefs and justification is not required in order to be counted as a realist either as common folk or as a scientist. What he calls conjectural truth nomination or falsity ascription is all that is necessary. Anything may be said to exist if one does so tentatively and leave open the fact that in future in might be proven either true or false. Mäki’s strategy of tentative truth nomination takes care of PI discussed above. We no longer have to worry whether a theory is true and the fact that past theories cannot be held as true merely strengthens a minimalist realist attitude. Let us explore Mäki’s (2005b) solution to mind-independence. One should rather speak of theory-independence or science-independence. The latter means that things exist independently from their consideration by scientists practicing science. Social things probably exist mind-dependently (such as the institution of marriage and trade) but their existence normally does not depend on those practising science and the practice of science. Sometimes the situation might be more complex where the practice of science actually constitutes the entities of the social world merely because social science forms part of the social world (Mäki, 2005b, p. 245). For the moment one may distinguish between conceptual and causal dependence (Mäki, 2005b). Elsewhere Mäki (2011, p. 7) calls this constitutive dependency which means that “saying so makes it so”: science or any epistemic activity cannot bring things in the domain of social phenomena into being. Casual dependence means that what science says about social phenomena might influence people to believe those things and behave and act as if they were real—social actions and beliefs thus depend causally on scientific proclamations, but it is people that do the acting and believing. A minimalist realism should aim for conceptual scienceindependence. Social reality is mind-, representation- or concept-dependent (Mäki, 2005b, p. 243). These are first-order dependencies. However, on a second and higher level, these objects can be studied and are thus science-, mind- and conceptindependent. What the scientist studies, is not created by her! An interesting consideration is whether Mäki actually succeeds in pinpointing the meaning of scientific objectivity as expressed by the principle of SR, namely, mindindependence? The intention of mind-independence is to circumvent idealism in

Minimal Scientific Realism

105

the process of asserting the primacy of ontology: this is exactly what Mäki wants to express with conceptual-independence or constitutive-independence.

Unobservables Mäki (2005b, p. 247) questions the viability of talking about unobservables in localised versions of realism. Moving beyond physics and the issue of unobservables will require a reformulation of what is regarded as unobservable in a specific science. The example of dinosaurs and electrons might suffice and by implication, the things psychology talks about, such as agency, can be regarded as localised versions of unobservables. Thus, the next classical requirement that can be relaxed is that of unobservables. One should keep in mind that the unobservables physics talk about, such as electrons, black holes and so on are in some way known by their effects— they are things! Traces of particles can be found in a cloud chamber and the postulated existence of a black hole by the behaviour of matter around it. The idea of an unobservable is qualitatively different in psychology and it probably entails a double unobservability, namely, concept-dependent and latent. Agency is a good example of an unobservable construct and as social constructionism has shown, many so-called entities such as personality characteristics operationalised by psychometric tests exist because of the labels we have created. Other sciences are in a similar situation and one should not hold the apparent lack of proven unobservables against their explanatory value. Mäki (2011) argues for a position of tolerance: claims of what exist can be done tentatively and it is the task of the particular local science to demonstrate its realness in the end. This immediately puts the psychosocial and human sciences on a better footing. They are allowed to postulate and assume until such time it is proven that a construct does or does not exist. We may thus retain its explanatory function until it is proven not viable. The minimal requirement for being a realist in scientific matters is that our theories claim some things to be real even if approximately so.

Commonsensibles Mäki regards economics as part and parcel of the social sciences, thus, his science theory, departing from SR, has important things to say about social science. He went some way beyond Bhaskar with applying realism to social science. Flowing from the above, is the idea of commonsensibles, a term he (2005b, p. 247) coined earlier and based on the idea of different images of sciences. A distinction can be made between the scientific image and the manifest image, which is similar to scientific and folk ideas of science, and in this case, psychosocial science and folk psychology, and physics and folk physics. When speaking about the scientific image of, for instance a table, it would consist of talk of atoms and so on, thus, very specific unobservables,

106

5 Realism in Psychological Science

whilst the folk image of physics would talk about the same table very differently (Mäki, 2005b, p. 247). The commonsensibles are the entities referred to in discussing the table. In social science the commonsensibles are not very different from each other in both realms of the manifest psychosocial and the scientific psychosocial. The question is what is the utility for talking about commonsensibles? Does Mäki plead for tolerance toward immature sciences meaning that in the end they will be on the same footing as physics? What is the benefit of letting go of mindindependence? Does it advance science and our knowledge of reality? Mäki (2011) makes a distinction between practicing (social) scientists, philosophers and laypeople, all whom deal with and speak about the economy. Commonsensibles are used by laypeople and scientists alike, and the reality of these are usually undisputed, while philosophers might challenge their ontological grounding. In a sense this is helpful and reassuring to realise that the mind-independence of commonsensibles is usually beyond dispute but does this really solve problem? Mäki assumes that the commonsensibles scientists/folk talk about exist to a large extent, but what if they and Mäki are plainly wrong? Take the example of the “cult of the unicorn”: is the unicorn a commonsensible for both folk and unicorn theologians? Both folk and theologians assume the unicorn exists, and it features as an entity with causal powers. But is this the task of the philosopher, to point out that the entities scientists and folk talk about might not exist? The question remains how the discussion of existing things differs between natural and social domains. According to Mäki (2005b, p. 248) the crucial difference is that social scientists have ready access to the commonsensibles they study because commonsensibles are part of the domain under investigation. Mäki (2005b) states that social scientists share the common ground of folk concepts of a field, but concepts or commonsensibles are changed, refined and so on, to suit the needs of a very localised science practice: “They are usually not concerned about the existence of those commonsensibles or about which non-commonsensibles to postulate” (Mäki, 2005b, p. 249). Why is this? According to Mäki (2005b, p. 249) “The major ontological issues in actual social science are thus not about the existence of the objects and properties depicted by their theories, but rather about their causal role and relevance in the functioning of social systems.” How would economics proceed as a science? Economics is a theoretical endeavour involving a number of reworked commonsensible variables where researchers construct models of the economy and with these, try to predict real world events. The models are sufficiently simple and, in a sense, more sophisticated than the folk version of relationships between commonsensibles (Mäki, 2011, p. 8). This means that models deliberately try to avoid the clutter of real-world complexity by modelling the relationships between a select number of refined commonsensibles. Models provide epistemic access to reality and according to Mäki (2011, p. 10) it is “notoriously difficult.” Models often fail or succeed partially, and epistemologically hard work is expected from the scientist to make necessary adjustments. Mäki (2011, p. 10) correctly states that this is an epistemological undertaking, rather than an ontological one, acknowledging the fallibility and also the socially conditioned nature of theories and models.

Minimal Scientific Realism

107

Models Issues of truth of models or theories are also relaxed in MSR. One need not, from the start, claim the truth of a theory (i.e., that it correctly reflects the state of affairs) because theories might possibly be true and only later need to be tested or subjected to claims of truth (i.e., that things are or are not so). As we will see below, this conception of truth is similar to Anderson’s view of truth. Mäki (2011, pp. 9–11) makes an interesting observation: MSR requires possible truths but these will not suffice on their own because truths must be relevant. So the scientist is not merely after all truths or easy truths but only those that have a “chance of being true” and relevant to the current situation (Mäki, 2011, p. 10). An interesting question for the scientist is whether it is at all possible to avoid the topic of truth or speaking about how well a theory reflects what is actually going on? Accordingly, MSR does not require claims to be true from the outset, but only that they have a chance to be true. In my opinion, this requirement is not all that different from Bhaskar’s position: constructing models or analogues of mechanisms are postulated in the process of doing science, and claims to truth, i.e., whether a proposed mechanism is true is a matter of testing. We will return to this point later in the discussion of critical realism. The way Mäki formulates the softer requirement creates the impression that SR, from the start, requires a theory to be true, and I do not think this is the case. However, the explicit truth-securing delay built into MSR is made explicit and allows other sciences than physics access to scientific realism.

Conclusion Mäki (2011) discussed the requirements for a minimal realism which consists in (1) revising the requirement of mind-independence and clarifying scienceindependence (2) revising the issue of unobservables and clarifying commonsensibles, and (3) theories need not be approximately true form the start; they need to have chance of being true. It is not so much the existence of constructs that concerns the social scientist but the causal effect of the postulated commonsensibles. This view, however, exempts the social scientist and the for-ever-reifying-psychologist from critically examining their postulates and according to Michell (1999) relaxing truth claims from the start can lead to pathological blindness, for instance, to the problems of measurement in psychology and the social sciences. I think that the conceptualisation of minimal scientific realism is important to Mäki to enable economics to draw upon scientific realism and by proxy be regarded as a science. What does Mäki say for our movement towards a metatheory for psychology? The encouraging matter is that various sciences about the reality we live in are not all on the same developmental footing or even trajectory. One can be tolerant of psychology’s claim to be a science although its subject matter on most levels are concept-dependent or not mind-independent. This still does not provide us with an

108

5 Realism in Psychological Science

answer of what a science is and should be but at least one need not be too much concerned if truth, representation and reality do not overlap sufficiently or at all. For Mäki the truth of a theory means how well it represents reality and this overlap, even if partial, qualifies as realism. It is the realist attitude that counts and this minimalist attitude solves a number of problems, amongst others, the PI problem and in the process the mind-independence issue gets clarified as concept-independence.

Critical Realism We now have in hand a minimalist requirement for science to be realist, essentially boiling down to an attitude that things the scientist theorise about might be real or eventually proven to be real. In addition, we acknowledge the intention of the SR mind-independence clause, namely, that things we claim to be real is non-constitutively real to someone. However, some things are real and minddependent. So it is important to distinguish the senses in which we will be using concept- and mind-dependence and when which applies. The major issue is that we recognise that some things are also mind-dependently real. Roy Bhaskar published his version of critical realism in the 1970s. He did his doctoral studies in philosophy under the guidance of Rom Harré in the UK and had a very strong Marxist slant. The first two major works Bhaskar (2005, 2008) published were a theory of natural science called “A realist theory of science” and this was later complemented by the “The possibility of naturalism” which focused on the social sciences. Both these works had an enormous influence even many years after their publication on many social scientists (Archer, Bhaskar, Collier, Lawson, & Norrie, 1998). Bhaskar gained a strong following, an Institute was established along with a journal and a series of publications saw the light. Bhaskar moved to an explicitly Hegelian philosophy and logic with later work. A number of authors support the CR movement and the various fields and disciplines CR influenced are numerous (Collier, 1994, 1995, 1997, 1998; Cruickshank, 2004, 2012; Danermark, Ekström, Jakobsen, & Karlsson, 2002; Fleetwood, 1999; Groff, 2000a, 2000b, 2004; Harré, 2009; Hartwig, 2007; Lawson & Staeheli, 1991; Lopez, 2003; Lopez & Potter, 2002; Peacocke, 1984; Sayer, 2004; Shotter, 1992; Sims-Schouten, Riley, & Willig, 2007; Viskovatoff, 2002). Despite his obvious enormous influence, CR did not really sit well with SR although many overlaps are readily apparent. Mäki is almost the only one engaging with Bhaskar albeit in a cursory manner and despite the complaint that SR managed to keep social scientists outside their particular brand of realism. Bhaskar passed away in 2014 but his first two works remains strongly influential in critical realist circles within social sciences and the humanities (“In memoriam: Roy Bhaskar 1944–2014,” 2015). Our focus is on his (2005, 2008) two major works in the theory of science because it established the basic principles well-known in CR circles. One major attraction of CR is its aim to avoid positivism and strong social constructionism. Talking about postpositivist debates, Bhaskar (1998a, p. x) says:

Critical Realism

109

“A problem for all these trends was to sustain a clear concept of the continued independent reality of being—of the intransitive or ontological dimension—in the face of the relativity of our knowledge—in the transitive or epistemological dimension” (Bhaskar’s emphasis). In the debates between quantitative and qualitative methods discussed above and in the clash between realism and relativism in the 90s, CR has much to contribute although not many authors utilised Bhaskar’s CR explicitly. Some other issues that made his philosophy attractive are his emphasis on the primacy of ontology, the epistemic fallacy, his acknowledgement of the fallible nature of knowledge and so on. The issues that arose from postpositivism were incorporated in his theories. Bhaskar’s critical realism was initially called transcendental realism with reference to the natural dimension of reality, then critical naturalism with reference to the social dimension and eventually critical realism reflecting the development of his thought in his theory of natural science and of social science (Bhaskar, 1975, 1998b).

The Transitive Dimension, Epistemic Access and the Epistemic Fallacy As we have seen above a few basic principles apply to a metatheory in order for one to call it realist. CR shares these principles but have one or two additional requirements. These may be summarised by the principles of reality as stratified, differentiated and structured. Before we explain the meaning of these principles one first might ask the question of how do we know that reality is a certain way? In fact, all that we have in order to make reality intelligible to us are our minds and we constitute knowledge with understanding by means of language. So at least on this point the social constructionists are correct: the only way we can understand reality is by means of our language. There is no other way. However, where they are wrong is about the following fact: because we can talk about things and because we can construct, frame or talk differently about things, it does not mean that these things do not exist and are only constructed in our talk. Of course, we can construct things out of thin air; this is part of our ability to speak! However, CR’s issue is that we only have epistemic access to reality through language. This language-dependence, or as we have seen above with Mäki, concept-dependence, brings about the numerous problems as metatheories, such as constructionism and positivism, have to deal with. The main problematic consequence with having only our language and its associated concepts available to access reality is what Bhaskar calls the epistemic fallacy. Although numerous philosophers pointed out the problem of the epistemic fallacy, Bhaskar (2008, p. 21) gave an appropriate example of this problem: two persons are looking at a sunrise—Kepler sees the horizon move away whilst Tycho Brahe sees the sun rise. The first point of the example is that there is something they both see and the second is that they understand it differently. Over and above the question that the reader might pose, namely, which interpretation is correct and how

110

5 Realism in Psychological Science

would we know, the two elements of our dealing with reality is delightfully given to us: a reality independent of us and our understanding of that reality. In the context of science the first Bhaskar (2008, p. 11) calls the intransitive dimension of science and the second the transitive dimension. Our theories, understanding, interpretation and constructions form the transitive dimension of reality. The epistemic fallacy rears its head as soon as we turn the order of dimensions around, namely, regard what we understand as real rather than acknowledging our understanding as transitive. The epistemic fallacy can be summarised with the simple principle: thought is not to be confused what is real. Psychologists and other social scientists commit this fallacy when reifying constructs, or when our ability to think about things misleads as to view them as concrete. We will consider the intransitive dimension in a minute. Let us first consider the transitive dimension further. This dimension of science consists of knowledge but not knowledge produced from nothing: knowledge is always produced from knowledge (Bhaskar, 2008, p. 176). This means that science, as a social activity, consist of a body of theories encapsulating knowledge about reality. The transitive object of science are these theories that it uses and transforms (Danermark et al., 2002, p. 23). From the nature of the transitive dimension of science, namely, consisting of verbalised material only accessible through language, it should be obvious that its constitution is a social activity. It means that agents construct theories but only as a social scientific community. The transitive dimension is designated by its social constitution (or construction if you will), its fundamental language substrate and lastly, because of these two aspects, its fallibility. With the latter its transient nature is indicated but also the possibility that it can be wrong. Knowledge cannot be infallible and certain—the certain Archimedean point sought by many for scientific knowledge has long dissipated. As we have seen above, this is one of the issues with which anti-realists skewer realists and we need to clarify how we ensure valid knowledge. The question is if theories are fallible and thus wrong in the past, it means that the current theories also might be wrong. Antirealists put this emphatically: realism is then wrong about the truth value of current theories; thus, realism is not sustainable. Below we will see how CR solves this problem.

The Intransitive Dimension of Science The transitive dimension of science assumes the intransitive object of science which is reality itself. It is relatively easy to pick out the intransitive objects from natural science’s perspective, namely, tables, quarks and powers. These things are durable and can be either observed or detected as the SRs indicated. However, Bhaskar provides a uniquely useful distinction that simultaneously reveals the difference between realism and versions of empiricism such as Logical Positivism. This is the idea of reality as differentiated. Different levels of reality can be distinguished and if the scientist does not focus on the correct level, her theory, more often than not, might be wrong in accurately describing reality but also explaining it.

Critical Realism

111

Reality as Differentiated Let us explain it in terms even undergraduate psychology students can grasp: almost every research methods textbook focusing on statistical analysis at some point says that one should be careful when interpreting correlation coefficients. It is taken as common knowledge that correlation or co-variation does not imply causation and this is simply based on the easily demonstrated fact that two variables might covary incidentally (Shadish, Cook, & Campbell, 2001, p. 7). A number of issues can be pointed out from this observation. First, correlation is based on observing regularities of events. Now events happening may be the only handle we have on suspecting something might be going on. Just as easy as it is to demonstrate that coincidences might look like regularities, it is likewise easy to demonstrate that the lack of regularities might obscure a causal connection. This means that the symptoms or lack thereof we observe are underpinned by a thing’s casual mechanism: events merely show that the causal mechanism is at work and cannot be equated with the mechanism (Bhaskar, 2008, p. 4). The effect of gravity was not apparent whilst Newton sat under a tree when the apples stayed in the tree but a lack of falling apple events does not mean gravity as the mechanism behind falling things does not exist. The abundance of events are the surface symptoms that causal mechanisms are at work or present. According to CR events or surface symptoms cannot be equated to a causal explanatory mechanism. This leads us to suspect that reality is differentiated, namely, that the causal mechanisms responsible for things happening in the world lie a bit deeper than what observation allows for. CR makes a distinction between three levels, namely, the empirical, the actual and the real (Bhaskar, 2008, p. 2). The real is where the causal or explanatory mechanisms operate, while the actual is where actual events take place. As soon as a causal mechanism is triggered—like the apple’s stalk disintegrating—an actual event takes place. Thus, the apple hits Newton on the head because gravity was allowed to execute its power. Of course, the apple might, and they actually do, fall without people observing them! The level of the actual means that things happen irrespective of people observing them. However, Bhaskar points out that when people observe things happening the empirical level is constituted. As soon as we take this level of observed events as the primary level of the real we step into the correlation fallacy. Bhaskar (2008, p. 1) says that this regularity concept of causation is rooted in Hume’s view of the constant conjunction of events as requirement for establishing empirical laws. Both empiricism and positivism restrict the real to this empirical level.

Reality as Stratified and Structured Stratification refers to different levels of reality or strata of explanatory/generative mechanisms and not that there are only two levels, namely, appearances and mechanisms (Danermark et al., 2002, p. 59). Thus, CR supports a depth ontology, depth understanding, or vertical movement of science. Lower levels explain higher

112

5 Realism in Psychological Science

levels without reduction or doing away with higher levels (Bhaskar, 2008, p. 160). Strata refer to among others the chemical, biological, psychological and sociological levels (Bhaskar, 2008, pp. 198–199). It also implies that strata are hierarchically organised and that properties of higher strata emerge from lower ones without being reducible to the lower ones, i.e., a lower level does not explain higher levels away. For instance, neurobiological mechanisms underlie events on the psychological level, but discovery of neurobiological mechanisms does not make the psychological level of explanation and description redundant. I must put this a bit more cautiously. Usually the principle should apply but we are sometimes not sure how to characterise the psychological: it is amenable to reification and proliferation of constructs. It might turn out that, for instance, we can redefine the so-called psychological construct of say intelligence in terms of the efficacy of certain brain network functioning so that it might be operationalised and measured qualitatively differently than how we are doing it currently (assuming per Mäki’s concession that such a commonsensible exists) (Hampshire, Highfield, Parkin, & Owen, 2012). Something similar to intelligence will then be constructed differently but might still find expression on a psychological level. Thus, we might still say that John acts intelligently meaning he is able to plan ahead, anticipate events accurately, solve problems quickly and so on but his psychological behaviour will be an expression of the variability and interconnectivity of brain networks and modules (Zhang et al., 2016). What does structured mean? Bhaskar (1998b, p. 12) says that objects of science are structured thus, “irreducible to patterns of events” and elsewhere he (Bhaskar, 2008, p. 172) contrasts complexity and structure with atomistic and event-like. This means that structure refers to things not being atomistic but sufficiently structured to generate other things, thus, having generative mechanisms (Bhaskar, 1998b, p. 187, 2008, p. 42). That reality is structured, differentiated and stratified describes the fact of reality as having generative mechanisms from different perspectives. These perspectives express the complex multileveled nature of reality as opposed to collapsing the real into what is observable.

The Nature of Intransitive Reality: The Transcendental Argument How do we know that reality is structured, stratified and differentiated? Bhaskar uses a transcendental argument (TA) to show what reality must be like in order for science to be possible (Bhaskar, 2008, p. 210). This sort of argument is also called retrodiction. For our purpose, one may distinguish between different forms of argument, namely, inductive, deduction, abduction/retroduction and retrodiction. Induction is a form of argument based on empirical premises and leading to a general conclusion. Deduction involves general and specific premises leading to a specific conclusion. Abduction or retroduction is also an inferential process and provides a possible scenario for given set of circumstances, so inferring causes from effects (Hartwig,

Critical Realism

113

2007, p. 195; Niiniluoto, 2018). Transcendental argumentation as a particular form of retroduction, as Bhaskar uses it, asks what the situation must be for x to be possible (Hartwig, 2007, p. 257). Bhaskar uses a TA to consider the conditions for the nature of reality if science is to be possible and given that it is possible, reality must be a certain way. Philosophy can therefore . . . tell us that it is a condition of the possibility of scientific activities φ and ψ that the world is stratified and differentiated, X and Y. But it cannot tell what structures the world contains and how they differ. These are entirely matters for substantive scientific investigation (Bhaskar, 1998b, p. 6).

Bhaskar (2008) takes experimentation as the hallmark of natural science and identifies the characteristics of reality in order for experimentation to take place (seeing that it does). I will come back to the point of experimentation as a crucial characteristic of science because as we will see experimentation cannot do the same work for CR in social science, so Bhaskar needs an alternative to motivate social science. However, for the moment one can assume that experimentation plays a crucial role in some of the sciences. Of course, there are natural science disciplines that cannot rely on physical experimentation. Bhaskar (2008, pp. 23–25) points out that when a scientist such as a physicist does an experiment, namely, constructing a closed system where only one mechanism can be triggered by the experimenter, resulting in the effect of the generative mechanism, a number of aspects can be identified about the nature of reality which made the experiment possible. Put in another way: if a distinction could not be made between open and closed systems, between events and generative mechanisms, between levels or strata of reality, and between objects and persons then experiments would not have been possible. Bhaskar (2008, p. 23) says “The intelligibility of experimental activity presupposes not just the intransitivity but the structured character of the objects investigated under experimental conditions.” Thus, the objects investigated exist independently from humans (intransitive) and they cannot be identified with a constant conjunction of events (structured). Stratification, structured nature, differentiation, intransitivity and change of natural reality can be assumed philosophically although the specifics of strata and mechanisms can only be identified by individual disciplines. The possibility of scientific experience demonstrates the nature of reality. Unfortunately, Bhaskar and CR’s starting point, namely, experimental activity forced his hand with social science. CR holds that social systems are fundamentally open which means that experimentation is not possible, along with what one normally associates with experimentation, namely, statistical analysis of systems. Some CRs soften the impact by saying that the possibility of closure forms a continuum, and the different disciplines can be indicated on this continuum. Of course, this is true: it is possible to close systems on a chemical and physical level, it is a bit more difficult on a biological level, more so on a psychological level and thinking of sociological phenomena such as marriage, communities, organisations, states, prison systems and so on, it becomes almost impossible to think of how to close those systems and conduct experiments. CR seems to think that the supporting levels or strata of a discipline allows experimentation and closure more easily than in

114

5 Realism in Psychological Science

other disciplines. For instance, psychology is sandwiched between the biological or neurophysiological stratum below and the social stratum above. One might argue that the biological level makes experimentation possible to some extent in so far certain biological systems can be closed. However, this is not true and harbours a remarkably restricted assumption about closure. Although one might agree that some biological systems such as dogs and people are complex systems, thus their complexity precludes closure, experimentation on humans and dogs are taking place and not only on a bio-physiological level. Cognition, emotions, perception and a host of other psychological acts are investigated in terms of controlled or partially controlled experiments (Irie, Hiraiwa-Hasegawa, & Kutsukake, 2018; Takagi et al., 2016; Vogel & Dussutour, 2016; Xu, Spelke, & Goddard, 2005). It is indeed much more difficult to control conditions and manipulate variables, effect triggers and so on a social level but maybe not impossible. CRs correctly, along with Bhaskar, say that one can find a proxy for experimentation in social science and more specifically on the level of social systems and processes. Such proxies are similar to what neuroscientist have to rely on when inferring brain function. It is possible to experiment on human beings by systematically destroying brain regions to see what behaviour or processes the destroyed regions are responsible for, but such acts are totally ethically unacceptable! Thus, observation of patients that have selective and specific brain damage for whatever reason is the only way to legitimately and ethically infer mechanisms. Similarly, the social scientist can rely on periods of breakdown and crisis to infer certain mechanisms in social systems. Heidegger’s hammer presents a very useful and creative metaphor for determining function in social systems. Only when a hammer breaks and cannot fulfil its function, can one surmise what it is supposed to do. Danermark et al. (2002, p. 100) discuss a few strategies which are a combination of conceptual and empirical processes to uncover generative mechanisms, namely, counterfactual thinking, social experiments, studies of pathological cases, studies of extreme cases and comparative case studies. The one interesting empirical strategy that they discuss is social experimentation. The traditional experiment takes place in a closed system and the traditional CR objection is that social systems cannot be closed thus precluding experimentation. However, the social experiment Danermark et al. (2002) talk about is discussed with an example from ethnomethodology. It does not require closure, merely disruption of current patterns of behaviours. For example, by giving an unexpected answer when someone asks “how are you?” the usual assumed behaviour is disrupted and people’s consequent responses (like regarding the off-beat answer as a joke) tells the observer much about the conditions for maintaining relationships. While Danermark et al. (2002) point out that this experiment consist of the researcher’s behaviour that might trigger particular responses, it does not pretend to have achieved closure. This type of experiment is nothing new and is known as field experimentation in research methods text books (Christensen, Johnson, & Turner, 2015). It does not require closure and takes place in field or natural situations. Although Danermark et al. (2002) within the CR framework support the Bhaskarian principle of the impossibility of experimentation in social science, his use of the example of social

Critical Realism

115

experimentation illustrates a very important point. First, the way Danermark et al. (2002) frame social experimentation is in line with Bhaskar’s (1998a, p. 52) view of a proxy for experiment or closure, namely, that situations of crisis or breakdown reveal social mechanisms—we can call it the (Heideggerian) breakdown principle. All the strategies, including Mäki’s (2005a) use of a model as experimentation, in some way are based on difference with the usual state of affairs brought about conceptually or empirically, thus, breakdown or disruption. However, in the social or field experiment it is the researcher or protagonist who actively causes the disruption. In this sense it is similar to the classical experiment. Secondly, as any undergraduate psychology student would remember from her research methods course, quasi-experimentation is also a mode of experimentation without full closure (Shadish et al., 2001). It is used in psychology and social science and has a rich theoretical and empirical history. In the light of Bhaskar’s transcendental argument one should ask: given that quasi-experimentation and social/natural experimentation do happen, what are the characteristics of social reality that allow this? The requirement for experimentation is closure, among others, but I am not convinced that one can postulate a hiatus between natural and social reality such as Bhaskar does, namely, that all reality is open but only natural reality can be closed. In addition, science is not defined by experimentation.

The Nature of Social Reality Bhaskar’s (1998b, p. 15) transcendental argument about social reality led him to the following question: “what properties do societies and people possess that might make them possible objects of knowledge for us?” We know that natural reality is mind-independent and consists of deeper levels housing generative mechanisms or in Bhaskar’s words, it is intransitive and structured. Can the same be said of social reality? At first glance social reality is concept-dependent thus definitely not mindindependent. Social reality is dependent on the thoughts and actions of people: if there were no people there would be no social reality in contrast to natural reality. As we have seen above this is precisely the point Mäki battled with while trying to make realism respectable for social sciences. It will be remembered that Mäki proposed social reality to be science-independent to a large extent which means that social reality is not determined by those studying it. However, for Bhaskar (1998b, p. 51) the concept-dependence of the social becomes doubly so because of the influence of social scientists studying the social. Although he allows for the possibility of the social changing by being studied—in fact we have transitive knowledge formed of transitive knowledge—Bhaskar (1998b, p. 51) makes a distinction between causal interdependency (of social objects and investigation) and existential intransitivity. Objects formed in the social sphere, thus depending on causal interdependency, exists independently as objects and in this sense can be studied. This is exactly the point Mäki made: despite concept-dependence social objects remain existentially intransitive and can be studied. Thus, we are saying that it is this existential

116

5 Realism in Psychological Science

intransitivity that should be the focus of a definition of realism: it applies to both the social and natural domains. Existential intransitivity implies a certain sense of mindindependence and this is what one should retain for realism. The characteristics of social reality then according to Bhaskar (1998b, p. 42) are threefold: (a) Social objects regulate certain activities but do not exist apart from these activities. An example would be the object or social structure of a judicial system that governs the actions of judges. Or take another example expressing the relational nature of social objects: a soldier makes only sense within a structure or institution such as the military, or “the cashing of a cheque [implies] a banking system” (Bhaskar, 1998b, p. 30). If there were no such structures as the military, government, civil service and so on the activities one normally engages in do not make sense. Of course, natural objects do not cease to exist when their activities stop. Bhaskar (2008, p. 7) formulated an important principle which marks the difference between CR and forms of positivism and empiricism: “Tendencies may be possessed unexercised, exercised unrealised, and realized unperceived (or undetected) by men . . . .” CR prefers to refer to laws as tendencies to distinguish the realist conception of a law from the empiricist one. The latter, as we have seen, is purely regulatory based, i.e., observing a constant regularity of events. The situation where these events are observed is usually in an experiment where a system was closed. The implication of an experiment is that (1) it is the experimenter triggering the law and (2) the law (as observable events following a sequence) was observable only in the closed situation. Clearly, laws do not depend on people neither on closed situations because they ought to apply without people and in open situations. As soon as one shifts the burden of explanation to a deeper, usually unobserved, level then the things or mechanisms responsible for surface events become our target for understanding reality. Mechanisms or structures may or may not execute their powers. This does not mean that mechanisms do not have those powers or that they do not exist—the vehicle able to reach 200 km/h but never gets the opportunity to do so is a case in point. If they get exercised then they might not realise—think about gravity exerting its power but confounded by the chair I am sitting on. If the mechanism realises its power and it gets exercised it might not be observed like the flower blooming deep in the Kalahari where no one will ever see it. CR talks about tendency to express the concept of a law which is not based on a regularity causal principle. Tendencies, furthermore, encapsulate the ability of things to act and have effects on other things, i.e., tendencies are their mechanisms or structures. (b) Social objects or structures depend among other on the agents’ conceptions of the objects. This is not the case in nature. (c) Social structures or objects are relatively enduring while those in the natural sphere are more or less permanent. The conclusion is that the concept dependence of social objects cannot exclude the social domain from scientific study. A

Critical Realism

117

simple example should suffice. Every student knows that a class timetable is very real even though one cannot equate the timetable to the printed hardcopy I have in my hand. It is a social object, the reality of which was constituted by the reciprocal interaction between individuals and social structures. What is the relationship between the individual and society? We know that social constructionism, especially strong versions, reject the existence of the individual and negate agency and intentionality of the individual (Kvale, 1992). Primacy is given to the social and its influence on the so-called individual. Like social constructionism, Bhaskar rejects atomism on ontological and methodological levels. Atomism as the belief that individual and separate events govern reality and that our task is to find event regularities underlies empiricism and positivism. However, unlike social constructionism, CR places equal emphasis on the individual and society. Both are different strata of reality, constituting each other but irreducible to the other. The lesson is that realism acknowledges the importance of the intentional agent: Society, then, is an articulated ensemble of tendencies and powers which, unlike natural ones, exist only as long as they (or at least some of them) are being exercised; are exercised in the last instance via the intentional activity of human beings; and are not necessarily space-time invariant (Bhaskar, 1998b, p. 39).

Bhaskar (1998b, pp. 34, 36) provides a wonderful example, which is based on the way language functions, of the “transformational” model he proposes. A person is born into a culture and language—language pre-exists the individual and the individual cannot speak (communicate) without this pre-existing language. Although language pre-exists, it can be changed in the course of speech, thus, developing new phrases and ways of expressing that are taken up in the tradition of the particular language. The relationship between the speaker (individual) and language (social phenomenon) is hereby clarified. It is not only a case of dialectical interaction, i.e., one creates and the other forms because it presupposes a direct relationship between two ontological things, viz., people and society. Rather, “Society stands to individuals . . . as something they never make, but exists only in virtue of their activity” (Bhaskar, 1998b, p. 34). People and society are two distinct ontological things, but the one exists because of the other. For instance, if there is no speaker then language does not exist, but language is required for speaking. Hence, social phenomena pre-exist individual actions but require actions for their existence. Bhaskar also says that people do not create society but transform it, which is illustrated by the language example above: language can be transformed by speech by incorporating speech forms over time. That society pre-exists but requires human action for its existence, relates almost conversely to the role of the individual: people produce and transform social things but whilst doing that they enable the social things to exist. Bhaskar calls the first the “duality of structure” and the second “duality of praxis,” and summarises the relationship between structure and praxis within their dual functions as follows: “Society is both the ever-present condition (material

118

5 Realism in Psychological Science

cause) and the continually reproduced outcome of human agency. And praxis is both work, that is, conscious production, and (normally unconscious) reproduction of the conditions of production, that is society” (Bhaskar, 1998b, p. 35). Interestingly, the unintended consequences of individual actions or praxis are a reproduction of the conditions of society: we do not speak to keep language alive but to communicate, but a consequence of speaking is a perpetuation of language. The transformational model of social action (TMSA) acknowledges the interdependence between society and individuals but maintains their ontological distinctiveness. Thus, reduction is avoided. It also allows for history by securing the continuity of social structures whilst granting change by means of transformation to take place (cf. Bhaskar, 1998b, p. 37).

The Process of Science Bhaskar views science not as a deductivist process, i.e., utilising laws and particular instances of events to infer what caused a state of affairs. A simple example of this hypothetico-deductive process is: (a) Water freezes at 0  C, (b) the water in the bird bowel froze, (c) thus, it was 0  C outside. A deductivist explanation does not explain much or help us understand what is going on except that it focuses on a sequence of observable events and not on a deeper mechanism enabling us to understand why water freezes as 0  C ceteris paribus. In order to get to a deeper level one cannot but move to a form of argumentation beyond induction and deduction allowing us to discover the explanatory mechanisms. The scientist uses retroduction or abductive reasoning to find a possible explanation for a state of affairs (Hartwig, 2007, p. 195). The process of science starts with an observed pattern of events, whereafter a plausible explanation is formulated. This formulation can be in the format of a model or theory, which then has to be tested empirically for its validity. This three-step, deceptively simple process, describes the actual practice of a realist scientist (Bhaskar, 2008, p. 4). According to Bhaskar (2008) the empiricist stops with the first step, namely the observation of a constant conjunction of events while the transcendental idealist stops at the model formulation step. Only the realist tests the model and may move deeper into levels of reality uncovering generative mechanisms. The first thing to notice about this process is the interplay between empirical reality and conceptual work. Above we said the only access we have to reality is via language constructing plausible conceptual models of how things work. This much the social constructionist got right. However, the ability to confront reality with our models and test them is where the realist makes headway. Some ideas might hold up, others might be revised and in the end we formulate a better understanding of how reality is constituted and operates. The process of formulating models—some call them metaphors—is a retroductive process where one, based on existing knowledge, formulate a plausible conceptual explanation of a generative mechanism or structure. Model formulation or even theory formulation, depending on the conceptual specificity of the

Critical Realism

119

framework, is essential in the investigative process of science. The model is constructed in an abductive way in order to account for the observed evidence much like the detective works (Niiniluoto, 2018, p. 4). The following step is to empirically test the model in order to determine its validity as an explanation, focusing on the generative mechanism. (Bhaskar, 2008) formulated the three-step process based on experimental work in natural science, but a similar process is involved in non-experimental natural sciences and in the social sciences. Let me clarify. Some clues were already given above and one clue relates to CR’s strict adherence to the open-closed system characteristic. Experimental work should be conceptualised differently than to what we are used to. The principles underlying experimental work, namely, identifying one mechanism responsible for observable events through closure, are reflected in both conceptual and empirical situations. As mentioned earlier, one should distinguish between true experiments and quasi-experiment meaning that some of the requirements for a true experiment can be relaxed. However, as also said above, natural empirical events such as disasters, brain damage and periods of crisis also reveal generative mechanisms. Thus, when one requirement of a true experiment, namely closure and manipulation by a human investigator cannot be met there are so-called proxies to utilise. Mäki’s view of a model as experiment comes close to what I have in mind as a conceptual experiment when empirical closure and/or manipulation are not possible.

Problem with Closure in the Natural and Social Sciences Bhaskar (2008) based his identification of closure-open system characteristic on an analysis of experimentation in natural science. In principle I do agree that closure is possible in some situations, more difficult in others and almost impossible in other situations. One should remember that experimentation depends on manipulation and closure. Hence, it is not only closure that is impossible for Bhaskar in the social sphere, but also the manipulation of a mechanism. However, it should be remembered that Bhaskar analysed the empiricist/positivist understanding of experimentation, i.e., the necessity of effecting closure in an empirical experiment based on the regularity view of causality, namely, the constant conjunctions of events. In the process of his analysis he also pointed out the realist take on what an experiment says about reality, namely, its structured, differentiated and stratified nature. In principle reality is open! This means that causal mechanisms rarely operate in phase so that we can easily pick them off the surface of our experienceable environment. Causal mechanisms work against each other, sometimes a number of mechanisms are involved in bringing about an effect, sometimes their powers lie dormant and sometimes their effects are unrealised. Our world is thus firstly, open but also complex, rarely linear and atomistic. This applies to both the social and natural domains of reality. In order for us to gain epistemic access to generative mechanisms, we have to do hard work. Indeed Bhaskar (2008, p. 239) is correct in saying that scientific training

120

5 Realism in Psychological Science

is a underanalysed area in philosophy. The necessity of training and the practice of science show that reality is stratified, structured and differentiated. Bhaskar (2008, pp. 23–26) took one activity (experimentation) as indicative of the practice of science and made conclusions about the nature of reality. Unfortunately, the result of this narrow focus is that closure became a wedge between the natural and social dimensions. Although some CRs allow a closed-open continuum in natural and social reality, experimentation incorrectly became another divisive trademark of proper science, similar to quantification, thus driving the wedge deeper between natural and social domains. Experimentation is not the hallmark of science. A realist metatheory correctly directs us in the direction of discovering generative mechanisms, the discovery of which, among others, can be obtained by experimentation when (semi)closure is possible. It is false to say that natural reality is open but in principle can be closed whilst in social dimension this is impossible. Sometimes (partial) closure is possible in the psychosocial domain, but again this possibility does not indicate a defining characteristic of science. The fact that CRs preclude experimentation from the psychosocial dimension reiterates the mythic understanding of science based on the received view of positivism. The result is that another principle merely encourages a dualism between natural and social reality. I say “another” because we have examined a number of these divisive principles that are symptoms of us saying that psychosocial reality is qualitatively different from the natural domain. Collier (1994, p. 121) says “A stratified world, as we have seen, is an open world—a world which does not naturally produce closed systems. But it is also a world in which we can produce closed systems at some strata.” The same applies to the social part of our world.

Lessons Learned CR clarifies the realist principle of mind-independence and what it means for both the natural and social dimensions of reality. Mind-, concept- or scienceindependence express the same sentiment, namely, that there is something, ontologically real, that can be critically investigated by people. Even a concept-dependent reality becomes mind-independent or intransitive when confronting the inquiring mind. Secondly, CR makes a valuable distinction about the nature of social reality that seems in step with what we experience in contrast to social constructionism and other relativist forms of postmodernism. Equal emphasis is given to the individual agent and the dimension of social reality. Both are acknowledged as real, interdependent but irreducible to each other. It actually serves as a good heuristic model for viewing how strata of reality work that lie next to each other and are mutually interdependent. Bhaskar’s transformational model is one example of such a model whilst Archer (2010) provides a finely-grained and advanced model. The relationship between agency and structure illustrates CR view of emergence, mutual interdependency but ontological irreducibility of layers or reality.

Situational Realism

121

Thus, we can acknowledge a number of aspects of social constructionism without endorsing its relativism. Language is crucial for understanding reality, actually this is the only access we have to reality (in science). Secondly, we recognise the ephemeral nature of our constructions of reality: our theories are fallible and tentative but against the social constructionist we say that our constructions are about something. It is this “something” we would like to understand and we avoid the epistemic fallacy of reifying our constructions. We do not have to downplay the importance of the individual but neither need we fall into individualism and atomism. Bhaskar’s (2008) analysis of the experiment, although invaluable, incorrectly places an ontological limit on experimentation. If there is a qualitative difference between natural and social dimensions then it lies with the role of conceptdependence and not so much with closure/openness. At the very least one can acknowledge that reality is open.

Situational Realism Some additional lessons can be learned from other versions of realism. Although as we can see realism comes in many guises as Harré (1986) pointed out but certain basic principles underlie most of them. We have seen that CR advocates a stratified, differentiated and structured reality. It also allows for social reality to be studied scientifically like natural reality but in a different way. However, the difference can become a wedge threatening an ontological divide between natural and social reality because of Bhaskar’s insistence that closure is not possible in social reality. Before we thus fall into the trap of ontic dualism, we have to find a way to affirm the nature of reality. One realist philosopher, John Anderson (1893–1962), a Scottish-Australian and contemporary of Wittgenstein, emphasised an egalitarian ontology. Anderson’s brand of realism is currently known as situational realism and he is not very well-known outside of Australian academia (Hibberd, 2011, p. 120; Mackay & Petocz, 2011a, p. 28). A number of authors and scholars are making his work available internationally (Baker, 1986; Boag, 2008; Hibberd, 2010, 2011; Passmore, 1962; Petocz & Mackay, 2013). It is believed that Anderson’s realism can serve as a fruitful metatheory for psychology and a collected bundle of essays were dedicated recently to this aim (Mackay & Petocz, 2011b). I will not embark on a full discussion of situational realism and not even attempt a critical overview but would like to point out some principles that can serve the purpose of establishing realism as a metatheory for social sciences and psychology. Anderson’s work is easily misunderstood and one must be careful not to misrepresent his work especially since it is difficult to locate his philosophy among his contemporaries (Cole, 2012). His relative isolation in Australia, development of thought without necessarily being influenced by international developments, such as the linguistic turn in philosophy, along with focusing on teaching and publishing in local journals, probably restricted his influence (Cole, 2012; Hibberd, 2011, p. 123; Mackay & Petocz, 2011a, p. 29). He reacted mainly against the prevailing idealism

122

5 Realism in Psychological Science

in philosophy and in the process established principles for the scientific investigating of social and natural reality (Mackay & Petocz, 2011a, p. 28). The Hegelian or objective idealism he reacted against mainly held that mind constitutes reality and he subsequently firmly emphasised the mind-independence principle. However, nature and mind do not form independent ontological domains. We do not have material and immaterial substances like Descartes believed (Mackay & Petocz, 2011a, p. 35). Whatever the nature of psychological, social or natural reality, they are all part of one reality or “one single way of being” (Anderson, 1962a, p. 48). Thus, reality is one; the reality we find ourselves in spatiotemporally is all there is. Consciousness and mind are part of this one natural reality as is social reality or whatever other dimensions we can imagine such as Popper’s third world. There are no second or third worlds: only one. Hence, one can characterise his version of realism as espousing ontological egalitarianism. There are thus no more real levels or less real levels. Bhaskar (2008) makes much of reality as stratified and differentiated and as we will remember a distinction is made between three domains, namely, the empirical, actual and the real. These distinctions need not go against the grain of an ontological egalitarianism although Mackay and Petocz (2011a, p. 35) regards CR’s stratification as contradicting an egalitarian ontology. Reality having strata or three domains still means it is one reality exhibiting certain characteristics. Our main concern in this section, however is making a distinction between two separate ways of being, namely, natural and the social. In contrast to Bhaskar’s wedge we can affirm, along with Anderson’s situational realism, that ontology is egalitarian or one. Any explanation we have about social and natural domains should not differ in basic structure or require us to invoke more than one ontological domain. Anderson does not expect us to treat social and natural objects the same and is well aware that different methods might apply. In fact, he makes a strong case for qualitative methods in social and psychological science but denies that their structure differ from that of methods used in natural sciences (Hibberd, 2011, p. 124). Anderson denies the principled distinction between ideographic and monothetic approaches in social and natural science because these approaches are supposedly built on an ontological distinction between different realities (Hibberd, 2011, p. 124). Reality is one and epistemic access should be based on subject matter. This means that the positivistic informed mythic image of science’s requirement of empirical experiment and mathematical description for social science need not be fulfilled so that social and psychological sciences may qualify as “science.” Ontological egalitarianism does not imply methodological unity based on the physical sciences, as required by versions of positivism and empiricism. Hibberd (2011, pp. 124–125) points out the structural similarities of the scientific process underlying natural and social science: Both are underpinned by the same set of general categories, both aim to study types of situations or occurrences in space and time, systems of interacting processes, i.e., both aim to discover what is the case. In both there are practical obstacles to objectivity, including psycho-social-political interests and forces and the sheer complexity of the phenomena

Situational Realism

123

studied. Both involve the making of interpretations, both have a unifying method, viz., critical inquiry, and measurement is not essential to either.

Anderson’s realism, as does CR, holds that epistemology cannot guide ontological enquiry. Ontology is primary and the epistemic fallacy should be avoided. For Anderson starting with what we can know is a Kantian mistake leading to idealism and must be avoided at all cost. One mistake idealism and variants make is to employ the principle of constitutive relationships. Thus, the reality of things is established in the mind of the observer. A constitutive relationship denies the three terms of any relationship, namely aRb where a and b are two relata and R the relationship. Constitutive relationships indicate that the existence of either a or b is dependent on the relationship with the other. Anderson does not deny that a and b are connected, but they remain distinct thus relationships are fundamentally non-constitutive. Non-constitutive relations mean that something cannot stand in a relationship to itself, i.e., aRa. According to SR then what is called an internal relation say, of object a, is non-constitutive relation between parts of a, i.e., another situation/state of affairs embedded in the state of affairs object a forms part of (Hibberd, 2011, p. 136). Hibberd (2011) and others show how the principle of non-constitutive relationships can be applied to some theories in psychology, especially cognitive psychology and representationalism. Suffice to say that Anderson and situational realists deny the validity of representationalism or the fact that there is a Humean idea or representation in the mind that mediates interaction between the world and the individual mind. However these issues need not defer us now; the interest reader is referred to Mackay and Petocz (2011b) and Charles (2012). The important issue for Anderson is that the principle of non-constitutive relations implies things can be connected because they are distinct. This principle that things are distinct yet connected flows from non-constitutive relations. Because whatever exists in space-time is related to other things, yet not constitutively so, this does not mean that relations are not real. They are real because things are distinct yet connected. This principle of distinct yet connected has the same function as the transformative relationship between agent and structure formulated by Bhaskar (1998b) and Archer (2010). I will come back to this point below. Things are connected in various ways: the cow stands before the barn or the timetable tells me when class starts. None of the elements of these relations or connectedness establishes the other. If I remind myself about the time my class starts then I have not established the reality of the timetable in my mind: timetables and people are separate but in certain situations they are connected although distinct. Anderson’s concept for things and events that are real is this situation: Whatever exists are occurrences or situations in space-time. There is nothing less than the situation and nothing but situations exist. Anything that exists or occurs, whether it be a stone, an emotion, an individual, a social movement, an event, a war, a Federal election, a causal process, etc., consists of situations or occurrences (Hibberd, 2011, p. 126).

The concept of a situation is thus the expression for what is the case, and what can be claimed to be the case, spatiotemporally; it is imbedded in other situations, it is

124

5 Realism in Psychological Science

complex, it expresses a single way of being, in a situation things are connected yet distinct, related but not constituted by relations (Hibberd, 2011, pp. 126–127; Mackay & Petocz, 2011a, pp. 35, 39–40). A thing such as a cow cannot occur a-spatiotemporally and not outside of a particular context or situation. The relations the cow enters into is determined by the situation. The claim “the cow stands in front of the barn” expresses the complex state of affairs or situation, and it is related to other complex situations such as “I cannot see the cow before the barn because a tree is in the way.” One can probably call a situation an occurrence or a context, but the concept of situation in situational realism expresses the complex nature of whatever exists. Situations exhaust the real, i.e., these are all there are and what can be used to describe the real coherently. The properties of objects within a situation have an effect on other things and powers influence events. A situation is embedded or nested in others which cannot be reduced to these other situations but remains connected. Cows, natural objects and powers are not the only things forming situations. Hibberd (2011, pp. 127–128) says “Social conventions, practices, institutions, movements and ways of life are no different either. They consist of myriad situations and, as irreducible social complexes, have their own qualities or characteristic ways of working not possessed by their components . . . They too are distinct but connected, pluralistic complexes grounded in space and time.” What is telling about this statement by Hibberd (2011, p. 128) is that social things are “irreducible social complexes, have their own qualities or characteristic ways of working not possessed by their components.” Thus, situations forming larger or smaller nested or embedded situations means that situation A can be formed from component situations although situation A is distinct and cannot be reduced to its components. Situation A similarly can form a component of another situation B. Hibberd (2011, pp. 127–128) emphatically states that nested or embedded does not imply levels of reality in the manner of Bhaskar’s three domains or even strata such as psychological-social strata or levels. One can understand that situational realists want to maintain the egalitarian sentiment of situational realism by shying away from saying that embeddedness refers to a process of emergence because this would imply higher and lower levels. However, “irreducible social complexes” having qualities not possessed by their components perform the same function as Bhaskar’s (1998b) transformational model of social activity (TMSA): individuals and societies are ontologically distinct but related. As against atomism and holism, Bhaskar’s emergence theory allows us to conceive of real, irreducible wholes which are both composed of parts that are themselves real irreducible wholes, and are in turn parts of larger wholes, with each level of this hierarchy of composition having its own peculiar mechanisms and emergent powers. (Collier, 1994, p. 117)

Anderson and those describing Anderson’s situations talk about events and objects (Hibberd, 2011; Mackay & Petocz, 2011a). We have been sensitised by CR to regard events with suspicion especially when using them in a Humean manner to characterise causality. However, situational realism’s view of causality is on par with that of CR. Situational realism rejects a regularity view of causality and Anderson says causality is not linear and also not singular (Mackay & Petocz, 2011a, p. 39). Causality and causal relations are complex, happens within a causal

Situational Realism

125

field, comprises action and effects at all points. The production of events in one causal field might not apply in another according to Mackay and Petocz (2011a, p. 39), which reminds one of CR’s view of causal mechanisms having “Tendencies [that] may be possessed unexercised, exercised unrealised, and realized unperceived (or undetected) by men . . .” (Bhaskar, 2008, p. 7). Although situational realism does not utilise the language of mechanisms, it provides a rich concept of causality within spatio-temporal situations. Situational realism might have some difficulty endorsing the idea of a mechanism because it implies depth ontology as opposed to an egalitarian one but this can be analysed at another time. However, I do not think that Bhaskar’s depth ontology and an egalitarian ontology is incompatible as mentioned above. Although much more can be said of Anderson’s theory, I would like to point out one last point for our purposes. Anderson, contra Kant, regards the categories not as structures of the mind but of reality. A category such as causality is not imposed by the mind on reality, it is a characteristic of reality. One of the important points flowing from this is that Anderson views logic as ontological. Thus, when employing logic one is not just exercising a epistemological heuristic. Reality complies to the basic four syllogistic statements of logic and again Anderson’s views should be interpreted in relation to the rich analytic philosophical tradition, a discussion we need to defer for now. However, the implication of his view is that when making a claim about reality, namely x is y, one is expressing something that can be challenged. Thus, someone can say x is not y but a. Herein lies the root of scientific criticism, the ability to make claims and counter claims. According to Mackay and Petocz (2011a, p. 43) science can be seen as “critical inquiry” but not . . . working within a paradigm (an idea largely inspired by the relativist thesis that paradigms generate their own conditions of truth), nor is it the application of a set of techniques; it is not the application of the experimental method, nor is it quantitative analysis, though it may involve any of these. It neither provides the imprimatur of certified truth (a thesis often wrongly attributed to realism), nor is it merely another social activity, for it must be understood as a cognitive activity.

Although Mackay and Petocz (2011a, p. 43) say that situational realism’s view of science as critical inquiry is not a principle specific to realism—a correct claim based on what we found in social constructionism—but situational realism certainly provides an ontological basis for science as criticism. Anderson’s view of logic and separability as rooted ontically means that our epistemic claims about this reality can be challenged and may be wrong. Knowledge is fallible and corrigible. It is this principle of claim/counter-claim that makes criticism possible and forms the structure of scientific method which I referred to above and underlies methods in all sciences. For Anderson both empirical and conceptual work conform to this structure aptly expressed by Mackay and Petocz (2011a, p. 43): Importantly for psychology, the element of critical evaluation in science is not merely empirical: Conceptual examination of theories’ claims and inferences is just as much part of science, and as much part of the testing of empirical hypotheses, as are the various empirical and observational tests which follow. The principal implication for psychology . . .

126

5 Realism in Psychological Science

is that the examination of the conceptual bases of psychology is just as much science as empirical testing, indeed no amount of the latter can substitute for the former.

Anderson’s egalitarian ontology is worth pursuing for the major reason that it emphasises a single way of being. Thus, despite the different ways we categorise and describe reality, natural, psychological or social, everything forms part of a single natural reality. We have to manoeuvre our philosophising within this single, natural way of being. The single way of being allows us to endorse the independence of things unequivocally. Everything that exists are independent of minds or as CR says, intransitive: Realists are united by the claim that the world exists independently of mind: Mind does not constitute the world. Of course, mental processes are part of the world, and it is analytically true that if there were no cognising organisms there would be no mind or mental processes. But neither the non-mental parts of the world, nor the mental processes within the world, exist by virtue of being thought about (Mackay & Petocz, 2011a, p. 34).

This claim must be read against what we have found about mind-independence from minimal realism and critical realism. A first-order claim is that some things exists because of minds and should minds disappear these to some extent will also (I say “to some extent” because some concrete evidence would remain such as money lying around, books and so on). A second order claim is that conceptdependent things can be studied as mind-independent things. Because of this mind-independence claims can be made about everything that exists; we can even make claims about thoughts, minds and consciousness. Making a claim about things means expressing something true about reality that can be contested, corrected and tested. The possibility of claim and counter-claim constitutes criticism, and criticism underlies the scientific process. One may claim that criticism forms the structure of the scientific process.

Nuovo Realismo The last stop in this journey for now is a relatively new realism originating surprisingly in continental philosophy, called new realism. I say “surprisingly,” because continental philosophy is mainly the recruiting ground for qualitative and likeminded researchers’ anti-science and postmodern interests. Different philosophers asserted the importance of ontology for philosophy and science and can be gathered loosely under the label of continental realism. Two main trends can be distinguished, namely, speculative realism and new realism with the latter also called common sense, or positive realism (De Sanctis & Santarcangelo, 2015). The philosophers are united in terms of their origin, namely, continental philosophy and their aim, namely, the rehabilitation of the real. Among others, they include De Landa (2002), Ferraris (2015a), Gabriel (2015), Harman (2002) and Meillassoux (2008). A number of recent publications appeared by these philosophers and are well worth exploring especially by those interested in science theory and realism

Nuovo Realismo

127

(De Sanctis & Santarcangelo, 2015; Gratton, 2014). An interesting observation is that the trend in continental philosophy was to oppose analytical philosophy especially on grounds of the early twentieth century empiricist hegemony in science. Continental philosophies and philosophers were recruited for anti-empiricist and anti-positivist views and especially in qualitative methods developed along the lines of the ideographic-monothetic tradition. For our purpose we can make a distinction between new realism and speculative realism within the larger context of continental realism. I will briefly discuss the work of Ferraris (2014, 2015a, 2015b) who can be called a new realist and whose work on realism predates that of a number of other continental realists. The main aim of the new and speculative realists, despite differences between them, is to dethrone the subject in an even more drastic manner than aimed for by postmodernists. Ferraris (2014, 2015a) would like to counter Derrida’s (1997, p. 163) claim that “there is nothing outside the text.” As we have seen earlier the implication of this statement amounts to strong constructionism, namely, that the world is created or constructed by language. Language becomes the medium whereby social realities are constructed. Language expresses thought so if the community believes what is expressed then that becomes real. Social reality is thus fundamentally conceptdependent. The strong thesis implies that the world we live in is constructed by virtue of a community of like-minded people. Thus, conceptuality or epistemology and ontology are collapsed. Ferraris (2014, p. 19) calls this the being-knowledge or the transcendental fallacy which is the same as Bhaskar’s epistemic fallacy and Anderson’s principle of non-constitutive relations. The continental realists would like to assert the primacy of ontology by unseating epistemology radically, namely, by focusing on the implications of a world of objects that precedes humanity and which can even take humanity out of the epistemological equation. The universe existed for billions of years of which a million or two is devoted to some kind of human consciousness and mentality, a very small portion of the total history of the universe. This sharply illustrates the hubris of humanity that thinks consciousness and mentality are primary and also places Kant’s transcendental move of reality conforming to mental categories in sharp relief. It is not denied that mind arose from being, but pitted against a world and reality that happily existed and evolved for billions of years, its puny nature is revealed. Thus, according to Ferraris (2014, p. 30) Kant’s Copernican revolution was more of a Ptolemaic one because he re-centered the universe on humanity and its cognitive capacity. In order to illustrate the particular nature of reality not depending on conceptuality Ferraris (2014, pp. 28–30) describes his slipper thought experiment: imagine a slipper on the carpet. Five objects encounter the slipper, namely, a human being, a dog, a worm, an ivy and a slipper. If I should ask you to pass me the slipper, you would be able to do it almost without thinking, because the slipper is there on the carpet. Its presence is rather obvious and different from us debating the advantages of Brexit. However, a common household object like a slipper makes its presence felt to us both because we know what a slipper is and what it is used for. We share a conceptual framework of something that is a slipper in an almost natural manner. If I

128

5 Realism in Psychological Science

ask the dog to fetch my slipper he would be able to do this easily if trained. His conceptual framework is different than mine and he probably does not know what slipper means or what it is used for but he is very much able to bring me my slipper. Similarly, the worm encounters the slipper and has two choices: to go around it or over it, but it still has to negotiate the slipper. The same goes for the ivy. Even another slipper thrown on the one on the carpet will encounter some resistance. To the range of beings, the slipper resists, from those with conceptual schemes and developed senses to those with none, the slipper is just there. The point of Ferraris (2014) slipper story is that some things tend to resist other things irrespective of conceptual frameworks and even the ability to sense obstacles. The fundamental characteristic of reality is what Ferraris (2014, p. 34) calls unamendability or “the fact that what we face cannot be corrected or changed by the mere use of conceptual schemes.” Unamendability characterises a reality that resists changes. At some stage in our confrontation with reality, it pushes back and refuses to budge. Like the slipper it is there not only for humans with conceptual frameworks, but for other things not sharing the same conceptual framework or maybe none at all. It posits itself and resists my efforts of construction. Of course, I can incinerate the poor slipper with a blowtorch and say “how’s that for resisting?” but still it was something posited which are now ashes and soon to be swept under the carpet. I can fondly remember my slipper and no one can deny that it existed at some stage—in fact I have proof it existed because I have documented its incineration and posted it on YouTube. The fact of my slipper existing in the past still resists in some way—I will be lying to you if I say I did not have a slipper. Ferraris slipper story that I have elaborated rather generously upon teaches us a number of things about reality. First, its unamendability as something concrete and then as something remembered. It is important to realise that historical facts, like fossils, are real (Ferraris, 2014, p. 49, 2015a). Interpretation is possible but not denial. Ferraris (2014, p. 52, 2015a) categorises the objects of reality into four classes, viz., natural objects, artefacts, social objects and ideal objects. Natural objects includes things like slippers and gravity; artefacts are objects existing in space-time that were produced by people such as cellphones and notepads; social objects are marriages, money, contractual agreements, timetables and governments; ideal objects are numbers that exist outside of space and time (Ferraris, 2015a). Artefacts and social objects differ although both are produced by humans. Social objects would cease to exist when there are no people (Ferraris, 2015a). As with Bhaskar’s condition for the possibility of the social, Ferraris regards relations as constitutive of social objects (note that this claim is different from Anderson’s non-constitutive relations principle. In Anderson’s case aRb means nothing in thing a or in thing b constitutes either, but a standing in a relation to b constitutes a new situation, thus allowing for social objects to be created: thus a paints something and b says “what lovely art, can I buy it?”). At least two parties need to agree on something for a social object to exist. This agreement usually is documented and in this sense the social object is concretised. If I do not know that I am married (as what happens in the South African illegal identity document trade), then I will happily carry on as an unmarried person until I try to get married and

Nuovo Realismo

129

some document shows that I am (Ferraris, 2014, p. 53). As Sheldon in the Big Bang Theory says, “this changes everything.” Ferraris formulated an interesting theory on social objects and social reality that in a sense can support Bhaskar’s transformational view of agency and structure. The former denies collective intentionality which Bhaskar can accommodate by means of non-reductive emergence (Ferraris, 2014, p. 60). Ferraris (2014, pp. 56, 60) proposes a plausible mechanism for the operation of Bhaskar’s (1998b, p. 34) TMSA, namely, the fact of documentality, i.e., the agreement inscribed in some way between two parties, allows people to act in accordance to the various agreements. Both the fact of relations and documentality constitute social objects. If no trace of any agreement exists then societies and agents cannot be held to it. Ferraris (2015a, 2015b) makes much of the fact that our postmodern age never saw the end of writing as predicted in the previous century. In fact, we write even more, in blogs, on phones and so on. The fact of a life documented became a reality with Facebook, YouTube and Instagram. People document life intensely by taking photos, videos and texting, and this process that Ferraris calls documentality maintains the social. Bhaskar’s transformational process of learning a language also shows that traditions and culture in more “undocumented” forms also qualifies as documentality: memory, verbal transfer, learning and teaching traditions and customs is a form of making things concrete (Ferraris, 2014, p. 60). Unamendability is not the only characteristic of the real. By resisting our efforts, reality also offers something positive, namely possibilities. Ferraris (2015a) calls this the positivity of the real and also labels his version of realism as positive realism, a term we will shy away from, for now at least, while dismantling the mythic image of science (Ferraris, 2015b). New or positive realism implies that the reality we are confronted with provides real possibilities but resists others. Ferraris (2015a) recounts a conversation between Umberto Eco and Richard Rorty where the latter said a screwdriver can be constructed as an earbud which Eco denied. A screwdriver resists certain uses but invites others. It is more appropriate to use as a weapon than an earbud! Reality’s unamendability and positivity shows that meaning inheres in reality. Of course, this claim is controversial because we have been conditioned by constructionists that it is the mind that gives meaning to things. However, Ferraris points out that before the existence of human beings, water invited nourishment and life: some things kill you, others nourish. Thus, ontology is primary and meaning inheres in reality. Let us pause for a moment and reflect on what CR would say: as we have seen, weak constructionism is compatible with CR. CR allows Tycho Brahe and Kepler to see different things when looking at a sunrise. Of course, interpretations differ. However, both are looking at something and one might be shown to be correct. This last point is what situational realism also would say: claims about what there is can be fallible as long as it is a claim about something, but crucially we make claims about things thinking that they are true. Thus, Anderson’s ontological grounding of logic, i.e., the situations when claims are true or false, reflects the intention of Ferraris’ positive realism. Reality will resist some but not other claims. In fact Ferraris goes further and says reality invites or affords some but not other

130

5 Realism in Psychological Science

interpretations. Ferraris’ (2014, p. 2) polemical thrust strikes at the heart of the postmodernist à la constructionists’ motto, namely, that “there are no facts, only interpretations.” Both Ferraris and Anderson rehabilitate facts. At some point, sticking your finger in fire has consequences: pain is not in the fire, not merely an interpretation, but as the slipper “found out”: it is a fact that the blowtorch obliterates (Ferraris, 2014, pp. 19, 36). Let us take a step back. In the discussion of Anderson I have mentioned the new realists at the turn of the last century namely Edwin Holt and colleagues (Charles, 2012). The intention of their denial of all forms of idealism was similar to the current speculative realists’ unseating of epistemology seated in the subject, albeit more radical. However, both positions have direct realist or non-representationalist implications for cognitive psychology as Hibberd (2011) pointed out. Ferraris’ realism rehabilitates perception particularly, direct access to reality (De Sanctis & Santarcangelo, 2015). It is perception that provides direct access to the unamendability of reality. Anderson’s principle of non-constitutive relations is one way of expressing the lack of correlation between mind and reality by means of intermediary representations. It is not too difficult to reformulate a realist theory of cognition along the lines of non-constitutive relations: meaning inheres not in the mind nor in the relation between the mind and objects, but in objects (Ferraris, 2015b). Objects afford certain interpretations and the mind appropriates directly, thus the relation expresses something real but cannot be substantialised. Thus, between Anderson’s realism, the erstwhile new realism and speculative realism there are some lines of commonality worth exploring especially in terms of realist implications for perception and cognition in psychology. However, this exploration cannot be done here. Underlying both versions of new realism then is the transcendental/beingknowledge fallacy. Its effect is establishing a mediated experience of reality: The consequences [of the being knowledge fallacy] are manifold and define the scene in which the modern and postmodern constructionist operates: what we see is made to depend on what we know; it is postulated that the mediation of conceptual schemes operates everywhere; and, finally, it is asserted that we never enter a relationship with things in themselves but always and only with phenomena (Ferraris, 2014, p. 31).

Importantly, the being-knowledge fallacy flowing from Kant’s “intuitions without concepts are blind,” implies that for Kant “concepts are necessary in order to have any experience, so that we need a concept even to slip on a patch of ice” (Ferraris, 2014, p. 24). Unamendability shows that Kant’s dictum cannot apply to natural reality but as Ferraris (2014, pp. 24, 56) says, it certainly is true of social reality and Derrida’s statement can become “there is nothing social outside the text.” Ferraris (2015b) thus, acknowledges the concept-dependence of social reality and objects but this fact does not reverse the primary order of being-knowledge for the realist! What have we learned from this discussion of Ferraris? The issue of ontology as something that resists our efforts to change it conceptually is primary. Unamendability expresses the intention of CR’s intransitivity very well. In principle,

Nuovo Realismo

131

reality is not mind- or concept-dependent. However, this must be a qualified claim. There is a part of our reality that is concept dependent but we must understand that even this sphere of social objects is intransitive. Thus, intransitivity indicates the non-I nature of real things. For the scientist this means that real things can be studied, be they natural, social or ideal as Ferraris (2015b) categorised real objects. The debate about the reality of ideal objects has been raging for decades and we do not need to analyse it now. Suffice to say is that according to Ferraris (2015b), ideal objects such as numbers exist independently of humans despite some of the things we have invented to work with numbers (such as a plus sign). Our main concern however, is social objects. Despite their concept-dependency they do present resistance. Think about the example we have used above, namely, the fact that someone might not know they are married (for whatever reason). The fact of a legal and binding record of a marriage does not deter the individual if she does not know about this fact. Her actions, interactions and relationships are unfettered by this particular social fact. However, as soon as she wants to wed her current fiancé, the fact of her marital status will certainly surface when she wants to register her status. Given this new knowledge about a particular social fact, her current actions will flounder upon the unamendability of her marital status. Note that we have a double concept-dependency, namely, our friend’s newly discovered knowledge of her marital status, and the societal constituted phenomenon of marriage. The force of the latter makes itself felt and she can no longer plead ignorance. This social fact is real, thus unamendable, albeit concept-dependent in more than one way. It will only be dissolved if society no longer regards marriage as a fact or if humanity disappears. Thus, whilst intransitivity refers to the ontological independence of knower and known, unamendability, expresses this independence forcefully as non-epistemological. Thus, a fact such as a timetable is intransitive and unamendable despite its existential concept-dependency. Although the timetable is the creation of a community of individuals and institutions, the class room timetable resists and exerts its grip on the student relentlessly. The third characteristic of reality pointed out by Ferraris is its positivity or the possibilities it affords. The same goes for concrete natural objects but also for social and ideal objects. Numbers can only be managed in a certain way, timetables require certain actions but not others, screwdrivers can be utilised in a number of ways but not others. Something that looks like a chair can be utilised for sitting even if it is a rock, custom-made chair or lump of sand. Similarly, the strong constructionist may think that the flame waits for the community to characterise it as something real, but like fossils existed as animals before humankind, so did fire incinerate its environs long before people labelled things.

132

5 Realism in Psychological Science

“Real” Science From where do we get the idea that criticism is the fundamental hallmark or characteristic of science? There is still one last step to take. We have established the independence of ontology and epistemology and also the primary nature of ontology. The world comes first and its understanding second. The understanding of the world and its interpretation is a very human activity. With the evolution and emergence of humans and their minds, new things were established that depended for their existence on those humans. If humankind disappears, some objects will disappear immediately and others might, like fossils, remain for others to find and interpret. While humans lived on earth they developed rather sophisticated ways to deal with their reality over and above their daily experience of reality. In order to survive they needed to solve certain problems and soon enough realised the only way to do this is to be careful with their common sense intuitions. Whatever the story of the origin and development of science, one can be assured that it was based on common sense confronting reality. Reality resists and invites certain but not other strategies and those valuable for survival and problem-solving can be discovered by making claims about reality and testing them. I do not even try to imagine how primitive man at some stage developed critical and systematic strategies, but one fact is certain: if humans did not manage to overcome resistance and find possibilities they would not have survived. The same goes for every living thing on earth, as dinosaurs would attest to. The point I would like to make is simple: reality is the root of science. Humanity developed cognitive abilities for various reasons and one may just assume that enhanced cognitive abilities, mind and speech had survival value. It allowed humans to be more flexible in adapting and surviving. Of course, I am not saying that science is the only route to valid knowledge and it is obvious that humankind can survive without our current understanding of what science involves. What I am saying is that our ability to make claims about reality along with reality’s unamendability and positivity, provide the fundamental structure of science. Science does not spontaneously emerge; it requires both reality and cognitive input but our fallible confrontation with reality’s resistance is what is required for the social development of science as process and institution. This confrontation can be captured by the concept of criticism. Criticism is thus not only one of the hallmarks of science in a similar way that systematicity or objectivity is. It is the fundamental structure upon which our problem-solving approach and methods are based.

References Ainsworth, P. M. (2010). What is ontic structural realism? Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics, 41, 50–57. https://doi. org/10.1016/j.shpsb.2009.11.001.

References

133

Anderson, J. (1962a). Realism and some of its critics (1930). In J. Anderson (Ed.), Studies in empirical philosophy [A machine readable transcription] (pp. 41–59). Sydney: Australian Digital Collections. Anderson, J. (1962b). Studies in empirical philosophy [A machine readable transcription]. Retrieved from http://adc.library.usyd.edu.au/view?docId¼anderson/xml-main-texts/ander son049.xml Archer, M. S. (2010). Morphogenesis versus structuration: On combining structure and action. The British Journal of Sociology, 61(Suppl 1), 225–252. https://doi.org/10.1111/j.1468-4446.2009. 01245.x. Archer, M. S., Bhaskar, R., Collier, A., Lawson, T., & Norrie, A. (Eds.). (1998). Critical realism: Essential readings. London: Routledge. Baker, A. J. (1986). Australian realism: The systematic philosophy of John Anderson. Cambridge: Cambridge University Press. Bhaskar, R. (1975). A realist theory of science. London: Verso. Bhaskar, R. (1998a). General introduction. In M. Archer, R. Bhaskar, A. Collier, T. Lawson, & A. Norrie (Eds.), Critical realism: Essential readings (pp. ix–xxiv). London: Routledge. Bhaskar, R. (1998b). The possibility of naturalism: A philosophical critique of the contemporary human sciences (3rd ed.). London: Routledge. Bhaskar, R. (2005). The possibility of naturalism: A philosophical critique of the contemporary human sciences (3rd ed.). London: Routledge. Bhaskar, R. (2008). A realist theory of science. London: Routledge. Boag, S. (2008). ‘Mind as feeling’ or affective relations?: A contribution to the school of andersonian realism. Theory and Psychology, 18(4), 505–525. https://doi.org/10.1177/ 0959354308091841. Cartwright, N. (1998). Causation. Routledge encyclopedia of philosophy. Retrieved from https://0www.rep.routledge.com.innopac.up.ac.za/articles/causation/v-1/ Chakravartty, A. (2007). A metaphysics for scientific realism: Knowing the unobservable. Cambridge: Cambridge University Press. Charles, E. P. (2012). A new look at new realism: The psychology and philosophy of E.B. Holt. New Brunswick, NJ: Transaction Publishers. Christensen, L. B., Johnson, R. B., & Turner, L. A. (2015). Research methods, design, and analysis (12th ed.). Boston: Pearson. Cole, C. M. (2012). John Anderson. The Stanford encyclopedia of philosophy. Retrieved from http://0-plato.stanford.edu.innopac.up.ac.za/entries/anderson-john/ Collier, A. (1994). Critical realism: An introduction to Roy Bhaskar’s philosophy. London: Verso. Collier, A. (1995). Critical realism: An introduction to Roy Bhaskar’s philosophy. Radical Philosophy, 69, 48–49. Collier, A. (1997). Critical realism. International Studies in Philosophy, 29(2), 120–122. Collier, A. (1998). Critical realism. Routledge encyclopedia of philosophy. Retrieved from http:// www.rep.routldege.com/articles/thematic/critical-realism/v-1 Cruickshank, J. (2004). A tale of two ontologies: An immanent critique of critical realism. The Sociological Review, 52(4), 567–585. https://doi.org/10.1111/j.1467-954X.2004.00496.x. Cruickshank, J. (2012). Positioning positivism, critical realism and social constructionism in the health sciences: A philosophical orientation. Nursing Inquiry, 19(1), 71–82. https://doi.org/10. 1111/j.1440-1800.2011.00558.x. Danermark, B., Ekström, M., Jakobsen, L., & Karlsson, J. C. (2002). Explaining society: Critical realism in the social sciences. London: Routledge. De Landa, M. (2002). Intensive science and virtual philosophy (Paperback edition). London: Continuum. De Pierris, G., & Friedman, M. (2013). Kant and Hume on causality. The Stanford encyclopedia of philosophy. Retrieved from http://0-plato.stanford.edu.innopac.up.ac.za/archives/win2013/ entries/kant-hume-causality/

134

5 Realism in Psychological Science

De Sanctis, S., & Santarcangelo, V. (2015). Afterword. The coral reef of reality: New philosophical realisms (S. De Sanctis, Trans.). In M. Ferraris (Ed.), Introduction to new realism. London: Bloomsbury Academic. Derrida, J. (1997). Of grammatology (G. C. Spivak, Trans. Corrected ed.). Baltimore: Johns Hopkins University Press. Falcon, A. (2015). Aristotle on causality. The Stanford encyclopedia of philosophy. Retrieved from http://0-plato.stanford.edu.innopac.up.ac.za/archives/spr2015/entries/aristotle-causality/ Ferraris, M. (2014). Manifesto of new realism (S. De Sanctis, Trans.). Albany: State University of New York Press. Ferraris, M. (2015a). Introduction to new realism (S. De Sancti, Trans.). London: Bloomsbury Academic. Ferraris, M. (2015b). Positive realism. Winchester: Zero Books. Fleetwood, S. (Ed.). (1999). Critical realism in economics: Development and debate. London: Routledge. French, S., & Ladyman, J. (2003). The dissolution of objects: Between platonism and phenomenalism. Synthese, 136(1), 73–77. Frigg, R., & Votsis, I. (2011). Everything you always wanted to know about structural realism but were afraid to ask. European Journal for Philosophy of Science, 1(2), 227–276. https://doi.org/ 10.1007/s13194-011-0025-7. Gabriel, M. (2015). Why the world does not exist (English edition). Cambridge: Polity Press. Gratton, P. (2014). Speculative realism: Problems and prospects. London: Bloomsbury Academic. Groff, R. (2000a). The truth of the matter: Roy Bhaskar’s critical realism and the concept of alethic truth. Philosophy of the Social Sciences, 30(3), 407–435. Groff, R. (2000b). The truth of the matter: Roy Bhasker’s critical realism and the concept of alethic truth. Philosophy of the Social Sciences, 30, 407–435. Groff, R. (2004). Critical realism, post-positivism and the possibility of knowledge. Nursing Philosophy, 4(1997), 104–105. Groff, R. (2008). Introduction: Realism about causality. In R. Groff (Ed.), Revitalizing causality: Realism about causality in philosophy and social science (pp. 1–10). London: Routledge. Hacking, I. (1983). Representing and intervening: Introductory topics in the philosophy of natural science. Cambridge: Cambridge University Press. Hampshire, A., Highfield, R. R., Parkin, B. L., & Owen, A. M. (2012). Fractionating human intelligence. Neuron, 76(6), 1225–1237. https://doi.org/10.1016/j.neuron.2012.06.022. Harman, G. (2002). Tool-being: Heidegger and the metaphyics of objects. Chicago: Open Court. Harré, R. (1986). Varieties of realism: A rationale for the natural sciences. New York: Blackwell. Harré, R. (2009). Saving critical realism. Journal for the Theory of Social Behaviour, 39. https://doi. org/10.1111/j.1468-5914.2009.00403.x. Hartwig, M. (2007). Dictionary of critical realism. London: Routledge. Hibberd, F. J. (2010). Situational realism, critical realism, causation and the charge of positivism. History of the Human Sciences, 23(4), 37–51. https://doi.org/10.1177/0952695110373423. Hibberd, F. J. (2011). John Anderson’s development of (situational) realism and its bearing on psychology today. In N. Mackay & A. Petocz (Eds.), Realism and psychology: Collected essays (pp. 119–158). Boston: Brill. In memoriam: Roy Bhaskar 1944–2014. (2015). Journal of Critical Realism, 14(2), 119–136. doi: https://doi.org/10.1179/1476743015Z.00000000062 Irie, N., Hiraiwa-Hasegawa, M., & Kutsukake, N. (2018). Unique numerical competence of Asian elephants on the relative numerosity judgment task. Journal of Ethology, 37(1), 111–115. https://doi.org/10.1007/s10164-018-0563-y. Kvale, S. (1992). Psychology and postmodernism. London: Sage. Ladyman, J. (1998). What is structural realism? Studies In History and Philosophy of Science Part A, 29(3), 409–424. Ladyman, J. (2014). Structural realism. The Stanford encyclopedia of philosophy. Retrieved from http://plato.stanford.edu/archives/spr2014/entries/structural-realism/

References

135

Ladyman, J., & Ross, D. (2007). Every thing must go: Metaphysics naturalized. Oxford: Oxford University Press. Lawson, V. A., & Staeheli, L. A. (1991). On critical realism, human geography, and arcane sects! The Professional Geographer, 43, 231–232. Lopez, J. (2003). Critical realism: The difference it makes, in theory (pp. 75–89). School Sociology and Social Policy, U Nottingham: Routledge. Lopez, J., & Potter, G. (2002). After postmodernism: an introduction to critical realism. London: Athlone. Mackay, N., & Petocz, A. (2011a). Realism and the state of theory in psychology. In N. Mackay & A. Petocz (Eds.), Realism and psychology: Collected essays (pp. 17–51). Boston: Brill. Mackay, N., & Petocz, A. (Eds.). (2011b). Realism and psychology: Collected essays. Boston: Brill. Mäki, U. (2002). The dismal queen of the social sciences. In U. Mäki (Ed.), Fact and fiction in economics: Models, realism and social construction (pp. 3–34). Cambridge: Cambridge University Press. Mäki, U. (2005a). Models are experiments, experiments are models. Journal of Economic Methodology, 12(2), 303–315. https://doi.org/10.1080/13501780500086255. Mäki, U. (2005b). Reglobalizing realism by going local, or (how) should our formulations of scientific realism be informed about the sciences? Erkenntnis, 63(2), 231–251. https://doi.org/ 10.1007/s10670-005-3227-6. Mäki, U. (2011). Scientific realism as a challenge to economics (and vice versa). Journal of Economic Methodology, 18(1), 1–12. https://doi.org/10.1080/1350178X.2011.553372. Mäki, U. (2012a). Realism and antirealism about economics. In U. Mäki (Ed.), Philosophy of economics (pp. 4–24). Oxford: Elsevier. Mäki, U. (Ed.). (2012b). Philosophy of economics. Oxford: Elsevier. Meillassoux, Q. (2008). After finitude: An essay on the necessity of contingency (R. Brassier, Trans.). London: Continuum. Michell, J. (1999). Measurement in psychology: Critical history of a methodological concept. New York: Cambridge University Press. Niiniluoto, I. (2018). Truth-seeking by abduction. Cham: Springer. Passmore, J. (1962). John Anderson and Twentieth-century philosophy. In J. Anderson (Ed.), Studies in empirical philosophy [A machine readable transcription] (pp. ix–xxvi). Sydney: Australian Digital Collections. Peacocke, A. (1984). Intimations of reality: Critical realism in science and religion. Notre Dame: University of Notre Dame Press. Petocz, A., & Mackay, N. (2013). Unifying psychology through situational realism. Review of General Psychology, 17(2), 216–223. https://doi.org/10.1037/a0032937. Porpora, D. V. (2008). Sociology’s causal confusion. In R. Groff (Ed.), Revitalizing causality: Realism about causality in philosophy and social science (pp. 195–204). London: Routledge. Psillos, S. (1999). Scientific realism: How science tracks truth. London: Routledge. Putnam, H. (1975). Mathematics, matter, and method. London: Cambridge University Press. Sayer, A. (2004). Critical realism: The difference it makes. Sociology, 38(4), 856–857. Schaffer, J. (2014). The metaphysics of causation. The Stanford encyclopedia of philosophy. Retrieved from http://0-plato.stanford.edu.innopac.up.ac.za/archives/sum2014/entries/causa tion-metaphysics/ Shadish, W. R., Cook, T. D., & Campbell, D. T. (2001). Experimental and quasi-experimental designs for generalized causal inference. Boston: Houghton Mifflin. Shotter, J. (1992). Is Bhaskar’s critical realism only a theoretical realism? History of the Human Sciences, 5(3), 157–173. Sims-Schouten, W., Riley, S. C. E., & Willig, C. (2007). Critical realism in discourse analysis: A presentation of a systematic method of analysis using women’s talk of motherhood, childcare and female employment as an example. Theory and Psychology, 17(1), 101–124. https://doi. org/10.1177/0959354307073153.

136

5 Realism in Psychological Science

Takagi, S., Arahori, M., Chijiiwa, H., Tsuzuki, M., Hataji, Y., & Fujita, K. (2016). There’s no ball without noise: Cats’ prediction of an object from noise. Animal Cognition, 19(5), 1043–1047. https://doi.org/10.1007/s10071-016-1001-6. Viskovatoff, A. (2002). Critical realism and Kantian transcendental arguments. Cambridge Journal of Economics, 26(6), 697–708. https://doi.org/10.1093/cje/26.6.697. Vogel, D., & Dussutour, A. (2016). Direct transfer of learned behaviour via cell fusion in non-neural organisms. Proceedings of the Royal Society B: Biological Sciences, 283(1845), 20162382–20162386. https://doi.org/10.1098/rspb.2016.2382. White, G. (2013). Medieval theories of causation. The Stanford encyclopedia of philosophy. Retrieved from http://0-plato.stanford.edu.innopac.up.ac.za/archives/fall2013/entries/causa tion-medieval/ Xu, F., Spelke, E. S., & Goddard, S. (2005). Number sense in human infants. Developmental Science, 8(1), 88–101. https://doi.org/10.1111/j.1467-7687.2005.00395.x. Zhang, J., Cheng, W., Liu, Z., Zhang, K., Lei, X., Yao, Y., et al. (2016). Neural, electrophysiological and anatomical basis of brain-network variability and its characteristic changes in mental disorders. Brain, 139(8), 2307–2321. https://doi.org/10.1093/brain/aww143.

Chapter 6

The Realist Image of Science

As we have seen the mythic image of science created and sustained within the nexus of the three dichotomies, namely, qualitative-quantitative, scientist-practitioner and positivist-constructionist, was not based on actual Logical Positivism but on a version of naive realism and related misunderstandings Michell (2003b, p. 17). It is this constructed and mythic image that has to bear the brunt of criticism. The three dichotomies, as we have argued, lie on three levels of analyses, namely, methodological, applicative and metatheoretical. The fact that qualitative proponents, practitioners or constructionists build their case on a mythic image of science does not mean that positivism is not to blame for a number of science theoretical sins (Michell, 2003b). As I will briefly discuss in this chapter, one of the major problems for various critics of the mythic image of science is the issue of measurement. One should realise that the force of the quantitative imperative is real in psychosocial science, albeit based on the mythic image of science. Logical Positivism cannot escape blame: its fundamental empiricist view along with the demand for operationism stoke the fires of measurement fiercely and psychology all too willingly incorporated this false requirement for the scientific character of a science into its own perceptions of what psychological science is.

Science as Criticism In the last chapter I have claimed that the fundamental structure of science is criticism. Only realism can sustain a view of science as criticism. Looking at constructionism we have found that criticism is a crucial activity for the postmodernist to deconstruct and dislodge commonly assumed knowledge. At least on this point realism and constructionism agrees: the fruits of epistemological labour can never be taken at face value. As Ferraris (2014) points out, Kant’s principle of intuitions without concepts are blind applies especially well to the social dimension of reality. What we can know of the concept-dependent part of our reality is fallible, © Springer Nature Switzerland AG 2020 D. J. F. Maree, Realism and Psychological Science, https://doi.org/10.1007/978-3-030-45143-1_6

137

138

6 The Realist Image of Science

flawed, changeable, and often wrong. I would argue that the same goes for the natural part of our reality as well—it is not as if transitive knowledge suddenly becomes infallible when we are dealing with other parts of reality! The social constructionist at minimum can agree with the realist about the essential function of criticism in questioning the claims about social reality. Naturally, the two diverge about the nature of reality. The constructionist holds reality, at least social reality, to be created by the mind. As we have seen in our discussion, in a sense, this is true of our concept-dependent reality! The realist holds some areas to be mind-dependent but nevertheless real. Reality can be known but precedes epistemology even if it is concept-dependent. Thus, it is true that the realist holds reality to be intransitive, unamendable and positive despite how it comes into being. Gergen (1982), however, had to get rid of the intransitivity of social reality in order to get to the admirable point that reality affords possibilities and that people are not bound by their conceptions. However, here is the problem: constructionism cannot acknowledge that claims are about something. Only realism can do this, and because it sustains a consistent first-order mind-independence, i.e., because it avoids the epistemic fallacy, it can ground criticism as the fundamental structure of science ontologically. The principles we have extracted from the discussions of realism can be specified. Ontology is primary, epistemology secondary: As we have argued, realism encourages us to take ontology seriously: it asks us to tease out, reflect upon and elaborate our ontological presuppositions. It encourages us to ask: do we really think the world is like this? Clearly this involves avoiding the epistemic fallacy so that we do not confuse reality with our knowledge of reality and end up asserting that things are so because we perceive them to be so (Ackroyd & Fleetwood, 2000, p. 15).

Therefore, realism is based on the mind-independence principle or the claim that there are things not dependent on us for their existence. Simultaneously, although mind-independence is a requirement for realism, realism maintains that mind is just as much part of its repertoire as nature. Thus, in support of ontological egalitarianism, reality and what it produces are regarded as one and we are careful not drive a wedge between natural and social reality. Although mind-independence can be regarded as a first-order requirement for the reality of objects in natural domain, it only becomes applicable as a second-order requirement for those psychosocial objects dependent for their existence on human minds. Thus, it becomes a science or investigative independence allowing scientists and researchers to be confronted with its unamendability or resistance as real objects. Interestingly, the unamendability (and by implication, the possibilities) of psychosocial objects is felt by everyone, not only investigators, although not in the same way. The timetable exerts its power on those within a particular community, thus while it is minddependent on a first-order level, it still is very real although not in the same way the researcher investigates the interplay between behaviour and timetables. Thus, in a sense constructionists are correct in allowing some things to become real with the difference that there is then something to make claims about. The unamendability

The Mythic Image of Science

139

(possibility) of social objects making themselves felt in the behaviour and beliefs of people allows one to claim, even about social objects, x is y. The characteristic of unamendability (possibility) of reality allows us to find the fundamental structure of science. To find where and when reality resists, and where it affords possibilities and how, is characteristic of science. We should note that identifying the critical thrust when dealing with reality does not tell us how to find or unlock its resistance and possibilities. Popper defined it as problem-solving or conjecture and refutation, others as deconstruction and so on. What the mythic image of science lead us to believe was that our critical engagement with our realities requires quantification and experiment or some other method, but no one successful method exists. The nature of the resistance from reality provides a clue how we should critically investigate, sometimes by measuring, other times by, for instance, interpreting, unmasking, critically engaging or understanding. Thus, to drive these points home, let us recap the construction of the mythical image of science from its core, namely, from the metatheory of constructionism. The metatheory motivates the epistemological and ontological characteristics of the mythic image of science that was adopted from Logical Positivism and utilised in the methodological dichotomy and applicative split.

The Mythic Image of Science In order to free humankind from its own prison, Gergen had to point out how current and past traditions in science (and psychology) and socio-political culture collaborated to imprison communities and societies. It is no surprise that Foucault and related thinkers could so easily be recruited to support our postmodern deconstruction of modern society and science. However, the realist does not deny that we have embroiled ourselves in false binaries, dichotomies, polarisations, power struggles, suppression of voices, and collaborated in silencing minorities. In fact, the realist can acknowledge that the suppression of the poor and powerless arises from epistemological fallacies and that these understandings or concept-dependent objects we create, such as human trafficking, heterosexuality, marriage, various religious and cultural traditions, healing practices and so on are very real to those living within these communities. The realist need not fall back on any relativising strategy to convince anyone that there is nothing out there binding us, that there is no absolute truth which means that one’s situation is changeable and one can in principle be whatever one likes to be. One need not as realist convince people that depression is not real in order to point out different possibilities. Even if it is merely a label and it might turn out that we were wrong in labelling certain behaviours in a particular way, it will be reality (natural, social or ideal) that shows us the way out because we critically engage with the fallible claims we make about reality. Gergen (1982) blamed the image of science for the existential problems we landed ourselves in and I cannot agree more. What he sketches is a particular salient and clear view of what I am calling the mythic image of science. Gergen (1982, p. 7)

140

6 The Realist Image of Science

does not call this the mythic image of science, but the received view informed by the Logical Positivist image of science and he attributes much of what is wrong with social science to the LP-image of science. Modern science, according to Gergen (1982) developed, among others, to extract and provide certainty from the flux of changing experience. Because human experience is so changeable it is understandable that certainty in terms of controlling our future by means of predictable ways of managing our lives is set as a primary goal. Uncertainty fosters a lack of manageability and control, thus, science should be able to assist in identifying patterns, called laws, which in their turn enable us to advance and progress. The picture Gergen sketches of natural science is simultaneously mixed with valuable insights about the process of the sciences, at least in the natural sphere, but also with incorrect assumptions, which have been constructed over time. First, those aspects associated with the perceptions of Logical Positivism can be identified, secondly, aspects that in fact characterise science can be identified and lastly, Gergen’s juxtaposition of natural and social science can be identified. (a) The natural sciences are much more successful in identifying laws and establishing objective knowledge than the socio-behavioural sciences. The principles associated with the Logical Positivist image of science are the following (Gergen, 1982, pp. 6–11). Science aims at establishing certain knowledge by basing facts on observation. Facts thus established become incorrigible knowledge and serves as the basis for the ongoing progression of science. Science is thus cumulative and progressive. Empirical knowledge, i.e., facts based on observation serves to correct theories. In addition, the natural science paradigm or specifically physics serve as model for other sciences, thus enabling a unified science. Science provide rules of inquiry (deductive, inductive, hypothetico-deductive, and so on). At the basis of natural sciences is the quest to identify patterns usually based on observing regular events. Established patterns form laws and enable prediction of future events. Usually these patterns or laws can be represented mathematically. Laws apply universally. The mythic view of science is based on the stability of relationships. If there were no stable relationships or patterns of events then empirical testing and prediction would be impossible (Gergen, 1982, p. 11). Gergen says that most parts of natural science are applicable and useful especially when targeting the patterns of stability in reality. It is impossible, for instance to predict the movement of a grain of sand in an avalanche, but these types of events are not the central focus of natural science (Gergen, 1982, p. 13). Gergen also grants that the natural and social domains are not delivered to chaos and complete flux—survival would be impossible. Where regularities and predictability obtain, such as with the human body, its chemical, biological and neurological components, natural science disciplines can apply and work with no serious problem. However, the subject matter of social science and natural science differ to such an extent that predictability and regularity cannot or with difficulty be found in the psychosocial domains. The traditional natural science paradigm thus does not apply (Gergen, 1982, p. 13). Gergen (1982, p. 31) echoes

The Mythic Image of Science

141

Bhaskar’s characterisation of positivist science, namely, the image psychology tends to adopt: For example, traditional psychology generally adopts a Humean conception of causality along with a positivist orientation toward observation. The function of science, within this framework, is to locate reliable patterns of contingency among observable events. Consistent with this framework is the traditional stimulus-response paradigm.

However, Gergen (1982, p. 25) accidentally managed to strike realist gold in between the myths of empirical science: “In addition to specifying sequence, normal science typically entails an elaboration of the underlying mechanisms or processes.” This typical CR depth explanation of causality as opposed to the regularity concept of causation, which is Humean, is used by Gergen to demonstrate how explanations of social phenomena can be misleading and have detrimental effects on society’s understanding of these phenomena. An example might be constructing a psychological explanation for criminal tendencies in a community thus fostering perceptions of members of this community to be criminal rather than examine the socio-historical and cultural context which may have led to deviant behaviour. Gergen’s point must be taken seriously: explanation based on the available facts and information might be wrong. Alternative theories are possible and it is the task of the investigator and the scientific community to be sufficiently critical to unmask inappropriate explanatory frameworks. This is the very valuable deconstructive role the postmodernist plays. Gergen’s relentless unmasking of fallacies, univocal interpretations and dogmatic assumptions is exemplary to the realist scientist. (b) Social science cannot operate in a similar way than natural science because the principle patterning cannot be found. It is unstable, changing and complex. Of course, those based on biological and natural systems are much more stable and predictable. Gergen (1982, p. 13) says that human beings are constitutionally predisposed towards novelty and creative behaviour. In contrast to the traditional behaviourist paradigm, human behaviour is largely stimulus free except for some biological reflex patterns: “. . . human biology largely serves to establish the grounds and limits of human action. Certain functions must be fulfilled to sustain life, and there are biological limits over performance. However, between the poles of grounding essentials and physical limitations there is virtually unlimited potential for variation” Gergen (1982, p. 17). In addition, humans have the ability of symbolic restructuring. This means that they can interpret their internal and external environment with relative freedom. Their freedom from stimuli essentially derives from their ability to re-conceptualise and self-reflect. Thus, their own histories can be reinterpreted from the present and their past provides re-interpretive possibilities for the present. This capacity for conceptualisation provides almost boundless possibilities for action. Furthermore, the social scientist cannot rely on cumulative knowledge in the sphere of the social. Gergen (1982, p. 19) points out that patterns of behaviour are historical and contextual, thus even if one manages to establish some patterns in social behaviour, these may change as soon as time, locality or context change. It

142

6 The Realist Image of Science

seems as if certain acquired human tendencies work against establishing existing patterns of behaviour despite periods of stable and patterned behaviour. These tendencies are the push to free oneself from restraints, striving for uniqueness and barriers trying to be unpredictable and Gergen (1982, p. 13) makes a case that these tendencies work against the establishment of any predictable pattern that can be generalised. Ironically these three tendencies maybe some of the only enduring and recurring patterns of human behaviour! (c) The third aspect Gergen (1982, p. 13) claims about the social sphere that works against the ideals of traditional science is the difficulty of not influencing the phenomena of social science, i.e., people. The received view of science requires objectivity or as we have characterised it, intransitivity which is not possible in social science. This is the issue of first-order concept-dependence or specifically causaldependence rather than existential-dependence we encountered above. (d) Finally, Gergen (1982) elucidates the scientific construction of social reality. The influence of science as a social activity on society in forming its perceptions and actions is illustrated by various examples. Theories developed by science tend to influence what people believe about reality. Behaviourism is but one good example: social engineering, manipulation and control of society seemed a real possibility based on the S-R paradigm of human behaviour: “Research methodology congenial to the stimulus-response paradigm serves to sustain the image of the human as manipulable automaton” (Gergen, 1982, p. 32). The image of science influences both scientific and psychological practice and how society constructs social reality and itself. The realist supports Gergen’s analysis of the social construction of reality. Gergen (1982) convincingly describes the influence of socio-historical and scientific issues shaping scientific practices in psychology. The target of Gergen’s aim is the received view of science and how it helped shaped psychology’s view of how human behaviour should be explained. These views permeate society and whether they are a correct presentation of what psychological science and psychology is, is another question. Gergen’s mistake is to regard the received view as correct, i.e., how science should be viewed, at least for natural phenomena. He correctly points out what the ramifications of such a view for social science and our understanding of human beings and society are but unfortunately he is pandering the mythic view. It does not even apply to natural phenomena. Thus, the argument of this book is that the mythic image of science constructed our practices in psychology and what we believe science and related activities should be in society.

Quantification and Measurement in Psychology and Social Sciences In the chapters above, especially in the discussion about the methodological dichotomy, we have encountered the issue of positivism and measurement as a characteristic of the received view of science. Even in the discussion of the split between

Quantification and Measurement in Psychology and Social Sciences

143

practitioner and scientist, it became clear that the practitioner cannot accept that a fundamental human orientation can be reconciled with something as reductive as the quantification of human behaviour, feelings, meaning and understanding. The concepts of quantification and measurement are integral to empiricist approaches and one can scarcely imagine how experimentation can take place without measuring something! It is true though that empirical science in both social and psychological disciplines regard quantification of constructs as characteristic of their science. However, the ubiquity of measurement, clearly modelled on some natural science disciplines, should not mislead us to think that this is the characteristic of a mature science. The received view that became the mythical view of science indeed regards quantification as a sine qua non of its approach. However, as we have seen, the fundamental characteristic of science is criticism ontologically grounded. How epistemic access pans out depends wholly on the nature of the phenomenon or aspects of reality we are dealing with. It might be to measure or it might be to interpret, or both, but confronting the unamendability (and possibility) of reality requires a critical orientation which can be quantitative, qualitative or both. The reader should note that I am not using “empirical” for quantitative methods as is the usage to characterise empirical methods as exclusively quantitative. Empirical plainly means dealing with tangible data in some form and it includes the meaning expressed by people transcribed or captured in other forms. That empirical and quantification mean almost the same thing is also part of the mythic image of science. In this section we will look at the place of quantification and measurement in a realist conception of science. Suffice to say is that measurement means epistemic access to things that can be measured; as a method it does not define what science is. Michell (1997, 1999, 2000, 2003a, 2003b, 2005, 2009, 2011a, 2013) provided a thorough analysis of the problem of measurement in psychology from a realist perspective. Michell (1999) provides a very lucid and in-depth account of the phase in the history of psychology when psychology committed to a positivistic view of science involving the quantitative imperative. In principle, and based on realism, no conflict between qualitative and quantitative methods as methods of enquiry needs to be assumed (Michell, 2011b). Methods are driven or determined by the nature of the phenomena under discussion. Qualitative method supporters construct their case for qualitative methods in pointing to positivism’s inability to accommodate meaning. In fact, positivism is believed to incorporate the quantitative imperative and the argument is that quantitative methods are wrong because they cannot allow for meaning in objects/phenomena studied. Michell (2003b, p. 12) aptly calls this frequent but vacuous claim a “creation myth,” hence, supporting my characterisation of the mythic image of science. Michell (2003b) pointed out that there is truth in the positivistic link with psychology, because in reality psychology, at the turn of the previous century, had a solid view of science as positivistic and modelled itself on this view of science. Natural science was, indeed, seen as the template of rigorous and valid science, and measurement in natural science conveyed this sense of rigour. The irony is that positivism, according to some proponents such as Carnap, did not see measurement

144

6 The Realist Image of Science

as crucial for scientific rigour and even allowed a place for qualitative investigation (Michell, 2003b, p. 12). What Michell (2003b) calls the quantitative imperative, or the requirement of measurement as the hallmark of science, comes from a Pythagorean worldview that claims an underlying quantitative structure for all of reality (Michell, 1997, p. 462, 2003b, p. 19, 2011b, p. 244). Although there is always something to count in most situations, this fact does not mean that situations and attributes have a quantitative structure. Having quantitative structure is an ontological characteristic that the scientist needs to discover. Throughout the history of thought the idea of quantity as a category was held as fundamental. Kant viewed quantity as transcendental, i.e., a category or structure imposed by the mind on the world (Michell, 2003b, p. 8). Anderson (1962, 2005) viewed the categories as ontological and amongst them the category of quantity. A number of scientists, from psychologists to natural scientists, expressed the measurement–scientific rigour equivalency. Michell (2011b, p. 244) quotes Thorndike who said, “whatever exists at all exists in some amount.” Thus, the quantitative imperative was and still is fundamental to scientific thought. Psychology’s unfortunate development took this imperative on board and it is no wonder due to the prevalence of the idea in mainstream psychology, that qualitative researchers position themselves against this requirement. Michell (2003a) contends that this is a mistake and that the quantitative imperative is false on realist grounds. We need to acknowledge that some structures and phenomena are quantitative, while others are qualitative. The mere fact that we can talk of more or less in some situations, say, of attitudes, does not logically entail the quantitative structure of that phenomenon. If I resent my father deeply and Susan resents her father only a bit, then it does not mean that this ordinal-sounding construct has a quantitative structure. Michell (1997) points out that the positing of a quantitative structure should be an empirical matter in both natural and social sciences. The conjoint measurement theory of Luce and Tukey (1964), provides a logical and empirical basis for testing the conditions of quantitative structure of a construct (Michell, 2008a, 2009). I will not discuss the theory here, but it is worth the effort to expose psychologists and students to this theory. The mistake researchers often make–and this mistake is actually taken as valid practice in psychometrics–is to regard ordinal structure as quantitative, thus, measurable (Michell, 2009, 2012). Quantitative structure of a thing means that it is amenable to measurement and Steven’s (1946) defence of the levels of measurement (which we teach psychology students to this day) allows us to argue from ordinal structure to quantifiability. Usually, ordinal structure is believed to be interval or ratio so that parametric statistics can be utilised. However, the fact that there are more or less of something, or that things are countable–an issue part of realism’s ontological principles–does not imply measurability (Michell, 2009). Thus, attitudes are not quantities but weight is. Michell (2000, 2008b), on a number of occasions, pointed out the systemic and systematic ignorance of the problems of measurement in psychology so much so that he calls it a thought disorder, or pathology, of measurement in psychology. The problematic “assigning numbers to things according to a rule” of Stevens (1946) in

Quantification and Measurement in Psychology and Social Sciences

145

the early twentieth century to save psychology from the embarrassment of not being a quantitative science, had the effect of psychologists measuring almost everything, from intelligence to personality (Michell, 1997, p. 360). Of course, the same happened in other social sciences but the hallmark of science as quantification prevailed strongly in psychology to this day to such an extent that constructs are reified unboundedly and psychologists never bother to empirically investigate the quantitative structure of their constructs. Instead, the operationalist mistake of regarding the nature of a thing as similar to its measurement keeps thoughtless quantification alive. Realism, thus, requires the empirical investigation of the measurability of a construct and this means finding additivity to be part of its structure. In sum, it is the duty of the scientist to examine constructs empirically for quantitative structure (Michell, 2011b, p. 245). In fact, the concept of “construct” requires evaluation as well because its ubiquitous use in psychology stems from the same chapter in psychological history when Stevens and others tried to save measurement for psychology (Michell, 2013). However, a construct cannot be measured if it does not have the attribute that is internally structured as quantifiable. Furthermore, an instance of that attribute should have magnitude and this magnitude needs to be related to units of some sorts (Michell, 2003a, p. 522). Michell (1993) regards modern measurement theory as basically representational, i.e., the numbers attached to attributes are arbitrary and represent something to give them meaning. However, traditional measurement theory stemming from Aristotle and Hölder takes number to be real and part of the inherent structure of magnitudes (Michell, 1993, 1994). He explicitly argues for a realist conception of number, although for our purposes it is necessary to understand that measurement assumes real quantities and magnitudes. As a side note, this is what the problem of the idea of construct boils down to: it is a construction of the researcher assuming that anything that can be specified in sufficient detail and clarity can be measured, and this just is not true. The tradition in psychology to examine the validity of constructs bypasses the empirical imperative to examine a phenomenon’s internal structure (Cook & Campbell, 1979). Michell (2005, p. 287) notes that measurement in science is commensurable with the aim of science, namely, “to find something out” and this process fundamentally reflects the character of science, viz., critical enquiry. If measurement is a way to find something out, it follows that it should be appropriate to the subject matter or phenomenon under investigation. Thus, the investigative process entails (a) an empirical determination whether the phenomenon has a quantitative structure, and (b) an estimation of the ratio of the measured unit (Michell, 2005, 2011b, p. 245). The first issue is rarely, if ever, done and in the social science. Its neglect leads to invalid reification of variables and constructs, and of course, the confusion of relations with qualities or attributes. It is important to realise that a construct such as weight, which in the example above, was regarded as an attribute is actually a relation, i.e., between the mass of John and the earth. Thus, some relations are measurable as are some attributes, but it is the scientist’s duty to find out whether they have a measurable structure. It is obvious that many psychological constructs

146

6 The Realist Image of Science

such as IQ, vulnerability or attitude are under suspicion of describing or referring to nothing real and do not have a measurable structure. Having a measurable structure means that the thing under consideration is real. Thus, a measurable real thing has quantitative structure as a property or relation while it occupies a point in space and time. Clearly, it is historical. If the measurable part is an attribute, then it must be an attribute of something; if a relation, then a relation between distinctive phenomena. The example of weight is again appropriate: weight as a relation, namely, the force of attraction between two bodies, is measurable although not observable. It should be noted that weight is not the process of measurement as Bridgeman’s (1927) invalid operationism requires: a thing cannot become the measurement and in psychology this view is, of course, rampant. In the case of weight then, we have something unobservable, yet real and measurable. The conditions for being regarded as real according to situational realism is then, occupying a distinctive point in space-time as a thing, and it must be property of a thing or a relation between things (Michell, 2005, p. 287). The complexity of the situation and the causal field is assumed in the spatiotemporal instantiation. Michell bases his realist assumptions on the Andersonian concept of literal truth, viz., making a claim about a state of affairs. Thus, the attributes of things are taken as real and some of them can be measured. The external structure of attributes consists of properties and relations. The internal structure of an attribute is “measurability,” i.e., the relationship between specific instances of the attribute. Michell says that the relation should be one of ratio and linearity, namely, x times a constant should always be the same irrespective which instance is chosen. This linearity, thus, leads to a constant ratio between sets of units. In plain English, first of all, linearity means that if a constant, say 2, is chosen then any unit say 5 and 30 means that the amount doubles when multiplied by 2, thus 10 and 60. Ratio implies exactly the same, namely, proportions of 1:2 and 3:6 are exactly the same. Now, according to Michell (2005, p. 287), that a thing’s attributes are measurable implies that different instances of a class of that attribute (e.g., weight) maintain the relations of ratio and this means the attribute is quantitative. Determining whether an attribute is quantitative is an empirical task and required for measurability. Michell (2005, p. 287) defines measurement under the assumption of realism as “. . . the attempt to estimate the ratio between two instances of a quantitative attribute, the first being the magnitude measured, and the second being a known unit.” According to Michell (2005, p. 287), a realist position requires a commitment to a realist view of numbers as well, at least for those numbers that demonstrate sustainability of ratios between different instances of an attribute. The issue of methods benefits from a realist grounding. The division of labour between quantitative and qualitative methods is usually done on paradigmatic grounds but their ontological requirements are relaxed for the purpose of using mixed methods (Cresswell & Plano-Clark, 2007). The increasing popularity of mixed methods bodes well for a realist ontological egalitarianism but is not motivated along these lines. In a sense, ontological differences are suspended for pragmatic reasons, but we can safely strengthen an approach of a mix of methods on realist grounds but with the additional requirement that the subject matter dictates

Doing Realist Science in Psychology

147

the method and that method is irrelevant to the characterising of something as a science.

Doing Realist Science in Psychology This chapter can be concluded with three examples of realist science which is relatively different from each other and not at all representative of the typical research going on in the psychosocial science. The first example refers to so-called personality constructs, a phenomenon largely debunked by constructionists. An example of a multidisciplinary approach involving researchers from various disciplines might allow us to refine, change or invent new constructs more appropriate to what seems locally valid. I am not invested in defending personality theories largely criticised for their essentialising activities but it is worth keeping an open mind when tendencies of behaviour are minimally circumscribed by biological structures and functions. The second example refers to a typical empirical experiment in psychology where number sense or conception of babies are investigated in a creative way, especially by closing certain parts of reality and experience not thought possible by Bhaskar and critical realists. The last example focuses on conceptual analysis of social objects, i.e., constructions of phenomena and interpretations that have implications for larger social structures and entities. I will not be repeating all the principles discussed in previous chapters but hopefully the reader will recognise the qualitative difference between these three examples and those that are informed by the mythic image of science. One of the areas of longstanding debate involves the concept of personality in psychology. On the one hand, some researchers are staunch believers in the idea of personality consisting of a number dimensions that adequately captures a person’s propensity to act in a certain way in a certain context. On the other hand, constructionist criticism denies the reality of these so-called personality constructs and warns of essentialising and reifying concepts we might think are true of people. Cloninger (1987) developed a psychobiological model of personality. His research led to the construction of a questionnaire capturing a number of constructs. Four temperament dimensions were identified, namely, Harm Avoidance, Novelty Seeking, Reward Dependence and Persistence. Later, three dimensions of character were added, namely, self-directedness, cooperativeness and self-transcendence (Cloninger, Svrakic, & Przybeck, 1993). Much research followed and Cloninger’s model has been compared to a number of different personality models (Zuckerman & Cloninger, 1996). The most popular personality assessment framework is the so-called Big Five model based on linguistic description of persons’ assessments of themselves. Common variance between the different personality theories and models such as Cloninger’s and the linguistic-based questionnaires was found (De Fruyt, Van De Wiele, & Van Heeringen, 2000). My point is that one might, in a multi-disciplinary way, discover mechanisms underlying so-called personality constructs. Whereas we start off with patterns we observe in patients that allow us

148

6 The Realist Image of Science

to postulate explanations on a psychological level, current neuro-research might enable us to match patterns and tendencies of behaviour to simple physiological mechanisms. Something as basic as harm avoidance, having a basis in biological structures and behaviour activated by neuro-hormones, can easily develop, in interaction with genetic factors and social and environmental influences, into the complex behaviour of someone we would normally call an introvert (see Puttonen, Ravaja, & Keltikangas-Järvinen, 2005 for an example of similar issues). I would like to continue with the second example of, what I would call, a realist psychological study. Xu et al. (2005) examined the number representation mechanism for large arrays that makes counting difficult in infants. An infant is an open system par excellence and a difficult to control psychological system. The researcher does not have direct access to subjective material (such as the ability to verbalise inner states). In this case, the researchers had to resort to constructing a closed system in a creative way by utilising known mechanisms to uncover hidden mechanisms. As starting point they took the phenomenon of habituation as a known mechanism. Habituation works with infants as well and means the more familiar a stimulus is, the less time the infant spends exploring the stimulus visually. Consequently, changing a visual display after habituation was established had the effect of sudden prolonged visual inspection time. In this way, the researchers were able to determine that infants of 6 months can distinguish between 16 and 32 arrays of dots, i.e., a proportion of 2 but not less than 1.5 (they were not able to distinguish between 16 and 24 dots). By carefully controlling conditions such as number, spacing and size of dots along with the test conditions, they were able to uncover mechanisms that could explain the results. Other explanatory attempts were ruled out and the one that explained the data sufficiently was retained: “. . . the present findings add empirical support to the thesis that a common system of number representation, shared by humans and other animals, is present and functional in 6-month-old infants” (Xu et al., 2005, p. 98). To my mind, this experiment is an example of scientific psychological work. First of all, it shows that approximate closure can be achieved. If we want to alleviate our worry about the difficulty of closure in open systems, we can view the experiment as an empirical and conceptual model within which we manipulate some variables and hold others constant, in order to see what generative mechanism is activated. I think we can also clarify the generalizability of the model much better than with a “pure” empirical experiment: the model reveals tendencies, and from a situational realist perspective, allows us, as in qualitative research, to find the general in the particular. In fact, the general-in-the-particular is an ontic principle that grounds Bhaskarian tendencies, i.e., a tendency is not an immutable law but a transfactual propensity. Is it too simplistic to say the applicability of this ontic principle requires only a single Andersonian situation? A historical occurrence of a situation implies a point in space-time but has in itself the roots of necessity or contingency. The manner in which we uncover the generative mechanism depends on what it is. Sometimes repetition hints at a mechanism, sometimes not. Sometimes empirical closure is required, at others, conceptual closure, while a combination is also possible. Then, the accessibility of reality also plays a role in how we uncover mechanisms: usually

Doing Realist Science in Psychology

149

observability suffices, but more often than not, things are just not observable at all. Finally, what we currently see as nature and mind might be a continuum or different sides of the same ontological coin: who knows what future discoveries might hold? The third example is a wholly ideographic one utilising a qualitative approach to analyse social phenomena. Chiumbu (2016) examines the role of the media in constructing a particular narrative about the Marikana killing of 34 mineworkers in South Africa in 2012. The event grabbed attention locally and internationally and focused on mineworkers’ protests about low fees. At some stage during the protests mineworkers were killed by police and the unfolding of events was interpreted in a particular manner but also progressively revealed the full extent of the events. Chiumbu (2016) examined media stories’ construction of the events, thus, media news clips or individual stories were the data that had to be examined. Chiumbu (2016) randomly chose 60 stories from a database of 399 stories. Various sampling techniques originating from experimental empirical methods are usually regarded as positivist (i.e., a principle firmly embedded in the mythic image of science) but it is used here effectively and appropriately. The reason is that a decision to choose stories not based on any perspectival basis but randomly to ensure representativity is based on the nature of the data, in this instance, a set of countable but distinct things. Does the use of some so-called quantitative method make this a mixed-methods study? No, it does not; it is scientific because its methods are appropriate to its phenomena under investigation. Chiumbu’s (2016) analysis determined that the media largely conformed implicitly or explicitly to neo-liberalist narratives, namely, the events had a largely negative effect on the economy and its investors. Thus, despite the end of apartheid in 1994 in South Africa, its economic and other institutions remain embedded within racist structures, even after 20 or more years. Chiumbu (2016, p. 419) points out that capitalist structures closely associated with colonisation did not disappear with decolonisation. Thus, coloniality remains as structural principle in so-called decolonised structures and institutions. Chiumbu (2016, p. 420) thus, uses “. . . the concept of coloniality of power to problematise both the media and economic systems of South Africa.” The media, as part of the institutionalised capitalist structure represented the Marikana strike as detrimental to the economy and in the process utilised various rhetorical techniques to present the striking miners as dehumanised, violent, a threat to police, and even their cultural appropriation of traditional weapons and beliefs were derogated in the light of westernised ideals: “Broadly, the media promoted an acceptable and uncritical discourse of Marikana aimed at legitimising liberal capitalism and forestalling any critique of the system” (Chiumbu, 2016, p. 424). Thus, Chiumbu’s (2016, p. 430) method was largely critical, which “seeks out the contradictions and conflicts inherent in a social structure and questions existing structures.” In the light of various theoretical conceptualisations of social structures and processes such as colonisation, coloniality, neoliberalism, race, power and so on, very real events and their interpretation were contested, i.e., criticised and analysed. The perceptions of social objects and events (in this instance by the media) were countered with alternative interpretations. Thus, the unamendability of social objects and interpretations were tested, challenged and other possibilities uncovered.

150

6 The Realist Image of Science

Phenomena were correctly tackled with methods appropriate to their nature, conceptually by means of critical analysis and interpretation but also were required by counting and statistics. This project, therefore, also qualifies as science.

References Ackroyd, S., & Fleetwood, S. (2000). Realism in contemporary organisation and management studies. In S. Ackroyd & S. Fleetwood (Eds.), Realist perspectives on management and organisations (pp. 3–25). London: Routledge. Anderson, J. (1962). Studies in empirical philosophy [A machine readable transcription]. Retrieved from http://adc.library.usyd.edu.au/view?docId¼anderson/xml-main-texts/ander son049.xml Anderson, J. (2005). Space-time and the proposition: The 1944 lectures on Samuel Alexander’s space, time and deity. Sydney: Sydney University Press. Bridgman, P. W. (1927). The logic of modern physics. New York: Macmillan. Chiumbu, S. (2016). Media, race and capital: A decolonial analysis of representation of miners’ strikes in South Africa. African Studies, 75(3), 417–435. https://doi.org/10.1080/00020184. 2016.1193377. Cloninger, C. R. (1987). A systematic method for clinical description and classification of personality variants: A proposal. Archives of General Psychiatry, 44, 573–588. Retrieved from http:// archpsyc.jamanetwork.com/article.aspx?articleid¼494070. Cloninger, C. R., Svrakic, D. M., & Przybeck, T. R. (1993). A psychobiological model of temperament and character. Archives of General Psychiatry, 50, 975–990. Retrieved from http://archpsyc.jamanetwork.com/article.aspx?articleid¼496431. Cook, T. D., & Campbell, D. T. (1979). Quasi-experimentation: Design and analysis issues for field settings. Boston: Houghton Mifflin. Cresswell, J. W., & Plano-Clark, V. L. (2007). Designing and conducting: Mixed methods research. Thousand Oaks, CA: Sage. De Fruyt, F., Van De Wiele, L., & Van Heeringen, C. (2000). Cloninger’s psychobiological model of temperament and character and the five-factor model of personality. Personality and Individual Differences, 29(3), 441–452. https://doi.org/10.1016/S0191-8869(99)00204-4. Ferraris, M. (2014). Manifesto of new realism (S. De Sanctis, Trans.). Albany: State University of New York Press. Gergen, K. J. (1982). Toward transformation in social knowledge. New York: Springer. Luce, R. D., & Tukey, J. W. (1964). Simultaneous conjoint measurement: A new type of fundamental measurement. Journal of Mathematical Psychology, 1, 1–27. Michell, J. (1993). The origins of the representational theory of measurement: Helmholtz, Hölder, and Russell. Studies In History and Philosophy of Science Part A, 24(2), 185–206. Michell, J. (1994). Numbers as quantitative relations and the traditional theory of measurement. The British Journal for the Philosophy of Science, 45(2), 389–406. Michell, J. (1997). Quantitative science and the definition of measurement in psychology. British Journal of Psychology, 88, 355–385. Michell, J. (1999). Measurement in psychology: Critical history of a methodological concept. New York: Cambridge University Press. Michell, J. (2000). Normal science, pathological science and psychometrics. Theory and Psychology, 10(5), 639–667. https://doi.org/10.1177/0959354300105004. Michell, J. (2003a). Epistemology of measurement: The relevance of its history for quantification in the social sciences. Social Science Information, 42, 515–534. Michell, J. (2003b). The quantitative imperative: Positivism, Naïve Realism and the place of qualitative methods in psychology. Theory and Psychology, 13, 5–31.

References

151

Michell, J. (2005). The logic of measurement: A realist overview. Measurement, 38(4), 285–294. Michell, J. (2008a). Conjoint measurement and the Rasch paradox: A response to Kyngdon. Theory and Psychology, 18(1), 119–124. https://doi.org/10.1177/0959354307086926. Michell, J. (2008b). Is psychometrics pathological science? Measurement: Interdisciplinary Research and Perspectives, 6(1–2), 7–24. https://doi.org/10.1080/15366360802035489. Michell, J. (2009). The psychometricians’ fallacy: Too clever by half? British Journal of Mathematical and Statistical Psychology, 62, 41–55. Michell, J. (2011a). Normal science, pathological science and psychometrics. In N. Mackay & A. Petocz (Eds.), Realism and psychology: Collected essays (pp. 461–499). Boston: Brill. Michell, J. (2011b). Qualitative research meets the ghost of Pythagoras. Theory and Psychology, 21 (2), 241–259. https://doi.org/10.1177/0959354310391351. Michell, J. (2012). “The constantly recurring argument”: Inferring quantity from order. Theory and Psychology, 22(3), 255–271. https://doi.org/10.1177/0959354311434656. Michell, J. (2013). Constructs, inferences, and mental measurement. New Ideas in Psychology, 31 (1), 13–21. https://doi.org/10.1016/j.newideapsych.2011.02.004. Puttonen, S., Ravaja, N., & Keltikangas-Järvinen, L. (2005). Cloninger’s temperament dimensions and affective responses to different challenges. Comprehensive Psychiatry, 46(2), 128–134. https://doi.org/10.1016/j.comppsych.2004.07.023. Stevens, S. S. (1946). On the theory of scales of measurement. Science, 103(2684), 677–680. https://doi.org/10.1126/science.103.2684.677. Xu, F., Spelke, E. S., & Goddard, S. (2005). Number sense in human infants. Developmental Science, 8(1), 88–101. https://doi.org/10.1111/j.1467-7687.2005.00395.x. Zuckerman, M., & Cloninger, C. R. (1996). Relationships between Cloninger’s, Zuckerman’s, and Eysenck’s dimensions of personality. Personality and Individual Differences, 21(2), 283–285.