Download Theoretical Sociology in the 20th Century

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Social constructionism wikipedia , lookup

Actor–network theory wikipedia , lookup

Social rule system theory wikipedia , lookup

Social group wikipedia , lookup

Public sociology wikipedia , lookup

Sociology of terrorism wikipedia , lookup

Frankfurt School wikipedia , lookup

Social network analysis wikipedia , lookup

Structuration theory wikipedia , lookup

Sociology of culture wikipedia , lookup

Symbolic interactionism wikipedia , lookup

Social network wikipedia , lookup

Index of sociology articles wikipedia , lookup

Differentiation (sociology) wikipedia , lookup

Structural functionalism wikipedia , lookup

History of sociology wikipedia , lookup

Development theory wikipedia , lookup

Sociology of knowledge wikipedia , lookup

Postdevelopment theory wikipedia , lookup

Sociological theory wikipedia , lookup

Transcript
1
Theoretical Sociology in the 20th
Century
Thomas J. Fararo
University of Pittsburgh
[email protected]
Prepared for a Festschrift in Honor of Linton C. Freeman
Vancouver, B.C., April 16, 2000 [Revised: April 21, 2000]
ABSTRACT: This chapter discusses theoretical sociology in historical perspective:
From the classic tradition to postclassical efforts of synthesis that culminated in
multiple paradigms, to the situation today in which theorists are more and more
constructing formal models as essential components of their methodology. The
classical phase is treated very briefly and the discussion of the postclassical phase is
limited to two major theorists, Parsons and Homans, in terms of their common focus
on the Durkheimian problem of social integration. The bulk of the chapter deals with
developments in recent theoretical sociology. I describe models of structure and of
process before defining two types of models that combine a structural focus with
process analysis. Finally, I set out a general perspective on theoretical model
building and conclude with a discussion of standards in the assessment of such work.
Prelude
This paper is written in honor of the person who had the greatest influence in shaping
my professional career, Linton C. Freeman. It was Lin who hired me as a research
assistant back in 1960 at Syracuse University, for a project dealing with community
power structure that he and others were about to undertake. I had been a student in an
integrated social science program, mainly studying the history of social thought.
Earlier and even then, however, I was reading the literature of the philosophy of
science and this prepared me to become very enthusiastic about the process of
building knowledge. I transferred to the Department of Sociology to pursue a Ph.D.
under Lin's direction and to continue to work with him and others on research
projects.
Lin's own strong commitment to basic science was communicated in every context of
our interaction and further strengthened the more abstract lessons I was learning
through my reading. It was he who brought me into every phase of the community
power structure project from conceptual discussion to interviewing to data analysis to
writing of research reports and an article for publication in ASR. It was socialization
to basic science that, even today, few students acquire in such depth.
Among other things, when I became aware of a then recent paper by Rapoport and
Horvath (1961) dealing with a new way to analyze large sociograms, I followed Lin's
1
2
advice and embarked on learning enough about it to apply it to our data as my Ph.D.
thesis project.
After earning the doctorate, through Lin's influence, I was appointed to the faculty of
the department and assigned to very congenial courses, one or two of which dealt with
formalization problems. These courses allowed me to communicate my ideas about
the philosophy of science and to teach formal logic and axiomatics as well as finite
mathematical model building in sociology (e.g., finite Markov chains). I felt
inadequately prepared in classical (non-finite) mathematics and applied for and
obtained a three-year postdoctoral fellowship for the study of pure and applied
mathematics at Stanford University. Then, in 1967, through Lin's influence I was
offered a position at the University of Pittsburgh where Lin and I resumed our contact
for several years. I remained there for the remainder of my career, while Lin moved
on. Here now, in the twilight of a career that would have been quite different without
my connection to him, I am pleased to present some ideas about sociological theory
that developed over the years.
Overview
This paper discusses theoretical sociology in historical perspective: from the classic
tradition to postclassical efforts of synthesis that culminated in multiple paradigms to
the situation today in which theorists are more and more constructing formal models
as essential components of their methodology.
The tradition of sociological theory as a whole exhibits a mixture of three types of
sociological interests that I call theoretical sociology, world-historical sociology and
normative-critical sociology. I discuss this mixture in the classical phase and then the
remainder of the paper has theoretical sociology as its focus. This focus represents the
sort of basic science interest that Lin communicated to me about forty years ago and
to which I remained committed over the years.
I limit my treatment of the postclassical phase to two theorists, Parsons and Homans,
each discussed in terms of a shift in theory construction strategy as well as in terms of
their common focus on the Durkheimian problem of social integration.
In analyzing the recent phase of theoretical sociology, I first discuss the situation of
multiple theoretical perspectives and then draw attention to what I call mutations and
new combinations. I emphasize that the role of models has become a major part of the
tradition of theoretical sociology, describing models of structure and of process before
defining two types of models that combine a structural focus with process analysis.
Finally, I set out a general perspective on theoretical model building and conclude
with a discussion of standards in the assessment of such work.
2
3
Phases and Components of Sociological Theory
Three Phases. In the late 19th and early 20th centuries, a handful of scholars who by
and large worked independently, elaborated conceptions of sociology as a science.
Probably the most enduring contributions were produced by Weber, Durkheim, and
Simmel, and, mainly for his influence on later theorists, Pareto. In addition, Comte
and Spencer were important 19th century precursors. Finally, although neither George
Herbert Mead nor Karl Marx ever elaborated a conception of sociology, their writings
have been incorporated into the tradition. These various writings are commonly
referred to as "classical sociological theory" and comprise the first phase of the
tradition.
The second phase, which I will call "postclassical," began with integrative efforts
directed toward building a common theoretical framework for sociology. Influential
writers with this ambition included Talcott Parsons and George Homans, among
others. But a unifying framework did not emerge and an era of proliferation of
perspectives took hold under the conception of sociology as a multiple-paradigm
science.
In the recent third phase of sociological theory, the multiple paradigms or
perspectives continue -- with mutations and new combinations -- alongside renewed
efforts to consolidate theoretical ideas. Examples of commitments to the growth of
scientific theories in sociology compete with postmodernist and other viewpoints that
rest upon a repudiation of the entire idea of sociology as a science.
Three Components. In addition to this phase description, it is useful to interpret the
corpus of writings over these phases as comprised of three components, reflecting
different intellectual interests. In a simplified model, I suggest just three such
components pervade the entire tradition of sociological theory. (See Figure 1.)
3
4
Figure 1.
Three Components of Sociological Theory in the 20th Century
One is the elaboration of ideas relating to the construction of generalized frameworks
of sociological thought. I treat this aspect of classical theory as the first phase of
theoretical sociology, the first of the three component sets of interests.
A second component relates to an intellectual interest in world-historical social and
cultural forces in the creation of the modern world. Today, this type of worldhistorical interest has shifted from modernization to globalization and postmodernity.
One reason for distinguishing this focus from general theoretical sociology is that it
enables a distinction between the importance of a general theoretical problem and the
importance of the empirical instance studied in terms of that problem. Thus, the
evaluation of theoretical model can occur with respect to empirical instances that have
little importance outside this scientific context.
Finally, a third component of the tradition of sociological theory involves critical
normative ideas. It entails evaluation of social phenomena rather than their
explanation or historical interpretation. For instance, from Hegel and Marx to
Habermas, critical theory emphasizes the task of critique of society and culture in the
interests of human emancipation from what it treats as coercive structures of
production and consumption. Feminist theory also has a primacy of interest in social
critique. Although such theories draw upon general theoretical sociology, as does the
world-historical orientation, they foster a primacy of ideology that detracts from the
pursuit of an interest in basic scientific knowledge of social life. Nevertheless, the
three components tend to be interrelated in the literature of sociological theory. As a
result, any body of theory -- or even a single work -- can be regarded as a kind of
4
5
weighted combination of the three components. I will illustrate this point in my
discussion of the classical phase of sociological theory.
The Classical Phase
For brevity, I select just five classical theorists and present a compact and brief listing
of some of the key ideas of each of them: Mead, Weber, Simmel, Durkheim and
Pareto, organized in terms of the three components. The format serves to illustrate the
three types of intellectual interests that permeate the tradition of sociological theory.
Each theorist's main foundational contribution to theoretical sociology is also
highlighted at the outset of the listing of sample elements of the three components in
that theorist's work.
1. Weber: Social action as a foundation concept
o Theoretical sociology. For sociological purposes, social life consists of
complexes of social action that can be studied by the explicit use of
analytic procedures involving idealization.
o World-historical sociology. The history of the West is one of
increasing rationalization trend and its consequences.
o Normative-critical sociology. Modern rational capitalism is an iron
cage.
2. Mead: Social behavioral foundations of human action systems
o Theoretical sociology. Mind, self, symbols, and institutions are coemergent in natural evolution. The starting point for social psychology
is the social act as an organized social activity in which actors take
each other's attitudes.
o World-historical sociology. Human history is evolution on a smaller
time-scale in which variant institutional solutions arise in relation to
common social problems.
o Normative-critical sociology. The basic problem of human society is
how to have orderly change and the best answer to this, so far, is to
organize society along democratic lines.
3. Simmel: Interaction concept as essential for sociology
o Theoretical sociology. Society is interaction among individuals. The
subject matter of formal sociology consists of forms of interaction, a
kind of geometry of the social world.
o World-historical sociology. The history of the West is one of
increasing social and cultural complexity.
o Theoretical Normative-critical sociology. Because culture becomes so
complex, the modern individual is in danger of alienation with a
subjective self that is not truly cultivated.
4. Durkheim: Integration as a fundamental problem
o Theoretical sociology. Neural networks are to psychic facts as social
networks are to social facts. The domain of sociology consists of social
facts that require a distinctive sociological explanation. Emergent
social integration is a key problem of sociology.
5
6
o
World-historical sociology. The history of the West is one of
increasing social differentiation and its consequences, such as
increasing individuation.
o Normative-critical sociology. Modern societies are not in a healthy
state because their moral regulation has not yet caught up with changes
in social relationships, especially in the economy.
5. Pareto: System concept as a key tool for theorizing
o Theoretical sociology. Scientific theory is analytical, which means
abstract and, in basic science, involving the construction of idealized
models.
o World-historical sociology. History is a story of interdependent
economic, political and cultural cycles.
o Normative-critical sociology. Human non-scientific belief systems,
whether religious or secular, are all ideologies subject to detached
critique (the standards for which are broadly humanistic).
In what follows, my discussion of the postclassical and recent phases of
sociological theory will be limited to theoretical sociology. Figure 2 outlines
the phases of theoretical sociology, showing the foci of discussion in this
chapter.
Figure 2.
Three Phases of Theoretical Sociology in the 20th Century
6
7
Postclassical Theoretical Sociology
Although there are various streams of developments in theoretical sociology that can
be traced to the influence of the classical theorists, in this paper I focus on two
theorists with a common background and a common aspiration, namely Talcott
Parsons and George Homans. Both were at Harvard in the 1930s when the idea of
creating a general theoretical sociology was discussed in the famous Pareto seminar.
The keynote for Parsons and Homans was the creation of an analytical sociological
theory that was based upon the classical phase of sociology and on related empirical
research not only in sociology but also in related fields, particularly anthropology. In
each instance, we can partition the resulting career of theoretical work into two phases
marked by a shift in theory construction strategy.
Parsons: The First Phase. Theoretical sociology was a central but not exclusive
concern of Talcott Parsons and his first major work, The Structure of Social Action
(1937), played a major role in subsequent developments. He analyzed the writings of
the economic theorist Marshall as well as those of Pareto, Durkheim and Weber. His
objective was to show that these writers had expanded the scope of analytical social
theory beyond the traditions from which they emerged, with their more limited
perspectives.
Any analytical theory, Parsons argued, treats only selected aspects of a complex
reality, formulating two kinds of conceptual schemes. One such scheme specifies a
general structural account of the type of empirical system of interest. It key concepts
refer to parts and relations among them. The other type of conceptual scheme
presupposes some sort of structural analysis and goes on to specify an analytical
system, a set of variables and relationships among them. His convergence argument,
he noted, pertained only to structural analysis.
Following Weber and Pareto, Parsons initiates his analysis in terms of an action frame
of reference. He treats social entities such as groups as systems of social actions.
Hence, structural analysis, at this level, will focus on relations among types of acts so
as to describe "the structure of social action" as a prelude to an analytical theory of
such social action systems.
Parsons' basic structural concept is the means-end chain. Each such chain represents a
series of interconnected actions, a kind of path through an action space. This suggests
representing means-end chains by paths in a finite directed graph. An edge, denoted
(m, e), corresponds to an action in which certain means m are employed toward some
end e. Two such edges, (m1, e1) and (m2, e2) are adjacent when the end point of the
first is the means point of the next: e1 = m2. Paths intersect because some means are
employed toward the same end and some ends are means in various further actions.
The structure of this system of social action has three sectors. Think of the graph in a
vertical orientation, the lines point upward. At the bottom are points with no edges
directed to them: they are only means, never ends. They comprise what Parsons calls
the ultimate means sector of the structure of social action. Similarly, at the top are
7
8
points such that no edge is directed from them: they are only ends, never means. They
comprise what Parsons calls the ultimate ends sector of the structure. All other points
have edges entering and leaving them: they are both means and ends. They comprise
what Parsons calls the intermediate sector of the structure. Looking downward, in
which ends control the selection of means, we have a hierarchy of normative control
from the ultimate end sector to the intermediate sector to the ultimate means sector of
the social action system. Moreover, the various ultimate ends (some of which are
diffuse values) are not independent. Connections among them constitute the emergent
property of value-integration, as in the existence of value systems. When such values
are not only connected but are shared among actors, they are said to exhibit the
property of common value-integration.
Parsons argues that the classical phase of sociological theory converged on the thesis
that the emergent property of common value-integration is essential for social order.
His "sociologistic theorem" says that a necessary condition for social equilibrium
(social order) is the existence of a common value system. He argues that economic
and political theories have focused on the intermediate sector in which actions are
means to immediate but not ultimate ends. Some such ultimate ends are not even
empirical. In such cases, Parsons classifies the corresponding actions as nonrational
because there is no scientific way of saying that the means are inappropriate, in
intrinsic or causal terms, to the attainment of such ends. The actions may be
appropriate in some symbolic sense, as in ritual action. Thus, Parsons' conceptual
scheme links the existence of social order to the nonrational aspect of action systems
via the sociologistic theorem. Moreover, in defining sociological theory as only one of
the analytical sciences of action, Parsons associates it with the emergent common
value-integration property and the sociologistic theorem.
Parsons: The Second Phase. In his next major work, The Social System (1951),
Parsons elaborates on the psychological foundations for this idea, stating what he calls
“the fundamental dynamic theorem of sociology," drawing upon ideas from Freud
that support Pareto's focus on nonrational elements in social system dynamics. The
theorem states that the stability of social equilibrium requires the institutionalization
of a value system that is also sufficiently internalized in the personalities of members.
Between the first and the second books, Parsons had changed his theory construction
strategy. In the first work, the entire elaborate discussion of the means-end structure
of social action systems was regarded as a preliminary to the task of constructing an
analytical theory. With the conception of the scope of theoretical sociology as
focussed on the emergent property of common value-integration, the analytical
variables that were needed would be value pattern variables -- variables whose
combinations could be used to characterize the dynamics of social action. These led to
his famous "pattern variable scheme" involving such value polarities as universalism
versus particularism and affectivity versus affective neutrality. Combinations of
selections from these value alternatives define value patterns that are definitions of the
directions of action to be expected in social relationships. For instance, in a doctorpatient relationship, the value pattern that defines the relationship includes
universalism and affective neutrality, among other value elements. In dynamic terms,
such value patterns would function as "control parameters" that enable and constrain
actions in the hierarchy of normative control.
8
9
However, at some point, Parsons became convinced that a simplification of the theory
task would be required. This took the form of forgoing a true dynamic analysis with
derived equilibria in favor of a focus on a social system as a "going concern." For
each social relational nexus satisfying this condition, the corresponding value pattern
is treated as a pattern tending to be maintained despite disturbances. Thus the
problematic feature for theory was to describe the mechanisms that tended to
counteract disturbances and thereby to help maintain the pattern. The theory becomes
structural-functional with its focus on mechanisms of socialization and social control.
In turn, the elaboration of this structural-functional type of theorizing eventually led
to two related classifications, one of social structural parts and one of social functional
subsystems. The four types of social structural parts are values, norms, collectivities
and roles. For instance, in American society, there is a diffuse value of freedom with
its implementation in diverse norms, such as the normative conception of a free press.
In turn, this norm is embodied in numerous collectivities, such as news organizations,
that disseminate ideas through the specialized activities of people acting in such roles
as editor, reporter and the like.
To define and analyze functional systems, Parsons applies a general conceptual
scheme for functional analysis. The starting point is that any system of action is said
to have four key functional problems: (latent) pattern maintenance (L), integration (I),
goal-attainment (G) and adaptation to the environment (A).
As applied to what we might call a societal action system, the four types of functional
problems are specified, respectively, in the reverse order, in two steps. In the first
step, this action system is modeled as a system with four types of interdependent
functional subsystems: cultural systems (L), social systems (I), personality systems
(G) and behavioral systems that adapt to the biophysical environment (A). What is
normally called "the society" is the most inclusive social system in this analysis of the
societal action system, so that its environment consists of cultural systems as well as
the personality and behavioral systems of its members. Then the analysis of a society,
in this sense, proceeds by four-function analysis once again. In AGIL order, the
society as an integrative system of action (I) has four functional problems: economic
(collective adaptation to the action and biophysical environments, IA), political
(collective goal attainment, IG), social integrative (II), and fiduciary (maintenance of
the cultural traditions, IL). Corresponding to these four problems are four
interdependent functional subsystems of the societal action system: economy, polity,
societal community and fiduciary system.
Linking the two conceptual schemes for structure and function, for instance, yields
political values, political norms, political collectivities and political roles as the
structural components involved in the polity. These units interpenetrate with the
components of all the other systems because the same people who act in political roles
(e.g., voter) perform actions in other roles (e.g.. consumer). Thus, functional
connectivity characterizes the partial differentiated structures (assuming here a
modern differentiated society).
In this conceptual scheme, in principle, each analytical theory of action has a scope
that corresponds to one or more functional subsystems. The analytical focus of
theoretical sociology, in this four-function perspective, is the integrative subsystem of
9
10
any social action system. This is a "system of solidarities" to use the nice terminology
of Baum (1975). The focus is on the problematic integration of a social system. In the
limit, the system may consist of a single collectivity without subcollectivities. In
another direction, it may consist of a huge number of intersecting subcollectivities but
not itself form a single collectivity. Finally, it may be both "many" and "one," in the
sense that it is both a single collectivity and has plural subcollectivities within it. For
instance, at the societal level, the term "nation" points to a single solidary system, a
collectivity, but the nation will be comprised of intersecting subcollectivities. In short,
as in Durkheim, the fundamental theoretical problem of sociology is social integration
at any level of social life, the "double I" (II) problem in Parsons's four-function
scheme.
In general, Parsons seems to have an image of a tree of analytical theories, each
scope-defined:




Cultural theory (L)
Social theory (I)
o Fiduciary theory (IL)
o Solidarity theory (II)
o Political theory (IG)
o Economic theory (IA)
Personality theory (G)
Behavioral theory (A)
The analytical view of theoretical sociology as focussed on the problem of social
integration or solidarity (II) leads to a specification of a fundamental problem: What
holds society together? It has been framed as the problem of order. From Hobbes to
Parsons to Dahrendorf (1959) and into recent theory, the problem has a long history in
social theory. For instance, Dahrendorf criticized Parsons' approach to the problem,
revising Marx's coercion-based approach to make power relations central.
Subsequently, Collins (1975) attempted a synthesis of Durkheimian theory with this
Dahrendorf-type of conflict theory. Actually, Dahrendorf makes legitimacy a
fundamental feature of his treatment of power relations, thereby invoking values and
norms and taking the sting out of his critique of Parsons' treatment of the problem.
However, criticism of Parsons' theory came not only from those who favored a
conflict theory perspective but also from theorists adopting other perspectives.
Prominent among these was George Homans. His early work initiated another mode
of generalized synthesis in theoretical sociology that led to a second phase that itself
drew considerable critical reaction.
Homans: the First Phase. In the first of his two major theoretical works, The Human
Group (1950), Homans argued that the creation of theoretical sociology should begin
with a scope restriction to small groups, defined as those in which each member could
interact with every other during the time the group meets. In framing a general
theoretical problem, Homans asked: What makes customs customary? In other words,
how do we account for order? The problem is framed at the elementary level of
interaction and pertains to the emergence, maintenance or change of systems of social
relationships among persons. Order in the form of social integration is explained
through an emergent "internal system," given external conditions. Social bonds
10
11
among members and shared norms are generated by mechanisms that are described in
terms of specific hypothesized linkages among analytical elements pertaining to
activities, sentiments, and interaction. For instance, the more frequently people
interact with each other, the more similar their sentiments, normative ideas and
activities become. Far more clearly than in Parsons' work, a system of variables and
their relationships is specified so as to undertake the verbal equivalent of the sorts of
steps that are taken in the mathematical analysis of a dynamical system model.
Moreover, Homans synthesizes classical ideas within his theoretical framework. For
example, in treating social control, the Durkheimian idea of the ritual effects of
punishment is embedded in the discussion of the stability analysis of equilibrium
states. In addition, he delimits the scope of two seemingly opposing theories of ritual - those of Malinowski and Radcliffe-Brown, respectively -- before reconciling them,
i.e., integrating them. In short, Homans' social system theory is in the Durkheimian
tradition, although critical of functionalist arguments that do not specify mechanisms
that account for the emergence and stability of equilibrium states.
If Homans' analytical hypotheses or laws describe group dynamics and the build-up
(or dissolution) of a group, what explains the laws? In searching for this more
fundamental level of theorizing, Homans invoked a conceptual scheme from
behavioral psychology in the next phase of his theoretical work (Homans 1961, 1974).
Interaction is an exchange involving material and non-material goods, and social
approval is a fundamental category of social reward.
Homans: the Second Phase. Just as Parsons had changed theory construction
strategy between his first and second major works, so did Homans. From a theory as
modeled on a system of differential equations, he moved to theory as a system of
propositions forming a deductive system. The behavioral principles have the function
of covering laws in logical arguments that explain regularities in social life, including
the results of experimental social psychology as well as field studies of the sort
analyzed in the earlier work.
We can interpret the basic logic of this approach as reduction in the sense of
explanation of social life from a non-social foundation. This is somewhat analogous
to the explanation of molecular levels of existence from a purely atomic basis. Critics
might ask: What if atoms only could have the postulated properties they have if these
properties emerge out of molecular relations? Then this sort of organic relationship
makes reduction nonsense. Similarly, if individuals are socialized beings, how can
their interaction explain social order? You are simply presupposing what is supposed
to be explained.
But there is a response to this criticism. In Homans' behavioral theory, the
fundamental unit is not the person but the behavioral act. The person as a complex
socialized entity is not the subject matter of interest to Homans, although such a
system -- corresponding to Parsons' personality system -- is within the scope of the
behavioral theory. In other words, Homans has a tree of theory with a basic behavioral
or action theory at its root and with a number of branches. Given his commitment to
analytical theory, he pursues just one branch, namely the one that deals with the
problem of the integration of the actions of plural persons to form a dynamic social
system with emergent patterns of order.
11
12
In this interpretation, Homans can agree with the classical sociological theorist
Charles Cooley who argued that individual and society are "twin-born," in that the
person is socially constructed in social interaction and that a society is a system of
interaction. In practice, then, Homans took mind, self and symbols -- three important
elements from the Cooley-Mead standpoint -- as givens in the pursuit of a pure
theoretical sociology that would formulate and explain group processes.
In taking this approach, Homans accompanied his work with a polemical argument.
He took aim at Durkheim, who had argued that what explanation means for
sociological theory is a causal account that remains at the level of social facts. For
instance, to explain varying rates of deviance in groups, Durkheimian theory would
point to varying levels of solidarity: the greater the solidarity of the group, the lower
the rate of deviance from its norms. (See Figure 3.)
Figure 3.
Durkheimian Social Generativity via Homans and Coleman
What Homans argued was that such a proposition, if it is true, could be derived
logically from a behavioral foundation. For instance, in a highly solidary group,
members experience or can anticipate high costs in lost social approval for deviation
from group norms, while in a less solidary group, such costs are lower. Solidarity or
cohesion is "micro-translated," to use Collins' (1981) term, in such an explanation.
12
13
Then the behavioral mechanisms are able to explain why varying rates of solidarity
lead to varying rates of deviance. In this way, Durkheimian explanation is causeeffect explanation while behavioral explanation can be seen as providing the
mechanism that, logically speaking, is invoked through covering laws drawn from
behavioral psychology. Combining the two, we have what I have called "Durkheimian
social generativity" (Fararo 1989a: Ch. 2).
The Recent Phase of Theoretical Sociology
I will treat the recent phase of theoretical sociology in two steps. In the first,
preliminary step, I discuss the current state of the field in terms of the existence of
multiple theoretical perspectives inherited from the postclassical phase but
undergoing a process that I describe as producing mutations and new combinations. In
the second, more extensive step, I turn to the growing use of formal models in
theoretical sociology.
Theoretical Perspectives. During the postclassical phase of theoretical sociology and
amidst its proliferation of perspectives, a book appeared that shaped the way many
sociologists reflected upon theory in their discipline. Thomas Kuhn’s The Structure of
Scientific Revolutions ([1962], 1970) argued that a normal science is characterized by
a shared paradigm but that there are also revolutionary episodes in the history of
science involving paradigm shifts.
In application of the paradigm concept to sociology, commentators characterized the
field as one with multiple paradigms, usually called theoretical perspectives. By the
late 1970s, most texts reflected this consensus, featuring separate chapters on
functionalism (Parsons), conflict theory (both critical theory and the Dahrendorf
tradition), exchange theory (Homans), symbolic interactionism (Blumer),
structuralism (French and American versions), and phenomenology (social
constructionism and ethnomethodology).
To make the picture even more diverse, two other developments occurred. Feminists
launched a wide-ranging critique of sociological theory and helped to make the study
of gender a key research topic. Postmodernist sociologists attacked the project of
sociological theory as a continuation of the Enlightenment’s grand narrative with
scientific pretensions that could not succeed. Critics respond by attacking the
cognitive relativism of this approach.
Some commentators argue that there is no possibility of placing these paradigms
under a common intellectual framework, thereby seeing the discipline as permanently
fractured and at war with itself. Others regard the situation as a positive one,
emphasizing the importance of diverse viewpoints that could be brought to bear on
any particular feature of social life. Still others recognize the diversity but argue for
integrative theorizing, as we shall see below.
The most prominent mid-century efforts in theoretical sociology that aimed toward
generality and synthesis -- the ideas of Parsons and Homans described earlier and a
13
14
strong integrative effort by Blau (1964) - have failed on the criterion of acceptance as
the paradigm of general theoretical sociology. Yet the spirit of what they tried to
accomplish is not gone. We can call it "the spirit of unification," meaning a valuecommitment to generalizing synthesis efforts in episodes of consolidating components
of distinct theoretical systems (Fararo 1989b). Robert Merton emphasized this idea in
his often-cited paper "On Sociological Theories of the Middle Range" (included in
([1949] 1968). A middle range theory employs a general conceptual scheme with
analytical elements, but it is scope-restricted to some abstractly specified class of
empirical systems, e.g., thermodynamic systems. It explains intuitively very different
empirical systems using the same analytical elements and laws that do not exhaust the
content of the empirical system. In short, a middle-range theory is an analytical
theory. Its scope is limited in the sense of dealing only with certain analytical
elements, not in the sense of dealing only with a class of concrete entities as classified
in folk culture.
Recent Developments. A value-commitment to the construction of limited-scope but
abstract theories coupled with a recursive process of integration of such theories may
well be a plausible path for the advance of theoretical sociology. At present, this
approach is most strongly institutionalized in the field of research known as group
processes, in which theorists elaborate and integrate their theories over time in
connection with the construction of experimental situations that provide opportunities
for testing the implications of theories. Long-term theoretical research programs,
spanning decades, have characterized some of this work. For instance, expectation
states theory and exchange network theory are two such programs among others (see
Berger and Zelditch 1993). These programs are a kind of mutation out of the earlier
small group research of the 1950s and early 1960s, many of them influenced by the
work of Homans.
Other recent developments indicate other mutations in the paradigms and the
emergence of new paradigms. Three such developments may be noted:
neofunctionalism, social network analysis, and rational choice theory.
Neofunctionalism is a mutation of Parsons' theory. It departs from his four-function
paradigm in a number of ways that reflect the influence of external critiques. For
instance, Alexander (1985) tries to incorporate ideas from various perspectives,
including conflict theory and symbolic interactionism. Much of neofunctionalist
writing is focussed on historical and normative interests. The revised functionalist
framework is employed to interpret historical situations, e.g., in terms of a baseline
social differentiation trend (Alexander and Colomy 1990).
One of the new paradigms is social network analysis. Although social system theorists
such as Parsons and Homans employed the network concept as a metaphor, they did
not employ formal tools. By contrast, the social network paradigm incorporates a
strong mathematical and statistical foundation in a program of cumulative research on
the properties of social networks. In the following section of this paper, when I
discuss structural models, I will pick up on this discussion of social network analysis.
Rational choice theory has departed from the behavioral psychological foundation that
Homans advocated, often favoring a more mathematically tractable rational choice
approach. Coleman (1990) presented his foundations of social theory as directed to
14
15
resolving the micro-macro transition problem that Blau's (1964) earlier effort had
defined. On the one hand, he endorsed Parsons' action framework with its concept of
purposive action and repudiated the transition to structural-functional analysis. On the
other hand, he endorsed Homans' methodological individualism and repudiated the
transition to reduction in terms of behavioral principles.
In Coleman's theory, macro-level systemic givens constrain and enable micro-level
situations of actors. Making rational choices based on their internal preferences and
the situational constraints, actors then collectively shape macro-level outcomes. This
is not equivalent to Homans' reduction program. Among other things, it is a trade-off
of behavioral realism for the deductive fertility that optimization arguments enable.
Homans is much more attuned to the task of the scientific theorist: to explain
empirical findings. Coleman's theory, in part, is more in the classical tradition of
sociological theory as a whole in that it blends general theoretical, world-historical
and normative interests.
A key contribution of sociological rational choice theorists has been their sharp
theoretical focus on variants of the basic problem of order or integration, treating
solidarity, coordination, cooperation, and trust. At the same time, the synthesis of the
Durkheimian theory of solidarity with conflict theory undertaken by Collins provides
a different middle-range perspective on the problem of social integration. I have
pointed out the centrality of this problem in the tradition of theoretical sociology.
Now, with such explicit theories treating it, the problem may well constitute an
important locus of episodes involving theoretical unification. In addition, some of this
work illustrates the use of mathematical models in relation to sociological theory
(Doreian and Fararo 1998). Model building can fulfill a variety of goals, including the
clarification of concepts, the representation of processes, and the specification of
theoretical constructs that explain a variety of phenomena (Berger et al 1962). Such
formal model-building developments are of growing importance in recent theoretical
sociology, a topic to which I turn at this point.
Formal Models in Theoretical Sociology
A model is an abstract entity that functions as a representation of some system in the
world that is of sociological interest. My aim now is to discuss a variety of types of
models that theoretical sociologists have employed in the analysis of social structures
and social processes. Thereafter I will treat the philosophy and methodology of model
building in more general terms. It should be noted that the term "model" is used in
sociology in diverse ways. Very often it refers to statistical models employed in the
analysis of data. This usage is excluded from this discussion, which is focused on
models and model building in relation to sociological concepts and theories. Also
excluded is the diffuse idea of a general model of society as a kind of social organism.
In the present context, a model is a formal object functioning as a representation of
some structure or process of sociological interest. A type of model is constructed,
generally, as an implementation of a representation principle (Fararo 1989a: Ch. 1), a
claim that a certain category of phenomena can be modeled in some specified way. In
15
16
what follows, some important representation principles associated with the concept of
social structure are described, and then the discussion turns to models of social
processes.
Models of Social Structure
Sociologists have employed at least four different types of models in the analysis of
structure in social life. We may regard these as four representation principles under
the headings: structure as network; structure as distribution; structure as grammar, and
structure as game. One aspect of recent theoretical sociology is the use of
combinations of these models in developing theories. Thus, the four types of models
form a set of interrelated conceptual elements (see Figure 4).
Figure 4.
Four Interrelated Representations of Structure in Social Life
Structure as Network. The metaphor of a social system as a network, widely
employed informally in sociology, was transformed into a mode of model building
and analysis through a convergence of ideas and techniques from several traditions.
One such source was sociometry (Moreno 1934), involving the analysis of network
diagrams indicating relationships among people in a small population. A second
source was balance theory, which deals with configurations of positive and negative
sentiments. The theory was absorbed into social network analysis via the
formalization of the configurations of sentiments in terms of signed graph theory
(Harary, Cartwright an Norman 1965). A third source was the analysis of structures of
kinship, especially after the publication of an influential monograph by White (1963).
Sociometric models, balance-theoretic models, models of kinship structure, as well as
numerous other model-building efforts -- such as those treating social diffusion and
16
17
small worlds -- converged by the late 1970s and the term "social network paradigm"
was used to describe this whole area of model building (Leinhardt 1977). Over time, it
became common for measured properties of networks -- for instance, centrality
(Freeman 1977, 1979) -- to be employed in the formulation and testing of empirical
hypotheses about the behavior of actors. By the end of the 20th century, social
network analysis had become a mode of structural analysis with an extensive battery
of formal techniques at its disposal (Scott 1991; Wasserman and Faust, 1994). The
close connection between formal representation, concept formation, and application
makes it a domain of social science that strongly exhibits what Freeman (1984) has
described as "turning a profit from mathematics."
Structure as Distribution. However, social network analysis has been regarded by
most macrosociologists as not the sort of model required for the description of
macrostructure. Sociologists often speak, in the latter context, of such entities as
"occupational structure" or "income structure." These terms refer to distributions.
Blau (1977) proposed a systematic theory in which the key analytical properties of
such distributions, in relation to rates of intergroup relations, provide one type of
answer to the Durkheimian problem of the nature of the integration of a large complex
social system. Blau employed the concepts of heterogeneity, inequality, and
consolidation as such key parameters and formulated theorems relating them to the
extent of intergroup relations, e.g., rates of intermarriage.
A definite model that would represent such a macrostructure was not a part of this
theory, but subsequently Skvoretz and myself formulated a mathematical treatment
(see especially Skvoretz 1983). It drew upon developments in the application of the
theory of random and biased nets (Rapoport and Horvath 1963; Fararo and Sunshine
1964). Thus, structure as distribution is linked to structure as network. All the key
parameters of Blau’s theory are formally linked to key parameters of the biased net
model - in particular, the contact density, the connectivity of the network and, in a
later development, the strength of weak ties. (A summary of the formalization is
presented in Fararo (1989a: Ch. 4).) This development, which we call formal
macrostructural theory, was undertaken in "the spirit of unification" in theoretical
sociology (Fararo 1989b).
Structure as Grammar. A third type of model of structure emerges out of the
language analogy or metaphor employed in one wing of structuralist thought based
upon the work of Saussure (1966 [1915]) and Chomsky (1957). This form of
structuralism has been a perspective based on the idea that in some sense, that social
and cultural systems should be treated with a language-like model (Levi-Strauss 1963
[1958]). One implication of this idea is abstraction from time: the system exists as an
infinite totality to be analyzed by algebraic or other formal tools.
Another strand of such work has been more process oriented, employing the idea that
a set of finite recursively applied rules generates a system of symbolically mediated
interactions comprising a domain of institutionalized social action (Fararo and
Skvoretz 1984). The formalism is drawn from cognitive psychology (Newell and
Simon 1972). The resulting model can be studied from two points of view. On the one
hand, the finite rule basis and the institution stand to each as grammar and language:
the analysis is in the spirit of structuralism (Skvoretz and Fararo 1980). On the other
hand, the finite rule basis can be used to analyze a system of symbolic interaction as it
17
18
is generated locally and in real time (Skvoretz and Fararo 1996b). This type of model
is one among a variety of those that draw upon techniques from artificial intelligence
and cognitive science (Bainbridge et al 1994).
I pointed out earlier how structure as distribution was integrated with structure as
network in formal macrostructural theory. A similar effort, not discussed here (see
Fararo and Skvoretz 1986), links structure as grammar with structure as network,
drawing upon an abstract algebra of interpenetration framed in network terms to
formally represent hierarchical levels of institutional structure (Fararo and Doreian
1984).
Structure as Game. A fourth representation of structure employs game theory. A
play of a game is analogous to an utterance in a language, wherein the rules of the
game play the role of the grammar. Given such rules, a tree of possible sequential
plays of the game is implied, called the game in extensive form. However, as distinct
from grammatical analysis, the focus in game-theoretic analysis is on strategic
interaction, so that a model of rational choice usually supplements the game model.
The aim of the game-theoretic model-builder is to derive the consequences of rational
choices on the part of each player, often with a view of showing how outcomes
involve "perverse effects" (Boudon 1982). Thus, the game model is an alternative to
the grammatical model that emphasizes emergent order at the level of the tacit or
implicit rules governing institutionalized social action. The game model, by contrast,
emphasizes the way in which the structure, as represented by the game, produces
predictable but often-paradoxical effects from the conjunction of rational choices.
It turns out that structure as game has been linked to structure as network. A good
example is the use of game-theoretic ideas to arrive at theoretical predictions of
outcomes of network exchange experiments (Bienenstock and Bonacich 1992). This
type of theory actually combines structure as game and structure as network with
structure as distribution because the outcome of any exchange process in a network is
a distribution of resources among the players. The theory shows how and why this
distribution depends upon the shape of the network. Another example of the linkage
of game and network representations occurs in some of the work of Peter Abell
(1989).
Process Models
The postclassical theoretical sociologists Parsons and Homans (among others) were
committed to the project of bringing dynamic analysis into sociological theory. No
clearer example of this exists than in Homans’ treatment of the social system in The
Human Group (1992 [1950]). So clearly did Homans try to model his discursive
analysis of group phenomena on the set-up and analysis of a system of differential
equations that shortly after this book appeared it was formalized as such by Herbert
Simon (1952), including an early treatment of nonlinear dynamics with multiple
equilibria.
18
19
Coleman (1964), responsive to the needs of survey research with its discrete data
summarized as proportions, developed a family of dynamic models that are stochastic
processes in continuous-time. Each individual makes transitions from one discrete
state to another - for instance, shifting candidates during an election campaign - and
the group makes transitions among states representing the number of individuals in
each of the discrete individual states (e.g., the number of people favoring a particular
candidate at a particular time.) This Coleman methodology extends to the social
network context in which each individual’s transition is influenced by a composite
flow of influence from other individuals to whom the person is connected in some
social relationship.
The most general way of thinking about processes is in terms of the concept of a
behavior manifold (see Figure 5, upper part). This consists of a parameter space
together with a state space, both multidimensional. Given a time domain and a
generator, the parameterized process is the tracing out of trajectory in state space. For
any given value of the parameter, apart from transient states, there may be various
types of attractors (generalized forms of equilibrium) as well as repellors (unstable
equilibria). For instance, the nonlinear Simon-Homans model, under some conditions
yields a configuration of two attractor states separated by a repellor. Thus, when an
initial state is close to the repellor it departs from it toward one or the other attractor.
The whole subject of nonlinear systems is framed in terms of such generated
configurations in state space that vary with parametric conditions. Special cases of the
general dynamical system include topics catastrophes and chaos as well as classical
linear system dynamics where equilibrium, if it exists, is unique. For an extended
discussion, see Fararo (1989a: Ch.2).
19
20
Figure 5.
Process and Social Structure
Models Combining Structure and Process
Of particular interest in sociology are two types of process models that relate to the
concept of social structure. (See Figure 5, lower part.) In one type, a network or some
other model object represents the structure, and other phenomena, say X, are taken as
defining the state space. The aim is to show how the outcome of a postulated process
with respect to X varies with parameters descriptive of the social structure. In the
other type of model, the structure is treated as emergent. An interaction process model
involving "E-states" may serve to illustrate (Skvoretz and Fararo 1996). The process
involves the over-time construction of stable relationships among pairs of actors until
equilibrium, when the postulated rules lead to social reproduction of the relational
pattern. The process is the trip through a state space of possible forms of the emergent
local social structure. Which trip is taken, in terms of which network states are visited,
depends upon the initial state, the parameters, and the specific realization of the
stochastic process representation of the generator.
20
21
General Considerations
I conclude this paper with a presentation of a conception of how to think about
models in relationship to the knowledge process as involving theories, data and the
relationship between them. The following discussion relates to Figure 6.
Figure 6.
Theoretical Model Building
Framework, Problem and Model. Sociologists, like other social scientists, use the
term "theory" to cover both general frameworks and more specific formulations that
address particular problems. This double usage can be articulated to the model
concept. Namely, we think of a scientific theory as having two levels, a framework
level and a model level. The two are linked by theoretical problems that are addressed
by constructing a model within the framework.
Suppose that T is a general theoretical framework, comprised of general concepts and
principles. Often a formal theoretical framework will contain what I will call a
"template," meaning a general form of a model that needs to be "filled-in" with more
definite terms. For instance, the Newtonian framework contains the famous F = ma
formula that provides a template for mechanical models.
Associated with the T-framework will be various problems, for instance, phenomena
calling for an explanation, a "T-problem." The theorist will invoke the T-framework
to address a T-problem in terms of theoretical methods to generate a theoretical model
appropriate to the problem, call it a T-model.
21
22
In addition, investigators will employ empirical methods to generate data appropriate
to the problem. In particular cases, this is followed by such procedures as parameter
estimation and calculations of empirical predictions. The comparison of the latter with
properties of the body of data may show discrepancies that, in turn, may lead to
revisions of the theoretical model, to questioning the quality or relevance of the data,
to a reformulation of the problem, or a revision of the general framework itself. Even
the worldview is not immune from rethinking, although this would be a last resort to
resolve some intolerable inconsistencies not only between data and models but also
between different frameworks within a research tradition.
In the event of a favorable assessment of the theoretical model, a natural step would
be extension of the scope of the theoretical model through removal of analytical
restrictions that were introduced to facilitate a first theoretical approach to the
problem.
This sketch works best when the framework entails formal model building. Let me
illustrate with a sociological example. In the recent phase of theoretical sociology,
Coleman (1990) constructed a framework with both a metatheoretical template and a
theory template. The former consists of an already famous "boat" diagram in which -in one interpretation -- a given macro initial condition M0 produces an outcome
macro-state M1 via three linkages. First, there is a linkage from macro M0 to micro
m0, interpretable as an actor with socially induced preferences in a situation with
opportunities and constraints. Second, there is a link from m0 to m1, an act by that
actor, postulated, as a first approximation, to be a rational choice. Then, third, some
mechanism combines the acts of the various actors to generate the macro outcome to
be explained, a link from m1 to M1. This metatheoretical template serves to orient
theorists to construct models that explain macrosociological causal relations by
postulation of theoretical models that incorporate the three types of links.
Coleman's theory template is a generalization of the logical structure of general
equilibrium theory in economics, compactly represented in two matrices. First, there
is a matrix in which there is a distribution of rights of control of a set of resources
among a set of actors. Second, there is a matrix of the distribution of each actor's
interests over the same resources, where the interests are parameters in a CobbDouglas utility function. The template then invokes an exchange process to carry the
state of the control matrix from its initial state to an equilibrium state. In this process,
each actor's utility function is maximized subject to constraints. Thus, to create a
theoretical model based on this framework means to specify the actors, the resources,
and the initial control relations and interests. Thus, in Coleman's structure of theory,
the exchange theory template satisfies the metatheory template and in turn exchange
models created within the framework are designed to satisfy the theory template. For
instance, one theoretical problem that Coleman poses is: How do norms emerge? The
theoretical model he proposes employs the theory template to address this problem.
Coleman includes conditions necessary for "the demand for a norm" to arise and also
conditions necessary for effective enforcement of the emergent norm.
A somewhat different and briefer example may be given in terms of the use of the
construct "E-state," mentioned earlier. E-state structuralism is a theoretical method,
functioning as the basis for model of network dynamics in which actors hold evolving
expectation states with respect to each other, as in a small group discussion setting
22
23
(Skvoretz and Fararo 1996). In this instance, the framework is the core of expectation
states theory itself with its principle that relational expectation states arise out of
social behavior and then come to form stable bases for the control of such behavior as
indicated by differential rates of participation in group discussion. Given the problem
of describing such an interactive process in detail, the E-state structuralist method, as
combined with several other ideas, yields a model that makes detailed predictions.
Preliminary tests of the model were undertaken but it was clear that far more detailed
interaction data were required. In turn, this led to further empirical inquiry to generate
this more appropriate body of data to permit more refined tests of the predictions
yielded by the dynamic model. This example, then, illustrates some of the dynamic
aspects of the interplay of two modes of implementation of a theoretical framework,
one involving theoretical methods that aid in the construction of theoretical models
and the other involving empirical methods that aid in the collection of appropriate
data.
Although these ideas about theoretical frameworks, theoretical problems and
theoretical models were devised with formal theories in mind, they also enable us to
interpret the logical structure of theoretical work that is not formal. For instance, let T
be a structural-functionalist theoretical framework. One T-problem is to explain the
universality of stratification, which is understood within the T-framework to refer to
rewards, especially prestige, assigned to positions in a social system. The famous
Davis-Moore theory of stratification can be interpreted as a T-model proposing a
theoretical solution of this problem. Employing some ideas about motivation, what
the authors do is equivalent to proving a theorem about the T-model: If a social
system, a system of interrelated positions, is stable, then that system is stratified.
Hence, stratification is a necessary condition for social order.
To derive empirically testable claims that can be compared with appropriate data, a
formalized functional approach would be helpful. For instance, Stinchcombe (1968)
represents functional arguments in terms of a negative feedback or homeostatic
system. The problem of appropriate data for the empirical assessment of such
functional models is addressed by Faia (1986).
Representation, Idealization and Approximation. The connection between
sociological frameworks and formal model-building will become much closer as
theoretical sociologists become more explicitly oriented to three basic aspects of
theoretical model-building, namely representation, idealization and approximation
(Fararo 1989a: Ch.1). Representation is the core idea of model building and,
therefore, in the context of constructing effective theoretical frameworks in sociology
an essential aspect of theory development. Berger et al (1962) set out an important
statement of the linkage between theoretical goals and model building. In their terms,
there are three basic goals that motivate the construction of a model: to explicate a
concept of a theory, to represent a recurrent process, or to formalize a theory in terms
of some theoretical construct.
The role of idealization in sociology was recognized by the classical sociologist Max
Weber, who used the term "ideal type" for "model." A model is based upon an act of
abstraction. Only selected aspects of a concrete reality are represented. Even key
features of reality may be omitted in this process in order to study a "pure" case. For
instance, economic theorists define and study general equilibrium models of perfectly
23
24
competitive economies. As of the late 20th century, however, this important role of
idealization has not yet found its way into most theoretical work in sociology. An
exception occurred in the work of Coleman (1990) in his adoption of the economic
approach, formulating the concept of a "perfect social system" and employing a
general equilibrium theory.
The role of approximation is closely related to the logical derivation of properties of a
theoretical model. For instance, a complex mathematical expression may be
approximated by a simpler one that enables deductions that would not otherwise be
obtained.
Standards in Theoretical Model Building. Implied in this entire discussion of
formal models in theoretical sociology is some conception of cognitive standards for
the construction and assessment of models. Lave and March (1975: Ch. 3) have
produced a lucid discussion of such standards. Two groups of standards they explicate
under the headings of truth and beauty, respectively.
I will discuss one of the standards of truth. There is general agreement among
philosophers of science that a model is not really a theoretical model unless it can be
shown to be wrong in relation to the world. This is what Lave and March call "the
importance of being wrong." The comparison operation mentioned earlier bears upon
this aspect of model building. The standard may be called "truth," but idealization and
approximation have to be taken into account. The more precise the prediction made
by a model the more likely it is to be untrue in the strict sense. The real point is that
the development of our collective grasp of the world in respect to the problems we
pose is under empirical control as well as informed by conceptual schemes and
theories. Another important point is that the generality of a framework enables
alternative models to be constructed. In addition, it may be that a given problem can
be re-framed so as to enable a quite different framework to be employed in the
construction of an alternative model. In principle, this could lead to critical
experiments to compare and judge two models.
Beauty is the other evaluative category. Fertility and surprise are two of the standards
of beauty in model building. A poor model, in respect to fertility, is one that has no
logical consequences we consider worthy of noting. A good model is fertile in the
deductive sense and it is an even better model if some of the consequences are
surprising, not at all obvious in the setting-up of the model.
A second aspect of beauty is simplicity. Complexity in models is to be sought at the
level of derived consequences, not at the level of postulates. In a process model, a few
simple rules of transition can lead to enormous complexity in the concatenation of
these rules over time and in regard to distinct actors in a system. Model builders
usually urge that their readers wait and see what the results are before abandoning a
model because a specific assumption seems too idealized or even "wrong," as if it
were an empirical generalization.
Let me add two ideas about standards in model building. Both relate to what seems to
be required to have an explanatory model. On the one hand, from a formal point of
view what seems essential is some kind of mechanism or rule-set that generates the
phenomenon to be explained. A postulated process literally shows how the
24
25
phenomenon arises, deducing it from premises or computing it in a simulation of the
postulated process. On the other hand, and this is perhaps more controversial, from an
interpretive point of view what seems important is that this generativity should be
based upon premises that refer to understandable human action. Simplicity of
postulates about the actions of agents, with complexity of generated systemic
outcomes: that is the standard that sums up these two ideas.
Realization of this standard is now in progress in theory-driven simulation studies,
often grouped under the rubric of computational sociology (Hummon and Fararo
1995). Simulation methods enable process models to be constructed that are based on
nonlinear parallel processing by a diverse set of actors in a dynamic network. The
consequences of the process rules are generated in "runs" of the computational model
rather than logically derived. However, there is still a place for the analytically
simpler types of models that enable the derivation of crisp theorems. Theoretical
sociologists are among those on the frontier of these new computational developments
even as older formal methods continue to be important.
Summary
The tradition of sociological theory exhibits a mixture of three types of intellectual
interests that I have called theoretical sociology, world-historical sociology and
normative-critical sociology. This was illustrated in the classical phase of the
tradition, where we see the beginnings of theoretical sociology as well as a focus on
world-historical trends and their normative assessment. I discussed the postclassical
phase of theoretical sociology by reference to the common aim of generalized
synthesis found in the successive works of Parsons and Homans. Between their earlier
and later phases of theorizing, each shifted theory construction strategy while
maintaining continuity of sociological ideas, Parsons to his four-function paradigm
and Homans to his program of behavioral reduction.
In the most recent phase of theoretical sociology, I have emphasized that formal
models have become a major part of the tradition. I have tried to highlight major
developments in terms of models of structure and of process. I also discussed two
types of models that combine a structural focus with process analysis. Finally, in the
context of treating a number of general considerations about formal models in
sociology, I described a general perspective on theoretical model building in relation
to theoretical frameworks.
In sum, this paper has provided an historical perspective on theoretical sociology from
the classic tradition to postclassical efforts of synthesis that culminated in multiple
paradigms to the situation today in which new developments -- that are productive -more and more rely upon formal models as essential components of their
methodology.
25
26
References
Abell, Peter. 1989. "Games in Networks: A Sociological Theory of Voluntary
Associations." Rationality and Society 2: 259-282.
Alexander, Jeffrey C. 1985. Editor. Neofunctionalism. Newbury Park, CA: Sage.
_____ and Paul Colomy. 1990. Editors. Differentiation Theory and Social Change:
Comparative and Historical Perspectives. New York: Columbia University Press.
Bainbridge, William, and Edward Brent, Kathleen Carley, David Heise, Michael
Macy, Barry Markovsky and John Skvoretz. 1994. "Artificial Social Intelligence."
Annual Review of Sociology 20: 407-436.
Baum, Rainer. 1975. "The System of Solidarities." Indian Journal of Social Research
16: 307-352.
Berger, Joseph, and Bernard P. Cohen, J. Laurie Snell and Morris Zelditch, Jr. 1962.
Types of Formalization. Boston: Houghton Mifflin.
Berger, Joseph and Morris Zelditch, Jr. 1993. Editors. Theoretical Research
Programs: Studies in the Growth of Theory. Stanford, CA: Stanford University Press.
Bienenstock, Elisa J. and Phillip Bonacich. 1992. "The Core as a Solution to
Exclusionary Networks." Pp. 231-243 in David Willer (editor), Special Issue on the
Location of Power in Exchange Networks. Social Networks 14: Nos. 3-4.
Blau, Peter. 1964. Exchange and Power in Social Life. New York: Wiley.
_____. 1977. Heterogeneity and Inequality: A Primitive Theory of Social Structure.
New York: Free Press.
Boudon, Raymond. 1982. The Unintended Consequences of Social Action. New York:
Macmillan.
Chomsky, Noam. 1957. Syntactic Structures. The Hague: Mouton.
Coleman, James S. 1964. An Introduction to Mathematical Sociology. New York:
Free Press.
_____. 1990. Foundations of Social Theory. Cambridge, MA: Harvard University
Press.
Collins, R. 1975. Conflict Sociology. New York: Academic Press.
Collins, Randall. 1994. Four Sociological Traditions. New York: Oxford University
Press.
26
27
Dahrendorf, Ralf. 1959. Class and Class Conflict in Industrial Society. Stanford, CA:
Stanford University Press.
Doreian, P. and T. J. Fararo. 1998. Editors. The Problem of Solidarity: Theories and
Models. Amsterdam: Gordon and Breach.
Faia, Michael A. 1986. Dynamic Functionalism: Strategy and Tactics. Cambridge:
Cambridge University Press.
Fararo, Thomas J. 1989a. The Meaning of General Theoretical Sociology: Tradition
and Formalization. New York: Cambridge University Press.
_____. 1989b. "The Spirit of Unification in Sociological Theory." Sociological
Theory 7(2): 175-190.
_____ and Patrick Doriean. 1984. "Tripartite Structural Analysis." Social Networks 6:
141-175.
_____ and John Skvoretz. 1984. "Institutions as Production Systems." Journal of
Mathematical Sociology 10: 117-181.
_____ and John Skvoretz. 1986. "Action and Institution, Network and Function: the
Cybernetic Concept of Social Structure." Sociological Forum 1(2): 219-250.
_____ and Morris Sunshine. 1964. A Study of a Biased Friendship Net. Syracuse:
Youth Development Center and Syracuse University Press.
Freeman, Linton C. 1977. "A Set of Measures of Centrality based on Betweeness."
Sociometry 40: 35-41.
_____. 1979. "Centrality in Social Networks: I. Conceptual Clarification." Social
Networks 1: 215-239.
_____. 1984. "Turning a Profit from Mathematics: the Case of Social Networks." Pp.
125-142 in Thomas J. Fararo (editor) Mathematical Ideas and Sociological Theory.
New York: Gordon and Breach.
Granovetter, Mark. 1973. "The Strength of Weak Ties." American Journal of
Sociology 78: 1360-1380.
Harary, Frank, and Robert Z. Norman and Dorwin Cartwright. 1965. Structural
Models: An Introduction to the Theory of Directed Graphs. New York: Wiley.
Homans, George C. 1992 [1950]. The Human Group. New Brunswick, NJ:
Transaction [Harcourt, Brace & World].
Hummon, Norman P. and Thomas J. Fararo. 1995. "The Emergence of Computational
Sociology." Pp. 145-159 in David Heise (editor), Special Issue on Sociological
Algorithms. The Journal of Mathematical Sociology 20: Nos. 2-3.
27
28
Kuhn, T. 1970 [1962]. The Structure of Scientific Revolutions. University of Chicago
Press, Chicago.
Lave, Charles and James March. 1975. An Introduction to Models in the Social
Sciences. New York: Harper and Row.
Leinhardt, Samuel. 1977. Editor. Social Networks: A Developing Paradigm. New
York: Academic Press.
Levi-Strauss, Claude. 1963 [1958]. Structural Anthropology. New York: Basic Books.
Merton, R. 1968 [1949]. Social Theory and Social Structure. New York: Free Press.
Moreno, J. L. 1934. Who Shall Survive? Beacon Press.
Newell, Alan and Herbert A. Simon. 1972. Human Problem Solving. Englewood
Cliffs, NJ: Prentice-Hall.
Parsons, Talcott. 1937. The Structure of Social Action. New York: McGraw-Hill.
_____. 1951. The Social System. New York: Free Press.
Rapoport, Anatol and William J. Horvath. 1961. "A Study of a Large Sociogram."
Behavioral Science 6: 279-291.
Saussure, Ferdinand de. 1966 [1915]. Course in General Linguistics. New York:
McGraw-Hill.
Scott, John. 1991. Social Network Analysis: A Handbook. London: Sage.
Simon, Herbert A. 1952. "A Formal Theory of Interaction in Social Groups."
American Sociological Review 17: 202-212.
Skvoretz, John. 1983. "Salience, Heterogeneity and Consolidation of Parameters:
Civilizing Blau’s Primitive Theory." American Sociological Review 48: 360-375.
_____ and Thomas J. Fararo. 1980. "Languages and Grammars of Action and
Interaction: A Contribution to the Formal Theory of Action." Behavioral Science 25:
9-22.
_____ and Thomas J. Fararo. 1996a. "Status and Participation in Task Groups: A
Dynamic Network Model." American Journal of Sociology 101: 1366-1414.
_____ and Thomas J. Fararo. 1996b. "Generating Symbolic Interaction: Production
System Models." Sociological Methods and Research 25(1): 60-102.
Stinchcombe, Arthur. 1968. Constructing Social Theories. Chicago: University of
Chicago Press.
28
29
Wasserman, Stanley and Katherine Faust. 1994. Social Network Analysis. New York:
Cambridge University Press.
White, Harrison C. 1963. An Anatomy of Kinship. Englewood Cliffs, NJ: PrenticeHall.
29