Вы находитесь на странице: 1из 64

Dublin City University

Inside the Demon Box


Science, communications, culture and the
transformation of entropy

Submitted in partial fulfilment for the Masters in


Science Communication

Warren J. Whitney

September 2006
I hereby declare that this thesis is of my own work, unless otherwise stated.

Warren Whitney
September 2006

ii
Physicists cease to be physicists unless they hold that the law of Entropy includes Gods
and men as well as universes.

Henry Adams, A Letter to American Teachers of History

He found in entropy or the measure of disorganization for a closed system an adequate


metaphor to apply to certain phenomena in his own world….He found himself, in short,
restating Gibbs’ prediction in social terms, and envisioned a heat-death for his culture in
which ideas, like heat-energy, would no longer be transferred, since each point in it would
ultimately have the same quantity of energy; and intellectual motion would, accordingly,
cease.

Thomas Pynchon, Entropy

iii
Acknowledgements

The credit for inspiring the idea of a dissertation on entropy must go jointly to Michael
John Gorman (whose lecture and discussion session in the Science and Society module set
me thinking about artistic uses of entropy) and my mother (for an unconnected query about
entropy shortly thereafter which further).

Brian Trench was the person who finally encouraged me to choose the topic, and went on
to supervise the dissertation. I am grateful for his advice and support, particularly in the
difficult early stages.

Numerous other people at DCU have been helpful to me in the course of my work. Eve
Merton and Pádraig Murphy provided invaluable feedback and general advice during the
development of the proposal (and afterwards). I would also like to thank the staff of DCU,
Trinity and UCD libraries for their assistance.

Amongst my classmates, I would like to thank Marieke Houben and Rachel Brennan
specifically for their helpful suggestions in relation to this dissertation. I would also like to
thank everyone else – Carmel, Claire, Bernadette, Eros, Danielle, the three Brians,
Gráinne, Karen, Patrick – for their friendship and support during what were often difficult
times for me. Sarah Hilley, the honorary Science Communicator, deserves special thanks
for her interest in and support for my work.

Although he was not directly involved in this dissertation, I owe a lot to Jim McGovern,
now of DIT but ten years ago the person who first attempted to explain entropy to me, as
the lecturer on my introductory thermodynamics course at Trinity College. Needless to say,
he is not responsible for any of my scientific errors.

Finally, thanks to my mother, father and brother for being so supportive throughout, and to
anyone I may have forgotten.

iv
Abstract

This dissertation examines the process by which the idea of entropy, having originated in
the field of thermodynamics, became established in cultural (particularly literary) contexts
during the twentieth century.

Initially, entropy and key associated concepts are introduced. This is followed by a brief
description of the historical background to the emergence of the idea of entropy and its
subsequent development as a scientific concept. The theories of metaphor and knowledge
production (which will be applied to the case of entropy) are then outlined.

The history of entropy’s transition from science to culture is summarised, particular


attention being given to its emergence as a cultural concept in the mid-twentieth-century
United States. Within this process, special attention is given to certain aspects, such as
communication theory’s adoption of entropy and the usage of the idea in science fiction
and, in the broader literary context, by specific writers such as Thomas Pynchon.

This transition is then discussed in terms of metaphor (including both general theories of
metaphor and a specific previous analysis of entropy as metaphor) and knowledge
production.

Conclusions are drawn on the basis of the analysis; it is concluded that theories of
knowledge production provide a better overall explanation of entropy’s migration from
science into culture than do theories of metaphor. Within the field of metaphor, the
metaphorical aspects of communication and cybernetic theorists’ use of entropy appear
more important than any inherent metaphorical characteristics of earlier scientific writings
on the subject. Areas having potential for deeper analysis are identified, and attention is
drawn to the “metaphorical popularisation” involved in factual writing for general
audiences which uses entropy in contexts other than thermodynamics; it is possible that
this represents an aspect of scientific popularisation deserving of further attention.

v
TABLE OF CONTENTS

Page
1 Introduction 1
1.1 Overview 1
1.2 Definitions 3
1.2.1 Thermodynamic entropy 3
1.2.2 The Second Law of Thermodynamics 3
1.2.3 Information entropy 5
1.2.4 Maxwell’s demon 5
1.3 Historical background 7
1.3.1 Overview 7
1.3.2 The beginnings of entropy: Clausius and Thomson 7
1.3.3 Further development of entropy: Helmholtz, Gibbs,
Boltzmann et al. 10

2 Approaches to entropy 13
2.1 Overview 13
2.2 Metaphor 13
2.2.1 General theories 13
2.2.2 Zencey’s analysis of entropy as metaphor 14
2.3 Knowledge production 15
2.4 Miscellaneous analyses of entropy 15

3 The transformation of entropy 18


3.1 Introduction 18
3.2 Early interdisciplinary usages 18
3.2.1 The science of human degradation: Henry Adams and
history as entropy 18
3.2.2 The beginnings of literary entropy 21
3.2.3 All order is doomed: Nathanael West and entropy 21
3.3 Post-war America and the spread of entropy 22
3.3.1 Information, choice and uncertainty: Claude Shannon and
communication entropy 22
3.3.2 Progress in a universe running downhill: Norbert Wiener,
cybernetics and entropy 25
3.4 Maxwell’s demon and interdisciplinary entropy 28
3.5 Emergence of literary entropy 29
3.5.1 Background 29
3.5.2 A final state of total, absolute kippleization: Entropy in
science fiction 30
3.5.3 Entropy’s wider use in literature 32
3.5.4 A metaphor not only verbally graceful, but also
objectively true: Thomas Pynchon and entropy 33
3.6 Conclusion 38

4 Discussion 39
4.1 Overview 39

vi
4.2 Entropy as metaphor 39
4.2.1 General considerations 39
4.2.2 Zencey’s suggested explanations 41
4.3 Entropy and knowledge production 45
4.4 Summary 48

5 Conclusion 50

References 53

TABLE OF FIGURES

Page
Figure 1 A Maxwell’s demon 6
Figure 2 Shannon’s model of a communication system 23

vii
1 Introduction

1.1 Overview

Entropy is a somewhat rarefied scientific concept, originating in the mid-19th century


within thermodynamics (the science of heat transfer), which has rather improbably been
adopted by a variety of non-scientific fields, including (but not limited to) literature,
communication theory and management science. The process by which this adoption took
place in the case of literature is the primary subject of this dissertation.

In the following sections of this chapter, I give more elaborate definitions of entropy and
the key concepts associated with it, along with a brief outline of its historical background.
For the time being, it suffices to define entropy, in the thermodynamic context, as energy
not available for work or as disorder (in a relatively strict statistical sense). The Second
Law of Thermodynamics states that (in an isolated system) entropy can only increase;
energy tends towards more disordered, less usable states, where, ultimately, it will be
impossible to utilise without inputting more energy than is produced.

In practice, the concept of entropy is almost always used in the (broad) context of the
Second Law when applied in non-scientific discourse; however, the word “entropy” itself
appears to be used far more widely than explicit references to the Second Law. Aside from
the historical context of the ideas’ original development, entropy and the Second Law are
effectively synonymous for the purposes of this dissertation.

The concept of entropy is not always readily understood in its original thermodynamic
sense; even an introductory thermodynamics textbook for third-level engineering students
acknowledges that “it is not easy to develop an intuitive grasp of…entropy” (McGovern
1996 p166). In the light of this, it may seem surprising that it would be so widely adopted
across other disciplines.

1
Part of the explanation may lie in the fact that entropy, in most of its usages, is by now
somewhat detached from its original meaning. As Zencey (1991) noted, it has “exercised a
fascination on thinkers far removed from the realms of physics” and has achieved the
status of “a world-ordering or root metaphor”. Tanner (1971 p141) points out that, at least
in the literary context, the word “is used now with a looseness which any scientist would
deplore”. In at least some of its contemporary uses, entropy has come to mean nothing
more than a kind of generalised decay. As Nicholls (1993a p386) phrases it:

Since the 1960s, to the annoyance of some scientifically minded people, the
concept of increasing entropy includes holes wearing in socks, refrigerators
breaking down, coalminers going on strike, and death.

My objective in this dissertation is to give an overview of how an arcane thermodynamic


concept came to be transformed into something which most users of the term would regard
more as a literary or cultural idea. However, since this process was neither especially rapid
nor direct, this also requires examining the concept’s progress through various other fields
en route, not to mention the tortuous process of its emergence within thermodynamics.

For the time being, I am only attempting to tell the story of entropy, with particular focus
on its transition from science to literature, and analyse how this “jump” occurred. In
support of this analysis, I make use of theories of metaphor and knowledge production,
which I discuss in chapter 2 of this dissertation. Chapter 3 outlines the history of entropy’s
transition, and I describe my actual analysis in chapter 4. Conclusions are contained in
chapter 5.

Inevitably, there are things which have had to be left out. I leave detailed analysis of, for
instance, how the accepted meaning of entropy might have changed in this transition, for
subsequent work. I also regret the fact that, by and large, I have had to exclude usages of
the concept of entropy which are neither thermodynamic nor literary; I have, however,
referred to some such usages where they represent important stages on the idea’s journey
between these two fields.

2
1.2 Definitions

1.2.1 Thermodynamic entropy

In the original thermodynamic sense, entropy measures the amount of energy irrevocably
dissipated in a heat transfer process; in the words of the Concise Oxford English
Dictionary (2004 p477), it “represent[s] the unavailability of a system’s thermal energy for
conversion into mechanical work”. Rudolf Clausius (1822-1888), a German mathematician
and physicist, introduced entropy as a named concept in 1865, though he (and others) had
conceived the idea of such a quantity as early as 1850 (see 1.3.2 below).

In the context of statistical mechanics (the application of statistical methods to molecular


behaviour), the state of a body can be described in two ways. On a large scale, the “macro-
state” deals with external properties such as temperature and pressure, while the smaller-
scale “micro-state” describes the position and velocity of the body’s individual molecules.
A large number of different micro-states can correspond to a specified macro-state. The
Austrian physicist Ludwig Boltzmann (1844-1906) proved in 1872 that entropy is
proportional to a quantity known as W. This indicates the number of possible micro-states
corresponding to a single macro-state. It can be interpreted as indicating the macro-state’s
probability (obviously, the more ways in which it can occur, the more likely it is to occur);
thus, higher-entropy states are more probable. Furthermore, “disordered or random
arrangements” of molecules are more probable than highly ordered ones, so increased
entropy can be equated with increased disorder on the molecular level. The connection of
entropy to disorder has been influential in its spread outside scientific contexts.

1.2.2 The Second Law of Thermodynamics

The Second Law was first stated by Clausius in 1850, simultaneously with his
identification of the quantity that was to become entropy. At this point, the law was
phrased to the effect that heat would only flow from a hotter to a cooler body without some
external work input. It complemented the earlier First Law of Thermodynamics (the
conservation of energy within a closed system).

3
When Clausius introduced entropy as a named concept in 1865 (as further described
below), he restated the Second Law in terms of the new idea as “the entropy of the
universe tends towards a maximum” (Clausius 1865 cited in Jammer 2003 p114). Since
then, the Second Law has become intimately associated with entropy, to the point that – as
previously mentioned – the ideas are used interchangeably in non-scientific contexts.

In terms of statistical mechanics, the Second Law can be phrased in terms of a tendency to
increased disorder or towards a more probable state (as discussed above, the two are
equivalent). Boltzmann stated that, as entropy increases, “the system tends towards a state
of maximum probability” (Jammer op. cit. p115). While the relationship of entropy to
probability has had some influence on its subsequent non-scientific uses, the idea of a
physical law dictating that disorder can only increase has had a particularly captivating
effect on non-scientists.

The “one-way” nature of the Second Law has also led to its becoming associated with the
unidirectionality of time itself (the so-called “arrow of time”). Since entropy increases
irrevocably over time, reversing the direction of time would violate the Second Law. As
Zencey (op. cit. p48) phrases it, the Second Law is “the only physical law of universal
content that describes an irreversible process” and [o]ur subjective experience of time is
intimately connected” with it. This may also have been a factor in the adoption of the idea
of entropy in non-scientific disciplines.

It is important to remember that the Second Law is only valid within a closed system (i.e.
no energy flows in or out). While the whole universe can be regarded as a closed system, a
single planet, or an individual organism, cannot. Disputes about the applicability of the
Second Law to a particular case may turn on this; for instance, Rothstein (2003), in
response to creationists’ claims that the Second Law contradicts the increased degree of
order involved in evolution, points out that “evolution does not take place in a closed
system”. The idea of a closed system has also been taken up in some metaphorical usages
of entropy.

4
1.2.3 Information entropy

The idea of entropy was first applied to information theory by Claude Shannon in 1948, as
is described in greater detail in chapter 3 of this dissertation. In the context of his model of
human communication as a channel or conduit through which a message is transmitted,
Shannon proposed a quantity as a measure of the randomness and uncertainty introduced
into the message by the transmission process. Since this was governed by similar statistical
laws to those which Boltzmann had observed for thermodynamic entropy, he named the
quantity “entropy”; however, as will be seen, the exact relationship of the two entropies
remains contested.

The usage of entropy in information theory is of significance in the overall context of this
dissertation not only because it represents an instance of the interdisciplinary migration of
entropy but also because it played a significant role in attracting the attention of writers
and of the general public to the concept. Furthermore, it was important to attempts to
address the problem posed by Maxwell’s demon (see below); this process in turn attracted
further interdisciplinary attention.

1.2.4 Maxwell’s demon

In the early years of entropy’s existence as a concept, it remained subject to question and
challenge, particularly in respect of micro-scale phenomena. It was in this context that the
Scottish physicist James Clerk Maxwell (1831-1879) proposed a thought experiment in
1867, subsequently introducing it to a wider audience in his 1871 book Theory of Heat
(Leff and Rex 1990). This posited a box divided into two compartments by a partition
containing a door or valve. Each compartment would contain air molecules at uniform
temperature (and thus with a uniform average velocity, though the velocities of individual
molecules would be highly varied). An intelligent creature would observe the velocity of
each molecule as it approached the partition, and open the door where necessary so as to
concentrate higher-velocity ones in one compartment and lower-velocity ones in the other
(Figure 1). This creature was ultimately to become known as Maxwell’s Demon; in fact

5
William Thomson (1990 [1874]), later Lord Kelvin, rather than Maxwell himself, was the
first to use the name “demon”.

Figure 1 A Maxwell’s demon (Schneider and Sagan 2005)

This outcome clearly represents an increased degree of order. Since it is assumed that the
door is frictionless and so the demon does no work, it follows that the box – obviously a
closed system – is progressing towards a more ordered, less entropic state without work
input, and thereby violating the Second Law. (It may also be seen as permitting a perpetual
motion machine, usually forbidden by the Second Law, since the temperature difference
between the two compartments could be used to drive an engine.)

Much attention has been given to Maxwell’s Demon over the years, and the proposed
solutions to the problem have generally involved some form of communication entropy, as
is further explained later in this dissertation.

6
1.3 Historical background

1.3.1 Overview

The objective of this section is to outline the process of entropy’s development in its
original thermodynamic usage, in order to expand on the definitions given above and
support my subsequent analysis of the idea’s interdisciplinary spread. For the most part, I
concentrate on the roles of a limited number of prominent individuals, but without
implying the insignificance of others’ contributions. In some instances, I give special
attention to aspects of their work – such as the religious/philosophical bent of certain of
Thomson’s writings – which may have influenced the way in which entropy was
subsequently applied.

1.3.2 The beginnings of entropy: Clausius and Thomson

The science of thermodynamics began to emerge in the early 19th century, in parallel with
the evolution of the steam engine. Jammer (op. cit. p113) credits Nicolas Léonard Sadi
Carnot (arguably the founder of thermodynamics in the modern sense) with being the first
to observe that “part of the stored energy is always dissipated” in a heat-transfer process.
This observation was ultimately (if tortuously) to lead to the emergence of the concept
which became known as entropy.

Following on from Carnot’s work, Joule, Mayer and Helmholtz established the First Law
of Thermodynamics, i.e. the conservation of energy within a closed system. As this
obviously conflicted with Carnot’s findings – energy could not be simultaneously
dissipated and conserved – it appeared that further work was necessary (ibid.).

Rudolf Clausius was to make the decisive contribution to establishing entropy as a


scientific concept. Clausius was twenty-eight, having completed his doctorate on the
subject of optics just two years earlier, when he delivered his best-known paper, Über die
bewegende Kraft der Wärme (variously translated as “On the driving power of heat” or
“On the motive force of heat”) to the Berlin Academy in 1850 (O’Connor and Robertson
2000).

7
In this paper, Clausius proposed that heat could only flow (without external work input)
from hotter to colder bodies. In doing this, he established what became known as the
Second Law of Thermodynamics. According to Truesdell (1980 p204), “[t]here is no doubt
that with this paper Clausius created classical thermodynamics”. He also recognised the
existence of a “quantity that remains invariant during changes of volume and temperature”
in the same paper (O’Connor and Robertson op. cit.). However, he did not yet name this
quantity.

William Thomson (1824-1907), later Lord Kelvin, had been one of those who observed the
necessity of further work in thermodynamics prior to Clausius’s paper. In its aftermath, he
took up the German’s ideas enthusiastically. His first major post-Clausius paper on heat
transfer, On the Dynamical Theory of Heat (Thomson 1851) endorsed Clausius’s work
(along with that of Rankine) as “[i]mportant contributions to the dynamical theory of
heat”(ibid. 5). While developing his own statement of the Second Law (which he claimed
had occurred to him before he knew of Clausius’s work), he acknowledged that “the merit
of first establishing the proposition upon correct principles is entirely due to Clausius”
(ibid. 14). The issue of which of Clausius and Thomson had precedence appears still
unresolved; Smith and Wise (1989 p327) suggest that both came to the same conclusion
via “largely separate” paths.

Thomson followed this with On a Universal Tendency in Nature to the Dissipation of


Mechanical Energy (Thomson 1852) In this paper, he addressed the “absolute waste of
mechanical energy available to man when heat is allowed to pass from one body to another
at a lower temperature”. Thomson went on to assert that “[a]s it is most certain that
Creative Power alone can either call into existence or annihilate mechanical energy”, this
waste “must be some transformation of energy” (ibid.). He drew a series of conclusions
which may constitute the first inkling of entropy’s blossoming into a general philosophical
principle:

1. There is at present in the material world a universal tendency to the


dissipation of mechanical energy.

8
2. Any restoration [Thomson’s italics] of mechanical energy, without more
than an equivalent of dissipation, is impossible in inanimate material
processes, and is probably never effected by means of organized matter,
either endowed with vegetable life or subject to the will of an animated
creature.
3. Within a finite period of time past, the earth must have been, and within a
finite period of time to come the earth must again be, unfit for the habitation
of man as at present constituted, unless operations have been, or are to be
performed, which are impossible under the laws to which the known
operations going on at present in the material world are subject.
(ibid.)

This is a far cry indeed from Thomson’s discussion of “the loss of power experienced by
steam in rushing through narrow steam-pipes” earlier in the paper. Jammer (op. cit. p113)
refers to it as a “sweeping generalization” of entropy, or at least the Second Law (entropy
as a concept not yet existing). Though Clausius had been responsible for originating the so-
called “entropy principle”, Thomson was arguably the one who started it on a journey that
was to lead to some unexpected places, and with which my paper is largely concerned.

As described by Smith and Wise (op. cit. pp497-502), Thomson clearly saw his findings as
constituting a universal cosmological principle, and, moreover, one which accorded with
his religious views. It was his belief that “[t]he creation of matter and energy…are
contingent upon God’s will” (Thomson 1854 cited in ibid. p634) and that mind and matter
were fundamentally separate, so that life could not arise from matter alone (ibid. p644). As
may be suggested by the reference to “Creative Power” quoted above, he believed the
dissipation of energy to represent the depletion of a stock provided in the act of creation
and not replenishable short of another similar act. In a preliminary draft for On the
Dynamical Theory of Heat, Thomson himself phrased his ideas as:

The material world could not come back to any previous state without a
violation of the laws which have been manifested to man, that is, without a
creative act or an act possessing similar power (Thomson 1851 cited in
ibid.)

It thus appears that Thomson’s religious ideas played a part in shaping his views on the
Second Law’s implications, and indirectly in rendering the concept of ever-increasing
entropy more attractive to other disciplines. However, it is important to bear in mind that,
as Smith and Wise (ibid. p501) point out, the principle of dissipation was not religious in

9
origin, but “Thomson’s commitment [Smith and Wise’s italics] to it as an inviolable axiom
received its extraordinary strength from [its] place in his theology of nature”.

It is worth reiterating that at this point there was still no concept of entropy per se. Clausius
continued his work in the area; in 1854 he “formulated … the rudiments of the theory…of
[a] measure of transformation equivalence” (O’Connor and Robertson op. cit.), developing
his earlier observations of such a quantity. However, it was not until 1865 that he named
the concept.

Clausius gave the concept the symbol S; by analogy with U, the internal energy, or heat
and work content, of a system, he initially called it “transformation content”
(Verwandlungsinhalt) (Jammer op. cit. p114). He justified his final choice of name for the
quantity in these terms:

But as I hold it to be better to borrow terms for important magnitudes from


the ancient languages, so that they may be adopted unchanged in all modern
languages, I propose to call the magnitude S the entropy of the body, from
the Greek word τροπή, transformation. I have intentionally formed the word
entropy so as to be as similar as possible to the word energy, for the two
magnitudes to be denoted by these words are so nearly allied in their
physical meanings, that a certain similarity in designation appears to be
desirable. (Clausius 1865 cited in ibid.)

Thus entropy, a decade-and-a-half in the making, finally came into existence. Having
established the concept, Clausius proceeded to state the Second Law in terms of entropy, as
previously mentioned. The inexorable increase of entropy may be the most commonly-
known statement of the Second Law to those outside the field of thermodynamics, and it is
the one that underpins the wider usage of entropy.

1.3.3 Further development of entropy: Helmholtz, Gibbs, Boltzmann et al.

Hermann von Helmholtz (1821-1894), a German physicist, took up some of Thomson’s


themes (acknowledging “the sagacity of Thomson” (Helmholtz 1854 cited in Smith and
Wise op. cit. p500)) in 1854. He proposed the concept of “heat death” as the ultimate state
of the universe, where all available heat energy has been transferred such that everything

10
has reached a uniform temperature, and no more work can be done (because heat cannot be
transferred between bodies at the same temperature). This is obviously a state of maximum
entropy, as no heat energy is available for work, and, as Helmholtz (1854 cited in Jammer
op. cit. p115) phrased it, the universe would be “condemned to a state of eternal rest”. The
idea of heat death has recurred in subsequent metaphorical applications of entropy, and
may be considered as adding another string to the idea’s bow by enhancing its apocalyptic
appeal.

The American Josiah Willard Gibbs (1839-1903) was the next major contributor to the
field; indeed Atkins (p169) considers him “responsible for most extending [the] range” of
thermodynamics. Gibbs, with a background in mechanical engineering, was one of the first
to apply thermodynamics to the analysis of chemical reactions. In the present context, his
most notable achievements were to confirm in a different form (as a tendency to ever-
lower “free energy”) the inexorable increase of entropy, and to establish the principles of
explaining molecular behaviour in statistical terms, what came to be known as “statistical
mechanics”. The latter was itself, in time, to give rise to new ways of expressing entropy.

The concept of entropy appears to have achieved widespread scientific acceptance within
Clausius’s lifetime. However, this was not universal. While there do not appear to have
been – at least by the 1860s – many serious objections to its application on a macroscopic
scale, some challenged the validity of entropy at small scales. The case of Maxwell’s
demon, already described, is the best-known instance of this. (It must, however, be noted
that Maxwell did not frame his doubts in terms of entropy per se; in fact, according to Leff
and Rex (op. cit.) he defined entropy in opposite terms to Clausius, namely as energy
available, rather than not available, for work. Furthermore, it is unclear whether he actually
intended to challenge the Second Law, though such was the practical effect of the Demon.)

Scientific discoveries in the late 19th and early 20th centuries began to broaden the scope of
entropy somewhat. As Jammer (op. cit. p114) puts it, “it was felt unsatisfactory that a
science of such astounding generality…should be based on engineering experience”. The
focus turned to finding more fundamental ways to define the idea of entropy at a molecular
level, and to extending the field of application of thermodynamics.

11
Possibly the most notable contribution in this regard, however, was made in 1872 by
Ludwig Boltzmann. Boltzmann’s research in statistical mechanics built on some of
Maxwell’s work on the distribution of gas molecules’ velocities. His most important
achievement in the present context was in demonstrating that entropy is proportional to a
quantity known as W. Hence, as entropy increases, “the system tends towards a state of
maximum probability” (ibid. p115).

In terms of entropy’s wider career, Boltzmann’s importance was (at least) threefold.
Firstly, he made the connection between entropy and disorder, one which proved important
to its subsequent spread (even if it was not always accurately used). Secondly (building on
Gibbs’s work), he connected entropy to statistics, and thus opened the door to using it to
describe quantities which were not thermodynamic in nature but were governed by the
same statistical laws as the properties of molecules. Thirdly, according to Jammer (ibid.
p118), Boltzmann’s subsequent work in the 1890s led to the first linkage of entropy with
the “arrow of time”, as discussed above. Many after him were to build on these ideas,
sometimes with surprising consequences.

12
2 Approaches to entropy

2.1 Overview

On the basis of a review of the history of the concept of entropy, with particular attention
to its migration into literary uses, I intend conducting an analysis of this process. My
primary objective in this analysis is to determine why entropy should have undergone such
a transformation from science to literature, and whether any of the characteristics of
entropy, of thermodynamics in general, or of the overall field of scientific writing
contributed to this. I will also give some attention to possible explanations for the time and
place (post-1945 United States) of entropy’s emergence in culture.

In respect of these issues, I intend to examine various theories that have been proposed in
the past, whether referring specifically to the case of entropy or having more general
applicability to science and/or culture. In this chapter, I attempt to provide a brief summary
of these ideas. I will not deal here with the detail of how they might be applied to the case
of entropy, which will be covered in the discussion chapter of this dissertation.

Here, I concentrate on three particular areas. The first is the general theory of metaphor,
the second Eric Zencey’s analysis of entropy as metaphor, and the third the theory of
knowledge production proposed by Michael Gibbons and others. I also briefly describe
some explanations proposed within the literary field for entropy’s popularity in the mid-
twentieth-century United States.

2.2 Metaphor

2.2.1 General theories

Lakoff and Johnson (1980) classify metaphors as structural (seeing one concept in terms of
another, orientational (directions or positions such as up/down) and ontological (treating a
concept as a substance, container etc., personification and metonymy, i.e. an entity
standing for a related one, such as a part representing the whole). They outline two

13
possible approaches to metaphor (ibid. pp106-107). The first, abstraction, involves “a
single, very general and abstract concept” (e.g. entropy) which is “neutral” between
different usages (in entropy’s case, the scientific, communication theory and generalised
cultural cases), which constitute “special cases of the same very abstract concept”. The
alternative view, homonymy, holds that individual usages are in fact “different and
independent concepts” whose “meanings may be similar in some respects”. Lakoff and
Johnson hold that neither can fully account for the real-life application of metaphors.
Abstraction falls down because, inter alia, no single concept can explain the variety of
ideas which some words (e.g. “up”) can stand for. Homonymy, on the other hand, may
postulate similarities where none exist (again, not all the concepts “up” stands for can be
shown to be meaningfully similar). The authors are of the opinion that metaphor itself
creates new similarities between the concepts involved (ibid. pp 147-155) by highlighting
selected aspects of both. They ultimately reject the concepts of objectivity and subjectivity
as applied to metaphors and hold that experientialism, which argues that objectivity is
possible but only within a “conceptual system and a set of cultural values” (ibid. p227), is
the solution.

2.2.2 Zencey’s analysis of entropy as metaphor

Zencey (op. cit) provides a review of the general prevalence of entropy as a metaphor
(including, but not specifically concentrating on, literary applications). He points out that,
although – as we have seen – entropy is considered a difficult idea to grasp, it “has enjoyed
widespread popularity as metaphor almost from the moment of its inception” and
“exercised a compelling attraction to thinkers and theorists beyond the confines of physics”
(ibid. p48). Zencey suggests that this may be attributable to the “esemplastic” (unifying)
ability of entropy, which relates to its nature as a basic or “root” metaphor. He
characterises entropy as similar to the “root metaphor of generating substance”, rejected by
earlier theorists of root metaphors as “inadequate”. In “entropism”, whatever is opposed to
entropy (e.g. energy or information) is the generating substance and the Second Law is
“the primary principle of change” (ibid. p49).

14
Zencey (ibid. pp51-55) follows his analysis with a lengthy list of arguments attempting to
explain entropy’s popularity as a metaphor. As I intend addressing these, where relevant to
the present case, in the discussion chapter, I will not enumerate them here.

2.3 Knowledge production

Gibbons et al. (1994) have proposed a theory of “knowledge production”, originally


concerned with science but expanding to take in humanities research and even the arts.
They distinguish two “modes” of production, Mode 1 and Mode 2. In essence, Mode 1 is
based on specific disciplines (the traditional mode of knowledge production), while Mode
2, which is characteristic of the late 20th century, is interdisciplinary in nature. Mode 1 is
“generated within a disciplinary, primarily cognitive context” (ibid. p1), “governed by the
largely academic interests of a specific community” (ibid. p3). Mode 2 is “created in
broader, transdisciplinary…contexts” (ibid. p1) and “carried out in a context of
application” (ibid. p3).

The specific applicability of the theory of knowledge production to entropy lies in the fact
that the migration of the concept from thermodynamics into literature, via intermediate
fields, may be interpreted as an instance of interdisciplinarity. Furthermore, the
transformation of entropy took place in an era when knowledge production in general was
beginning to make the transition between Modes 1 and 2. It is possible that this change
may have facilitated its spread, and this is one of the issues which I will examine.

2.4 Miscellaneous analyses of entropy

Various writers on the subject of literary entropy have also tried to explain the factors
involved in the concept’s emergence, often with particular reference to the place and time
in which it attained prominence. Tanner (op. cit. pp150-151) addresses this issue briefly
with regard to modern American literature. Lewicki (1984 pp 75-77) deals with it at
greater length in a broadly similar context. The explanations offered are numerous and
varied. However, the principal factors proposed by Tanner can be summarised as:

15
 Modern American literature often concerns “people who [are] turning themselves
into ‘isolated systems’ [taking] in a decreasing amount of information, sensory
data, even food, [so] that the sense of their own personal entropy is heightened”.
However, an individual who is not a closed system is capable of anti-entropic
activity (as Norbert Wiener’s theories, invoked by Tanner and discussed in the
following chapter of this dissertation, propose).
 The “unprecedented amount of force” and “unique diffusion of energies into…an
unparalleled amount of available space” in US history “produce writers who are
obsessed with the basic mystery of force and energy”.
 The fluidity of US society creates a “sense of mysterious powers at work behind the
visible concretions of the everyday world”, and entropy is one way of
characterising such powers.

Lewicki takes a somewhat broader historical sweep, coming up with a generally different
set of possible explanations, though there is some overlap. Lewicki’s main points can be
summarised as:

 American literature’s previous dominant philosophy of naturalism created “an


interest in the laws and processes of nature, [making] American writers more
receptive to notions related to the natural sciences”.
 The “general knowledge of and fascination with science” in the US were “markedly
higher than in Europe”.
 Specific aspects of US conditions, such as large-scale agriculture and rapid
urbanisation, were “particularly conducive to entropic visions”.
 The “continuing interest in apocalypse in American culture” encouraged interest in
entropy.

Clearly, some of these explanations may be more applicable than others to a consideration
of entropy’s transition between disciplines rooted in the scientific, rather than literary,
sociological or historical, background to the process. I will discuss them further at a later

16
stage in this dissertation, in conjunction with the other ideas I have set out in the present
chapter.

17
3 The transformation of entropy

3.1 Introduction

The objective of this chapter is to outline the process of entropy’s migration from scientific
into cultural (particularly literary) uses in the 20th century. As such, it primarily covers the
period from the end of World War II, when interdisciplinary interest in entropy began to
grow, until the mid-1970s, by which time entropy had become firmly established in literary
usages. However, I have also addressed some notable (and possibly influential) instances
of interdisciplinary applications from earlier in the century; these are dealt with in section
3.2. Section 3.3 considers the emergence of entropy in communication and cybernetic
theory during the immediate post-war years, and section 3.4 the development of
interdisciplinary solutions to the Maxwell’s demon problem. Section 3.5 then addresses the
subsequent emergence (or re-emergence) of entropy in literature.

3.2 Early interdisciplinary usages

3.2.1 The science of human degradation: Henry Adams and history as


entropy

While scientists continued to broaden entropy’s scope of application and deepen its
theoretical underpinnings, there were stirrings of interest in the concept from outside
science by the early 20th century. One of the earliest prominent individuals to express such
an interest was Henry Adams (1838-1918).

Adams was to make, as Tanner (op. cit. p148) puts it, “[t]he most famous application of the
law of entropy to society and human history”. His background was highly distinguished –
as the great-grandson of the US’s second President and grandson of its sixth – but not in
any way scientific. Adams’s career included periods as a diplomat, journalist and professor
of history; his interest in entropy stemmed from the third of these callings.

18
Adams had long taken an interest in scientific matters, and was a close friend of the
geologist Clarence King (Nadel 1999 pp.xi-xii). Much of his science-related writing
constitutes an effort to develop a scientific basis for history. In this connection, entropy
was not the first scientific concept Adams had turned to. He appears to have taken an
interest in Darwinism (Adams 1999 [1918] pp190-200) gravitation (ibid. pp314-315) and
electrical energy (he compared “the Dynamo and the Virgin” (ibid. pp317-326), equating
energy’s and electricity’s influence over turn-of-the-century life to religion’s in the Middle
Ages). As described by Conder (1970 p55), Adams’s subsequent book Mont-Saint-Michel
and Chartres develops the idea of religion’s waning influence in an entropic manner, with
the movement from literal to symbolic interpretations reflecting “a movement away from
emotional fervor in favour of rational analysis” presumably analogous to the degradation
of energy.

Adams’s use of entropy is probably his best-known, and almost certainly his most
influential, deployment of science. Though his autobiography, The Education of Henry
Adams (Adams op. cit.) is by far his best-known work, his most important thoughts on
entropy are contained in A Letter to American Teachers of History (Adams 1919). (In fact,
as Conder (op. cit. p22) points out, Adams “was ignorant of thermodynamic entropy until
after he completed the Education”, notwithstanding that various commentators have
associated his thoughts on entropy with that work.)

In this essay, Adams enthusiastically endorses Thomson’s On a Universal Tendency in


Nature to the Dissipation of Mechanical Energy (“this young man of twenty-eight thus
tossed the universe into the ash-heap…” (ibid. p142)). He believes that life, or “vital
energy”, and by extension the “social energy” of entire nations, can be treated analogously
to heat energy. In this way, history can be analysed thermodynamically:

[T]he analogy between heat and Vital Energy…was insisted upon by the
physicists in accents that became sharper with each generation…(ibid.
p149)

If life was to disappear, the form of Vital Energy known as Social Energy,
must also, presumably, go to increase the Entropy of the Universe, thus
proving – at least to the degree necessary and sufficient to produce
conviction in historians – that History was a Science. (ibid. p150)

19
[S]ixty years of progress in science have only intensified the assertion that
Vital Energy obeys the law of thermal energy.(ibid. p150)

Adams applies this philosophy to his own era, seeing decline and dissipation in many
things (notably the work of sociologists such as Emile Durkheim; “philosophical schools
founded on the supposed failure of society” (ibid. p188). He also returns to the theme of
energy’s role in technological progress, which he had dealt with in The Education:

[E]very gain of power – from gunpowder to steam – from the dynamo to the
Daimler motor – has been made at the cost of man’s – and of woman’s –
vitality. (ibid. p236)

Adams concludes that “the historian will have to define his profession as the science of
human degradation” (ibid. p195) and that reason and will “must submit to the final and
fundamental necessity of Degradation” (ibid. p208). He also appears to see entropy as
contradictory to evolution, on the grounds that the latter implies that “vital energy” must
increase over time (ibid. pp152-154); this may relate to the disillusionment with
Darwinism which he expressed in the Education.

Clearly, entropy was of great importance to Adams; Zencey (op. cit. p49) describes it as
“an organizing metaphor” in his philosophy of history. It is also readily apparent that
Adams’s use of entropy drew heavily, though not exclusively, on Thomson. (Brooks
Adams, brother of Henry, also attempted to apply entropy to history, according to Zencey
(ibid. p49); however, his approach was apparently more literal than Henry’s.)

There are obvious limits to how far a scientific principle can be stretched in usages such as
this one. Daub (1990 [1970]) points out that Adams’s understanding of thermodynamics
was severely limited in some respects; for instance, he misunderstood Maxwell’s Demon
as aggressively agitating molecules rather than passively sorting them. However, it seems
that his metaphorical usage of entropy was extremely influential on subsequent writers,
and – as will be discussed later – there are many indications that Adams’s writings played
an important role in the emergence of entropy in American fiction over the decades
following his death.

20
3.2.2 The beginnings of literary entropy

It is unclear exactly when the concept of entropy was taken up in the literary world, and –
given that it is often used in a relatively ill-defined way – probably impossible to be
entirely certain. Tanner (op. cit. p142) believes that “[t]aken in its broadest sense as
meaning …increasing disorder, finally arriving at total inertia” entropy avant la lettre can
be found in past works, back at least as far as Alexander Pope’s Dunciad, which predates
Clausius by over a century.

Nevertheless, some writers were beginning to deploy entropy more explicitly even before
its generally-accepted emergence as a literary idea in the 1950s US. Entropic themes
appear to go back a considerable way in American literature. Lewicki (op. cit. p78) detects
the idea of entropy, though not the word itself, in Herman Melville’s 1853 short story
Bartleby the Scrivener. Tanner (op. cit. p143) sees the “valley of ashes” which is a key
setting of Scott Fitzgerald’s 1925 novel The Great Gatsby as a manifestation of entropy
(perhaps the “ash-heap” in Adams’s description of Thomson’s predicted fate for the
universe?).

3.2.3 All order is doomed: Nathanael West and entropy

Nathanael West (1903-1940) was an influential (if somewhat obscure in his lifetime) US
novelist who is of interest in this context as an early user of the concept of entropy in
fiction. West made use of entropy throughout his relatively small body of work, referring
explicitly to it in his 1933 novel Miss Lonelyhearts:

Man has a tropism for order. Keys in one pocket, change in another.
Mandolins are tuned G D A E. The physical world has a tropism for
disorder, entropy. Man against Nature…the battle of the centuries. Keys
yearn to mix with change. Mandolins strive to get out of tune. Every order
has within it the germ of destruction. All order is doomed, yet the battle is
worth while. (West 1983 [1933] p104)

Lewicki relates West’s application of entropy in Miss Lonelyhearts to the protagonist’s


“desire to improve the world by imposing a better structure on it”, ending in his defeat by
“the entropic chaos inherent in the world” (Lewicki op. cit. p81). He also sees West’s

21
subsequent, and better-known, novel The Day of the Locust as entropic in that it deals with
“gradual but inescapable death” (e.g. in references to a continually growing dump, or the
decline of the ageing human body).

It might legitimately be asked why West’s usage of entropy did not draw the attention that
other writers’ would three or four decades later, or lead to its becoming widespread.
Lewicki’s suggested explanation is that West’s ideas may have failed to appeal to the
mood of the US at the time (in the immediate aftermath of the Depression); his works were
rediscovered, along with entropy, in “an era that seems more resigned to the inevitability of
[entropy’s] implications” (ibid. p82).

3.3 Post-war America and the spread of entropy

3.3.1 Information, choice and uncertainty: Claude Shannon and


communication entropy

Notwithstanding various attempts at co-opting it into other fields such as those we have
just seen, entropy remained the almost exclusive preserve of thermodynamics, statistical
mechanics and associated disciplines for the bulk of its first century. The critical point in
its adoption on a wider basis arguably came when Claude Shannon (1916-2001), then of
Bell Laboratories, introduced it to communication theory in his 1948 paper A
Mathematical Theory of Communication (Shannon 1948).

Shannon modelled communication (Figure 2) as a linear progression from information


source to destination via a transmitter (which “encodes” the message), a “channel”, and a
receiver (which “decodes”) it. He defined a quantity to which he gave the symbol H as a
measure of “information, choice and uncertainty”; observing that “[t]he form of H will be
recognized as that of entropy as defined in certain formulations of statistical mechanics”,
he proceeded to name it “entropy” (ibid. p11). As Fiske (1982 pp10-11) describes it,
communication entropy is “the opposite of redundancy” in a message, and the result of
“low predictability”. A highly entropic message is more unpredictable and randomised
(just as a highly entropic physical state is less ordered). (However, Morowitz (1991 cited in

22
Newhagen and Fox 2003) suggests that information and thermodynamic entropy are in fact
opposed because information entropy reflects uncertainty, whereas Boltzmann’s entropy
has an opposite significance – presumably because highly entropic states are more
probable.)

Figure 2 Shannon’s model of a communication system (Shannon 1948)

This appears to represent entropy’s first decisive venture outside thermodynamics, though
of course Henry Adams (and others) had tentatively explored its possible metaphorical
applications in the past, without the idea becoming truly widespread. Perhaps inevitably,
the first usage of entropy in an information-related sense is contested; Watanabe (1985
p216) claims to have used the idea of entropy “as a measure of indeterminacy and
correlation independent of the thermodynamic context” in 1939. Nevertheless, A
Mathematical Theory of Communication was influential in a way which previous work was
not.

In relation to the factors which may have played a part in Shannon’s decision to give the
name “entropy” to his concept, it may be noted that he, by contrast with (say) Adams, had
a background in the sciences, with an undergraduate degree in mathematics and electrical
engineering and postgraduate degrees in both fields (O’Connor and Robertson 2003a).
Thus, it might be expected that he would not use the name lightly.

23
Apropos of this, it is occasionally claimed that the mathematician John von Neumann
encouraged Shannon to use the term “entropy”, possibly over Shannon’s objections,
because, in von Neumann’s (alleged) words, “no one knows what entropy is, so in a debate
you will always have the advantage” (American Heritage Book of English Usage 1996). In
fairness to von Neumann, his own first degree was in chemical engineering (O’Connor and
Robertson 2003b), which would appear to imply an even more intimate familiarity than
Shannon’s with the concept of entropy in its original thermodynamic usage.

The exact nature of the relationship between the communication and thermodynamic
manifestations of entropy remains in dispute. According to Newhagen and Fox (op. cit.
p2), “[w]hether the use of entropy by Shannon was coincidental or intentional is in itself a
topic of both technical and even philosophical interest”. Newhagen and Fox point out that
thermodynamic entropy was not directly alluded to in Shannon’s original paper, but that a
reference was inserted when it was republished as a monograph (ibid. p4). In their opinion,
both Boltzmann’s and Shannon’s uses of entropy involve “slippage when the leap is made
from a description of the physical system to that of the behavior of its constituent parts”
(ibid. p12). In the case of communication theory, this slippage limits the idea’s
applicability in a human context where psychological factors must be taken into account.

Ritchie (2004) believes the relationship to be essentially metaphorical. While he


acknowledges that Shannon’s adoption of entropy is appropriate to his particular purposes
– “it is difficult to imagine a better measure” (ibid. p44) – he believes that Shannon did not
prove it to be the only possible way of measuring the variance in a message. To Ritchie,
the relationship between the two entropies is “primarily metaphorical even as it is based on
a literal linkage” (ibid. pp45-46), the linkage in question being the transmission of signals
as “organized energy states” subject to degradation just as other forms of energy are.
Ritchie further observes that information and entropy have both “mutated
into…quasimystical concept[s] empowered by the aura of apparently impenetrable
mathematics” (ibid. p47).

24
3.3.2 Progress in a universe running downhill: Norbert Wiener, cybernetics
and entropy

In 1948, the same year as Shannon’s paper, the US mathematician and philosopher Norbert
Wiener (1894-1964) published Cybernetics (Wiener 1961 [1948]). Wiener himself coined
(or re-coined) the term “cybernetics” (from the Greek χυβερνήτης, “steersman” or
“rudder”) to describe the study of systems, communication and control, with particular
reference to feedback processes. He appears to have been heavily inspired in his cybernetic
work by the research he conducted during World War II on control systems for anti-aircraft
guns (O’Connor and Robertson 2003c).

Wiener followed Cybernetics with The Human Use of Human Beings, published in 1950
with a revised edition in 1954 (Wiener 1968 [1954]). This book was essentially a
popularised treatment from a cybernetic viewpoint – Wiener (1961 [1948] p.vii) called it
the “small popular companion” to Cybernetics – of issues relating to human consciousness,
language and automation. As will be seen, it has been widely credited with a decisive role
in the popularisation of entropy, in particular as it related to the writers who were to
introduce it to literary use in the United States of the 1950s and 1960s.

The Human Use’s treatment of entropy initially emphasises the concept’s relationship to
probability, which Wiener sets against the certainties of Newtonian mechanics:

[I]t is, I am convinced, Gibbs rather than Einstein or Heisenberg or Planck


to whom we must attribute the first great revolution of twentieth century
physics. (ibid. pp13-14).

[P]hysics no longer claims to deal with what will always happen, but rather
with what will happen with an overwhelming probability. (ibid. p14)

He attempts to generalise this beyond physics by relating the “irrationality” of “incomplete


determinism” in post-Gibbs/Boltzmann physics to “Freud’s admission of a deep irrational
component in human conduct and thought” (ibid. p14) and the “fundamental element of
chance in the texture of the universe itself” to St. Augustine’s idea of evil as
incompleteness (ibid. pp14-15).

25
Wiener proceeds to address communication entropy, acknowledging Shannon, before
developing the concept of both organisms and machines (at least those controlled by
feedback mechanisms) as anti-entropic:

It is the function of [a machine’s control and feedback mechanisms] to


control the mechanical tendency toward disorganization; in other words to
produce a temporary and local reversal of the normal direction of entropy.
(ibid. p25)

[T]he physical functioning of the living individual and the operation


of…communication machines are precisely parallel in their analogous
attempts to control entropy through feedback. (ibid. p26)

The machine, like the living organism, is…a device which locally and
temporarily seems to resist the general tendency for the increase of entropy.
By its ability to make decisions it can produce around it a local zone of
organization in a world whose general tendency is to run down. (ibid. p33)

To Wiener, a system not in equilibrium permits local reductions in entropy. He justifies


Maxwell’s demon in these terms, claiming that the demon requires light to detect
molecules, so the box contains a mixture of gas and light, which must not be in equilibrium
because an equilibrium state would imply equal light from all directions and no possibility
of distinguishing objects. However, over the long term, the collisions between photons and
gas molecules tend to bring the system to equilibrium, so the demon cannot operate
indefinitely (ibid. p29). The origins of this approach to the problem are dealt with
separately below. Life itself is anti-entropic, and human progress, according to Wiener, is
the result of this anti-entropic capability (or “homeostasis”):

There are local and temporary islands of decreasing entropy in a world in


which the entropy as a whole tends to increase, and the existence of these
islands enables some of us to assert the existence of progress. (ibid.p35)

Wiener, however, does not appear to count himself among the “some of us”. His personal
attitude to progress is that:

The best we can hope for the role of progress in a universe running downhill
as a whole is that the vision of our attempts to progress in the face of
overwhelming necessity may have the purging terror of Greek tragedy. Yet
we live in an age not over-receptive to tragedy. (ibid. p39)

26
In this vein, he proceeds to challenge the universality of a belief in progress, emphasising
how much of an exception the latter-day US “worship of progress” is both historically and
among present-day belief systems (ibid. pp39-44). The idea of entropy as setting a limit to
human progress appears to echo some of Henry Adams’s opinions.

Wiener also sees entropy, and specifically its application to information, as an argument
against the commodification of information (for him, another defining characteristic of the
mid-20th-century US):

Just as entropy tends to increase spontaneously in a closed system, so


information tends to decrease; just as entropy is a measure of disorder, so
information is a measure of order. Information and entropy are not
conserved, and are equally unsuited to being commodities. (ibid. p102)

In addressing the effects of automation, he calls on the concept of entropy yet again in
order to make the point that machines – which, at least from Wiener’s point of view, can
only handle certainties rather than probabilities – are as yet far from being capable of
superseding humanity:

The dominance of the machine presupposes a society in the last stages of


increasing entropy, where probability is negligible and where the statistical
differences among individuals are nil. Fortunately we have not yet reached
such a state. (ibid. p157)

Overall, Wiener’s use of entropy is philosophical and moralistic; in this sense, it may invite
comparison with Thomson’s. There are, of course, differences in that Wiener reaches out
further in a metaphorical direction, where Thomson was merely pushing physics to (and
perhaps a little beyond) the boundaries of philosophy, and in the fact that Wiener was
actually philosophically trained.

Wiener referred to entropy in many of his works, and, according to Watanabe (op. cit.),
“often brought up the problem of Maxwell’s demon”. Wiener’s other works refer to some
of the issues raised in The Human Use, including the demon and “the entropic properties of
light” (Wiener 1985a (1952)), and life as an anti-entropic process, with particular reference
to photosynthesis as being, like the demon’s activity, facilitated by light (Wiener 1985b
(1950)).

27
(Curiously enough, Cybernetics itself gives rather less prominence to entropy. This may be
explained by the assumption that the intended audience would be more familiar with the
concept and so it would need less explanation, or alternatively by The Human Use’s more
sweeping application of cybernetic ideas, which may have required a more effective
metaphor to explain effectively.)

3.4 Maxwell’s demon and interdisciplinary entropy

Norbert Wiener’s usage of entropy, and the role in it of Maxwell’s demon, bring us neatly
back to the progress of scientific attempts to square the demon with the Second Law.
According to Leff and Rex (op. cit.) the “life” of the demon entered its second phase in
1929, with Leó Szilárd’s pioneering work on entropy and information. Szilárd, a
Hungarian physicist, examined a modified version of Maxwell’s thought experiment and
concluded that “[o]ne may reasonably assume that a measurement procedure is
fundamentally associated with a certain definite average energy production, and that this
records concordance with the Second Law” (Szilárd 1990 [1929], p124). His ideas “met
with mixed response” (Leff and Rex op. cit. p16), partly because of their linking of entropy
and information, which (for instance) Karl Popper criticised as “spurious” (ibid.).

The next breakthrough was a theory developed by Léon Brillouin (1889-1969) and others
in the late 1940s and early 1950s (more-or-less contemporaneously with Shannon’s and
Wiener’s work). Brillouin, a French-born physicist working in the US, was seeking “ a
scientific framework in which to explain intelligent life” (Leff and Rex op. cit. p2). His
paper Life, thermodynamics and cybernetics (Brillouin 1990a [1949]), was published the
year after Wiener’s Cybernetics; this paper endorsed Wiener’s thoughts on such matters as
information entropy and the anti-entropic nature of biological processes, and briefly (in a
footnote) discussed the demon, pointing to its need of a light source.

Brillouin followed this two years later with a second paper, the uncompromisingly-titled
Maxwell’s demon cannot operate (Brillouin 1990b [1951]), developing his ideas on
Maxwell’s demon and observing that, if the system is indeed in equilibrium, there cannot

28
be a light source present and so the demon cannot observe the molecules (this is basically
the approach adopted by Wiener in The Human Use, as discussed above). He also argued
for the equivalence of thermodynamic and information entropy, proposing that the
demon’s activities convert negative entropy or “negentropy” (i.e. the highly-ordered
energy of the light source) into information, which is reconverted into negentropy by using
it to sort the molecules and thus reduce their entropy.

The information aspects of the demon problem were later developed further with the
emergence of the discipline of computer science. Developing Brillouin’s work, Rolf
Landauer argued in 1961 that erasing a computer’s memory is an entropic process; in turn,
Charles Bennett built on this to argue that, as the demon must repeatedly erase its
observations from its memory, the information entropy gained compensates for the
thermodynamic entropy lost in the process of observing and sorting (Leff and Rex op. cit.
pp3, 21). For Leff and Rex, this represents the “third phase” of the demon’s existence.

The overall story of Maxwell’s demon is clearly a fascinating one in itself, and I regret
being unable to cover it in greater detail. Its primary significance to the subject matter of
this paper lies in the fact that it has been characterised by a progressive convergence of
thermodynamics and information theory. This is of interest per se in examining the
dissemination of the idea of entropy – it obviously suggests that thermodynamics has
gained something tangible from the usage of its ideas in non-scientific fields – but also
because this convergence may have assisted the transference of the idea into the cultural
sphere which was taking place over a similar timescale. It is to this process that I now
return.

3.5 Emergence of literary entropy

3.5.1 Background

The United States of America, in the decades immediately following World War II, were
to see Clausius’s “transformation content” undergo its own fundamental transformation. It
was, of course, the time and place of the publication of Shannon’s, Wiener’s and other

29
works extending entropy’s scope into communication and beyond. Some of this material
found its way into general discourse, for instance through Wiener’s The Human Use of
Human Beings.

Simultaneously, scientific education was making the general public aware of entropy and
the Second Law in their original contexts (Lewicki op. cit. p76). Indeed, Tanner (op. cit.
p141) quotes a critic of C.P. Snow’s well-known “two cultures” comments (which, inter
alia, identified the Second Law as something non-scientists were insufficiently familiar
with) to the effect that the law is something “which every American schoolboy knows”.
Lewicki’s description of the greater public interest in and knowledge of science in the US
(see 2.4 above) is also of relevance.

Whatever the causes (which will be discussed in more depth later in this dissertation),
entropy was about to make the transition from science to general cultural use. Perhaps
unsurprisingly, science fiction was one route by which it achieved this.

3.5.2 A final state of total, absolute kippleization: Entropy in science fiction

The first reference to entropy in science fiction cited by Nicholls (op. cit.) is in the title of a
1952 novel, House of Entropy, by Roy Sheldon. However, the concept itself does not seem
to have been widely deployed until the 1960s. Its first widely-known usage was in Isaac
Asimov’s short story The Last Question (Asimov 1987 [1956]), published in Science
Fiction Quarterly in 1956. According to Seiler and Jenkins (2004), this became Asimov’s
personal favourite of all his stories. The Last Question is of interest in its fairly extensive
application of the concept of entropy. Furthermore, as will be seen, Asimov’s usage of
entropy has influenced that of at least some subsequent writers.

In the story, a succession of human characters in settings ranging from the year 2061 to ten
trillion years into the future discuss the idea of entropy and the irreversibility of the Second
Law – “[y]ou can’t turn smoke and ash back into a tree” (ibid. p226). Each time, they ask a
computer (on each occasion a more powerful descendant of the original machine) whether
it is possible to reverse “entropy” (i.e. the Second Law). On each occasion, the response is

30
that there is “insufficient data for a meaningful answer” – until the last time. By now, the
final computer has outlasted space and time themselves, existing in hyperspace “only for
the sake of the one last question that it had never answered” (ibid. p230). It finally works
out the solution and implements it in a rather dramatic fashion:

And AC [the computer] said “LET THERE BE LIGHT!”


And there was light –
(ibid.)

at which point (perhaps faintly echoing William Thomson’s belief in “Creative Power” as
the sole antidote to entropy) the story ends.

On one level, this story represents a competent usage of entropy as a plot device which
incidentally helps to explain the idea to the general reader; given Asimov’s background in
chemistry and biochemistry, this is perhaps not surprising. However, it also hints at deeper
connections with the mid-20th-century development of entropy.

The implication that sufficient data would permit the reversal of the Second Law may
suggest that Asimov is alluding to the concept of information entropy; consider Ritchie’s
(op. cit. p47) critical summary of the linkage between the two entropies:

[T]he two concepts are frequently linked…information is equated with


negative entropy and credited with the power to neutralise the unpleasant
consequences of the second law…[including] even the eventual descent of
the universe into total disorder.

It is unclear how aware Asimov may have been of theories of information entropy.
However, the likelihood that he was reasonably familiar with Norbert Wiener’s work
seems high. Asimov had used the term “cybernetics” in a 1948 short story, shortly after the
appearance of Wiener’s book of that title (Scrivener 2003). Furthermore, Asimov and
Wiener gave a joint lecture in 1956, and even collaborated on a science-fiction story which
“never came to fruition” (American Institute of Physics 2000). This suggests that the story
may be at least hinting at an equivalence between thermodynamic and information entropy,
and Asimov, though he may have been the first writer of fiction to take up this point,
would be far from the last.

31
In relation to the further spread of entropy within science fiction, Nicholls (op. cit). refers
to Philip K. Dick as “the first to popularize it” in the 1960s. Dick’s most notable usage of
the concept was in his 1968 novel Do Androids Dream of Electric Sheep?, where he coined
“kipple” as a synonym for entropy. His characters’ description of kipple touches on several
entropic ideas, including the Second Law, heat death and Wiener’s concept of local
decreases in entropy due to human activity:

Kipple is useless objects…When nobody’s around, kipple reproduces


itself…It always gets more and more.

No one can win against kipple…except temporarily and maybe in one


spot…It’s a universal principle operating throughout the universe; the entire
universe is moving towards a final state of total, absolute kippleization.
(Dick 2005 [1968] pp56-57)

Also beginning in the 1960s, the so-called “New Wave” of science fiction writers, notably
J.G. Ballard and Michael Moorcock, began to adopt entropy. The “New Wave” was
characterised by:

a pessimism about the future that ran strongly counter to genre [science
fiction]’s traditional optimism, often focused on the likelihood of disaster
caused by overpopulation and interference with the ecology, as well as by
war (Nicholls 1993b p866).

Entropy and the Second Law obviously reflected the preoccupations of this movement, and
entropy was to reach its peak of popularity as a science-fiction concept in the 1970s (ibid.).
It may also be noted that the “New Wave” writers, the most prominent of whom (such as
Ballard and Moorcock) were British, appear to represent the most notable usage of the
concept of entropy in fiction outside the US.

3.5.3 Entropy’s wider use in literature

However, by the time this development was underway, entropy was already beginning to
become part of the vocabulary of mainstream literary writing, particularly in the US. By
1971, Tanner (op. cit. p141) was able to list Norman Mailer, Saul Bellow, John Updike,
John Barth, Walker Percy, Stanley Elkin and Donald Barthelme as amongst modern

32
American writers making explicit use of entropy; Tanner also mentions William Burroughs
and Susan Sontag as users of the concept though not the word itself. Lewicki (op. cit.
pp103-108) adds William Gaddis to this list. Above all, however, Tanner gives special
attention to Thomas Pynchon’s use of entropy throughout his works, and I will examine
Pynchon’s relationship to the concept in greater detail shortly.

The factors leading to the emergence of entropy as a significant literary concept in the US
around the beginning of the 1960s seem unclear at this point, and I will return to the matter
in my analysis. However, Tanner quotes both Henry Adams and Norbert Wiener
extensively (Tanner op. cit. pp144-150). He believes that the relationship of entropy to
information makes it inherently relevant to the writer, and contends that (since entropic
states are more probable) the “increasing probability of mass-media messages” has
compromised “the ability of any language to transmit significant information” (ibid.
pp145-146), and connects this to latter-day American writers’ struggle to find a style of
their own. Tanner also relates Wiener’s vision of entropy and disorganisation as evil to
such writers’ “obsession with plots, agents, codes…a measure of the paranoia induced by
American life” (ibid. p148).

These comments are particularly true of Thomas Pynchon, and his name is possibly that
most readily associated with entropy within literary fiction. Given this, it is worth
discussing the origins and development of Pynchon’s interest in entropy at greater length.

3.5.4 A metaphor not only verbally graceful, but also objectively true:
Thomas Pynchon and entropy

Just as entropy was beginning to emerge as a literary concept, a short story entitled simply
Entropy (Pynchon 1984a [1960]) appeared in the spring 1960 issue of the Kenyon Review.
It was the fourth published story by the then relatively unknown Thomas Pynchon, a
twenty-two-year-old Cornell University English graduate and technical writer for Boeing.
Entropy deals with the inhabitants of two neighbouring apartments in Washington DC.
Downstairs, Meatball Mulligan hosts a chaotic party “moving into its 40th hour” (ibid.
p81), attended by a mixture of sailors, students and beatniks; some of the guests discuss

33
information theory – “[n]oise screws up your signal, makes for disorganization in the
circuit” (ibid. p91). Upstairs, Callisto (an enthusiast for entropy as metaphor and admirer
of Henry Adams) and his girlfriend Aubade occupy a sealed room containing a “hothouse
jungle” (ibid. p83). Outside, “for three days now…the mercury had stayed at 37 degrees
Fahrenheit” (ibid. p85); for Callisto, this presages heat death, as does his inability to
transfer his body heat to a sickly bird. Finally, as Meatball struggles to avoid chaos
downstairs (pairing and separating guests, perhaps in the manner of a Maxwell’s demon?),
upstairs the bird dies. Aubade, seemingly interpreting this as a sign of heat death, “the
single and unavoidable conclusion to all this”, smashes the window to bring the room into
equilibrium with the outside world, then awaits with Callisto the moment when “the
hovering, curious dominant of their separate lives should resolve into a tonic of darkness
and the final absence of all motion” (ibid. p98).

Like most of Pynchon’s work, Entropy generates widely varying interpretations (including
matters as fundamental as whether the heat death is real or imagined), and I will not deal in
detail with these here. Clearly, there are entropic suggestions in many of the images
(Callisto’s closed system versus the open system of Meatball’s party, Meatball as demon,
the “countless false signals and instances of communication breakdown” (Tabbi op. cit.
p64) at the highly-entropic party), alongside the numerous explicit references to the idea.

For Tanner (op. cit. p142), Entropy “provides the preface to [Pynchon’s] subsequent
novels”. While I cannot deal in any detail with the latter here, it is important to address
some aspects of the way in which they approach entropy. V, Pynchon’s first novel, was
published in 1963, and makes no explicit mention of entropy, though critics see the idea as
ever-present in its treatment of “decay and decline”, “dead landscapes of every kind” and
life’s struggle with the inanimate (Tanner op. cit. p157). His subsequent work (and
probably his most widely read), The Crying of Lot 49, published in 1965, uses the concept
far more openly.

The Crying of Lot 49 nominally deals with its protagonist Oedipa Maas’s attempts to
discover the truth about an unofficial postal system, the Tristero, which she becomes aware
of in the course of acting as executor to the estate of her wealthy ex-lover Pierce Inverarity.

34
Maxwell’s demon is referred to explicitly at length; an engineer explains the concept to
Oedipa (Pynchon 1996 [1965] pp59-60) and informs her of an inventor (Nefastis) who has
developed a working version operable by certain “sensitive” people who can transfer
information telepathically to the demon. Nefastis, in turn, explains the relationship of
entropy and information:

She did gather that there were two distinct kinds of this entropy. One having
to do with heat-engines, the other to do with communication. The equation
for one, back in the ’30s, had looked very like the equation for the other. It
was a coincidence. The two fields were entirely unconnected, except at one
point: Maxwell’s Demon. As the Demon sat and sorted his molecules into
hot and cold, the system was said to lose entropy. But somehow the loss
was offset by the information the Demon gained…

‘Entropy is a figure of speech, then,’ sighed Nefastis, ‘a metaphor. It


connects the world of thermodynamics to the world of information flow.
The Machine uses both. The Demon makes the metaphor not only verbally
graceful, but also objectively true.’
(ibid. pp72-73)

Oedipa is unable to work the machine, and doubts whether the demon is any more than
metaphor.

Aside from these direct references, much of the remainder of the novel is commonly seen
as relating to entropy. Oedipa’s ultimate inability to gain useful information about the
Tristero – even whether it is reality or an elaborate hoax – has been interpreted as a
manifestation of information entropy (Dutta 1995), and the prevalence of binary “either/or”
imagery in the novel (such as the aforementioned) as related to the binary sorting role of
Maxwell’s demon (Stuart 1997).

Pynchon’s best-known novel, Gravity’s Rainbow, was published in 1973; as has been seen,
Pynchon’s interest in entropy had already drawn significant attention from critics by this
time. Gravity’s Rainbow itself was to attract even more attention, frequently being cited as
one of the most important post-World War II novels in English. The plot (insofar as it
exists) deals with the German rocket programme towards the end of World War II and the
subsequent struggle amongst the Allies over the technology, though it reaches back into the
past and forward to a variety of imagined futures, while ranging over numerous themes.

35
A full treatment of scientific themes in Gravity’s Rainbow would be impossible in the
space available, and in any case has been attempted by others on several past occasions.
Friedman (1983 pp70-71) characterises the novel’s use of science in terms of three
“visions of the universe”; Newtonian mechanics, statistical physics and quantum physics.
It is the second of these which relates to entropy, and, according to Friedman, “provide[s]
some of the most effective imagery in Gravity’s Rainbow – time-reversal and Maxwell’s
Demon”.

In Friedman’s view, the random rocket impacts portrayed in the novel (and which several
characters devote significant efforts to predicting, sometimes by very unorthodox methods)
parallel the collisions of gas molecules in a sealed container, as studied by statistical
mechanics. Furthermore, images of time-reversal (such as the fact that a supersonic rocket
can only be heard approaching after it has hit the target), representing apparent challenges
to the Second Law, are omnipresent (ibid. pp84-85), as is the idea (reminiscent of Wiener)
of life as challenger to entropy, “biological order springing up from disorder” (ibid. p85).
Maxwell’s demon is again invoked on more than one occasion (“Destruction, oh, and
demons – yes, including Maxwell’s – were there deep in the woods…” (Pynchon 2000
[1973] p239).

With the publication of Gravity’s Rainbow, entropy could be said to be firmly established
in literature. At this point, before leaving the historical treatment of the subject altogether,
it is time to turn to the possible origins of Pynchon’s (and, to some extent, other writers’)
interest in entropy. Pynchon’s name has become a byword for reclusiveness in the literary
world, and comparatively little is known of his background. What biographical information
is available indicates that he studied physics or engineering (Royster (2005), apparently
the fullest biographical source available, states “physics and engineering”) at Cornell
University before changing to, and ultimately graduating in, English. It would be
reasonable to assume some connection between this and his interest in entropy, but critics
tend not to emphasise this particular angle.

36
Pynchon himself addresses the issue of entropy in his sole published piece of
autobiographical writing, his 1984 introduction to Slow Learner, a collection of his short
stories (including Entropy). He acknowledges the general perception of his relationship to
the concept and gives some indication of where his interest in it may have originated:

[P]eople think I know more about entropy than I actually do. [The writer]
Donald Barthelme has suggested…that I had some kind of proprietary
handle on it.

[E]ntropy got picked up on by some communication theorists and given the


cosmic moral twist it continues to enjoy in current usage. I happened to read
Norbert Wiener’s The Human Use of Human Beings…at about the same
time as The Education of Henry Adams, and the “theme” of [Entropy] is
mostly derivative of what these two men had to say.
(Pynchon 1984b, pp12-13)

Tabbi (1984) sees the references to Wiener and Adams as “play[ing] down the thematic
aspects of science” in the story; the absence of any reference to Pynchon’s academic
background may strengthen this argument. However, if this was Pynchon’s approach in
1984, it may not have been at the time of Entropy’s writing. Tabbi also mentions (op. cit.
p58) that “[when] Pynchon first submitted Entropy to an undergraduate writing class, one
of his teachers can recall trying to get [him] to change the title to something that didn’t
sound so scientific”.

It will also be noted that Pynchon apparently (and not uniquely) misattributes Adams’s
thoughts on entropy to the Education instead of the more obscure Letter to American
Teachers of History, but it is unclear whether this says anything useful about Pynchon’s
own interest in entropy.

Whatever Pynchon may be leaving out or distorting in this account, it is clear from what is
actually present that Adams, Wiener and Asimov played important parts in his adoption of
entropy as a metaphor. It is reasonable to believe that these figures had similar impacts on
at least some of the other writers who took up the idea, especially in the United States of
the post-war years, though, of course, every individual’s circumstances will have been
different. It is also reasonable to believe that Pynchon’s education had at least some effect,

37
notwithstanding that it is not foregrounded in his account. Pynchon himself may have had
a significant influence on other writers’ adoption of entropy, though it is likely that this
would have come to prominence too late to explain the initial take-up of the idea.

3.6 Conclusion

This chapter has examined some of the more prominent occurrences of entropy in literary
and other non-scientific applications over the course of the twentieth century. While it
cannot be a comprehensive survey, it does indicate something of the scale of the concept’s
usage. It is also clear that there has been considerable interaction between the
interdisciplinary spread of entropy within fields which are not part of science, but also
distinct from culture (e.g. communication theory) and its usage in literature, so that it is
questionable whether a straight migration of entropy from science into literature is an
adequate description of the process in question. The importance of certain figures in
spreading the idea does emerge, but this is not to imply that they were the first or only
individuals to extend entropy’s sphere of application in this way.

38
4 Discussion

4.1 Overview

In this chapter, I will consider some factors which various past commentators have
suggested as playing a role in the use of entropy outside its original scientific context, with
a view to determining to what extent they can explain the process and results observed in
the present case. I work within the overall framework set out in chapter 2, i.e. examining in
turn metaphor (both in general and in Zencey’s specific application to the case of entropy),
knowledge production and miscellaneous approaches to the issue.

4.2 Entropy as metaphor

4.2.1 General considerations

The fundamental idea of entropy, in its original thermodynamic sense, can in some ways
be interpreted as metaphorical. I have already referred (see 2.2.1 above) to the idea of
“container” metaphors; Lakoff and Johnson (op. cit. p30) refer to the fact that substances,
as well as abstract concepts, can be metaphorically seen as containers. Clausius’s concept
of “transformation content” clearly sees the substance possessing the quality of entropy in
such terms, though of course it is not an original coinage but an analogical extension of the
idea of work and energy content. The word “entropy” itself, as has been seen, was coined
by analogy with “energy”. Thus, it seems that, while there were aspects of metaphor in the
original development of the concept of entropy, analogy may have had a stronger
influence.

It is also possible to apply the alternative theories of metaphor – abstraction and


homonymy – discussed by Lakoff and Johnson (see 2.2.1 above) to the case of entropy.
Abstraction would hold that the thermodynamic, communication and general cultural
senses of entropy referred to a single concept, whereas homonymy would see them as
unrelated ideas with some points of similarity (such as their connections with disorder). In

39
practice, it appears from the evidence thus far that cultural uses of entropy are generally
distinguished from scientific ones by reasonably informed commentators, which suggests
that homonymy is followed in this case. In terms of the relationship between
thermodynamic and communication entropy, however, the picture is more complicated,
and the question of whether these constitute a single concept or two related (e.g. by similar
mathematical expressions) ones appears still unresolved. Thus, neither abstraction nor
homonymy dominates in the latter case. It will be remembered that Lakoff and Johnson
concluded that neither approach is universally valid; their contention in this respect that
metaphor can actively create similarities may have some relevance in the case of the two
entropies.

To return to the idea of an inherently metaphorical element in scientific writing, while we


have seen that the metaphorical angle of early works on entropy was rather limited, it is
possible to consider communication and cybernetic uses of entropy as metaphorical (if the
viewpoint that the two entropies are distinct is accepted). Indeed, as Reddy (1993 [1979])
points out, the idea of human communication as a conduit (essentially the basis of
Shannon’s model) is itself a metaphor. (Reddy also suggests (ibid. p175) that the “debate
and confusion” around the relationship between the two entropies may result from the
limitations of the conduit metaphor, which he sees as conflicting with the Second Law.)

The metaphorical nature of (for example) Shannon’s and Wiener’s usage of entropy may
then have been what attracted writers. Certainly, in Thomas Pynchon’s case, we have
already seen his attribution of the “cosmic moral twist” in entropy’s usage to
communication theorists, not to mention his portrayal of the connections between
thermodynamic entropy, communication entropy and Maxwell’s demon in The Crying of
Lot 49. While it is important not to associate literary entropy exclusively with Pynchon, the
fact remains that he has been probably the most prominent user of the concept, and a
significant influence on others. Similar connections can also be discerned in the case of
other writers; the cross-fertilisation between the work of Asimov and Wiener has already
been alluded to.

40
Mention has already been made of the possibility that the tone of some of Thomson’s
writings on entropy and heat death influenced subsequent non-scientific users of the
concept. However, it is difficult to see Thomson’s views as metaphorical in nature (though
they are definitely coloured by non-scientific, and in particular religious, ideas), and so
they cannot realistically be analysed in terms of metaphor.

4.2.2 Zencey’s suggested explanations

As has been previously mentioned, Zencey (op. cit.) has proposed a series of reasons for
entropy’s popularity as a metaphor. Here I attempt to apply them to the particular case of
entropy in literature, in view of what has come to light in the preceding chapter, and of the
theories proffered from purely literary viewpoints (e.g. Tanner’s and Lewicki’s) as to why
entropy came to prominence. I follow the order and numbering used in the original, from
which all quotations are taken.

1. “Entropy connotes disorder” and its use is associated with times in which
“accepted modes of thought” are being questioned. This may go some way
towards explaining its popularity during the “counterculture” era of the 1960s
and early 1970s, though the process of its transition from science to culture had
begun much earlier. It could also be argued that Norbert Wiener questioned
accepted modes of thought in some parts of The Human Use of Human Beings,
but this was not the primary purpose of his writings.
2. “Entropy is pessimistic” and “articulat[es] a sense that things are running
downhill, falling apart, getting worse” related to a “cult of nostalgia”. The
applicability of this explanation appears mixed. The “cult of nostalgia”
certainly does not accurately describe the attitudes of many (if any) of the
writers who adopted the idea of entropy. Zencey associates this concept with
(for example) Reaganism and the commodification of the past, as practised by
Disney. However, in at least some cases, nostalgia may spring from a
disaffection with the idea of progress. We have seen how Wiener used entropy
as a critique of the American belief in progress, and this may suggest that some
literary usages of the concept represent pessimistic reactions against the modern

41
world, just as the “cult of nostalgia” does, even if the motivations and
ideologies are different. Furthermore, it could reasonably be argued that Henry
Adams’s attraction to entropy arose at least partially from such sentiments, and
Adams’s writings on the subject clearly played an important role in entropy’s
transition from science into culture.
3. “Entropy is a vague idea, difficult to understand” because of the difficulty of
visualising it. Vagueness in turn “suits the temper of the times”, since “more
and more of our policy questions reveal themselves to be true dilemmas”, and
presumably metaphorical usage of entropy catches the general mood. The
applicability of this explanation to literature is uncertain. It is possible to argue
that the post-World War II world represented a transformed intellectual
landscape which challenged existing preconceptions and thus created a culture
receptive to vague ideas, but this argument does not seem to have been put
forward by any commentator on literary entropy. However, as has been seen,
Tanner makes a similar argument with regard to the characteristics of American
society in general rather than any particular time period. Again, it is possible
that Henry Adams, rather than his literary successors, is closest to the spirit of
Zencey’s ideas. In the Education, Adams perceived a continual increase in the
complexity of human activity in general, in line with technological
development, which would pose severe difficulties for education and science
(Adams 1999 [1918] pp407-414). His subsequent interest in entropy may owe
something to this perception.
4. Entropy reflects the concerns of “an energy-conscious age”. The relevance of
this factor is obvious in relation to usage of entropy from the mid-1970s on,
when consciousness of energy and the finite nature of its sources increased in
the wake of the 1973 oil crisis and the growth of environmentalism (the Club of
Rome’s influential The Limits to Growth was published in 1972). However,
entropy’s “breakthrough” in literary uses appears to have come over a decade
before this, so the applicability of such an explanation is questionable.
Nevertheless, there may be some connection in the case of the New Wave
science fiction writers, who, as previously mentioned, related entropy to ideas
of overpopulation and environmental destruction. It could also be argued that,

42
as Tanner proposed, the history of the United States encouraged an unusual
interest in force and energy amongst its writers, though this is considerably
harder to prove.
5. Energy-consciousness likewise leads to the spread of energy as metaphor, and
“where there is energy, there is entropy”. This explanation is subject to the
same provisos as the preceding one.
6. “The use of entropy as metaphor lends the authority of science to an argument
or image”, invoking “the main stock of apparently certain knowledge in our
culture”, even for those hostile to science, who can claim that “science itself
can be made…to suggest that a culture based on science is in decline”. This
argument has, in fact, sometimes been made by detractors of the use of entropy
in literature, who argue that writers applying the idea are attempting to blind
their audiences with science. For instance, Gore Vidal (1980 cited in Schuber
(1983) pp47, 59) dismisses Thomas Pynchon (and, by implication, similar
writers) as “writing about Entropy and the Second Law of Thermodynamics
and a number of other subjects that he picked up in his freshman year”.
According to Vidal, “English teachers without science like this sort of thing
while physicists are tempted to write excited letters to literary journals”. The
degree to which literary use of entropy actually constitutes the kind of trickery
Vidal alleges is debatable, though undoubtedly, as it became more widespread,
there were some such instances. (Von Neumann’s supposed comments to
Shannon, if true, suggest that he saw the usage of entropy in communication
theory as serving a similar purpose.) In science fiction, the “authority of
science” argument is obviously true up to a point, since traditionally (though
not universally) writers in the genre would call on scientific fact to support their
fictions. Nevertheless, it hardly explains why the New Wave writers, who
“inclined more towards psychology and the soft sciences” (Nicholls op. cit.
p866), should have been the most enthusiastic adopters of entropy within
science fiction. In wider literary contexts, entropy might be seen as a science-
driven critique of the science-driven optimism prevalent in, say, the post-war
US. Wiener’s views on progress again have relevance here; Lewicki’s
comments on the degree of public interest in science in the mid-twentieth-

43
century US may also help to explain the popularity of entropy in this particular
place and time.
7. “Entropy is truly a ubiquitous process, lying behind all change in the physical
world” including death and decay at one extreme and eating or cleaning at the
other. Insofar as this implies that entropy is a versatile metaphor (perhaps too
versatile in that, as mentioned at the beginning of this dissertation, it comes to
mean simply a kind of generalised decay), it seems applicable to the present
case. Some writers’ usages of entropy (such as Philip K. Dick’s “kipple”) have
emphasised this angle of ubiquity.
8. Entropy as metaphor is “a sort of compromise between scientific rationality
and…inscrutable magic” appealing to our taste for myth. This factor could be
applied to some instances of literary entropy – Pynchon’s Gravity’s Rainbow,
for instance, makes repeated reference to religious and mythological concepts
alongside its usage of entropy and other scientific ideas – but may not be valid
as a general rule.
9. “[T]o see [entropy] at work in the affairs of humans is to suggest, subtly, that
those affairs are beyond our control. What we cannot control, we cannot have
responsibility for…entropy as metaphoric mirror reflects our sense of
helplessness”. While it is difficult to see literary entropy as a disavowal of
responsibility, there are undeniable connections between entropy and the idea
of control. In Wiener’s view, control is anti-entropic, but (notwithstanding that
cybernetics dealt in large part with control), it can only win out “locally and
temporarily” (Wiener 1968 [1954] p33). Thus, the Second Law could from this
viewpoint be seen as a tendency for things to become less and less controllable.
Even if this is accepted, however, it is still unclear to what degree it contributed
to writers’ adoption of entropy.
10. “Humans have always worshipped a god of Order and feared the dark agency of
Chaos”, and social change involves merely an appeal to a different form of
order. We “cast our vision of the ideal order” in terms of “dialectical pairs”
(such as God/Satan), and energy/entropy fits into this pattern. Once again, this
factor may be more applicable to such usages as Adams’s rather than latter-day
cultural uses in general. However, Lewicki’s observations on entropy as a

44
continuation of an American fascination with apocalypse could be seen in such
a context.
11. Entropy “functions as a talisman” by virtue of making the “forces of darkness”
susceptible to examination, if not prediction. It is difficult to determine the
relevance of this factor; it may be of more value in explaining the increased
prominence of entropy across a range of different contexts in the 1970s, when
“forces of darkness” were to the fore in general debate as a result of (for
example) the rise in environmental and energy consciousness mentioned above.
12. “[T]he success of humanity’s domination of nature” leaves us to confront “the
chaos of our own devising”. In preference to this, we seek refuge in the
alternative visions of chaos which entropy offers, “recaptur[ing] something of
the psychic economy of the walled city...tell[ing] us that the elements of culture
are statistically improbable and temporally parochial pockets of order”. Again,
“the chaos of our own devising” can be interpreted in a number of ways. It is
certainly possible to relate this to environmental consciousness; furthermore,
Zencey himself cites the “potential for instantaneous disorder” inherent in
nuclear weapons, which would perhaps have been a more prominent concern at
the time when entropy was making its transition. In a wider sense, the idea of
an obsession with entropy as a response to “humanity’s domination of nature”
also chimes with some of Tanner’s and Lewicki’s comments on the peculiarly
American factors underlying its emergence in US literature.

4.3 Entropy and knowledge production

In chapter 2 of this dissertation, I gave a general outline of the theory of knowledge


production put forward by Gibbons et al. (op. cit.). It will be recalled that the core element
is the distinction between disciplinary “Mode 1” and interdisciplinary “Mode 2”
production. Here, I attempt to apply some of the principles of this theory to entropy (with
specific reference to its transition from science to culture). In essence, I am seeking to
determine whether entropy possessed some fundamental interdisciplinary nature which

45
may have eased this transition. A number of possible factors suggest themselves in this
respect.

Firstly, thermodynamics itself may possess some of the features of Mode 2 knowledge
production, notwithstanding that it long predates the general emergence of Mode 2. It
arose, not within traditional scientific disciplines, but out of the practical problems of
steam power. The major contributors to its early development came from diverse
backgrounds – Clausius from optics or Gibbs from the study of gear systems, for instance.
It was, in fact, criticised within physics for being excessively driven by engineering
applications rather than pure science; Jammer (op. cit. p114) mentions one such critic’s
comment that “the steam engine did much more for science than science has done for the
steam engine”.

However, Gibbons et al. (op. cit. p4) have to some extent forestalled this argument. They
discuss certain disciplines characterised by “research carried out in the context of
application”, such as aeronautical engineering or computer science. While their work is
oriented towards applications, they are not applied sciences because they fill gaps in the
existing science. They have, however, become integrated into the traditional disciplinary
structure. While they “share with Mode 2 some aspects of the attribute of knowledge
produced in the context of application”, they have a less complex context, lacking the
“diverse set of intellectual and social demands” that characterises Mode 2.

Thermodynamics shares some traits with the disciplines referred to (in being an area of
pure research that evolved to support applications not adequately served by the traditional
disciplines) but also differs, notably in that it is not always regarded as a scientific
discipline unto itself, at least in the academic context. Thus, it would appear that it lies
somewhere between what Gibbons et al. are describing and true Mode 2 activity.

Secondly, the context in which the usage of entropy was extended to communications after
World War II also reflects some of the characteristics of Mode 2. Claude Shannon’s
linkage of entropy with information theory originated at Bell Labs, an example of the type
of interdisciplinary, corporate-led research establishment that began to supplant

46
disciplinary academic research as Mode 2 emerged. Likewise, Léon Brillouin’s work on
Maxwell’s demon was done at IBM. It is conceivable that the interdisciplinary nature of
these surroundings may have encouraged intellectual cross-fertilisation and thereby
spurred the transfer of entropy into new contexts.

However, it is important not to make too much of this. For one thing, Norbert Wiener (the
importance of whose writings has already been indicated) conducted his work in a more
traditional university context, though there were nevertheless aspects of interdisciplinarity
in his own background as a mathematical philosopher and in his collaboration with, for
example, engineers (O’Connor and Robertson 2003c).

Thirdly, the development of Mode 2 has itself paralleled that of literary entropy. Both are
primarily creatures of the post-World War II era; both have benefited from the impact of
trends such as wider access to education, which not only helped create the conditions for
Mode 2 but, as previously mentioned, widened scientific understanding and thus increased
the applicability of ideas such as entropy. Gibbons et al. (op. cit. p108) themselves refer to
the:
many intriguing correspondences between the growth of modern culture and
the transformations of industrial society in which science and technology
appear to have been triumphant…[these] are more than simple co-
occurrences; they are co-constructions of meanings and cultural symbols…

They quote the parallel development of the Industrial Revolution and movements such as
Impressionism in the 19th century, and Einsteinian relativity and surrealism, atonal music
and the modernist novel in the early 20th century, as examples, with the post-industrial
economy (as exemplified by Mode 2) and postmodernism as the latter-day equivalent. In
this context, it is clear that Mode 2 and the rise of entropy (amongst other trends) in
literature can be seen as part of the same general process.

Fourthly, Gibbons et al. (ibid. p83) refer to “the deliberately decentred diversity and
incoherence associated with postmodernism” as one of three ways (along with “the
ceaseless subdivision of knowledge” and the acceptance of “wider definitions of
knowledge”) in which “the knowledge which is [the university’s] chief commodity has
become diffuse, opaque, incoherent, centrifugal”. Several of the authors most closely

47
associated with the literary usage of entropy (certainly Pynchon and Barthelme, arguably
some of the New Wave science fiction writers) would generally be classed as
postmodernists. It may be that the emergence of literary postmodernism opened the way
for a wider definition of what forms of knowledge and culture were legitimate subjects for
literary reference and metaphor, and that entropy was a beneficiary of this. In fact,
Gibbons et al. (ibid. p84) refer to a breaking down of the distinctions between, inter alia,
knowledge and culture under the pressure of the changes in education and research
resulting from the rise of Mode 2 knowledge production, and entropy in literature could be
interpreted as an interpenetration of knowledge and culture.

It should also be noted that Gibbons et al. explicitly address the humanities’ (including
culture’s) place in the production of knowledge (ibid. p110). They feel that there are
problems with direct application of their general analysis, since the humanities in some
senses stand outside society in general (e.g. providing critiques), but in other ways they can
be said to be “deeply implicated” in the new order (through the “powerful, even hegemonic
images” of the “culture industry”, and the effects of mass higher education). Entropy in
culture tends to be associated with “high” rather than “mass” cultural forms, so it would
seem to fit easier with the role of the humanities as critic; indeed, Gibbons et al. list
“doom-laden prophecies” as one manifestation of this, and this is obviously one prominent
way in which entropy is used as social criticism. Entropy is thus both (in at least some
ways) a product of Mode 2 and an aid to critics of the society in which Mode 2 knowledge
production takes place. As referred to in the discussion of Zencey’s ideas, the fact that
entropy is associated with the ideas (whether scientific progress or Mode 2 knowledge) on
which society is based may actually make it a more attractive basis for a critique of them.

4.4 Summary

In this chapter, I have analysed entropy and, in particular, how it came to be adopted by
writers of fiction, from the viewpoints of metaphor and knowledge production.

48
In terms of metaphor, I have addressed the idea that a metaphorical element in the work of
the original scientific exponents of the concept may have played a part in the wider
adoption of entropy. I have also briefly outlined how major theories of metaphor could be
applied to the case of entropy. Furthermore, I have discussed Zencey’s comments on
entropy as metaphor extensively with regard to the specific case of its use in literature.

In relation to knowledge production, I have placed the ideas of thermodynamics generally,


and entropy in particular, within the historical/cultural context described by Gibbons and
his co-authors. Theories of knowledge production appear particularly promising as an
explanation of entropy’s interdisciplinary transfer in the wider context of mid-twentieth-
century knowledge and culture.

49
5 Conclusion

This dissertation has attempted to cover in reasonable depth one particular aspect of
entropy’s development and interdisciplinary spread. I have been necessarily limited in the
scope of my research by virtue of the scale and timeframe of this dissertation, which have
required concentration on specific aspects of the history of entropy. As I have already
mentioned, my decision has been to give particular attention to the process by which
entropy-as-metaphor became widely adopted in the literary world.

No one person was solely responsible for this process. Certain individuals did play
important roles; for instance, Henry Adams and, later, Claude Shannon were instrumental
in spreading the idea of entropy beyond its original field. Norbert Wiener was particularly
important in popularising the concept of communication entropy, and thus attracting the
attention of many – including, it appears, at least some of the writers who were to utilise it
in their fiction – to the potential of entropy outside scientific usages. Explicit use of the
concept within literature goes back at least to Nathanael West. Isaac Asimov played an
important role in bringing it to a wider audience, and, from about 1960 onwards, it rapidly
became widespread in both the science fiction of the “New Wave” and certain branches of
literary fiction, particularly amongst US authors and generally (though not exclusively)
those considered postmodernists.

While I have avoided going in detail into the subject – itself fascinating – of how the
scientific community came to adopt entropy, it is clear that some aspects of its early usage
have influenced subsequent interdisciplinary dissemination of the idea. In particular, the
strong philosophical angle in William Thomson (Lord Kelvin)’s writings on entropy had a
definite influence on the subsequent attempts of Henry Adams to apply the concept in non-
scientific contexts, though there is no evidence that later users of entropy referred
explicitly to Thomson.

50
What factors were involved in this transition from science into culture? I have given
particular attention in this dissertation to two areas. The first is the theory of metaphor; in
this regard, I examined both general theories and the specific explanations put forward by
Zencey (op. cit.).

I have investigated the idea that some inherent metaphorical character in scientific writing,
and specifically in the writings of early theorists of entropy, may have influenced its wider
adoption. However, it does not appear that this was so, though the metaphorical aspects of
entropy’s later application by communication theorists may well have played such a role.

In terms of alternative approaches to metaphor, the case of entropy appears to fit better
with homonymy (unrelated ideas with some similarities) than with abstraction (a single
concept). It is, however, uncertain exactly what this says about the process of its transition
between disciplines; the most that can be concluded is that this suggests that non-scientific
usages were indeed metaphorical and not a matter of discovering that other concepts were
somehow identical with thermodynamic entropy. In the case of the relationship between
entropy in thermodynamics and entropy in communication theory, neither approach fits
well.

Zencey’s analysis comprises primarily a list of possible factors in entropy’s popularity as a


metaphor; as might be expected from this structure, the applicability of these factors to the
specific case I am examining is mixed. In general, they better explain the attraction of
entropy for early interdisciplinary users of the concept, such as Henry Adams, than they do
its later adoption in literature. However, they do on occasion echo the arguments of others
who have assessed the relationship between entropy in literature (a primarily American
phenomenon) and aspects of American culture.

My second major area of analysis was the theory of knowledge production, as proposed by
Gibbons et al. (op. cit.). In general, I have found this highly applicable to the case of
entropy. It appears that the interdisciplinary characteristics of thermodynamics in general,
and of the way in which entropy was applied in the mid-20th century, go a considerable
way towards explaining entropy’s appeal in cultural contexts. Furthermore, it is possible to
draw parallels between relevant scientific/knowledge and cultural developments, just as

51
Gibbons et al. do, with specific reference to entropy. The general developments in culture
and knowledge arising from the transformation of knowledge production (e.g. the rise of
postmodernism) are also very relevant to the dissemination of entropy.

I have not explicitly touched on the role of popularisation in the dissemination of entropy.
It nevertheless becomes clear from my analysis that the use of the idea of entropy in works
aimed at a popular, or at least non-scientific, audience (as in the cases of Adams and
Wiener, for instance) has had an important role to play. Since these usages were essentially
metaphorical, it is difficult to represent them as part of a process of popularisation in the
conventional sense, but there are obvious parallels with the latter. Conventional
popularisation is rightly considered important to society’s perceptions of science, but the
example of entropy suggests that “popularisation of the metaphor” may play a key role in
certain instances.

Obviously, a project such as this one only touches on a very limited part of a substantial
subject area, and there are many aspects – such as how entropy is explained in the context
of education or popularisation, the detail of how the idea came to be adopted by scientists,
and the latter-day metaphorical usage of entropy in fields other than literature (e.g.
management, the visual arts) – which I regrettably had to exclude. It is clear that there is an
opening for a much more comprehensive study than I have been able to conduct of entropy
and its interdisciplinary manifestations.

In addition, there appears to be considerable potential for deeper analysis of how theories
of knowledge production could explain entropy’s interdisciplinary transition, and this type
of analysis could potentially be applied to other scientific concepts. Furthermore, the idea
of metaphorical usage as playing an equivalent role to popularisation also appears to
deserve examination in greater depth.

My journey through the history of entropy has been an eye-opening, if all too brief, one. I
hope that it has gone some way towards providing a new perspective on an idea that is
well-known in both science and literature, and that I have plausibly answered some
pertinent questions, while at the same time raising further ones.

52
References

Adams, H. 1919. A Letter to American Teachers of History IN: Adams, H. The


Degradation of the Democratic Dogma. New York: Macmillan.

Adams, H.1999 (first published 1918). The Education of Henry Adams. Oxford, Oxford
University Press.

American Heritage Book of English Usage. 1996. Science Terms: Distinctions,


Restrictions and Confusions – Entropy [Online]. Available from:
<http://www.bartleby.com/64/C004/024.html> [Accessed 17 June 2006].

American Institute of Physics. 2000. Finding Aid to the Papers of Norbert Wiener, 1898-
1966 [Online]. Available from:
<http://www.aip.org/history/ead/mit_wiener/19990053_content.html>
[Accessed 25 June 2006].

Asimov, I. 1987 (first published 1956). The Last Question IN: Asimov, I. Robot Dreams.
London: Victor Gollancz.

Atkins, P.W. 1994. The Second Law. Revised edition. New York: Scientific American
Books.

Brillouin, L. 1990a (first published 1949). Life, Thermodynamics and Cybernetics. IN:
Leff, H.S. and Rex, A.F. (eds.) Maxwell’s Demon: Entropy, Information, Computing.
Bristol: Adam Hilger.

Brillouin, L. 1990b (first published 1951). Maxwell’s demon cannot operate: Information
and Entropy I IN: Leff, H.S. and Rex, A.F. (eds.) Maxwell’s Demon: Entropy, Information,
Computing. Bristol: Adam Hilger.

Concise Oxford English Dictionary. 11th edition. 2004. Oxford: Oxford University Press.

Conder, J.J. 1970. A Formula of His Own: Henry Adams’s Literary Experiment. Chicago
and London: University of Chicago Press.

Daub, E.E. 1990 (first published 1970). Maxwell’s Demon. IN: Leff, H.S. and Rex, A.F.
(eds.) Maxwell’s Demon: Entropy, Information, Computing. Bristol: Adam Hilger.

Dick, P.K. 2005 (first published 1968). Do Androids Dream of Electric Sheep? London:
Orion.

Dutta, A. 1995. The Paradox of Truth, the Truth of Entropy [Online]. Available from:
<http://www.themodernword.com/pynchon/papers/dutta.html>
[Accessed 5 February 2006].

Fiske, J. 1982. Introduction to Communication Studies. London: Methuen.

53
Friedman, A. 1983. Science and Technology IN: Clerc, C. (ed.) Approaches to Gravity’s
Rainbow. Columbus: Ohio University Press.

Gibbons, M., Limoges, C., Nowotny, H., Schwartzman, S., Scott, P. and Trow, M. 1994.
The New Production of Knowledge: The Dynamics of Science and Research in
Contemporary Societies. London: SAGE Publications.

Jammer, M. 2003. Dictionary of the History of Ideas: Entropy [Online].


Available from: <http://etext.lib.virginia.edu/cgi-local/DHI/dhi.cgi?id=dv2-12> [Accessed
5 April 2006].

Lakoff, G. and Johnson, M. 1980. Metaphors We Live By. Chicago and London: University
of Chicago Press.

Leff, H.S. and Rex, A.F. 1990. Overview IN: Leff, H.S. and Rex, A.F. (eds.) Maxwell’s
Demon: Entropy, Information, Computing. Bristol: Adam Hilger.

Lewicki, Z. 1984. The Bang and the Whimper: Apocalypse and Entropy in American
Literature. Westport (Connecticut): Greenwood Press.

McGovern, J. 1996. The Essence of Engineering Thermodynamics. Hemel Hempstead:


Prentice Hall Europe.

Nadel, I.B. 1999. Introduction IN: Adams, H. The Education of Henry Adams. Oxford,
Oxford University Press.

Newhagen, J.E. and Fox, J. 2003. The importance of analytical level in theory building:
The use of entropy in communication theory and the problem of heat death [Online].
Available from: <http://jnews.umd.edu/johnen/research/heat_death.pdf> [Accessed 31
January 2006].

Nicholls P. 1993a. Entropy IN: Clute, J. and Nicholls, P. (eds.) The Encyclopedia of
Science Fiction. London: Orbit.

Nicholls P. 1993b. New Wave IN: Clute, J. and Nicholls, P. (eds.) The Encyclopedia of
Science Fiction. London: Orbit.

O’Connor, J.J. and Robertson, E.F. 2000. Rudolf Julius Emmanuel Clausius [Online].
Available from:
<http://www-history.mcs.st-andrews.ac.uk/history/Biographies/Clausius.html>
[Accessed 10 June 2006].

O’Connor, J.J. and Robertson, E.F. 2003a. Claude Elwood Shannon [Online]. Available
from:
<http://www-history.mcs.st-andrews.ac.uk/history/Biographies/Shannon.html>
[Accessed 20 June 2006].

54
O’Connor, J.J. and Robertson, E.F. 2003b. John von Neumann [Online].
Available from:
<http://www-history.mcs.st-andrews.ac.uk/Biographies/Von_Neumann.html> [Accessed
21 June 2006].

O’Connor, J.J. and Robertson, E.F. 2003c. Norbert Wiener [Online]. Available from:
<http://www-history.mcs.st-andrews.ac.uk/history/Biographies/Wiener_Norbert.html>
[Accessed 20 June 2006].

Pynchon, T. 1984a (first published 1960). Entropy. IN: Pynchon, T. Slow Learner: Early
Stories. Boston: Back Bay Books.

Pynchon, T. 1984b. Introduction. IN: Pynchon, T. Slow Learner: Early Stories. Boston:
Back Bay Books.

Pynchon, T. 1996 (first published 1965). The Crying of Lot 49. London: Vintage.

Pynchon, T. 2000 (first published 1973). Gravity’s Rainbow. London: Vintage.

Reddy, M.J. 1993 (first published 1979). The conduit metaphor: A case of frame conflict in
our language about language IN: Ortony, A. (ed.) Metaphor and Thought. 2nd edition.
Cambridge: Cambridge University Press.

Ritchie, D. 2004. Information as Metaphor: Biology and Communication IN:


Braman, S. (ed.) Biotechnology and Communication: The Meta-Technologies of
Information. Mahwah (New Jersey), Lawrence Erlbaum Associates.

Rothstein, D. 2003. Does evolution contradict the second law of thermodynamics?


[Online]. Available from: <http://curious.astro.cornell.edu/question.php?number=441>
[Accessed 18 June 2006].

Royster, P. 2005. Thomas Pynchon: A Brief Chronology [Online]. Available from:


<http://digitalcommons.unl.edu/cgi/viewcontent.cgi?article=1001&context=libraryscience
>
[Accessed 6 January 2006].

Schneider, E.D., and Sagan, D. 2005. The Cosmic Casino: Statistical Mechanics [Online].
Available from: <http://www.intothecool.com/cosmic.php> [Accessed 24 February 2006].

Schuber, S.P. 1983. Rereading Pynchon: Negative Entropy and “Entropy”. Pynchon Notes
[Online]. 13, pp47-60. Available from:
<http://w3.ham.muohio.edu/~krafftjm/pn/pn013.pdf> [Accessed 16 May 2005].

Scrivener, A.B. 2003. What Ever Happened to Cybernetics? Cybernetics in the Third
Millennium [Online]. 2(4). Available from:
<http://www.well.com/~abs/Cyb/4.669211660910299067185320382047/c3m_0204.txt>
[Accessed 25 June 2006].

55
Seiler, E. and Jenkins, J.H. 2004. Frequently Asked Questions about Isaac Asimov
[Online]. Available from:
<http://www.asimovonline.com/asimov_FAQ.html> [Accessed 24 June 2006].

Shannon, C. E. 1948. A Mathematical Theory of Communication. Bell System Technical


Journal [Online]. 27, pp379-423, 623-656. Available from:
<http://cm.bell-labs.com/cm/ms/what/shannonday/shannon1948.pdf>
[Accessed 31 May 2006].

Smith, C. and Wise, M.N. 1989. Energy and Empire: A biographical study of Lord Kelvin.
Cambridge: Cambridge University Press.

Stuart, Z.P. 1997. Entropy [Online]. Available from:


<http://www.pynchon.pomona.edu/entropy> [Accessed 23 January 2006].

Szilard, L. 1990 (first published 1929). On the Decrease of Entropy in a Thermodynamic


System by the Intervention of Intelligent Beings (tr. Rapoport, A. and Knoller, M.) IN:
Leff, H.S. and Rex, A.F. (eds.) Maxwell’s Demon: Entropy, Information, Computing.
Bristol: Adam Hilger.

Tabbi, J. 1984. Merging Orders: The Shaping Influence of Science on “Entropy”. Pynchon
Notes [Online]. 15, pp58-68. Available from:
<http://w3.ham.muohio.edu/~krafftjm/pn/pn015.pdf> [Accessed 7 May 2005].

Tanner, T. 1971. City of Words: American Fiction 1950-1970. London: Jonathan Cape.

Thomson, W. 1851. On the Dynamical Theory of Heat, with numerical results deduced
from Mr Joule’s equivalent of a Thermal Unit, and M. Regnault’s Observations on Steam.
[Online]. Available from:
<http://zapatopi.net/kelvin/papers/on_the_dynamical_theory_of_heat.html> [Accessed 10
June 2006].

Thomson, W. 1852. On a Universal Tendency in Nature to the Dissipation of Mechanical


Energy. [Online]. Available from:
<http://zapatopi.net/kelvin/papers/on_a_universal_tendency.html>
[Accessed 10 June 2006].

Thomson, W. 1990 (first published 1874). Kinetic Theory of the Dissipation of Energy.
IN: Leff, H.S. and Rex, A.F. (eds.) Maxwell’s Demon: Entropy, Information, Computing.
Bristol: Adam Hilger.

Truesdell, C. 1980. The Tragicomical History of Thermodynamics 1822-1854. New York:


Springer-Verlag.

Watanabe, S. 1985. Wiener on Cybernetics, Information Theory and Entropy IN: Masani,
P. (ed.) Norbert Wiener: Collected Works with Commentaries – Volume IV. Cambridge
(Massachusetts), The MIT Press.

56
West, N. 1983 (first published 1933) Miss Lonelyhearts IN: West, N. The Complete Works
of Nathanael West. London: Picador.

Wiener, N. 1961 (first published 1948). Cybernetics: or Control and Communication in the
Animal and the Machine. 2nd edition. Cambridge (Massachusetts), The MIT Press.

Wiener, N. 1968 (first published 1954). The Human Use of Human Beings: Cybernetics
and Society. Revised edition. London: Sphere.

Wiener, N. 1985a (first published 1952). Cybernetics (Light and Maxwell’s Demon) IN:
Masani, P. (ed.) Norbert Wiener: Collected Works with Commentaries – Volume IV.
Cambridge (Massachusetts), The MIT Press.

Wiener, N. 1985b (first published 1950). Entropy and Information IN: Masani, P. (ed.)
Norbert Wiener: Collected Works with Commentaries – Volume IV. Cambridge
(Massachusetts), The MIT Press.

Zencey, E. 1991. Some Brief Speculations on the Popularity of Entropy as Metaphor.


Metaphor and Symbolic Activity [Online]. 6(1), pp47-56. Available from: EBSCOHost
(TCD access)
<http://atoz.ebsco.com/home.asp?Id=1324> [Accessed 12 June 2006].

57

Вам также может понравиться