Download as pdf or txt
Download as pdf or txt
You are on page 1of 224

INTERACTION THEORY IN FOREST ECOLOGY AND MANAGEMENT

FORESTRY SCIENCES

Baas P, ed: New Perspectives in Wood Anatomy. 1982. ISBN 90-247-2526-7


Prins CFL, ed: Production, Marketing and Use of Finger-Jointed Sawnwood. 1982.
ISBN 90-247-2569-0
Oldeman RAA, et aI., eds: Tropical Hardwood Utilization: Practice and Prospects. 1982.
ISBN 90-247-2581-X
Den Ouden P and Boom BK: Manual of Cultivated Conifers: Hardy in Cold and Warm-
Temperate Zone. 1982. ISBN 90-247-2148-2 paperback; ISBN 90-247-2644-1 hard-
bound.
Bonga JM and Durzan OJ, eds: Tissue Culture in Forestry. 1982. ISBN 90-247-2660-3
Satoo T and Magwick HAl: Forest Biomass. 1982. ISBN 90-247-2710-3
Van Nao T, ed: Forest Fire Prevention and Control. 1982. ISBN 90-247-3050-3
Douglas J: A Re-appraisal of Forestry Development in Developing Countries. 1983.
ISBN 90-247-2830-4
Gordon JC and Wheeler CT, eds: Biological Nitrogen Fixation in Forest Ecosystems:
Foundations and Applications. 1983. ISBN 90-247-2849-5
Hummel FC, ed: Forest Policy: A Contribution to Resource Development. 1984.
ISBN 90-247-2883-5
Duryea ML and Landis TO, eds: Forest Nursery Manual: Production of Bareroot Seed-
lings. 1984. ISBN 90-247-2913-0
Manion PO, ed: Scleroderris Canker of Conifers. 1984. ISBN 90-247-2912-2
Staaf KAG and Wiksten NA: Tree Harvesting Techniques. 1984. ISBN 90-247-2994-7
Duryea ML and Brown GN, eds: Seedling Physiology and Reforestation Success. 1984.
ISBN 90-247-2949-1
Boyd JD: Biophysical Control of Microfibril Orientation in Plant Cell Walls. 1985.
ISBN 90-247-3101-1
Findlay WPK, ed: Preservation of Timber in the Tropics. 1985. ISBN 90-247-3112-7
Interaction theory in forest
ecology and management
by

ROLFE A. LEARY
School oj Forestry
Michigan Technological University
Houghton, Michigan, U.S.A.

1985 MARTIN-US NIJHOFF/DR W. JUNK PUBLISHERS


a member of the KLUWER ACADEMIC PUBLISHERS GROUP
DORDRECHT / BOSTON / LANCASTER
ISBN-13: 978-94-0 I 0-8779-7 e-ISBN-13: 978-94-009-5151-8
DOl: 10.1007/978-94-009-5151-8

Distributors

for the United States and Canada: Kluwer Academic Publishers, 190 Old Derby
Street, Hingham, MA 02043, USA
for the UK and Ireland: Kluwer Academic Publishers, MTP Press Limited,
Falcon House, Queen Square, Lancaster LAI lRN, UK
for all other countries: Kluwer Academic Publishers Group, Distribution Center,
P.O. Box 322, 3300 AH Dordrecht, The Netherlands

Library of Congress Cataloging in Publication Data

Leary! Rolfe A.
Interaction theory in forest ecology and management.

(Forestry sciences ; 19)


Bibliography: p.
Includes index.
1. Forests and forestry--Multiple use. 2. Forest
management. 3. Forest ecology. 4. Forests and
forestry--Philosophy. I. Title. II. Series: Forestry
sciences j v. 19.
SD387.M8L43 1985 634.9'28 85-18146

ISBN-13: 978-94-010-8779-7

Copyright

© 1985 by Martinus NijhofflDr W. Junk Publishers, Dordrecht.


Softcover reprint of the hardcover 1st edition 1985
All rights reserved. No part of this publication may be reproduced, stored in a
retrieval system, or transmitted in any form or by any means, mechanical,
photocopying, recording, or otherwise, without the prior written permission of
the publishers,
Martinus NijhofflDr W. Junk Publishers, P.O. Box 163, 3300 AD Dordrecht,
The N€therlands.
TABLE OF CONTENTS

Preface ix
Introduction 1
Part I: Setting the Stage, Exactly 6
1: The Science of Forest Science 7
A. Language, Construct, Object 9
B. Category Inventories 11
C. Within Category Relations 19
D. Between Category Relations 23
E. Research Problems 30
2: Interaction, "Interaction", 'Interaction' 31
A. Interaction - The Event 31
B. "Interaction" - The Concept 34
C. 'Interaction' - The Term 40
D. Research Problems 43
3: The Roles of Spontaneous and Induced Change in Some Forestry Concepts 45
A. Site and Related Concepts 45
B. Measures of Space Occupancy by Trees 51
C. Accretion 53
D. Assessing the Animal Producing Power of Forests 54
E. Discussion 55
F. Research Problems 56
4: Interaction Geometry 57
A. Fundamental Parts 57
B. Combining Fundamental Parts 60
C. Mapping Experimental Results 63
D. A Wrinkle in Interaction Space 65
E. Haskellian Coordinates as a Fourth Order Predicate 66
F. Research Problems 67
Part II: Analysis 69
5: Haskellian Coordinates in Analysis - Searching For Pattern 71
A. Competition (-, - ) 72
B. Prey-Predator (-, +) 79
C. Parasite-Host (+, - ) 84
D. Plant-Herbivore (-, +) 88
E. Pattern and Expectation 91
F. Research Problems 92
6: Patterns of Interaction in Mixed Forest Stand Dynamics 95
A. Historical Review 96
B. Role of Presuppositions in Study Design and Analysis 97
C. Example Mixed Stand Dynamics 98
D. Controlling Interactions by Thinning 103
E. Geometrizing Silviculture 106
F. Research Problems 107
vi Table of Contents

Part III: Synthesis 108


7: Haskell's Coordinate System in Synthesis 109
A. Synthesis in General 110
B. Integrative Principles 111
C. Forms of Scientific Synthesis 111
D. Haskell's Coordinate System and the Mathematical Form of Synthesis 112
E. Discussion 118
F. Research Problems 119
8: "Interaction" in a System of Concepts 121
A. Valence of "Interaction" 122
B. A cummulative level structure 126
C. "Interaction" and Concept Maps 131
D. Summary 134
E. Research Problems 134

Part IV: Structure Evolution 136


9: A Theory of Indirect Effects and Implications for
the l\atural Selection of Community Structures 137
A. Community, Structure, Natural Selection 137
B. The Set, the Operator, the Structure 141
C. A Theory of Indirect Interactions 143
D. Discussion of Theorems 145
E. l\atural Selection Levels in Community Interaction Structure Evolution 153
F. Community Interaction Structure Evolution 155
G. Summary 157
H. Research Problems 158
10: l\atural Selection of Interaction Structures
in Communities With Many Populations 161
A. Representing Structure Analysis Results 161
B. Interaction Structure Analysis Methods 162
C. Interpreting the Analysis 165
D. Toward a New Exclusion Principle 166
E. Discussion 169
F. Summary 170
G Research Problems 171
Ta ble of Contents vii

Part V: Decision-making 172


11: A Clarification and Extension of Multiple Use 173
A. 'Multiple use', "Multiple use", Multiple use 173
B. Analyzing of "Multiple use" 175
C. Extending the Concept of Multiple Use 176
D. Research Problems 178
12: Conservation Traditions 179
A. Economic Efficiency Tradition 179
B. Ecologic Tradition 182
C. Biocentric Tradition 183
D. Summary 184
E. Research Problems 184
13: Toward a Synthesis of Conservation Traditions 187
A. The Central Problem 187
B. Biocentric-Ecologic 188
C. Ecologic-Economic 189
D. Biocentric-Ecologic-Economic 189
E. Observations 195
F. Research Problems 196
Literature Cited 197
Su bject Index 213
Preface

As J understand it, a book Preface is where the author explains to the reader how
the book in hand came about, something of the personal reasons for having inflicted such
extended duress on one's self to complete the manllscript. and other items that are fit to say
but do not fit in the text.
This book had its conceptual beginnings in the 1970's wit h my 'studies in scientific
synthesis at the North Central Forest Experiment Station, St. Paul, Minnesota. Ours is,
clearly, the age of analysis. But, I felt, we must soon begin frameworks for synthesis, or a
synthesis would never be possible. In short, I hoped to develop 'interaction' as an integrative
principle in forestry. As work progressed on the manuscript, other subthemes developed.
First, there was the vague feeling on my part that the forestry profession was losing
ground in the contest to see who should manage the forests of the world. This was happening
not because foresters do not know how to manage forests in a reasonable manner, but because
the public seemed to be loosing faith in the judgement of foresters as professional, responsible,
wise land managers. Several well-known incidents of poor judgement in timber harvesting
methods on national forests in the United States did little to help the forester's image. Only
recently I was shocked while watching Jacques Cousteau's beautiful public television report
on the Amazon basin to hear him state that as new areas of the basin are opened, the first
people in are the miners and the foresters. Rven the very di"tinguished public seems not to
distinguish between foresters and loggers. Being associated with miners is cause for alarm.
Second, there is the need for an alternative to financial criteria when a forester needs
an arbiter to chose a good and right action in dealing with forests. The dominant arbiter
is man centered, human demand driven, with a heavy dose of wealth accumulation and
financial optimization. This seems ill-advised for the health of a profession that is supposed
to have a very long term view of things.
Third, foresters have little appreciation for philosophy in general, and even less appre-
ciation for its importance in their work. Even foresters with a doctor of philosophy degree
usually have had no formal instruction in it. Philosophers study, among other topics, how
people view and categorize nature. Every forester views and categorizes nature in some
manner, even if unarticulated. How a forester views nature affects how one approaches the
task of managing that bit of nature where trees abound. Forester Aldo Leopold's call for
a land ethic, published in the Journal of Forestry, seems so logical, yet foresters as a whole
seem oblivious to it. Or, perhaps, they don't know what to do with an ethic. There are at
least three options: accept it, repudiate it, or ignore it. Foresters, it seems, have ignored it.
But, in my view, forestry will not survive as a viable profession without philosophy, including
ethics.
lowe much to those persons who have helped and encouraged me the past several years.
First, I thank R. Brander, C. Berntsen, and C. Fasick, who a decade ago saw some value in
what were then my embryonic ideas on scientific synthesis. My apologies for taking 12 years
to nurture the ideas to their more mature current state. I thank A. Lundgren for maintaining
a special kind of research environment where it was okay to read and think about the ideas
x Preface

of Piaget, Haskell, Bunge, Margenau, and others, and for introducing the 'timber is trees
but trees are more than timber' ideas I make use of I thank G. Brand for his invaluable help
in completing tllP analyses in Chaper 6 and in clarifying my thoughts about the synthesis in
Chapter 1:1.
Several persons read and reviewed manuscript chapters: 1. Donoghue, K. Brown, B.
Hargraves, D. Dindal, W Lidicker, M. Bunge, .I. Stewart, E. Bakuzis, N. Lorimer, W
Mattson, and G Brand Thanks. J. lIaefner declined to revipw a single chapter when I
asked, but offered to, and did, review (at that time) the entIre manuscript, and was a great
help. Of course, any mistakes that remain are my own.
Two persons lowe special thanks arc Edward Haskell and Egolfs llakuzis. Edward
Haskell has generously shared with me over the past 12 years an important part of his life's
work. I hope that I have used it and extended it in a direction that is consistent with one of
his overall goals - scientific synthesis. My greatest debt lowe Egolfs V. Bakuzis, Professor
Emeritus of Forest Ecology, College of Forestry, Cniversity of Minnesota. It was in Egolfs'
1971 forest synecology class that I first heard of Piagpt, Haskell and Bunge. Over the years
()ur periodic consultations amounted to a 14 year post-cloc for me. Without Egolfs' constant
encuuragclllent to cuntinue, I wuuld likely have taken an easier route years ago.
I thank my wife Barbara Buckett Leary and our children Daniel and Eleanor for the
financial sacrifices they have endured which allowed me to complete the camera ready copy
of this manuscript. I promise you the evenings and weekends of the corning year, the summer
of 1985, and all those to follow.
L. Bartelli, School of Forestry, Michigan Technological University graciously provided
support for research and manuscript preparation. S. McCulley edited the manuscript, E.
Crittenden did an excellent job producing the camera ready copy. D. Clawson and D. White
illustrated my ideas admirably well. M. Wacek got me over the hump with her superb job
of typing much of the manuscript. lowe a special debt to Bart Childs and his students
Tim Holdridge and Scott Alderink at Texas A&M University. They took my Wordstar files
and produced this beautiful copy with the 1EX typesetting system. This was done on an
MV/8000 computer system using a Lasergraphix 1200 printer.
Houghton, 1985

nyc is a registered trademark of the American Mathematical Society.


Wordsta~ is a registered trademark of the MicroPro International Corporation
MV/8000 is a registered trademark of Data General Corp.
Lasergraphix is a registered trademark of Quality Micro Systems, Inc.
Introduction

Renewable natural resource managers are often forced to choose between actions that are
economically efficient and those that are ecologically and environmentally sound. Sometimes
an action may be both economically efficient and ecologically sound, but often the two action
criteria are incompatible. When incompatible, the choice between actions can be difficult for
a number of reasons. For one, citizens who become involved in resource use decisions often
are aligned with one of two broad categories for action, one financial or economic, the other
ecologic or enviornmental. The economic efficiency criterion for action presumes a particular
view of the world, one associated with commodities or items for trade or commerce. On
the other hand, the ecologic criterion for action probably presumes no financial values at all
among the affected object populations. Citizens favoring each of these two broad categories
of actions have different value systems. Groups sharing the same or similar value systems
often form special interest groups and vie for effect on public renewable natural resource
decisions through litigation. Final decisions on management actions are increasingly made
by courts of law rather than by professional resource managers.
Another reason for difficulty concerns not ordinary citizen's views of the world and
ordinary citizen's value systems, but rather the world view of a profession at the center of
many important natural resource management questions. Such a profession is forestry.
It is my view that forestry has become too associated with the economic criterion for
action, or commodity view of the world. Included is lumbering in the tropical forests of
the world, inattention to regeneration of forests harvested, and, in general, an exaggerated
concern for profit and wealth accumulation at the expense of continuance of forest dependent
human communities and the associated plant animal communities. A commodity view is alien
to a very large number of citizens whose concerns are more often associated with ecologic
or environmental critera for action. Consider the large and rapidly increasing membership
in organizations with an avowedly noncommodity orientation: Sierra Club, Friends of the
Earth, Natural Resource Defense Council, and other groups (Hendee 1984).
For brevity, and to integrate it with other topics discussed later, I call a commodity
view of the world a commodity ontology. Ontology, a branch of philosophy, is the study of
order and structure of reality in a very broad sense.
A commodity ontology was introduced to forestry in North America nearly as early as
forestry itself. Gifford Pinchot set the direction with his "greatest good for the greatest
number in the long run" Clearly, the "greatest number" refers to the greatest number of
humans, not, for example, biotic creatures. Pinchot advocated saving the nation's forests
for many humans rather than letting them benefit a few, such as the lumber barons of the
late 19th and early 20th centuries. John Muir and others in the conservation movement, on
the other hand, advocated going further and saving some forests for many humans as well
as saving fellow living creatures from the ravages of uncontrolled forest cutting.
The anthropocentrism that a commodity ontology breeds permeates the forestry pro-
fession, and to some extent the agency Pinchot first headed, the U.S. Forest Service. The
latter organization has been accused of enshrining the anthropocentric view in the multiple
2 Introduction

use concept. Here again the view is really multiple human use. Multiple human use ignores
the fact that other organisms use the forest as well. Lately, talk has been about the notion
of resources. But, again, the implicit modifier "human" is left out. Trees are resources to
more than just humans. If foresters perpetuate this very narrow view of forests as supplier
of goods and services only to man, the forestry profession is in danger of losing jurisdiction
over much of the world's forest lands.
Foresters must change. But, in what way? Hendee (1984) suggests that the forester
emphasize positive scenarios that cover the full range of the public's environmental interests.
Included are to humanize forestry's image, emphasize economics, emphasize non commodity
forest uses, and show forests as growing places for people. My suggestion is that foresters
develop an ontological repertoire, i.e., several world views that allow them to accommodate
the concerns of humans of many persuasions. Included must be those human groups that
represent the well-being of many noncommodity biotic components of the forest.
There appear to be four fairly distinct ontologies that foresters ought to be familiar wi th
and be able to use effectively. The first two ontologies, commodity and natural object, may
be considered thing based ontologies. The last two may be considered relational.
Beginning with the familiar, there is the commodity ontology. Begun by Pinchot and
perpetuated to this day mainly be economists, a commodity ontology has a forester looking
at the forest through financial glasses. Every thing in the woods must have a financial value
to participate in this ontology, so it might as well not exist if it has no financial value. Aldo
Leopold was among the first to object to this ontology when he became concerned about
"circumlocutions" needed to assign financial values to song birds. Of course, not all is bad
with a commodity ontology. Financial values are useful in selecting efficient cources of action.
It is the tendency to force every natural object population into the financial value mold that
is objectionable.
The second ontology that foresters must make use of might be called natural object.
Here the concern is with, say, the tree, not as a supplier of building materials but as a natural
object or a population of natural objects. Period. No concern is paid to whether or not the
object has financial value. It is sufficient for the object to exist to be of concern. Times
change. Our knowledge of the natural world changes. To focus attention only on natural
object populations with present positive financial values is to adopt a very short range view.
And, according to Pinchot, forestry is supposed to be distinct in having the very long view
of things.
A third ontology that foresters should make more use of is a relational one. The idea
behind it has been stated elsewhere by John Muir:
When we try to pick out anything by itself we find that it is bound fast by a
thousand invisible cords that cannot be broken to everything else in the universe
(quoted from S. Fox (1981): John Muir and His Legacy).
In a similar vein, the 19th century poet Francis Thompson wrote "That thou canst not stir
a flower without troubling a star" ( Mistress of vision. Stanza 22, 1897).
The notion of interconnectedness of natural object populations is an important ingre-
dient in the modern systems approach to understanding nature. Typically more than one
level (trophic or otherwise defined) of natural object populations is involved. Both within
and between level interrelations are important to this view. The pioneering work of Lud-
wig von Bertalanffy in general systems theory did much to advance this view over the past
several decades. Nonetheless, foresters trail other neighboring disciplines in adoption and
implementation of this view.
A fourth ontology that foresters should make use of, the one toward which this book is
directed, is, like the third, a relational one. However, here the relations are dominant and the
natural object populations are subordinate to the relations. Such an ontology is suggested
in two pl"aces:
Introduction 3

All the world's a stage. And all the men and women merely players, They have
their exits and entrances .. (William Shakespeare. As you like it. Act II, Scene 7)
The second is given in the title of G. E. Hutchinson's book: The Ecological Theater and the
Evolutionary Play. I propose to develop in some detail an ontology that can be represented
with a theater-play metaphor.
This view holds that the evolutionary play is running at the ecological theater on a
stage of interactions. The "cast" of object plant and animal populations has changed greatly
over time but the stage and theater arrangement remain time invariant. All the plant and
animal populations are merely players. They each have their exits and entrances. It can be
argued that the theater stage has a very special floorplan patterned after a mathematical
coordinate system. Character population pairs are assigned positions, move on stage along
somewhat fixed trajectories, and exit (exclusion) is not stage-left or stage-right, but through
the floor.
Summarizing, foresters need a broader conception of reality. Consider these four possi-
bilities:
reality
thing based relations based

commodity natural forest objects interactions


object as anchors for as a stage for
interactions forest populations

I II III IV

ontology type

Today there is little dispute that foresters most prevalent reality is thing based and
commodity oriented (timber is trees). I argue that we should not only assume the more
neutral, long-term view associated with a natural object ontology (trees are more than
timber), but we should also accommodate the systemic view of a relations based ontology
(everything is connected to everthing else). A still longer term view is to subordinate
things to the relations between things (all the world's a stage, and all the plants and
animals merely players) so that relations are on an equal footing with things in forestry
decision-making. We should do this for a very fundamental reason: there are no things
without relations between things, and there are no relations (interactions) without things.
Neither came first and neither is more important than the other. In short, I seek to help
foresters obtain a better ontological balance in their work.
One reason the Type IV relational ontology is not widely supported is that it has been
argued prosaically; witness the theater-play metaphor. The argument has not been endowed
with the rigor that comes from using elements of logic and mathematics. In this book, I
make an argument for the importance of a Type IV relational ontology and attempt to
make it rigorous by a) appealing to elements of philosophy, especially the exact philosophy
of Mario Bunge, b) using concepts and techniques from qualitative as well as quantitative
mathematics, and c) using the mathematical coordinate system of Edward Haskell. I attempt
to spell out how a relational ontology can be used in forest ecology and management. My
goal is to show how relations and relations-based forest science can be put on an equal footing
with things and thing based forest science.
To implement a Type IV ontology, representing constructs (concepts, laws, and theories)
are needed. But,- to cover the range of topics needing attention and to do it in a fairly small
space, I am limited in the depth to which I can examine anyone construct. In this situation
I am reminded of Karl Popper's comment:
4 Introduction

Theories are nets cast to catch what we call the world: to rationalize, to explain,
and to master it. We endeavor to make the mesh ever finer and finer (Popper 1959).
My mesh may be too coarse for some people's tastes. The latter may well be the referents
of Heyerdahl's aku aku who observed:
in order to penetrate ever further into their su bjects, the host of specialists narrow
their fields and dig down deeper and deeper till then can't see each other from hole
to hole ... (Heyerdahl 1958).
I will consider my objective met if forest scientists and managers can find new places to dig
for having read this book.
Chapter 1 outlines the apporach that will be taken in later chapters. Following a
brief introduction to the science of forest science, I attempt to synthesize points from the
first four volumes of Mario Bunge's Treatise on Basic Philosophy that are important to
my objectives. Later chapters make use of the distinctions Bunge makes among objects,
constructs, language, ontological categories, and representing constructs. Chapter 2 takes
a general look at interaction the event, interaction the construct, and interaction the term.
Here I point out the importance of what Bunge calls spontaneous and induced change in
a population. Constructs to represent change and interaction are axiomatized, and terms
used to designate interaction related constructs are examined for meaning. Chapter 3 is an
analysis of several forestry constructs that are based on the notions of spon taneous and/or
induced change.
Chapter 4 contains my interpretation of the evolution of Edward Haskell's mathematical
coordinate system. The form of the system I use differs somewhat from what I have used
elsewhere and what Haskell uses. Previously referred to as the Periodic coordinate system,
I feel it is more appropriate that it bear Haskell's name. It should be emphasized that the
interpretation of the coordinate system's evolution and development is my own. Haskell's
explanation can be found in references in Chapter 4. It is also shown here that the coordinate
system has a "wrinkle" in it that allows some interesting trajectories between interactions,
and these provide a basis for evolution of interaction type taken up again in Chapter 9.
The next two chapters (5 and 6) examine the role of Haskell's coordinate system in
analysis. In Chapter 5 I show how it can be used in the search for pattern in the results of
interaction experiments. "Classical" interaction studies due to Gause and Utida, and other
more recent studies by Mattson and Addy, and others are plotted in the coordinate system
and resulting trajectories scrutinized for keys to patterns that help in forming anticipations
about the biological system.
Methods of characterizing mixed population dynamics are applied to mixed forest stand
dynamics in Chapter 6. It is argued that trajectory patterns in Haskell's coordinate system
may provide earlier cues to the need for action (e.g., thinning) than when other coordinate
systems are used. Further, since some silvicultural measures are designed to control inter-
action intensity, it is speculated that there is an opportunity to abstractly geometrize some
aspects of silvicultural systems.
Chapters 7 and 8 treat interactions and Haskell's coordinate system in synthesis. In 7 I
emphasize synthesis in general and show a transformation needed to use Haskell's construct
in a Type IV ontology. Here I outline the evolutionary "stage" arrangement and show how
character pairs are assigned their stage positions. An example using Dindal's expanded
coaction cross tabulation is geometrized and mapped into Haskell's coordinate system. The
idea that advanced sciences have concept systems is proposed in Chapter 8 where I see
what concepts can Join with interaction in forming a system. Two notions are introduced
here: concepts can be assigned valences to show their capacity to participate in concept
systems, and such concepts must have active sites, "places", where other concepts bond to
form a system. Also shown is that in addition to mapping events in Haskell's coordinate
system, o-ne can map concepts, forming concept maps. An example is shown for a small part
of silviculture, forest pathology, and forest entomology.
Introduction 5

Chapters 9 and 10 examine indirect interactions in community structure evolution.


Community structure is taken to be the system of interation types and intensities among
community populations. Structure evolution is scrutinized by axiomatizing a theory of in-
direct interactions and deducing three theorems: self-help, self-annihilation, and helpless.
The latter condition is shown to be the least desirable from the perspective of an individual
population's future well being. In Chapter 10 I extend the methods of Chapter 9 to commu-
nities made up of many populations. The role of a population in a community is suggested
to be of two types: a receiver of effects and a giver of effects. Or, as is sometimes stated,
populations have competitive abilities and competitive influences. I suggest a new exclusion
principle that combines measures of competitive ability and influence. Exclusion is shown
to occur at a particular interaction, a particular place on stage.
Chapters 11, 12, and 13 contain my attempts to apply the theory and methods of pre-
vious chapters to a greatly simplified problem in multiple use forest management. Chapter
11 contains an extension of the multiple use concept into a form more suitable for analysis.
Resources are broken into populations of natural objects and populations of users of natural
object popuations. The user population is then broken into human and other biotic popu-
lations. In all, the multiple use concept is extended to where human is just a special type
of use. In Chapter 12 I review Joseph Petulla's categories of human activities and attitudes
toward nature, what he calls conservation traditions. The economic efficiency tradition is
interpreted in terms of traditional multiple use economics. The ecologic tradition is inter-
preted in light of expressed beliefs that particular community interaction structures do or
will predominate in nature. The beliefs of biocentric tradition principals, especially those
of forester Aldo Leopold, are then reviewed. The final chapter, 13, contains my attempt
to integrate the primary (certainly not all) concerns of each tradition into an integrated
multiple use decision-making framework. The biocentric-ecologic concerns are integrated
around the concepts of extinction and exclusion. The ecologic and (multiple use) economic
traditions are integrated via their information organizing constructs, the phase plane and
the product-product graph. Since extinction is the last exclusion, and exclusion has been
shown to occur at a particular place in the phase plane, I show that exclusion (also ex-
tinction) can be prevented by limiting production to a constrained region of production
possibilities space. The integrated system, together with supporting theory, should permit
non commodity natural object populations to participate in a well established, commodity
based, decision-making framework. A beginning is thus made toward development of an
ethical, yet efficient, approach to forest management.
It is my hope that foresters will seriously consider the proposals made here, improve on
them, and attempt to apply them in practice.
Part I
Setting the Stage, Exactly
1: The Science of
Forest Science

Science is a many-faceted enterprise, hence it may be studied from a number of perspec-


tives. Of course, not all perspectives can be examined simultaneously. Narrowing the view,
as I do here, may give the erroneous impression that perspectives not in focus are unim-
portant or less important. This is not my feeling. Some views of science have received lots
of attention; others, little. In this book I will pay attention to a more logical and semantic
perspective of science.
Bunge (1967a) outlined a structure of the science of science as follows:

psychology of scientists
external sociology of science
history of science
science of science
logic of science
internal methodology of science
philosophy of science

This chapter ignores the upper branch, although a well-rounded introduction to the
science of science should treat both. The upper branch is not excluded because the topics
are uninteresting. Far from it. Knowing them may increase one's conversational prowess in
discussions with colleagues, but in my view they offer little in the way of technical guidelines
for exactness that can be used when faced with the difficult question, "How and where do I
begin?." Stated more positively, scientists featured in history of science books probably got
there because they excelled in an internal aspect of science. (Maslow (1966), Kuhn (1970),
Pelz and Andrews (1966), and Gingerich (1975) provide useful places to begin the study of
the upper branch.)
My focus is on the lower branch with special attention to the logical and semantic per-
spectives. But, I skip the methodological aspects of science - the problem, hypothesis,
observation, test, revision sequence. These ideas are addressed in books by Bunge (1967a,
1967b), Wilson (1952), Beveridge (1957), and Davies (1973). There is, of course, a correspon-
dence between methodological and logical aspects. A declarative sentence may, when viewed
methodologically, be called a hypothesis, whereas when viewed semantically or logically is
called a proposition. I adopt the logical mode throughout this chapter.
The serious student of science should, in my view, begin his or her study with the seman-
tics and logic of science rather than with its methodology. This is because the methodology of
science takes too much for granted. Its lowest level element is the hypothesis. The semantics
and logic of science· provide the means to penetrate the proposition (hypothesis) to reveal
the formal and factual predicates (concepts) from which they are comprised. Figure 1.1
illustrates what I mean.
8 The Science of Forest Science

form conclusions _ _ _ _ _

~ analyze rVidence

analyze data 1 2 gather evidence for tests 3

t
gather data ~-----
form one hypothesis form several hypotheses

~Ipro~ statementl~
C p

proportion n
·:",·, ) C'"t'"' ,) .".';0".'
propottion n (con~ePts~.~\
therY 4 5 therY ~ their degre;

deductions deductions their order

falslfibation criteria falsification crite+ria their reference


j I class
evidence evidence

~ form conclusions------------
Figure 1.1. Schematic representation of the scientific method. Loops 1, 2, and 3 are in the
traditional epistemological mode. For loops 2 and 3 the hypothesis is the most elemental
unit. Loops 4 and 5 are in a more logical mode. What were hypotheses in 2 and 3 are now
called propositions. Loop 5 shows that the most elemental unit is the predicate (concept).

Across the top (loops 1,2,3) are variations of the traditional theme of research methods,
the hypothesis-test-revision sequence. Persons whose "science" is typified by loop 1 have
inflated views of the value of statistical methods in research. Loop 2, wherein one hypothesis
is formed at an early stage, is required in the most elemental form of scientific method. Others
(Chamberlin 1897, Platt 1964) have advocated forming several hypotheses before a testing
sequence is developed (loop 3). Loops 2 and 3 represent scientific research in its traditional
epistemological mode. Loops 4 and 5 represent an approach to scientific research rooted
in logic (both propositional and first order predicate). This approach is more fundamental
because what is a hypothesis in loops 2 and 3 is now viewed as a proposition. Using methods
from predicate logic and semantics I can penetrate the proposition and analyze its factual
and formal predicates. This, in turn, allows me to clarify, strengthen, and if needed, invent
new concepts purposefully. Methodology (loops 1,2,3) alone will not provide the means to
transform an entire domain of inquiry from protoscience to science proper, protoscience being
data gathering without seeking or finding reproducible patterns rooted in constructs. For a
field of inquiry to rate "science proper" it must, at a minimum, possess

a~ a concept set containing,


i. concepts shared with other fields of inquiry and
ii. concepts unique to the field
b~ scientific laws (corroborated hypotheses), and
c. some scientific theories.

Much of the structure of this chapter, and its reason for being included, can be sum-
marized in a schematic of the process of measurement. What is measured? Bunge (1967b)
suggests it is:
The Science of Forest Science 9

a a a a a
particular numerical quantitative property concrete
value variable concept of system
of designating representing

That is, the seemingly simple process of reading a value from a measuring device is
(or should be) preceded by some deep thought. In this chapter, I sketch the relations
of designation and representation. I also pay particular attention throughout to concepts
and ontological questions, to what I consider is real, the categories we use for classifying
reality, and the properties of real objects by which we "know" them. Forestry and ecology
need something other than common sense ontological categories (Hull 1980). Poorly formed
categories unnecessarily complicate scientific research, and in many cases, significantly reduce
its effectiveness. The emphasis on concepts and ontology here reflects my bias that forestry
suffers not so much from lack of methods for evaluating claims to knowledge as from an
inconsistency in forming fruitful knowledge claims.
Forest science differs from science in general simply because it is focused on a fairly
restricted number and kind of objects commonly found in, or associated with, forests. One
would not expect the psychology of forest scientists, the sociology of forest science, etc. to
be significantly different than when the focus is on different systems. Because of the long
life span of trees, some of the methods are probably different than in other sciences, but the
overall process of science is no doubt very similar among sciences. This has led Bunge (1973)
to argue that the sciences are unified methodologically.
I examine the place of interaction theory in forest ecology and management. Doing so
will require that I treat interaction the event and interaction the concept. To treat them and
communicate the results here, I am forced to make use of interaction the term. A detailed
examination of these three items, the event, the concept, and the term, is given in Chapter 2.
A thorough examination requires the use of tools beyond what are normally used in science.
The purpose of this chapter is to begin the development of tools to scrutinize these items
and their interrelations.

A. Language, Construct, Object


First, a survey of the categories - object, construct, and language - is needed, and then
an examination of their interrelations. Of course, in one short chapter I cannot examine such
broad categories in depth. Rather, I propose to do three things:
a. inventory each category and point out those items in each that are of special concern,
b. examine selected within-category relations that I feel are important to forest science for
each category, and
c. examine four important between-category relations.
Some of the general categories, as well as relations among categories, are studied in the
various special sciences, whereas other categories and relations normally come in the domain
of philosophy. Each item discussed is related to chapters or sections later in the book where
it is applied, or shown to have been applied, elsewhere. In a sense this chapter functions
somewhat as an index to the book as a whole. It is needed because I feel the average reader
interested in interactions may benefit little from the subsequent chapters without some of
the perspective I attempt to develop here.
A "modern" description of what I propose here might be that it examines the scientist's
language-construct-concrete object "system." The quotation marks around system indicate
that I am using it· in a somewhat novel sense. For example, Bunge (1974a) specifies that
systems must be categorically pure, either all natural objects or all constructs. Thus, my
use of the general notion of system is not in the usual or recommended sense, yet it seems
10 The Science of Forest Science

an appropriate term for the material as it is presented. In current systems thinking, wherein
a system is specified by the 3-tuple

< composition, environment, structure>,

language, constructs, and concrete objects form the composition. Symbolically I show the
composition as three nodes of a graph (Figure 1.2). The system structure is indicated by the
arcs connecting the nodes. For example, the arc connecting object with itself indicates there
are object-object relations that are important in science. The study of natural objects and
natural object interrelations is taken by some to be science. The other arrows in Figure 1.2
indicate that my view is that science is more than just natural objects and their interrelations.

Figure 1.2. The "system" within which science operates.

The numbers in Figure 1.2 indicate the order in which the items will be discussed.
Items 1, 2, and 3 expand the specifications of the system's composition. Each takes the
form of an inventory. I stretch this somewhat in the construct category wherein the notion
of predicate is discussed in some detail. Items 4, 5, and 6 treat intracomponent relations.
For example, item 5 examines the various relations between elements inventoried in item
2, the construct component. Items 7, 8, 9, and 10 deal with the kinds of relations that
exist between components. Much of the material in this chapter has been gleaned from the
various writings of Mario Bunge, especially from the first four volumes of his 'I}eatise on
Basic Philosophy. The exact nature of the components is not as general as Bunge's. Some
of the relations are not treated in the 'I}eatise, and the order of treatment is different.
The Science of Forest Science 11

B. Category Inventories
1. Language.
An important element forming the scientist's "system" is language. Language categories
of interest are two: natural (e.g., English) and artificial (e.g., mathematics). Whether natural
or artificial, languages must have the following:
a. a vocabulary, or set of allowed symbols,
b. a syntax, or grammar rules, and
c. a set of semantic rules.
If artificial, the language will also have a logic specified. The permissible sentences and
formulas in the language are those strings of symbols contained in the vocabulary that have
been combined according to the language's syntax rules.
Although an obvious place to begin the quest for scientific truth, natural languages do
not have a structure suitable for the kind of precision and scrutiny typical of good science.
Mathematics is often employed instead of prose to gain a greater degree of precision and
clarity. On the other hand, many ideas fundamental to science and the scientific approach
are rooted in natural language, for example, predicate, hence they should be dealt with in
this treatment of the language category.
Listed in Table 1.1 are the two types of language under consideration here. Each can
be broken down in a number of ways. As noted above, each needs at least a vocabulary,
syntax and semantics. At this point my interest is in natural language syntax, so it is listed
in expanded form. In each division under syntax the items of most concern are identified
with an asterisk. Thus, in the part of speech category I am more interested in nouns, verbs
and adjectives than in, say, interjections and prepositions. Also, in parts of sentences I
am primarily interested in subject, predicate, and phrase. Sentences of declarative and
interrogative kind are of most interest in this category. When discussing arrows 7, 8, 9, and
10 (Figure 1.2), I will address the relation of semantics to my objective.

Table 1.1. Simplified inventory of the language category.


vocabulary vocabulary
syntax semantics
semantics syntax
logic
Sentences Parts of sentences Parts of speech
-word groups -words
-words

Structure subject * noun *


simple predicate * number
compound direct object case
complex modifier gender

connectives person
Uses independent elements pronoun

declarative * verb *
imperative adjective *
interrogative* adverb
. exclamatory phrase * preposition
clause conjunction
12 The Science of Forest Science

There is an interrelation between the language and object categories. If the focus is
primarily on things and qualities of things, that is, exhibits a thing ontological bias, then
a more frequent use of nouns and adjectives would be expected than if the focus was on
processes and change, reflecting a process ontological bias. In the latter case, one might
expect more need for verbs and adverbs (Rescher 1962).

2. Construct.
The second category, construct, is, following Bunge (1974a), thought to have three
major subcategories: concept, proposition, and context. The latter subcategory may be
further divided into theory and non-theory, or what Bunge calls context. Here I propose to
do a simple inventory of each of the three subcategories.
The frequent occurrence of 'concept' and 'theory,' in recent literature, suggests wide
concern about and use of these elements. For example:
1. Education research scientists are finding that it is important that students of various
abilities and at all levels of formal education be given explicit instruction about concepts
in general as well as concepts specific to factual knowledge (Ausubel 1966; Stewart et
aI., 1979; Klausmeir and Harris 1966).
2. Elementary and se<;ondary science education seems to be placing greater emphasis on
instruction centered around concepts, e.g., Humphreys 1978; Novak 1966, 1977.
3. In 1977 the prestigious journal The American Naturalist changed its subtitle from "De-
voted to the Advancement and Correlation of the Biological Sciences" to "Devoted to
the Conceptual Unification of the Biological Sciences."
4. With increasing frequency there are claims in the forestry literature to the invention of
new or expanded concepts, e.g., Boyce and Cost 1978; McClure, Cost and Knight 1979;
Shigo 1979.
5. Numerous books, especially about ecology, claim to be about concepts, unifying con-
cepts, and theories, e.g., Kormondy 1969; May 1976; Christiansen and Fenchel 1977.
These facts are all the more interesting for forestry because in 1968 Bentley questioned the
traditional natural object-event orientation of forestry education, suggesting that more em-
phasis be given to "packaging" information around concepts (Bentley 1968). His questioning
drew not a single response in the literature or in personal communication (Bentley 1979).
The frequent use of terms 'concept' and 'theory' is, upon deeper examination, somewhat
misleading. For example, although claiming to be about concepts of ecology, Kormondy's
book is actually about natural objects and events referred to by the concepts or the proper-
ties of natural objects represented by the concepts (Kormondy 1969). On similar grounds,
the book Unifying Concepts in Ecology does not contain references to the literature which
would explain how "unifying" concepts differ from other concepts (van Dobben and Lowe-
McConnell 1975).
An important objective in this discussion of concept, proposition, and theory is to em-
phasize that they are the object of study by logicians and philosophers, and they have an
existence of their own apart from forestry, ecology or any other factual body of knowledge.
Ample evidence exists that forest scientists share little awareness of anything like a "taxon-
omy" of concepts or an "ecology" of concepts. This puts forest scientists in a primitive state
of understanding constructs in general, making purposeful development and use of constructs
unlikely. The remedy is, of course, a direct examination. I begin with taxonomy, delaying
until the construct-construct section a discussion of between subcategory relations.
a. Concept. Concept, the first construct I will examine, is far from a simple notion, being
one of the oldest terms in the philosophical vocabulary, and one of its most equivocal
(Edwards 1967). It has been defined as the fundamental unit of thought, and as a
fiction existent in a human mind for only as long as the latter exists. The implication
of statements such as "having a concept" and "forming or acquiring concepts" have
lead to long and exacting arguments among philosophers and scientists (Norreklit 1973;
The Science of Forest Science 13

Hospers 1967; Edwards 1967). I do not question statements of this kind, rather, I take
the existence and acquisition of concepts for granted and examine their nature, use, and
general role in science.
A taxonomy of concepts suggested by Bunge (1967a), with forestry examples, is as
follows:

name of kind examples


individual "General Sherman sequoia"; "x"
class "hardwood"; "softwood"
relation "is taller than"; "outgrows"
quantitative "basal area"; "total height"

What is regarded as an individual concept cannot be stated precisely unless the level
at which one is working has been spelled out. At one level "General Sherman sequoia" is
an individual concept, and at the same time an individual object. At another level, the
indeterminate individual concept "x" may be a proposition such as "red pine outgrows jack
pine on this site." Typically, a definite individual concept is designated by a proper name.
Symbolically, logicians/philosophers designate it by a lower case letter from the first part of
the alphabet. If it is an indefinite individual, it will be designated by a lowercase letter from
the latter part of the alphabet.
Class concepts apply to groups of individuals. Membership in the class is based on
the qualities possessed by individuals. Thus "hardwood" is a class concept in which the
membership is limited to trees, the stemwood of which is greater than some specified specific
gravity. Concepts of this kind play an important role in the business of classification, typical
of the early stages of science.
Relation concepts apply to relations among objects of some kind and are particularly
important in all formal and factual sciences. The simple examples given above, "is taller
than" and "outgrows," may be applied to things that possess propertjes of length (vertical
extent) and life, respectively. Because of the importance of relation concepts to factual
science in general and to the purposes of this book, I will return to them shortly.
Relations that pair elements of two sets in a unique way are called functions (Bunge
1967a). Quantitative concepts take their form as functions, i.e., one-to-one or many-to-one
mappings of elements in the domain into the range (normally the set of real numbers). For
example, basal area is a function B that maps the set of stem thicknesses (s.t.) into the set
of positive real numbers, or b.a. = B(s.t.) when expressed in mathematical equation form.
Table 1.2 gives a suggested logical structure of concepts.

Table 1.2. Logical structure of concepts (adapted from Bunge 1967a).

kind of concept logical structure


individual individual constants (a, b, c, ... ) or individual variables
(. .. x,y,z)
class predicates with a single ob~ect variable either constant
or variable, e.g., B(a), B(y), etc.
relation predicates with more than one object variable, e.g.
T(x,y), G(a,b,c)' etc.
quantitative numerical functors (for our purposes, functions)

A new term, predicate, has been introduced in Table 1.2. It is an important idea, hence
deserves further study. I lise it extensively later in this chapter and the book.
14 The Science of Forest Science

Both subject and predicate find use in four different areas of philosophy: grammar (see
the section on language), logic (the major use at the moment), epistemology and metaphysics
(Edwards 1967). Use in epistemology, and metaphysics can be safely ignored for my purposes.
Examining subject and predicate use in logic, the context is the simple declarative statement
that can be partitioned as follows:

Subject Copula Predicate


an object of typically some form a quality of
some sort of the verb "to be" the object

Example: Hickory IS a hardwood.

This traditional form, due to Aristotle, has been altered into the so-called modern form used
by logicians (Orenstein 1977), that is,

Subject "Modern" predicate


an object of some form of verb
some sort 'to be' applied to
a quality of the object

Example: Hickory is a hardwood.

The rule normally followed is to designate the modern predicate by a capital letter, e.g.,
H (for hardwood). This is called the predicate letter. The object to which the quality is
directed by the predicate is called the object variable and is usually designated by a lowercase
letter, e.g., c for Carya. Thus, H(c), cH or Hc all assert that hickory c has the quality
designated by the predicate letter, H. When the quality designated by the predicate letter
is a relation, the number of object arguments increases. The number of object arguments
determines the predicate degree. Bunge (1963) presents an analysis of other qualities of
predicates besides their degree, i.e., their order, metrical character and distributivity. The
stature of the predicate concept is such that it is usually used in place of class and relation
concepts. Thus, instead of saying that "outgrows" is a relation concept with binary predicate
structure, the typical statement is that "outgrows" is a binary predicate. Further analysis
of the qualities of predicates is deferred until the discussion of construct-construct relations.
b. Proposition. Proposition is the second construct I will consider. Although early stages
of investigating a problem may be replete with the questions: What is the character of?
What if? and Why?; the resolution of problems involves statements (sentences) in the
declarative mode. A proposition is the meaning of a declarative statement.
Simple categorical propositions are expressed with declarative sentences that express a
judgment, hence have a definite truth value true or false according to whether the subject
has the qualities attributed to it by the predicate. Thus in 'hickory is a hardwood,' the truth
value rests on specifying a criterion for membership in the class of hardwoods and whether or
not hickory meets the criterion. Classification of simple propositions has centered on their
quality and quantity. Quality refers to the inclusion of a subject object in the predicate
class (affirmed or denied). Quantity refers to the numerosity of subject objects included in
the predicate class. Quantity criteria are normally denoted singular or general (universal or
particular). Quality criteria are affirmative and negative. These criteria may be assembled
as follows (see, for example, Copi 1978):
The Science of Forest Science 15

example
singular
affirmative Cl is an H
negative Cl is not an H
general
universal affirmative all care H
particular affirmative C2 is an H
universal negative no care H
particular negative C3 is not an H

A more powerful approach to analyzing propositions focuses on the predicates contained


therein. The general nonspecific statement of a proposition is made more precise by a process
of quantifying the object or predicate variables, using existential or universal quantifiers.
Bunge (1967a) offers the following list of quantifiers, symbols for each, and an interpretation:

symbol read
(::3 x) There is at least one x such that ...
(::3 x)n There are exactly n x such tha· ..
(::3 x)U There are x in U such that ...
(x) Every x is such that ...
(X)"'EU Every x E U is such that ....

He then goes on to present the following table of elementary quantified proposition kinds
(See Bunge 1967a):

singular H(c) with Hand c constant


individual existential (::3 xhH(x)
proposition indefinite (::3 x)H(x)
general particular (::3 x)nH(x) with n = 1
universal bounded (x)"'EuH(x)
unbounded (x)H(x)

In factual science interest lies in more complex propositions than the ones like "hickory
is a hardwood." Consider, for example, the proposition expressed as follows:

XjY:: WjZ.

Verbally, this is read' X is to Y as W is to Z.' It may be converted into an equation


having identical meaning by solving for X, i.e.,

X = YWjZ

Technically, this is a propositional function or open sentence because to this point the
vocabulary jsemantic aspects of the mathematic language have been ignored in favor of its
syntax. Before a propositional function can be made a proposition (something with a definite
truth value), it is necessary to interpret the symbols. Any number of interpretations can
be given to the above set of symbols and have an affirmative truth value for the resulting
proposition (e.g., !he law of sines, the law of cosines, etc.). The particular interpretation I
make deals with the distribution of increment in a forest stand. Interpretation or specifica-
tion involves two steps: "predicate variable specification and object variable specification" ,
(Bunge 1967a).
16 The Science of Forest Science

Looking first at predicate variable specification we see that X and Y must be of the
same units, and Wand Z also. Thus, both 'X( )' and 'Y( )' designate the concept "standing
crop growth," and both 'W()' and 'Z()' designate the concept "standing crop." Considering
object variable specification,
'YU)' designates annual standing crop growth per unit area in stand, j,
'X(i)' designates annual standing crop growth in the ith tree, where i E j,
'W (i)' designates standing crop of the ith tree, and
'Z(j)' designates standing crop of the j stand.
Looking back to the original form,

X(i)/YU) :: W(i)/ZUL and

X(i) = YU)(W(i)/ZU)).
Of course, I do not divide the entire concept "standing crop" in the ith tree, W (i), by
the concept "standing crop" in the jth stand, Z(j); I only divide their respective numerical
values.
c. Context. A final kind of construct is what Bunge (1974a) calls a context. Context may
be open or closed. Open contexts are sets of propositions or mathematical equations
such that one or more of the propositions has no logical relation to any of the others.
Bunge symbolizes a context as follows:

C =< S,P, D >

where
5 is a set of statements or propositions
P is the set of predicates contained in 5, and
D is the domain or reference class of the context.
A theory is, according to Bunge, a special kind of context, one with a set of propositions
related by the logical relation of entailment.
Symbolically a theory is
T =< S,P,D,f->
where 5, P, and D are the same as for context and f- means "logical consequence."
As with 'concept,' 'theory' is not at all a clear notion in philosophy and science. The
term theory is sometimes used to signify the opposite of practice. Others use 'theory' in
an epistemological context when it seems 'hypothesis' is more appropriate. Still others,
e.g., Bunge (1974a)' consider theory in a logic context as a set of propositions related by
entailment. Further views about 'theory,' the relationship between law and theory, as well as
many references, can be found in the Encyclopedia of Philosophy (Edwards 1967). A recent
effort at clarification was made by Rapaport (1978).
So, for a bunch of formulas to qualify as a theory, each must be logically related to one
or more other formulas in the bunch. Theories have no place for stray propositions. Every
statement must either be
a) a premise (axiom or definition) or
b) a consequence of premises of the theory.
An example may help at this point. Consider a theory of population growth. One
premise is that change in population numbers is governed by the equation:

dN/dt = (b - d)((K - N)/ K).


Another premise is that the birth rate, b, and death rate, d, are constant. A third premise
is that change is continuous over time (i.e., not seasonal). There may be other premises
The Science of Forest Science 17

as well. One important consequence of the theory is that the population level has a time
history given by N{t). The consequence of the theory, the N{t) trajectory, is entailed by the
three premises.
My main interest is theory, not context. Sometimes, however, a heap of formulas may
take on the appearance of a theory because of their number or complexity but still do
not meet the relatively simple requirement spelled out above (e.g., Odum 1969). Thus,
complexity has little bearing on whether or not a set of propositions constitutes a theory.
Theories may be classified in at least six ways. Listed below are brief comments describing
the classifications:
1. phenomenological - black boxish vs. representational - white (translucent),
2. axiomatic - of the form: definitions + axioms + deduction => theorems + corollar-
ies vs. non axiomatic - no identification of presuppositions, consequents, or tests of
consistency in the deductions,
3. normative - describes ideal behavior vs. descriptive - describes actual behavior,
4. general - the reference class is very broad vs. specific - the reference class is limited
to a particular aspect of a real system,
5. formal - the reference class is a set of formal objects and the predicates are nonfactual
vs. factual - the predicates are factual,
6. abstract - the reference class is made up of arbitrary elements (abstract theories are
typical of modern logic and mathematics.) vs. interpreted - the reference class is made
up of specific items either natural or formal objects.
In Chapter 9, I present a phenomenological, axiomatic, normative, general, factual, inter-
preted theory of indirect interactions.
This is enough of theories, contexts, and other constructs. I move to another challenging
category-object.

3. Object.
The third element forming the scientist's "system" is object. I constrain 'object' to
denote real, material objects, not conceptual ones. Studying what can and should be con-
sidered "real" and developing means to categorize reality is the business of ontology. Thus,
the object node in Figure 1.2 is where ontology, a branch of philosophy, is centered.
Foresters, and ecologists as well, have rarely given explicit recognition to the important
role of ontology and ontological categories in their work. As a consequence, these groups
are generally unaware of the variety of possible ontologies. Of course, every scientist has an
ontology by which to work, but most have been content to approach it at a common sense and
subconscious level, improvising as new categories are needed. The difficult questions, Hull
(1980) argues, will require something more formal than common sense categories. Questions
about the unit of natural selection, the organization of nature (levels, hierarchy, etc.), and
the characterization of natural systems (using systems science methods) are of this kind.
In the Introduction I listed four types of possible ontologies for foresters and ecologists.
A commodity ontology dominated forester's views of forests for many years, but in recent
decades forest research has progressed toward a natural object categorizing of reality. One
need only examine the forestry scientific literature to see the shift in emphasis. I suspect
that foresters have also been influenced by their neighbors in plant ecology who have never
had a commodity orientation.
In passing, it is important to stop for a moment and mention that much of the contro-
versy about values and the valuation process has suffered from too little attention to ontology.
Once reality is categorized in an agreed-upon way, valuation may proceed, I think, in a fairly
straightforward manner. But if no agreement is made on how to view and categorize reality,
there is little hope for agreement on the results of valuation. In the last chapter I attempt
to integrate the three ontologies, commodity, natural object, and relational (Type IV in the
Introduction).
18 The Science of Forest Science

Bunge analyzes eleven ontological (object) categories: substance, form (property), thing,
possibility, change, spacetime, system, chemism, life, mind, and society (Bunge 1977, 1979).
Table 1.3 contains an inventory of selected items in this category along with specification of
subcategories and brief descriptions where appropriate. Not treated in Table 1.3 are change,
spacetime, chemism, life, mind and society because they are not considered particularly
relevant to this book. Change is treated in Chapter 2 where interaction the event, interaction
the construct, and interaction the term are discussed. By giving interaction, a subcategory
of change, its own chapter, I hope to impress upon the reader its great importance as an
ontological category. The "elevation" of interaction to a standing near thing is a crucial part
of a Type IV ontology. Repeating, there is no thing that does not interact with some other
thing, and there are no interactions without things. Both are absolutely essential.
In Chapter 3 I examine several forestry concepts that reflect a possibilist ontology.
The possibility involved is for a thing and its properties to change under a special kind
of interaction. Further discussion must await Chapter 3. This completes my inventory of
the categories of language, construct, and object. I now proceed to this chapter's second
objective, an examination of selected within-category relations.
The Science of Forest Science 19

Table 1.3. Categories of material objects of primary concern (Abstracted from Bunge 1977,
1979).
Category Subcategory In terpretation/comment
Substantial An entity that has only one basic property that of being
individual capable of associating with at least one other substantial
individual.
Property of Basic or The property is not reducible to any other properties.
a substantial Derivative The property is the conjunction of two properties.
individual
Note: Properties of substantial individuals should not be confused with attributes of individ-
uals. The latter are concepts. Properties are what attributes represent (see item 10-
Represen tation) .
In trinsic or The property applies to individuals or in the set of concern.
Mutual The property is shared by more than one individual in the
set of concern.
Scope The scope of a property of substantial individuals is the
collection of entities possessing the property.
Thing Roughly speaking, a thing is an entity endowed with all its
properties, including the property of changing in various
ways.
State of a The ordered n-tuple (vector) of all its individual, intrinsic
thing properties.
State space The collection of all possible states of an entity. A thing is
of a thing a substantial individual together with its state space.
Note: The three levels of concrete biological things important to my objective here are the
individual organism, the aggregate of individuals of the same kind, and aggregates of
individuals of different kinds, or .mixed populations.
System A set of interrelated items.
Concrete A concrete system is an aggregate of items that are things
thing and are interrelated in such a way that the relations make
some difference to the components.
Note: To specify a concrete system one must spell out three items:
i) composition of the system, i.e., the things that are strongly related,
ii) environment of the system, i.e., the things not related, or so weakly related, to
strongly related items as to not warrant inclusion in the composition,
iii) structure of the system, i.e., the network of inter-relations between things in the
composition.
----------------------------------------------------------

c. Within-Category Relations

4. Language-language.
The scope of this discussion is limited to sign-sign relations instead of the whole gamut
implied by language-language. (For a brief forestry example of a broader aspect of language-
20 The Science of Forest Science

language relations, see "Writing about writing (or the theory of types)" (Fontanilla 1978).)
My concern is with the assembly of signs into sign systems and with obstacles to the speci-
fication of sign meaning.
Quine states, "The less a science is advanced, the more its terminology tends to rest
on an uncritical assumption of mutual understanding. With increasing rigor this basis is
replaced piecemeal by the introduction of definitions" (Quine 1976). In order for forest
science to advance, it is necessary that it have a precise and adequate terminology.
Because of the nature of forest science, heavily dependent on other sciences, it is also
necessary that each supporting science have a precise terminology. For example, forest
ecology is dependent on general, plant, and animal ecology as well as soil science, plant
physiology, meterology, etc. What is the condition of the terminology for these sciences?
Can forest science ever be more advanced than its least advanced supporting discipline?
These questions and others naturally come to mind when discussing the adequacy of a
science's terminology.
Although nearly everyone agrees on the need for precise terms (see most any newer
textbook on ecology and references at the end of Chapter 2), there is little agreement on
the means to attaining that end. For example, Bunge (1974) lists ten different views on
meaning. A view receiving wide use in forestry and ecology is operationism; i.e., the view
that meaning is operation (computation or measurement) (see e.g., Boyce and Cost 1978,
Innis 1972). Bunge has claimed, however, that operationism leads to a beheading of science,
to theory-less science. Further, it is asserted that definition is a sign-sign correspondence, so
something that is a sign-operation (object) correspondence is, in fact, not a definition at all
(Bunge 1967a).
The goal of term or word meaning precision and definiteness is often obstructed by two
word attributes: ambiguity and vagueness. But they are not of equal difficulty to surmount.
"Ambiguity is ambivalent: on the one hand it enables us to economize signs ... [keeping down
vocabulary size] ... ; but ... [it] ... shelters confusion. Fortunately it can always be removed in
part or in full with the adjunction of further signs" (Bunge 1967a). (See also Quine 1960.)
Vagueness is another matter. It is a word disease of conceptual origin, not of linguistic origin,
the case with ambiguity. It follows that vague terms like 'entity' are explicated only after
analysis of the concept they designate. Examples of ambiguous and vague terms related to
biotic interactions are discussed in Chapter 2.
In the less advanced sciences perhaps the dominant way to combat indefinite meaning
of terms is to consult a dictionary or glossary of terms specific to that science (e.g., Ford-
Robertson 1971). This practice may be dangerous, however, because context-free definitions
often lack the precision necessary for scientific discourse. For example, the Terminology of
Forest Science, Technology, Practice, and Products (Ford-Robertson 1971) lists one definition
of 'density' as having a meaning " ... roughly the same as .. " 'stocking.' In Chapter 3 I show
that density and stocking can be very different because they are about different things (have
different reference classes).
Complicating the problem further is the wide variety of types of definition that can
be made. For example, Copi (1978) lists five: (a) stipulative, (b) lexical, (c) precising,
(d) theoretical, and (e) persuasive. Borsodi (1967) carries this to an extreme in his book
The Definition of Definition where he lists 27 "different" kinds of definition. Such excesses
would seem to lead scientists to the mistaken conclusion that all meaning questions can be
solved within language, i.e., with no appeal to constructs (especially concepts and theories)
or objects. Limiting ourselves to sign-sign relations limits our meaning-giving enterprise to
definitions. Albeit useful, and in some cases necessary, it appears that foresters have placed
too much reliance on definition as a meaning specification device. Some think meaning is
definition. It is but one among several, and probably not the most important in science.
Chapter 2 reviews several attempts to render a precise meaning to 'interaction' and terms
related to 'interaction.'
The Science of Forest Science 21

5. Construct-Construct.
Another important relation in Figure 1.2 is construct-construct. Two classes of this re-
lation can be addressed: within-subcategory (e.g., between concept and concept) or between-
subcategory (e.g., between concept and proposition). Our concern is with the latter class
under the relation 'is built from.' The first in each set of relata are concept and proposition,
and in the second, proposition and theory. Propositions are built from concepts, particularly
predicates. Simple propositions can be built from factual predicates alone. More complex,
hence more interesting, propositions require both factual and formal predicates (e.g., and,
or, not, etc.). Contexts are constructed from propositions, either closed under entailment or
open. Hereafter, I focus on the concept-proposition end of the relata spectrum.
From the point of view of structure, a predicate's place in a declarative statement is
one of association with an object and copula, or, in the modern view, just the object. Bunge
(1974a) explains a different view of predicates. I consider it a functional view because it helps
answer the question: What does a predicate do? In order to understand this view we must
be able to distinguish between a proposition and a propositional function. A proposition,
you will recall, is the meaning of a declarative statement that has a definite truth value.
Thus, 6 + 3 > 8 is a true proposition. On the other hand, x + 3 > 8 is a propositional
function, an "incomplete" proposition because it is missing a specific value for one of the
object variables. The view of predicates is that they are propositional functions. Specifically,
they map (in a mathematical sense) objects into statements about objects. The predicate
domain is objects, the predicate range is statements, or for our purposes, propositions, that
contain the predicate. In symbolic terms this may be expressed

H : c --. 5,

where
H is the predicate "hardwood,"
c is an arbitrary member of a set of trees belonging to the genus Cayra
"--." symbolizes a mapping (one to one or many to one) from the domain to the range,
and
5 is a set of statements containing H.
A binary predicate maps elements from the Cartesian product of domain objects into
statements containing the predicate. For example, a forestry use of the concept "outgrows"
is a binary predicate, symbolized

DC : A x B --. 5,
where
DC is the predicate "outgrows,"
A is an arbitrary set of trees of kind 1, e.g., A = {a,b,c,d} ,
B is an arbitrary set of trees of kind 2, e.g., B = {e,/,g,h} ,
A x B is the Cartesian product of sets A and B, i.e., {a,e},{a,J},{a,g}, ... ,{d,h}, and
5 is the set of statements containing DC.
The number of sets in the predicate's domain determines its degree. A third degree
predicate, e.g., "between", would be symbolized

B : A x C x E --. 5,

with interpretations here the same as above for the second degree predicate.
Another important dimension of predicates is called "order" Predicate order refers
to the nature of the domain elements. I shall say that when the domain elements are
individual natural objects the predicate is first order. When the domain elements are classes
of individuals, the predicate will be said to be second order. When the domain elements are
22 The Science of Forest Science

relations between classes, the predicate is third order, and if the domain elements are relations
between relations, the predicate is fourth order. Predicate order seems directly related to
level of abstraction (Quine 1972). In Chapter 4, I suggest that Haskellian coordinates have
fourth-order predicate structure.

6. Object-Object.
In the treatment of objects I excluded conceptual objects and limited myself to the
following general categories of material objects: substantial individual, thing, and system
(Bunge 1977, 1979). I now look briefly at object-object relations. Again both between-
subcategory relations and within-subcategory relations may be considered. I begin by ex-
amining relations between subcategories, i.e., substantial individual property of a substan-
tial individual; (substantial individual and property of a substantial individual)-thing; and
thing-system. These relations are briefly summarized in Table 1.4.

Table 1.4. Relations between subcategories of material objects (Abstracted from Bunge 1977,
1979).
OlJect category Object category
Comment
A B
substantial property of a The only function of substantial individual is to support
individual substantial the concept of property, i.e., to provide the domain of the
individual predicate "property".
substantial thing A thing is a substantial individual together with the n-
individual - tuple of values that its properties may take, i.e., its state
property of a space.
substantial
individual
thing system A system is composed of interrelated things. Things
strongly related are grouped together in the system com-
posltlOn. Things less strongly related to those things
strongly related form the environment. Relations be-
tween things is the structure.

Relations within subcategories of objects are summarized briefly in Table 1.5. Juxtapo-
sition and superposition of things will be used in Chapter 3 when considering forest growth
models, interaction is treated in Chapter 2, and hierarchy and levels are used in Chapter 8.
Note that actions and connections have been defined for things, not for properties of
things. The latter can be interdependent but not interacting.

This completes a brief highlighting of important within-category relations. They form


a part of the structure of a scientist's "system." I now proceed to the other part of "system"
structure, to between-category relations.
The Science of Forest Science 23

Table 1.5. Relations within subcategories of material objects (Abstracted from Bunge 1977,
1979).
Category of object- Name given
Comment/Explanation
object relation to relation
1. substantial individual- juxtaposition The physical sum of individual objects. Jux-
substantial individual + taposed objects retain their individual iden-
tity in the resultant.
superposition The physical product of individual objects.
:i: Superposed objects lose their identity in the
resultant.
2. property of a substantial precedence Property 1 of a substantial individual pre-
individual - property of cedes property 2 of the same individual if it is
a substantial individual necessary for, or generates, property 2.
3. thing-thing juxtaposition The physical addition of things.
+
part-whole The thing x is a part of thing y if x+y = y.
superposition The physical product of things.
x
acts on Thing x acts on thing y if the state space tra-
jectory for y is different when x is present than
when it is absent.
interaction Two different things x and y interact if each
acts upon the other.
4. system-system subsystem X is a subsystem of system y if and only if x is
a system and the composition of x is contained
in the composition of y, the environment of x
is contained in the environment of y, and the
structure of x is contained in the structure of
y.
hierarchy or A set of component systems ordered by the
system of dominance relation.
nested systems

level of systems A set of component things comprised in sys-


organization terns that are ordered by the precedence rela-
tion.

D. Between-Category Relations
I propose to address between-category relations indicated by arrows 7,8,9, and 10 in
Figure 1.2. These relations are among the most philosophized of the ideas discussed in this
chapter.
Contrary to my expectation, considerable disagreement exists among philosophers on
the nature of the relations and on a proper name for each. How bad the situation has
become is indicated by Geach (1968) who suggests that 'denote' (arrow 8 is denotation)
is "so battered and defaced a coin ... [that it should beJ ... withdrawn from the philosophical
currency .... " Discevering this situation leaves me ambivalent. Seeing another group wrestle
with the confusion resulting from words being used in different senses by the same author at
different times or by different authors at the same time (Lyons 1977) is somewhat comforting.
24 The Science of Forest Science

When this happens among philosophers, those self-proclaimed guardians of clear thought
and expression, at least one of whom has made much ado about similar problems among
scientists (Berlinski 1976)' an extra ounce of comfort ensues. On the other hand, the comfort
provided by someone else's confusion turns to frustration when attempting to catch a glimpse
of what all the fuss is about. Notwithstanding the disagreements and difficulties within
philosophy, there remains the need for forest scientists to know about the interrelations
between natural objects, constructs, and language. A sample reason for needing to know
about each interrelation follows:
1. The representation relation (arrow 10 in Figure 1.2) includes as a special case the de-
velopment of conceptual models of natural objects, and modeling is one of the most
frequent activities in many branches of forest research as well as practice.
2. Anyone who has invented, concocted or dreamt-up a factual theory, proposition, or
predicate should be able to precisely answer the question: What is your construct
about? Further, any claims to "new" or "expanded" constructs should be accompanied
by exact statements of the basis for such claims. The reference relation (arrow 9 in
Figure 1.2) establishes what constructs are about.
Although there is ample justification for learning about representation and reference,
it is more easily argued that the other two relations, denotation (arrow 8) and designation
(arrow 7), are totally the concern of philosophers and language analysts. To show why this
is not true, I examine designation and denotation.

7. Designation.
There is no particular magic in the name 'designation' for the relation under considera-
tion. Bunge (1974) suggests that 'stands for' or 'expresses' could be substituted for D in the
expression aDb. What, exactly, are the relata in this expression? a comes from language and
b comes from constructs, but I noted earlier that both language and constructs are them-
selves complex categories containing various subcategories. An unambiguous explication of
designation requires a definite statement of what subcategory in language has the capacity
to serve as designator for a construct subcategory. A simplified view of the various subcat-
egory relations via designation is shown in Figure 1.3. In sum, terms designate concepts,
copulae + qualities of objects designate predicates, predicate letters (capital letter) stand for
predicates, sentential functions express propositional functions, declarative sentences express
propositions, and entire languages are needed to express theories.

Language category (a) Construct category (b)

concept (individual)
Sy:~:~
?-
~ concept (class or relation)
Phrase~
predicate letter ~ a D b
predicate

sentential function ~ ~ propositional function


sentence / " ~ proposItIOn
language theory
Figure 1.3. Summary 01 the language-construct relations su bsumed under the relation of
designation.

In .certain of the designation relations, there can be confusion over whether, in a partic-
ular instance, one is speaking of, to pick an example, density the term or density the concept.
To avoid this confusion I follow the convention of Bunge and others by using:
The Science of Forest Science 25

'singJe' quotes when writing of symbols and terms (language category items) and
"doubJe" quotes when writing of concepts and predicates (construct category items).
While on the subject, and for the sake of completeness, I use no quotes when writing of
natural objects or events. Thus, the title of Chapter 2 is stated to be about interaction (the
event), 'interaction' (the term), and "interaction" (the concept).
The symbolism just described is simply a matter of convention, for, as Hayakawa (1972)
has observed, "human beings, by agreement, can make anything stand for anything else .... "
The use of quotation marks seems to be one area where philosophers have agreed on some-
thing. I now look at the other relation with language as its origin.
8. Denotation.
Denotation is a somewhat more involved relation than indicated in Figure 1.2. Bunge
(1974) suggests there are two routes from language to natural objects. Both, he suggests,
should be called denotation, but one route goes directly from language to natural object,
while the other goes to natural objects by way of constructs, i.e., through designation and
reference. In this way, when the objects under consideration are actually constructs, as
in formal science, designation and denotation are identical. Notice that when denotation
goes to objects through constructs, it goes through reference (arrow 9 in Figure 1.2), not
through representation. Denotation rules, then, give names to the referents of constructs.
Denotation rules and what Bunge (1974) calls semantic assumptions constitute so-called
semantic formulas. The latter play an important role in the process of interpreting an
abstract theory into a factual theory. The denotation rules "baptize" (Bunge's term) the
referents of the theory, while the semantic assumptions "link constructs to factual items by
indicating the traits (properties) of things that the constructs are supposed ... to represent"
(Bunge 1974).
The differences between the two parts of semantic formulas are significant. First, the
naming or baptizing operation is somewhat in the nature of a custom-it is a matter of
personal convention. The other part of semantic formulas, dealing with the relation between
constructs and factual items they are supposed to represent, is not simply a matter of
convention. Rather, they are hypotheses about the construct-natural object relation. The
distinction between these two parts of semantic formulas is often blurred in the scientific
literature by use of the third person singular, present indicative mode of 'to be,' 'is,' when
making the interpretation of the central proposition of a factual theory.
As an example of this confusion, consider the proposition expressed elsewhere about the
dynamics of forest stands (Leary 1979):

ASC//I,t = Pl1 - ,:xp(b/SC)],


where
SC is standing crop measured in square feet of basal area per unit land area,
P is the potential of the standing crop to increase in one unit of time,
[1 exp(b/SC)] is the modifier of the potential due to competition,
Pi1 exp(b/SC)1 is the resultant net increase in standing crop in one unit of time,
b is a numerical constant, and
expO denotes the Naperian base to the indicated exponent.
The statements following 'where' may be thought to constitute the factual interpretation
of the abstract theory of growth expressed mathematically III the above equation. As stated
earlier, 'is' blurs the true nature of the relation between symbol on the left and statement
to the right. To what extent are these statements simply naming operations, and to what
extent are they specific hypotheses about construct-natural object relations? To be more
specific, hence pre"tise, in interpreting the above equation (proposition), each occurrence of
'is' must be replaced by the more appropriate relation: designates, denotes, or represents. I
suggest the following:
26 The Science of Forest Science

Be designates standing crop measured in square feet of basal area per unit
land area,
P designates the potential of the standing crop to increase in one unit of
time,
[1 - exp(bjSe)i represents the modification of the potential due to competition,
Pll - exp( bj Se)] represents the resultant net increase in standing crop in one unit of time,
b denotes a numerical constant.
e denotes (also designates because it is a formal object) the base of natural
logarithms.
In the process of separating denotation (naming) rules from semantic assumptions, I
find that some of the items (the first two and the last two) are indeed denotation rules,
matters of convention, while the middle two are semantic assumptions, matters for empirical
test.
I consider this a sufficient introduction to the two arrows emanating from language in
Figure l.2 and proceed to examine the two arrows emanating from the construct category.
9. Reference.
Reference is the name for the relation between constructs and objects as a whole. The
reference relation itself deals with what a construct is about, factual if about real objects
or formal if about formal objects. Again, recall that each category, object and construct,
has several subcategories. For the purposes of this discussion I am little interested in the
subcategories of factual objects, but I am interested in the three major subcategories of
construct: predicate, proposition, and theory.
Some examples may clarify the notion of reference. In the case of predicate, e.g., "site,"
we ask what is "site" about? To what does it refer? In the case of propositions, "Red pine
outgrows jack pine" raises the same kind of question. Similarly, in the case of theories, we
may ask what is the theory of indirect interaction~ (Chapter 9) about? The set of objects
that each is about forms what are called reference classes. Thus, I shall speak of predicate
reference class and statement reference class.
One motivation for engaging in a reference class analysis of propositions and theories
is simply that before engaging in empirical tests of propositions and theories, the complete
and appropriate set of factual referents must be clearly in mind. Absence of this information
may lead to tests that refute a theory other than the one conjectured. Since a predicate
is never validated directly, the motivation based on empirical tests is absent. On the other
hand, I am interested in the reference class of predicates, such as "timber," because of their
occurrence in propositions.
The complete theory of reference presented by Bunge (1974) is much too complex to be
fully discussed here. However, let me sketch briefly what I feel are his main points
1. The notion of reference is heavily dependent on the view of predicate as a propositional
function that maps objects into statements containing the predicate. It is the domain
of the mapping that is of particular concern.
2. The referent of a predicate is the set theoretic union of its domain elements. Thus, if B
is the binary factual predicate, Bac, it has the form

B :A x e ---; Statements containing B, and its reference class is

R (B) = Aue,
that is, the elements of the union of sets A and C.
3. Reference has nothing to do with "truth." (Extension deals with truth.)
4. Reference is not the same as evidence. (Evidence is nearer the converse of reference.)
5. The referent of a proposition is little affected by what Bunge calls "the gross proposi-
tional structure" of a statement. (More on this later.)
The Science of Forest Science 27

6. When computing the referents of propositions and theories, the task normally gets back
to one of determining: i) the factual referents of the predicates and ii) determining an
appropriate set theoretic operation for their combination.
Working through two simple forestry examples should illustrate the reference relation:
Example 1. Consider the concept "competition." It has at least ternary predicate structure.
That is,
C : A x P x R --+ S.

This states that the predicate "competition" takes three kinds of objects into statements,
S, that employ the predicate. Suppose I say that object A is the set of individuals in one
population, P the set of individuals in another, and R the set of resources. The referents of
"competition" are, then,
R(C)=AuPuR,
the set-theoretic union of populations and resources. More on this in Chapter 2.
Example 2. What is the reference class of the simple proposition p, where p is Red pille
outgrows jack pine on all sites. Symbolically this may be written:

(lIa)(lIb )(lIs)( (Ja 1\ Rb 1\ S s) -> Gab),


where
II means "for all"
a, b designate arbitrary trees growing on site s
J, R, S designate the class concepts (factual predicates) jack pine, red pille, and site,
and
G designates the formal binary predicate "outgrows."
The formula for computing the reference class of a proposition is to form the union of
the reference classes of the factual predicates that go together to make up the proposition.
Thus

R (p) = R (J) u R (R) u R (S).


Now, predicate J takes trees into statements that use the concept "jack pine," R takes
trees into statements containing "red pine," and S takes places of any kind into statements
that make use of the concept "site." Thus, the reference class of p is given by

{jack pine trees} U {red pine trees} U {places}, or

R ()= {red pine jack pine


p trees ' Places} .
trees

This overly-simple proposition should, of couse, be made more realistic by adding two
items: a specification of the environment and specification of units of measure. Rather
than pursue this complexification, I make several observations relative to items 1 - 6 listed
above. First, the reference class of the proposition would be the same, had it read 'jack pine
outgrows red pine on all sites' (item 5). The reference class tells us what the proposition
is about, red and jack pine trees and sites, and offers no direct help in ascertaining its
truth value. In its overly-simple form, the proposition's domain of truth, its extension, is
indeterminant. Before it can have a definite truth value, the environment and measure
units must be specified. The range of these variables for which the proposition is true
will constitute its extension (item 3). Evidence for or against the proposition will come as
measures of jack -and red pine trees growlllg in nearly identical environments, so evidence
is quite different than reference (item 4). Computing the reference class of a proposition
required that I determine the reference class of each predicate contained in the proposition
28 The Science of Forest Science

and then form their set theoretic union (items 6, 1, 2). Of course, the real challenge is to
compute the reference class of factual theories. But this is very difficult if the theory has
not been axiomatized.
I will make greater use of the reference relation in Chapter 3 where I examine some
important forestry concepts to see if spontaneous change is d part of their reference class.

10. Representation.
Representation is a ubiquitous binary relation in the affairs of man. A brief search
of a card catalog subject index yields the following areas in which representation plays
an important role: art, political theory, mathematics, applied scientific research, cognitive
development in children, philosophy, and general systems theory. This brief discussion is
limited to the use of representation in applied scientific research and philosophy. My concern
is when representation is used to pair constructs with properties of natural objects. Recall
reference pairs constructs with entire natural objects.
Six aspects of the representation relation are used to order this discussion.
a. Parties to the relation. There are three basic categories of parties to the relation: natural
objects, scientific constructs, and artifacts. The pairings permissible are (adapted from
Bunge 1969b):

Description Example
1) artifact represents natural object drawing of a tree
2) artifact represents artifact flow diagram of a factory
3) artifact represents construct tree diagram of argument
4) construct represents natural object theory of evolution
5) construct represents artifact automata theory
6) construct represents construct coordinate of a point

Notice that natural objects represent nothing. The kinds of representation in later chap-
ters fall mainly under numbers 4 and 6, although there is some occasion to use drawings
of natural objects, item 1). When focusing on construct-natural object representations,
I make use of mathematical methods from analysis and group algebra; while focusing on
construct-construct representations, I use Cartesian and Haskellian coordinate systems.
b. Kinds of representations. Bunge (1969a) identifies 2 kinds of representation: formal
and substantial. Formal representation tends to be superficial in nature. That is, a
mathematical model of forest dynamics may be a formal representation of a forest if the
behavior of the model and the forest are similar. Substantial representation requires
that the parties to the relation share some parts or mechanisms. Thus, an old forest
plantation may be a substantial representation of a natural stand because both have
growing trees of the same species and age.
c. Properties of the relation. Following some of the symbolism and terminology introduced
earlier, I can say that representation is a binary predicate and symbolize it Re( ).
The "object variables" are either artifacts, scientific constructs, or natural objects. The
properties one can examine are of the following nature: Given Re(x,y), i.e., x represents
y, does it follow that Re(y,x), i.e., does y represent x Definitely not. Representation
is not a symmetric relation. The second property is contained in the answer to this
question: Is something a representation of itself? That is, is Re(x,x) true? The answer
IS yes. The third property involves the relation between 3 objects, say x,y,z. The
question is, if Re(x,y) and Re(y,z), is it true that Re(x,z)? That is, if x represents y,
and y represents z, does x represent z? The answer is yes. Thus, representation is a
nonsymmetric, reflexive and transitive relation. This makes it a subrelation of analogy
and .simulation (Bunge 1969b).
d. Strength of the relation. A deeper examination of the relation may allow an assessment
of its strength. This can be accomplished when representation holds between sets.
The Science of Forest Science 29

Strength is assessed by seeing how elements in two sets are paired. When Re(x,y) holds
between sets x and y, the representation is called plain if just some elements of x are
paired with some elements of y. The representation is said to be injective if every element
of set x is paired with a unique element in set y. It is bijective if the unique pairing is
true in both directions. The next greater degree of strength requires the existence of
a relation between the elements in x, i.e., the set must have a structure. A relation is
homomorphic if it maps every element of the set x onto some element of the set y and
it does so in a way that the structure in the set x is preserved. If the converse holds as
well, the relation is said to be isomorphic. Here is a summary of these ideas about the
strength of the representation relation:

Kind: plain ... injective ... bijective ... homomorphic ... isomorphic
Strength: ------------------------increasing

Bunge {1973b) suggested that plain representation, injective representation, and


structure-preserving injection (homomorphism) are the most common in science, and
that true isomorphism is found only in mathematics, hence it is an unattainable ideal
for applied science.
e. Heuristics of representation construction. Clearly, the methods used in constructing a
representation depend on the parties to the relation. For discussion purposes, I limit
attention to the relation between scientific constructs and natural objects, item 4).
Much of the current activity in this area is called mathematical modeling. Quantitative
mathematical methods, as opposed to qualitative mathematical methods, employing
mathematical equations are the dominant tools in use. Only a very small portion of
mathematical modeling employs methods from geometry, abstract mathematics (e.g.,
group theory)' and mathematical coordinate systems. Forest scientists should under-
stand that to mathematize something is not to quantify it, and to suggest that they are
the same evidences too early a termination of mathematics training.
It is not my purpose here to attempt a description of model building basics. Rather, I
would point out that mathematical models are only parts of representations, although
in some cases the representation effort begins and ends with a mathematical equation.
Instead of beginning with a mathematical equation, one may wish to follow the steps
suggested by Bunge {1974a):
l. The schema of a model object must be specified. This requires making a list of the
important properties possessed by the actual object.
2. A sketch of the model object must be put together. The sketch simply shows how
the properties of the actual object are inter-related in the model object.
3. A theoretical model or a specific theory must be specified. The representation must
take the form of a hypothetico-deductive system interpreted for the specific model
object at hand. Hopefully the hypothetico-deductive system will be axiomatized.
4. The general theory (of which item 3) is a special case) must be identified. Step four
is seldom stated in the final reports of representation efforts. The general ignoring
of this step in forest science can have the effect of misleading scientists into thinking
that their work is a special case of nothing it is totally new. With few exceptions,
this is not true. It can be very mind-broadening, and helpful in finding a place to
start model building, to be required to identify the generic theory. I will to return
to the idea of a generic theory when discussing Haskellian coordinates in Chapter 2.
f. Metaphysical requirements of representation. Typically several, say 3 or 4, alternative
representations of a natural object seem attractive, hence they become candidates for
"truth" testing. A rule is invoked, consciously or unconsciously, to select the single
"best" representation. When referring to the kind of representation called mathematical
model, the focus may be on the agreement between deductions from the model and
30 The Science of Forest Science

properties of the natural object being modeled. Bunge (1967a) lists three criteria to
assist in this:
1. goodness of fit of model (representation) to data (measured values of properties of
natural object),
2. theorification potential of the model, and
3. interpretability of numerical constants in the model.
Margenau (1950) suggests some broader criteria that may be used to "measure" con-
structs: 1) logical fertility, 2) multiple connections with other constructs, 3) permanence
and stability, 4) extensibility, 5) causality, and 6) simplicity and elegance. Perhaps sim-
plicity has received greatest emphasis, in part, at least, due to Ockham (Hutchinson
1978, Sober 1975, Bunge 1963).
In sum, let me pose the question: What represents what? Bunge (1974a) suggests the
following tableau:

representing construct represented natural object

predicate property of a system


set of singular or existential fact involving one or more systems
statements
set of universal statements pattern of system's composition,
structure or change
set of statements (e.g., a theory) system

This completes my summary and highlighting of the "system" in forest science. Greater
depth in each of the ten topics can be found in Bunge's writings and the philosophy literature
in general. I now attempt to apply some of the ideas just discussed to real problems of forest
research.

E. Research Problems
1. Examine the results of several research studies with which you are familiar and assign
them an approriate loop number from Figure 1.1. What is the mean loop number?
Relate mean loop number to the stage in one's career. Is there a trend?
2. Assemble propositions (laws) about forests or trees, and identify the constituent formal
and factual predicates. Identify the range and reference class of each proposition.
3. Assemble theories of forest ecology and forest management. Identify the constituent
propositions, and, in turn the predicates in each proposition. What is the reference
class of each theory?
4. Classify the theories identified In problem 3 using the 6 criteria identified in this chapter.
5. Consult recent issues of Forest Science and Canadian Journal of Forest Research for
examples of mathematical formula interpretations. In each case decide whether desig-
nation, representation, or denotation is a more appropriate name for the relation.
6. What are additional criteria for choosing among alternative mathematical representa-
tions of processes being modelled? Alternate problem: Identify the many dimensions of
simplicity as a criterion for representation selection (see Bunge 1963).
7. Articulate the difference between an event and a process. Find concepts that represent
events. Find ones that represent processes.
8. Do forest natural object populations, e.g., trees of different species, juxtapose or super-
pose when interacting?
9. For each theory identified in problem 3, identify its generic theory. What other species
are .there? For additional insights, examine Shive and Weber (1982).
10. Find at least one example where a forestry concept has been mathematized but not
quantified.
2: Interaction,
"Interaction", 'Interaction'

Doing exact scientific research is like driving in a big city; maneuvering among streets
that have more than one lane. Scientists must be at ease in any of three lanes: the
event/thing, the construct, ana the language. This is because whatever qualifies as an
ontological category may have three "faces." One deals with the actual physical-biological
event/thing. Another deals with the mental constructs used to think about the event/thing.
The third deals with the words used to communicate thoughts about thoughts and about
the event/thing. Not every word has two other faces, nor is every conceivable event/thing
referred to by a concept that is designated by a word.
I begin this chapter by briefly outlining Bunge's treatment of interaction as an onto-
logical category, as a subcategory of change. Following this extension of the discussion of
ontological categories begun in Chapter 1, I discuss the construct "interaction," with special
emphasis on biological interactions. I then look at the term 'interaction' and skim the lit-
erature that discusses what is perceived as inadequate attempts to give the term a precise
meanmg.

A. Interaction
The interaction event is an ontological category in the same sense as thing, property,
state, etc. outlined in Chapter 1. A basic difference is that interaction is built from these
more basic categories. In no way, however, does its derived character diminish the reality of
interactions.
What are the bare essential ontological categories of which interactions are made? Let
me list some of the important ones in abbreviated form as I abstracted them from Bunge
(1977):
• A thing is a substantial individual endowed with certain properties.
• A thing's properties can be used to characterize it.
• The collection of properties of a thing constitute the state of a thing.
• Things change.
• When things change, their properties change.
• When a thing's properties change, a point representing the thing in state space moves
to a different position.
• Change in a thing is represented by a thing's trajectory in state space.
• A thing's entire trajectory constitutes its history.
• A segment of a trajectory is an event or process.
• The state of a thing is relative to a reference framework used to characterize its state.
• In order to be useful, the reference framework must not affect the thing's state, and the
thing's state must not affect the reference framework.
32 Interaction, "Interaction", 'Interaction'

• Things change either spontaneously or by being induced to do so.


• When a thing is being acted upon by another thing, its trajectory is different than if it
is changing spontaneously.
• Induced change can operate reciprocally, i.e., thing x can act on thing y, and thing y
can act on thing x.
• If reciprocal induced change is true, thing x and thing y may be said to interact.
• The size of an action is given by the amount of distortion effected by the agent on the
behavior of the patient.

Several observations are in order:

1. Only things interact. Recall we had three major categories in Chapter 1: objects,
constructs, and language. The claim here is that subcategories within language, e.g.,
words, parts of sentences, etc., do not interact, nor do subcategories within construct,
e.g., concepts, propositions, and theories. In the case of constructs, it is their referents
that interact: for language, their nominata. Elsewhere (USDA Forest Service 1977,
page 429) one finds statements expressing views about 'resource interactions.' Sticking
to our observation above, this statement is correct only if 'resource' is a thing. If, say,
resource is a relationship between things (a predicate)' then it is not appropriate, in
this ontology, to speak of resource interactions.
2. Among scientists studying biological interactions, there has been an increasing lack of
concern with change in population properties when populations are not affecting each
other. This type of change is called spontaneous change. Apparently, the view is that
when seeking to learn about interacting things, it is sufficient to study things interacting.
This view is typified by Williamson (1972) who stated, in effect, that spontaneous
change is clearly of no interest. The fact is that spontaneous change forms the reference
framework from which induced changes are measured. No reference framework and
induced change is adrift. Things never found by themselves (i.e., obligatory) can be
expected to have a negative level of spontaneous change from a hypothesized nonzero
initial condition. Facultative things have zero or positive levels of spontaneous change
from nonzero initial conditions.
3a. Bunge (1977) reserves the name interaction for reciprocal induced change. This bears
a striking similarity to the view suggested by Haskell (1947), but is more restrictive.
The similarity is based on the notion of spontaneous and induced change. Haskell used
Einstein and Infeld's (1938) idea of a process proceeding at a "normative rate" that
can be affected by another's action. By requiring reciprocally induced change, Bunge
limits the coverage of the event to less than what is typically included within biotic
interactions. For example, commensalism would not be considered an interaction, yet
it is a well-established relation in which one party's spontaneous change is enhanced
and the other's is not affected. Other relations that are typically included as biological
interactions would be excluded also.
b. The length of time over which a set of parties is monitored for spontaneous and induced
change can greatly affect a diagnosis. Perhaps in physical sciences reactions to actions
are sufficiently instantaneous as to present few difficulties. However, in biology, and
forestry in particular, a change induced at one time may take a long time to become
evident.
c. It would seem that the specification of interaction as action plus reaction masks the fact
that both participants may act on the other population as well as react to it. So, in
reality, there are two sets of actions and reactions in an interaction. For example, it is
not uncommon for the grazing action of an insect on plant foliage to initiate a change
in leaf chemistry so that grazing a few moments later is less beneficial to the insects
(Haukioja and Niemela 1979). Unless, of course, the grazer can change its metabolism
to counteract the materials the plant has produced in reaction to its actions.
Interaction, "Interaction", 'Interaction' 33

d. What I have in mind is that a net action is in fact a two-, rather than one-stage process.
Namely:
agent patient
action
A ----i.~B

!
reaction of B
to action of A

B'
Of course, the schema must include the action reversed:

agent patient
action
A .....f - - - - B
reaction of A
to action of B
A'
~
e. Not every thing or class of things has an equal ability to initiate a reaction to an action,
so an enumeration of all possible combinations of actions and reactions would reveal the
totality of possibilities.
f. Net agent action is basically the agent's action's effect plus or minus the amount the
action effect is changed by the patient's reaction. Schematically,

agent patient
action

net reaction
agent
action

g. Typically, when A's action is detrimental to B, the reaction of B is to decrease the


action, i.e., make the net action effect less than the action effect. If the schema above
were a triangle, the reaction of B would be to change angle ABB' from 90 0 to where
angle ABB' is less than angle BB' A.
h. It is generally assumed that if the action of A is to enhance the spontaneous change in
B, there is little if any reaction on the part of B. (This is an assumption that requires
further scrutiny.)
1. Net interaction then, can be symbolized

action

net
A •B reaction of
action
AB ~! B'
B to A

• A

~!
net reaction of
action A to B'
B'A
A'

and said to be the ordered 2-tuple


34 Interaction, "Interaction", 'Interaction'

< net action AB, net action B' A > .


4. If change in a thing is judged by changes in the numerical values of quantitative variables
used to represent properties of the thing, it follows that an appropriate set of properties
must be selected. Typically we focus on a limited set of properties of the thing and
keep close watch on them. Bunge (1967a) calls this the earmark property set. Useful
properties are those that clearly reflect the action of another thing. For example, when
studying forest tree interaction with other forest trees, dimensions of the central stem
may be adequate properties. Not so when studying tree-insect relations where the
various plant metabolites quickly reflect an action by an insect. Selection of an earmark
set is extremely important. Inclusion of too many properties leads to unneeded effort,
omission of a key property may lead to an inappropriate verdict.
5. Bunge's size of an action is analagous to the biologist's intensity of an effect. Although
"the amount of distortion effected by the agent on the behavior of the patient" can be
expressed qualitatively rather straightforwardly, it is a problem to express size of an
action mathematically, especially quantitatively. Typically, more than one element of
an earmark property set of an individual will be altered in an induced change. The
size of such a change, then, amounts to quantifying an array of distortions, or vector of
effects. If a scalar is desired, the immediate problem of an appropriate vector norm is
encountered.
Now, consider the construct "interaction." For the most part, the ensuing topics and
discussions are focused on interaction the construct and term, as viewed from a forestry
perspective. However, I shall feel free to revert to the philosophized topics discussed up to
this point whenever doing so clarifies one of the numerous issues to be raised.

B. "Interaction"
Recall that the construct category has three subcategories of interest: concept, propo-
sition, and context or theory. In this section I briefly examine the standing of interaction in
these subcategories.

1. Concept.
A conceptual analysis examines, among others, the following aspects of a concept:
1. its logical structure, i.e, is it an individual concept, class or relation concept?
2. its predicate degree, i.e., how many argument places are there for the predicate?
3. its predicate order, i.e, are the predicate arguments things or populations (classes)
of things, or are the arguments themselves lower order predicates?
4. its factual reference class, i.e., what material objects is the concept about?
I consider these in the indicated order. First, concerning the logical structure of "in-
teraction," there is little doubt that "interaction" is a relation concept. It should be noted,
however, that out of relation concepts class concepts may be made. Hence, it is expected
that "interaction" harbors a number of less general relation concepts, e.g., mutualism and
predation, so it is in fact a class concept as well.
Second, interaction has no less than fourth degree predicate structure. Two popula-
tions, each undergoing spontaneous and induced change are the minimum arguments for the
predicate.
Third, my concern is with the predicate order of "interaction." First order predicate
logic " ... is part of symbolic logic ... in which the notions of all and some are applied only
to individuals and not also to classes or attributes of individuals" (Mates 1972). Recall
that all and some are existential qualifiers of individual variables and were discussed briefly
in Chapter 1. Although symbolic logic may limit itself to first order predicates, science
in general need not, and is not, so limited. When operating at more than one level of
predicate order, it is convenient to change the terminology slightly. In first order logic we
Interaction, "Interaction", 'Interaction' 35

have predicates (variable or constant) and individuals (variable or constant). The latter name
no longer applies if higher order predicates are admitted. I substitute the term 'argument'
for 'individual' and let it be variable or specific depending upon, for example, whether a
general class of things or a particular class of things is being discussed.
To determine the predicate order of "interaction," I must examine its arguments. "In-
teraction" is built from the lower level predicates "net agent action" (A) and "net patient
action" (P). In order to state "interaction's" order, it is necessary to determine the order of
A and P. Net agent action, the event, is represented by net agent action the concept (and
is designated by the predicate letter A), hence A deals with the relation between "agent"
and "patient," or symbolically A( , ). "Net patient action" deals with the relation between
"patient" and" agent," or P( , ). Clearly, both net agent action and net patient action are
relation concepts with binary degree predicate structure. But what is their predicate order?
This requires a commitment to state whether the "agent" is an individual or a population
(class) of individuals.
Throughout this book I am assuming that the interactions occur between populations of
individuals, hence both A and P are second order binary predicates. It is worth noting that
the arguments for A are patient populations under spontaneous and induced change, and
the arguments for P are agent populations under spontaneous and induced change. Finally,
"interaction" is a third order predicate because its arguments are "net agent action" and
"net patient action," themselves of binary second order predicate structure.
The fourth, and last, step in our conceptual analysis is to determine the factual refer-
ence class of "interaction." This is difficult because the existing techniques are directed at
first 'order predicates. How does one proceed with higher order predicates? I proceed in a
manner similar to mathematicians when converting a higher order differential equation to a
system of first order equations. By back substitution into J(A, P), I get J(A( , ), P( , )),
which indicates that at the second order, as low as I shall go since I am dealing with pop-
ulations of things, "interaction" is a fourth degree predicate. Here, the arguments are pop-
ulations of things under spontaneous change and change induced by the other population.
The propositional function form of the interaction predicate is:

J P B X Pi X AB X Ai ---> S,

a mapping from 4-tuples of populations (agent (A) and patient (P)) to statements containing
J. The factual reference class of J is the collection of its arguments at the second order level,
i.e., the collection of 4-tuples of populations under spontaneous and induced change.
Several observations can be made about the above analysis. First, it is the simplest
conceivable situation that meets the necessary requirements for an interaction according to
Bunge's ontology (Bunge 1977) and also meets conditions necessary to represent a certain
kind of interaction found in nature. Second, the previous formulation is by no means suf-
ficient. It is insufficient because many interactions are not between "agents" and "patients"
but are between two "agents" through another thing. Sometimes the latter is referred to
as a resource. However, there is a serious question about naming such things resources. I
substitute a more general name, mediary, dropping the redundant inter, to denote those
things through which the actions are effected. Third, once a mediary is permitted, it must
be presumed dynamic, both spontaneously or by induction from either population, patient,
or agent. This increases the degree of J (at the second order) from 4th to 7th (added is
mediary population under spontaneous change, change induced by the patient population,
change induced by the agent population). Fourth, the mediary need not be living, although
it may be.
36 Interaction, "Interaction", 'Interaction'

Fifth, three classes of interaction events are suggested depending on the referring con-
cept's degree and the biotic-abiotic nature of the mediary:

structure example
4th degree predicate, no mediary prey~predator
parasite~host
7th degree predicate, biotic mediary parasitoids~
host larvae
7th degree predicate, abiotic mediary competition for
light among
forest trees

Sixth, when no mediary is present, each interaction participant can be said to form part
of the other's environment (Niven 1980). Seventh, interactions with a biotic mediary come
close to indirect interactions treated extensively in Chapters 9 and 10. Eighth, not to be
overlooked is the possible combination of two of these classes, as well as the existence of two
different mediaries. An example is black walnut~alder interactions where alder is affected by
juglone secretions from walnut tree roots (juglone is the mediary) and the alder intercepts
phytosynthetic active radiation before it reaches the crowns of the shorter walnut trees.
In summary, interaction is not a simple concept, yet an examination of its properties
as a concept helps to unravel its complexities and to show how additional interaction events
can be subsumed under the concept in an orderly manner.

2. Proposition.
Intermediate between concept and theory is the construct category called proposition.
Propositions play a key role in the goal of many scientists who purposefully attempt to
answer difficult questions with statements of great generality (Leary 1985). I believe that in
substantive (as opposed to methodological) scientific research, only the first and third items
below are of importance:
a) the question being asked,
b) the methods used in answering the question, and
c) the answer statement.
Generally true propositions can be answers of great value. The framework in Figure 2.1
summarizes the relations between three important question types and the generality of an-
swering propositions.
Included in Figure 2.1 are gradients of worth progressing from least valuable in the
upper left to most valuable in the lower right. Included are two dashed lines indicating
most-frequent courses taken in an aspect of science as it matures and moves toward the
lower right corner. Route A is perhaps the traditional route. Scientists attempt to quantify
an idea and progress toward an explanation of a phenomenon using methods of an analytical
nature. If successful, they may then attempt to generalize their proposition to fit more
situations, moving down the right-hand column. Methods of developing generality quickly
have been suggested by Chamberlin (1897), Platt (1964), Popper (1959, 1963)' and others,
and are significant topics in the philosophy of science.
A second strategy for progressing from upper-left to lower-right in Figure 2.1 is to de-
velop a universal statement of a descriptive nature, and then to attempt to develop predictive
and explanatory relations. This book is a purposeful attempt to turn the corner in strategy
B, using as a starting point Haskell's universal descriptive proposition that is the coaction
cross-tabulation. Significantly, Haskell was able to get universality in his proposition because
he focused on relations between things. Those who focus on things are the traditional trekers
along route A.
Interaction, "Interaction", 'Interaction' 37

QUESTION
(answer name)

What Is?
What character? What if? Why?
(description) (prediction) (explanation)

- - Increasing difficulty to answer------+


ANSWER STATEMENT

:---------~- --- __ I
- - ---
singular

Indefinite existential
("In at least one case")
-
,.,
~II
- ~,- --1-- ~-=-=--~,~
\ )~~
r::
definite existential II \ I '~

..•
("In n cases") 01
01
r::
--\--1- - - - - -1---
bounded universal
("In a/l cases In
~\ I I
II
universe A") '''', 1
"r:: - - - - ->G~- - - - I - - - - -
universal
("in a/l cases") I I

Figure 2.1. Framework for assessing the research productivity of a scientist. The framework
is limited to the evaluation of substantive research (adapted from Leary 1985).

Generally true propositions that meet additional requirements (see Bunge 1967a,
Hempel 1965) may be termed laws of nature. One of the shortcomings of forestry as a
scientific discipline is that few relations are known that warrant a law label. A candidate
law of forest growth has been developed by Pienaar and Turnbull (1973). Based on Berta-
lanffy's anabolic-catabolic balance equation of biological growth, it has the form

dBjdt = aB). - ;3B.

A possibilist biological law of forest growth could have a similar algebraic form, but reflect
quite different ontological assumptions. The form

dB j dt = a' B(1 - ;3' B A')

is predicated on the idea that forest growth is exponential (a' B) until competition sets in
from other trees. Competition is expressed in the equation's second term.
Every law of nature is represented by a law formula that expresses a proposition. A
theory is a web of laws.

3. Theory.
Quantitative biotic interaction theory has been dominated by the writings of Verhulst,
Volterra, Lotka and Gause, and the very large numbers of modern day analyses and reanaly-
ses of their ideas. So wide spread are these analyses that I shall not discuss them here except
to call attention to the original works and Hutchinson's excellent treatment of the history
and growth of these ideas (Hutchinson 1978).
Most of this type of interaction theory development is taking place in ecology, not
forestry, yet it is informative to note that Gause acknowledges the research findings of V. N.
38 Interaction, "Interaction", 'Interaction'

Sukatschev, noted Russian forest ecologist, as an inspiration for his laboratory experiments
(Gause 1964).
Quantitative biotic interaction theory in ecology may be said to be far ahead of that of
forest science. One of the important reasons is that the "fathers," Volterra and Lotka, were
mathematicians in their own right who took an intere~t in population dynamics for a period
in their lives and in the process gave it a direction that has changed little since their seminal
works. Recent advances have moved the already advanced stage of development ahead even
further. The following observations still seem in order.
1. Recent efforts to axiomatize these quantitative theories have given a more clear under-
standing and appreciation of the basic presuppositions, axioms, and theorems (Lewis
1980). Failure to fully understand the presuppositions can lead to disputes over ex-
perimental procedures and designs when one attempts empirical tests of theorems (See
Ayala 1969, 1970; Gause 1970).
2. Problems in biology and ecology have spurred the development of new or litte devel-
oped branches of mathematics. A good example is the work of Rashevsky (1960), and
associates Rosen (1958) and Nahikian (1964). Of course there is always a tendency
for the existence of certain mathematical methods to spur the development of selected
aspects of ecology, e.g., diversity indices. Caution is needed, however, because at times
the power of the mathematical methods have exceeded the powers of biological under-
standing so that "ghosts" were required to justify the mathematical analysis of negative
population levels (May 1972, Roberts 1974).
3. Recent concern about alternatingly open and closed systems, phase plane trajectories far
from equilibrium, bifurcations, and dissipative structures (Bakuzis 1978) has provided
a needed balance to the traditional problems dealing with the mathematical analysis
of model ecosystems, closed systems, isoclines, equilibria, and other questionable artifi-
cialities. Caution is in order or mathematical ecology will join mathematical economics
as a formal science.
4. Recently, a new class of mathematical equations that uses a time lag has been devel-
oped that promises improved abilities to represent dynamics of actual populations (May
1976). The idea was first put forward some time ago (Cunningham 1954, Hutchinson
1954). For many years the same class of mathematical equations was repeatedly used to
represent population dynamics: constant coefficient, first order, logistic-based simulta-
neous differential equations. There remains the opportunity to use system identification
techniques to estimate a variable time lag (Bellman, et a!., 1967).
5. Forestry and ecology are still troubled with operationalism. Clark (1971) suggested that
population models employing second order differential equations would offer superior
representational capability, including hereditary effects. Innis (1972) criticized the idea
because, it was argued, no one had ever "seen" a second derivative.
6. Until recently there has been an imbalance in the structure of the science of ecological
population dynamics (see Figure 1.2), wherein so much emphasis has been placed on
the representation relation that little time has been left for ontology (e.g., developing
precise ontological categories), reference analyses, and construct-construct relations. For
scientists of a mathematical persuasion, the danger resides in rushing to elegance using
mathematical analysis methods with no parallel analysis of the resulting construct's
factual referents. The elegant construct may have a vacuous reference class because
no system of populations could possibly meet all the assumptions necessary for the
mathematics to be valid. A spelling out of important assumptions is a first step in
referent analyses. Granted, a measure of science is its predictive power, but of what
vall}e are predictions about smaller and smaller chunks of reality?
The state of quantitative population dynamics theory in forest science is much less de-
veloped than in ecology. Although deficient, good progress is being made. Some suggestions
Interaction, "Interaction", 'Interaction' 39

for accelerating progress are:


1. Most quantitative forest scientists need better mathematical tools for the first step,
which is stating theories of system dynamics. For many years the single linear regression
equation was the dominant tool used to represent forest dynamics. Several excesses with
computers occurred wherein the entire task of selecting a representation was given to
the computer with a single criterion to be used in the selection-best fit of model to
data (Grosenbaugh 1958; Furnival 1964,1971).
2. Applied mathematics instead of applied statistics appears to be a more fruitful source
for quantitative expressions of dynamics. Forests are full of "logical observation" or
data points, commonly called natural boundary conditions, that to help specify the
solution to governing differential equations and to aid in estimating numerical constants
via inverse problems.
3. Twenty years ago, forest tree population interaction theory was on the verge of a sig-
nificant breakthrough (Thrnbull 1963), but for some unknown reason, a retreat to a
one-dimensional nonlinear view ensued and the full potential of the original ideas is yet
to be rekindled.
4. Forest scientists working in population interaction theory must guard against those who
would wield the operationalist's guillotine. Grosenbaugh (1970) put it to use when he
argued that differential equations had no place in forest growth models because growth
is not observed or measured per se, rather, size is measured at different times. The
guillotine was readied else where: "Obviously, the first requirement was to develop an
operational definition for diversity of renewable resources" (Boyce and Cost 1978).
5. Like other scientific research areas, population interaction theory has a history of making
the object category item (things, populations of things, etc.) fit the representation tool
rather than the representing tool fit the unique characteristics of the object items. This
is a classic case of Kaplan's law of the instrument: "Give a small boy a hammer, and
he will find that everything he encounters needs pounding" (Kaplan 1964).
6. Forest scientists need to know the different ways to quickly establish generality for their
propositions about population interactions. More use is needed of the hypothetico-
deductive method of science wherein axioms, definitions, and other primitives are as-
sumed to be true (based on prior knowledge) and are used to derive theorems, corol-
laries, etc., and the latter are tested. Typically, for example, forest mensurationists
have "tested" only their primary axiom, the growth model, with some measure of fit of
predicted to observed growth and have paid little attention to testing all of its logical
consequents.
A few brief observations can be made about qualitative interaction theory:
1. By qualitative I do not mean that the theory is any less mathematical than the ones of
Verhulst, Volterra, Lotka, and Gause. Rather, they are different in that they employ,
for the most part, different branches of mathematics, e.g., special geometric coodinate
systems, graph theory, and relational mathematics. Of course, quantitative theories also
employ coordinate systems, but not of the same type used here. The qualitative theories
may also make use of numbers, and other fundamental arithmetic, trigonometric, and
geometric concepts. However, no use is made of "higher" formal concepts like derivatives
and integrals.
2. In quantitative theories, mathematical equations form the major axiom (premise), and
various theorems are deduced from them and supplementary axioms, based on various
methods of mathematical analysis, e.g., simple integration, stability theory for differen-
tial equations. In qualitative theories of biotic interactions, the major axiom does not
have the form of a mathematical equation. For example, in Levins's theory of evolu-
tion in communities, his major axiom is a signed digraph representation of population
structure in a community (Levins 1975). By performing what he calls "loop analysis,"
40 Interaction, "Interaction", 'Interaction'

he determines the logical consequents of his major and minor axioms. These deductions
form theorems that are scrutinized under empirical tests. Haskell's qualitative theory of
coaction (interaction) has as its major axiom a mathematical coordinate system. The
object system dynamics take the form of a trajectory in coordinate space. Because of
the way the coordinate system was formed (see Chapter 4), location and orientation of
a trajectory allows certain deductions to be made (theorems to be formulated) about
the object system. The theorems are then subjected to empirical tests. In the course of
other chapters, I shall expound a number of theorems deduced from this starting point.
Qualitative interaction theories have a broader span than the quantitative ones men-
tioned here. By span I mean they are capable of representing a number of varied trajectories
and structures that would require separate systems of mathematical equations had they
been quantified. The price for the increased span is reduced depth; the qualitative theories
may only predict, say, direction of change instead of amount of change. Levins (1966) has
addressed these issues in a more formal manner.
Although some ecologists have made use of very early stages of Haskell's coaction theory,
none appear to see any worth in its later forms if judged by level of use. Forest scientists have,
with the exception of my colleagues, Bakuzis, Mattson, and Addy, been totally oblivious to
it (Bakuzis 1974, Mattson and Addy 1975).

c. 'Interaction'
The third element of the triad of this chapter is interaction the term. Little has been
written about this term, perhaps because it is nicely ambiguous. Its mere mention conjures
up images of things bumping into one another or somehow having entangled destinies. Inter-
action can be a great mileage maker for some models of discourse where a judiciously placed
" ... complex nonlinear interaction ... " can reap a substantial rating improvement. Also, it has
the advantage of all ambiguous terms-it keeps vocabulary size down, until you really need
something more specific. That something can normally be had for a small term added fore,
e.g. 'biotic,' to give biotic interactions. More definite terms may warrant further scrutiny
that most often takes the form of an attempt at meaning specification. Milne (1961) made
a particularly clear and forthright case when speaking of competition, a particular type of
interaction:
'[WI' has been a familiar term in the vocabulary of biologists for at least a century. As
a scientific term, [WI ought to have only one meaning-clear, precise, and unambiguous, but
unfortunately this is not so .... So far the only comprehensive attempt to clear up a most
unscientific situation is [Xl's paper .... In the opinion of the present writer [Xl's paper ... does
not analyze the situation altogether satisfactorily nor does it provide an acceptable definition
of [WI.
The above excerpt exemplifies a wide belief that scientists have about terms:
a) a scientific term should have just one meaning,
b) the meaning should be clear and precise,
c) it is unscientific if a) and b) do not hold, and
d) what is needed is a good definition.
In the period when Milne made the above comments there was a large number of sug-
gested definitions, objections to suggested definitions, suggestions to discard entirely the
objectionable terms, suggestions to keep the terms but use them more carefully, and calls for
keeping the dictionary close at hand (Milne 1961; Andrewartha 1961; Birch 1957; Colinvaux
1973; Darlington 1972; Donald 1963; Hardin 1956; Harper 1961; Malcolm 1966; Williamson
1957, 1972). A more recent episode of a similar nature concerns so-called consumer terms
(Lubchenco 1978,1979; Lewin and Enright 197.8), although Lubchenco claims to have iden-
tified the principal problem-conceptual confusion. Historically, the explication of forestry
Interaction, "Interaction", 'Interaction' 41

terms and concepts has occupied an important but little-emphasized role in forest science
(e.g., see Tansley (1935) and the references he cites).
There is a basic premise in the above cited papers: Terms have meaning.
Bunge (1974b) argues convincingly that we should take a different view of terms. Let
me outline my interpretation of his assertions:
a) Terms get any meaning they have vicariously through the constructs they designate.
b) This vicarious meaning is called significance.
c) Terms have significance only if they designate a construct.
d) Terms that name an object have no significance because they are only conventions.
e) Constructs have meaning.
f) The meaning of a construct is the ordered pair of its sense and reference.
g) The sense of a construct is the set of logical ancestors and logical progeny of the con-
struct.
h) Sense is particularly easy to see in axiomatized theories wherein the construct hangs
from its axioms, e.g.,

axiom 1 axiom 2
• •
~I
logical ancestors of
definition 1
theorem 1 • • proposition in theorem 2

~Itheorem 2 • • axiom 3

'\/
• logical progeny of
theorem 3 proposition in theorem 2

i) The sense of the proposition expressing theorem 2 is the set


{logical ancestors, logical progeny}.
j) The meaning of the proposition contained in theorem 2 above is, then:
< { axiom 1, axiom 2, theorem 1, definition 1, theorem 3, axiom 3},
referents of propositions expressed in theorem 2>.
Since the sense of a theory is more easily seen if the theory has been axiomatized, let
me attempt a briefaxiomatization of part of the plant-animal portion of Haskell's coaction
theory as I make use of it. I shall focus primarily on identifying the logical ancestory portion
of sense, which is illustrated in Table 2.1. (One set of logical progeny is given in Chapter 9.)
The axiomatic format followed in Table 2.1 (after Bunge 1973a) could also be applied
to the quantitative theory of interaction expressed in what are normally called the Lotka-
Volterra competition equations:

dNI/dt == TlNl((K l - Nl - o.N 2)/ K l )

dNddt == T2N2((K2 - N2 - fNd/ K2).

Doing so would show that radically different formal assumptions are needed, sufficient
to cover differential equation theory, along with specific ontological assumptions, primitives,
and axioms. The Lotka Volterra theory has very specific axioms about spontaneous change
(logistic based), parameters for within population induced change (r and K), between pop-
ulation induced change (o.N2/ K1 and fNl/ K2), and biological laws of change. No such
assumptions are made in Haskell's theory. Of course, Lotka-Volterra theory purports to
represent some underlying cause of short term population fluctuations, while the coordinate
42 Interaction, "Interaction", 'Interaction'

Table 2.l. Axiomatic format of Haskell's qualitative theory of coaction as applied in this
book. [FA], [SA]' lOA]' and [BA] denote formal, semantic, ontological, and biological as-
sumptions, respectively.
Background Assumptions:
Formal (logical Presuppositions of arithmetic, trigonometry, and geometry.
and mathematical)
Semantic Presuppositions of semantics (reference, representation, denotation,
and designation) shown in Figure l.2 and discussed in Chapter l.
Ontological Presuppositions of biosystem, bioaggregate, and either juxtaposition
or superposition
Primitives: N [two population biosystem]
1'.1 [two population bioaggregate]
N [biosystem representative]
M [bioaggregate representative]
r [distance]
o [direction]
<, =, > [system aggregate comparator]
-,0, + [comparator representatives]
Axioms on:
TIME T is an interval on the positive real line [FA].
POPULATION N is a two-population biosystem lOA].
and SYSTEM M is two monospecific populations forming a bio-aggregate lOA].
N is a vector of non-negative real numbers [FA].
N is a vector of numerical values of attributes representing proper-
ties of populations in a two-population biosystem undergoing induced
change [SA].
M is a vector of non-negative real numbers [FA].
M is a vector of numerical values of attributes representing prop-
erties of two monospecific populations in a bioaggregate undergoing
spontaneous change [SA].
INTERACTION If at a given time, tET,NJ(t) < MJ(t) and N 2(t) < M2(t), then
the qualitative type of interaction between populations in biosystem
N is (-, - ) [BA]. [Other axioms follow for all combinations of the
comparators <, =, >.]
If at a given time tET, the state of the biosystem N is represented by
N and the state of the bioaggregate M is represented by M(t), the
intensity of interaction between populations of the biosystem is given
by a ratio of vector norms, i.e.,
r(t) = { IIN(t)-M(t)1I } [BA]
IIM(t)11

If at a given time tET, the state of a biosystem N is represented by


N(t) and the state of the bioaggregate M is represented by M(t), the
intensity of interaction between populations of the biosystem is given
by a ratio of vector norms, i.e.
O(t) = arctan {{ ~~~!l::~~~:l} [BA]}

system just represents the results of change. But, as will be shown later, interaction trajec-
tories in the coordinate system can hint at future events, specifically changes in type and
Interaction, "Interaction", 'Interaction' 43

intensity of interaction (see Chapters 5, 6, 9, 10).


Because of the widely different logical ancestors to the two competition constructs, it is
obvious that the constructs do not have the same sense, hence do not have the same meaning,
even though they may have the same referents and may be designated by the same word. It
follows that the terms also do not have the same significance because of the different senses
of the construct designated. In sum, the assessment made by Milne is incomplete.
A more exact approach to questions about terms is to adopt the following:
a) At a minimum, greet questions of term meaning with an eye to the complexity of
meaning speCIfication.
b) Try to identify the exact construct designated by the term.
c) Learn some construct and term husbandry to keep them alive and healthy. The alter-
native is dead ones, carcasses, that need disposing of.
d) Ferret out all the construct's presuppositions.
e) Separate them into formal, semantic, ontological, and biological/ecological assumptions.
f) Set your dictionary well out of reach. It is of little use in explicating scientific terms.
g) Reserve the use of definition for the purpose of formal introduction of new ideas into an
argument.
h) Recognize that a term can have more than one significance because the construct it
designates can have
i) more than one sense (logical ancestor, logical progeny) yet the same referents, or
ii) more than one referent yet have the same sense.
The road to exactness in forest science has three lanes: object, construct, and language.
The good driver knows what lane he is in at all times, changes lanes with skill and ease, and
knows what can be accomplished from each. So too the good scientist. Inattentive driving
and inexact science have much in common, intended destinations are seldom reached in an
efficient manner, if at all.

D. Research Problems
1. Assemble forestry concepts important to your area of interest. Separate them into
factual and formal predicates. Analyze each factual predicate for its degree, order, and
reference class.
2. Odum (1953, page 169) suggested that cumbersomely worded definitions plagued the
early history of ecology. Is the same true of the early history of forestry? Search the
recent and past forestry literature for definitions. What is the most cumbersomely
worded definition you can find? Improve it.
3. Identify an example of strategy A in Figure 2.1. Trace its historcal roots as far back as
possible. Include all important concepts and crucial developments that allowed math-
ematization and quantification. Who were the important scientists in moving from de-
scription to prediction to explanation? Who was instrumental in establishing generality
of the proposition? How was it established?
4. Identify a successful example of strategy B in Figure 2.1. Trace its historical roots.
5. The information value of a number characterizing an event or process is strengthened if
it can be associated with a relation between participants in the event or process. What
important physical constants apply only when certain relations exist.
6. Axiomatize a theory of importance in your work. Follow the categories in Table 2.1.
Consult Lewis (1980) for candidates.
7. Odum (1953) suggested a modification of Haskell's coaction cross tabulation to differ-
entiate between the effects when an interaction is "on" and when it is "off". Perform a
conceptual analysis of this interaction concept.
8. Interaction is a candidate multiordinal term. What are the problems and opportunities
affored by multiordinal terms? (See Korzybski 1958).
44 Interaction, "Interaction", 'Interaction'

9. Select another important notion in your work and analyze the term, the construct, and
the event.
10. Examine the new theory of change developed by Prigogine and associates (Prigogine
and Stengers, 1984). Is its central notion of time asymmetry characteristic of events in
forests?
3: The Role of Spontaneous and Induced
Change in Some Forestry Concepts

Every scientific discipline has a concept set that is to some degree specific to the disci-
pline. Bentley (1968) argued that for an applied discipline like forestry it is the concept set
that differentiates it from other disciplines, and the possession of the concept set differenti-
ates foresters from botanists, hydrologists, etc. It is through its concept set that a discipline's
propositions and theories are expressed. So, no valid concepts, no scientific knowledge, and
no scientific knowledge, no scientific discipline.
In Chapter 2 I focused directly on "interaction" as one member of the set of forestry
concepts. In this chapter I examine some other concepts widely used in forestry to see
if interaction the event has had a role in their formation, and, if so, to see if an explicit
recognition of this fact can be used to strengthen them, thus making them more rigorous
and extendable to other situations. The way I propose to examine these concepts is to
identify members of their reference class. Recall that reference class elements are ontological
categories, e.g., things, change (both induced and spontaneous).

A. "Site" And Related Concepts


Perhaps more has been written about the concepts "site," "site quality," and "site
productivity" than about any other single forestry concept. Carmean (1975) gives a com-
prehensive review of forest site quality evaluation work in the United States. Most text and
reference books on silviculture, silvics, and forest ecology devote major sections to what the
site concept refers to (e.g., Tourney and Korstian 1947; Sukachev and Dylis 1964; Assmann
1970; Spurr and Barnes 1980). Hagglund (1981) presents a review of research on site quality
evaluation published after 1973. It is not my purpose to review the literature on the referents
of "site." Rather, I wish to focus on what is designated by the terms 'site,' 'site quality,' and
'site productivity' and to examine presuppositions behind some of the methods of estimating
site productivity.
The great interest shown in these concepts is understandable because the thing called
site is the source or potential receptor of the abiotic flux that is transformed and accumulated
by the photosynthesizing trees, a part of Lotka's "world engine" (Lotka 1956). A thorough
understanding of the site concept is fundamental to a scientific basis of forestry. Various
attempts have been made at precisely stating a description of the thing called site. Typically,
use is made of definitions as the specification device.
Some definitions are:
1. A forest site ... is an area considered as to its physical factors with reference to forest
producing power; the combination of climatic and soil conditions of an area (Frothing-
ham 1921, after SAF Forest Terminology 1919).
2. A forest site is a piece of forest land, using the term 'land' in a broad sense to include
the soil with its underlying rock and the layer of atmosphere superimposed upon it
(Sparhawk, et aI., 1923).
46 The Role of Change in Some Forestry Concepts

3. [A forest site is] the sum of the effective conditions under which the plant ... community
lives (Heiberg and White 1956).
Although Heiberg and White claim the concept of site "has remained undisputed and
well understood ... ," it is apparent that an evolution of thought has taken place as shown
by comparing their description of the thing called site with both Frothingham's and that of
Sparhawk, et al.. To say "the effective conditions under which the plant ... community lives"
is to include everything biotic and abiotic that affects the plant community. Clearly, Froth-
ingham and Sparhawk make no reference to any biotic factors in their definition. Rather,
their focus seems to be on the physical factors that make a "piece of ... land" capable of
supporting a forest. In a sense, their view is geocentric.
When following Tourney's argument that the ecologist's habitat is the exact equivalent
of the forester's site, as Heiberg and White do, the focus switches from geocentric to phyto-
centric (Tourney 1924; Tourney and Korstian 1947). This is because the only way "effective"
conditions are separated from noneffective conditions is by doing a little (or big) experi-
ment: condition X "on," condition X "off," and compare the plant community under "on"
and "off." Is it different under "on" and "off?" If so, condition X is an effective one. If no,
it is not. In other words, habitats have no existence in their own right. They are always
habitats of some thing, in this case, of plants or plant communities. Hence my assertion that
the habitat argument is a phytocentric one.
The site concept is designated by the term 'site.' By itself, the latter is ambiguous.
Attempts to reduce ambiguity have taken their usual form, addition of terms, to wit, 'site
quality' and 'site productivity.' 'Quality' seems the less useful of the two modifiers. It is
in contra-distinction to quantity, hence connotes some sort of non metric measure of forest
producing power. It also is a reflective modifier. 'Productivity' connotes a relation between
the site and some other thing or system, because sites are not productive of sites. Further,
'productivity' does not connote a nonmetric measure. For these reasons, I shall use 'site
productivity' to designate the concept of the forest producing power of a site.
Examining for a moment the logical form of the site productivity concept, it can be
treated as a class concept, relation concept, or quantitative concept. The class concept
would enable certain groupings to be made: forest site-nonforest site; good-medium-poor
site. The relational concept "site productivity" would allow comparisons, grading from
"different than" to "better than." But, it would still not produce cardinal (as opposed tq
ordinal) numbers to be attached to the concept variable's numerical values. The quantitative
treatment of forest site productivity calls for developing numerical values on a cardinal scale
that allows an ordering of sites against an absolute scale of forest-producing powers.
Assuming two views on the best approach to site productivity estimation, phytocentric
or geocentric, there is need to add another dimension, that of method, in order to encompass
most of what is being done. I designate the methods as direct or indirect, although a contin-
uum of directness may be more realistic. Figure 3.1 shows the beginnings of a classification
of criteria arranged by the cross tabulation of views and methods.
The phytocentric view assumes that total stand volume or phytomass production is
the ultimate measure of a site's productivity. Because of the practical difficulties of direct
measurement of total stand volume and weight, it was suggested years ago that some measure
of tree height be substituted for volume or weight. There is, of course, a fair correlation
between tree height and tree volume. Methods employing tree height are considered indirect
phytocentric.
The geocentric view assumes that site productivity rests with the soil and climatic
factors, and an accurate knowledge of their workings supplies a basis for estimating the
amount of phytomass to be produced from an area. Direct methods assess energy flux into
an area as well as rate and quantity of matter cycling in the area. This approach is predicated
on the notion that large rates of flux and cycle indicate probable availability to plants. An
The Role of Change in Some Forestry Concepts 47

METHODS

VIEW direct indirect

--example crlterla--

phytocentrlc volume of wood tree height

measures of availability climatic Information


of chemical elements physiographic class
geocentric soil moisture land forms
photosynthetic active plant indicators
radiation

Figure 3.1. Example criteria, classified by view and method, used to assess the forest pro-
ducing power of a site.

example of the direct method is Czarnowski's work on productivity as a function of a locality's


soil and climate (Czarnowski 1964). Indirect methods in pursuit of a geocentric approach
may focus on larger land form features, such as physiographic class, as well as various forms
of lower vegetation that are presumed to be sensitive to the locality's matter cycling and
energy flux, that is, on species that have low ecological amplitudes. Since the vegetation
used in this method normally forms an insignificant fraction of total dry matter production,
it would not be thought an example of phytocentrism.
Actually, the early phytocentrists accepted the argument that the geocentric view is
more fundamental, but they countered with the argument that at the time (around 1915)'
American forestry could not wait for scientists to determine how the abiotic factors worked
together, and how they worked with biotic factors, to give observed plant production. Today,
more than 60 years later, the status of indirect phytocentric methods is so inflated that some
speak of direct and indirect methods, not of site productivity estimation, but of site index
estimation (Carmean 1975). This appears to be an unhealthy situation; what began as an
interim solution (site index) to a difficult problem (geocentric approach) should not now be
called the solution to the original problem.
The remainder of this section is addressed to the indirect phytocentric approach. Specif-
ically, I want to present some historical background material and develop an argument to
account for the dominance that this method and view has in the area of site productivity
estimation.
Historical Perspective of Indirect Phytocentric Method
Zon (1913) strongly advocated the geocentric approach. Because of the early date of his
remarks and the clear statement of views, it is instructive to examine some of his remarks:
The average height of the timber is usually considered the measure of the site; the
greater the average height of the stand of the same species the better is supposed to
be the site or quality class ... silvicultural units are determined not by composition
of the stand, but by the physical conditions of growth, such as climate, soil drainage,
and topography ...
Zon's arguments led him to a series of 13 conclusions, of which one is of particular concern:
48 The Role of Change in Some Forestry Concepts

The average height of the stand or site class may be the result of the interference
of man, fire, animals, etc., and for this reason cannot always be taken as the true
measure of the productive capacity of the soil.

Mean height, then, is simply influenced by too many factors to be a reliable indicator.
Roth, a leader in adapting German methods to problems of site productivity estimation
in the United States, revealed that the German experience showed top height (mean height
of the 100 largest trees per unit area) to give improved reliability in yield tables based on
height relationships (Roth 1916; Assmann 1970, pg 158-168). Frothingham (1921) added, "
... the choice of [dominant] trees for [height and age] measure is an important matter. The
trees selected must be normal, dominant and ... there must be sufficient evidence that they
have been in the upper crown cover from the start" It is not known what he meant by
"normal," although "giant," "dwarf," "wolf" and open grown trees were to be excluaed.
Following a spate of articles on the subject, there followed a hiatus in the literature
on qualities trees should possess before their height-age relationship is useful for estimating
site productivity. The situation, as of 1975, is shown in Table 3.1. A gradual evolution of
restrictive conditions leading to the current specifications of suitable trees is evident. Today,
most site productivity studies are based on individual tree stem analyses. By exammmg
the height age pattern, it is thought possible to identify trees that violate the conditions
summarized in Table 3.1.

Table 3.1. Evolution of specifications for selecting trees to be measured for height and age in
order to estimate productivity using site index. Based only on English language literature
from around 1905.

Author /Date Specification Basis

Zon 1913 mean height Height is less sensitive to stocking than other
measures of trees in a stand
Roth 1916 top height or height of Top height is less influenced by other trees
dominants except in extreme cases.
Frothingham 1921 forest grown, "nor- Even for dominant trees there can be variation
mal," dominant plue not evident from current social position
evidence tree has been
in upper crown from
start
Vincent 1961 no mammal, insect, or Zootic factors can cause anomalies m height
frost damage to affect age relationships
height age relationship
McQuilken 1975 age of site tree must When trees become established in small open-
be within ± 10 years of ings in a sapling stand, they tend to "catch"
mean stand age the surrounding trees height causing an over-
estimation of site productivity
-------------------------

Various kinds of representations of the idea underlying the use of tree height as a measure
of site productivity can be given. Listed below are categories of kinds of representations and
their interpretation, listed in order of increasing complexity and generality (Bunge 1974a):
The Role of Change in Some Forestry Concepts 49

Table 3.2. Completed abstract representations of the "forest producing power of a site"
concept as currently viewed by indirect phytocentrists.

OBJECT forest site

SCHEMA forest trees (subject and other), "lower" habitat (abiotic environment),
"higher" habitat (other plants and animals, man)
other trees
SKETCH other plant~ subject tree - - abiotic environment
animals

other trees _ _
THEOR- (2,1) abiotic
ETICAL other plants - - (3,1) - - subject tree - - (0,5) - - environment
MODEL (4,1) - -
animals - - - - -

other trees _____


GENERIC (2,0)
THEORY - - subject tree - - (+,O)--"lower" habitat
(3,0)--
"higher" habitar-

object- that which is being elucidated,


schema- a list of items that are used in elucidating the object,
sketch- a display in outline form of the relationships among objects
in the schema,
theoretical model- a spelling out of the sketch,
generic theory- a theory free from specifics but convertible into a theoretical
model (specific theory) upon being adjoined a schema.

Filling in the specifics on schema, sketch, model, and theory for the object forest site
produces the results shown in Table 3.2. The schema includes broad categories of things,
for example, the central thing and things affecting it. The sketch shows that the concern is
how the central thing is affected by other things, not, for example, how other things affect
other things. The theoretical model specifies the relationships that must be obtained between
subject tree and elements of "higher" and "lower" habitat. That is, the subject tree undergoes
induced change only from the "lower" habitat. A key to the success of this approach is the
insistence that the relationship denoted by the numeral 1 is neutral (symbolized as 0). This is
made easier by using height as the property of the subject tree that is affected by interaction.
The kind of change (spontaneous or induced) need not be specified for relations denoted by
numeral 2, 3, and 4. It could be enhancive (+), detrimental (-), or neutral (0). Of course,
the induced change denoted by the numeral 6 is considered enhancing, since this is the
driving force behind height growth. Another key item is the assumption about the effect of
subject trees on abiotic environment-particularly the moisture nutrient regime. Normally,
a neutral effect is assumed, but examples have been given where effect 5 is + (enhancing)
or - (detrimental) (e.g., Carmean, et ai., 1976). The property of the abiotic environment
that undergoes induced change is its capacity to support elements in the "higher" habitat
(Haskell 1970).
50 The Role of Change in Some Forestry Concepts

The generic theory, the last step of this representation process, is an approach to as-
sessing the intensity of an induced change on a subject object by its "lower" habitat. This
is accomplished only after neutralizing any induced changes the set of "higher" habitat el-
ements can exert. A physical science analog of this approach is the use of a test body to
assess the potential strength of a force field.
The theoretical model presented in Table 3.2 is a special case of the generic theory.
It is abstracted from ideas contained in the literature on site productivity measurement.
Presented in terms of interactions, it
a) makes explicit assumptions sometimes overlooked
b) allows one to systematically consider complications or opportunities that might arise if
not all assumptions are met (e.g., effect 5 is enhancing, or effect 5 is detrimental)
c) allows one to do (b) within the framework of an existing and successful theoretical
model.
The. concept of site productivity can be subjected to the same kind of logical/semantic
analysis as any other predicate. Applying Bunge's idea that a predicate is a propositional
function that maps objects into statements, the following formula results:

SP: DT x OT x ABF x OC -----> S,

where
DT designates the set of dominant trees undergoing change induced only by ABF,
OT designates the set of other trees undergoing either spontaneous or induced change,
ABF designates the set of abiotic fluxes (energy flow and matter cycles) under sponta-
neous change,
OG designates the set of other organisms under spontaneous or induced change,
S designates the set of statements that contain (make use of) the predicate site pro-
ductivity.
Because there is no change, either spontaneous or induced, outside of things, these two
components of change have not been included in the argument list, although they could have
been.
The reference class of the site productivity concept is the set formed by the domain
elements of the propositional function above:

R (SP) = {DT,OT,ABF,OC}.

I have purposely avoided using the term site index in this discussion. To endorse the
concept of tree height under conditions described in Table 3.2 as a measure of site produc-
tivity does not imply endorsing the use of a single number (ordinate (scalar) value at an
index age) to characterize the variety of possible patterns of height development over time.
The scalar assessment of tree height pattern seems very naive from a mathematical point
of view. At a minimum, quantitative estimates of site productivity based on tree height
patterns would seem to require a vector of numbers to adequately represent the process.
It appears that the indirect phytocentric view has achieved its current dominance be-
cause its adherents have, over the past 60+ years, hammered out a theoretical model of the
idea. For a wide range of conditions, the model appears to function quite adequately. What
is needed is to learn from the successes to date of the indirect phytocentrists and to use what
has been le.arned to progress toward the goal of a half century ago-a geocentric view.
I now switch attention to measures of space occupancy to see if or where spontaneous
and induced change enters their formulation.
The Role of Change in Some Forestry Concepts 51

B. Measures Of Space Occupancy

Forest scientists have long sought measures to characterize the state of the forest tree
subsystem of forests. Since it is not possible to assess more than a few properties of the
subsystem, the desire is to develop a measure that integrates as many properties as possible
into a single value. Numerous measures have been suggested. Bickford, et a!., (1957) reviewed
the status of two prominent measures: stocking and density. I consider both to be measures
of occupancy of physical space by trees in a stand. At one time stocking was more of a
modifying term than a concept in its own right; an example cited by Bickford, et a!., reads
" ... density of stocking in terms of number of trees, basal area, volume, or other criteria on
a per acre basis." These authors went on to suggest that the former situation had caused
much unnecessary confusion. " ... it is evident that stocking is a loose term at best. It has
been used in many senses, .... There are also varying shades of meaning as the term is
used in silviculture, management, and economics. Confusion in the use and interpretation of
[']stocking['] should not be surprising." Their recommendation was that stocking and density
be dissociated and assigned the meanings: "stocking is comparative to the stand desired and
density is expressed numerically." (Bickford, et a!., 1957).
From a logical point of view, this recommendation is inadequate. However, from their
ensuing discussion, it appears they wished to make this distinction:
(i) stocking is the actual quantity of standing crop in relation to an ideal quantity of
standing crop, and
(ii) density is the actual quantity of standing crop in relation to metric (meter stick) or in
relation to actual "normal" quantity, e.g., "full" or "normal" density (MacLean 1979).
This distinction places both on a comparative basis, one to an ideal stand, real or imagined,
and the other to a meter stick, or to "normal" or "fully stocked" stand.
At about the same time the Bickford report was published, there began to appear
findings on the relation of the individual tree to its growing space (Krajicek and Brinkman
1957; Krajicek, Brinkman, and Gingrich 1961; Smith 1963; Newnham 1964; Opie 1968; Stage
1969; Lin 1969; Gerrard 1969; Bella 1971; Keister 1972; Honer 1972; and many others). For
some the relation is expressed on a per unit basis, for others it is an index applicable to
a subject tree. In either case, the expression is basically different than both stocking and
density. For lack of a better name, I call this idea 'packing,' where
(iii) packing is the actual quantity in relation to possible quantity when undergoing sponta-
neous change.
The "possible quantity when undergoing spontaneous change" is often expressed as
"growth of a tree free from competition," or "open grown," or sometimes as "physical di-
mension of a tree when open grown." The interaction connoted in all these statements is
either reciprocal neutrality (0,0) or at least (O,±) in the relation below:

Subject Other
tree trees

For both "stocking" and "density," quantity refers to standing crop quantity. However,
in terms of crown competition factor (CCF) as developed by Krajicek, et a!., (1961) and
modified by Honer (1972), I must interpret "quantity" to be physical space (ground area),
52 The Role of Change in Some Forestry Concepts

in which case "packing" is as follows:

actual quantity area covered by projection of existing crowns to ground


level. If the stand is maximally stocked and occupies ex-
actly one acre, then actual quantitiy is 43560 sq. ft.
possible quantity = ground area that would be covered by crowns of trees grow-
when an interaction ing on the acre if each tree had its identical stem thickness
holds yet had crown dimensions as if it were growing free from
crown competition (under spontaneous change).

Table 3.3 Summarizes a variety of space occupancy measures.

-7'ti"
Table 3.3. A Classification of space occupancy measures.

based on human ideal, e.g.,


""bj,d'"
full stocking
well stocked
absolute - relative
(without ~o another stand) ~With reference to another stand)
based on based on
diameter, e.g., basal area
--------
induced 7hange spontaneous change
height, e.g., Wilson's 1) number of trees VS. mean 1) open grown tree dimen-
dbh, e.g., stand density sions e.g., CCF (Krajicek)
index (Reineke) CW /H (Honer)
diameter and height, e.g., 2) mean tree volume vs.
bole area (Lexen) trees/acre, e.g., relative
density index (Drew and
Flewelling)
3) relative basal area, e.g.,
(actual)/(normal) = .5

Employing the methods developed in Chapter 1 to analyze predicates, there are three
different propositional functions, as follows:

BT : AQ x IQ --> B

D : AQ x NQ --> B

P: AQ' x PQ --> B,
where
BT designates "stocking," D "density," P "packing,"
AQ designates actual standing crop under induced change,
IQ designates ideal standing crop under induced change,
NQ designates normed standing crop under induced change,
AQ' designates actual physical area occupied by the standing crop undergoing induced
change, and
The Role of Change in Some Forestry Concepts 53

PQ designates maximum possible physical area occupied by the standing crop under-
going spontaneous change.
The reference classes of the three concepts are easily formed:

R (ST) = {AQ,IQ}
R (D) = {AQ, NQ}
R (P) = {AQ',PQ}.
The packing concept requires the comparison of two quantities, one of which is tied to
spontaneous change. As a result it does several things:
a) it clearly identifies packing as something having fundamentally different presuppositions
than both stocking and density,
b) it ties the measure to an objectively determinable quantity, and
c) it permits inclusion of information from several levels of the organizational hierarchy
(tree morphology, physiology, genetics) into the expression.
The result is that packing is both a rich concept (in the information sense) and strong concept
(being objectively determinable). Density is either objective and barran, or, like stocking, is
subjective.
Let us now look briefly at the role of accretion in forest growth models.

c. Accretion
Gross change in the tree portion of a forest system can be viewed as having two parts:
change in the number of trees and change in the size of trees not coming into or leaving
from the system. Mathematically, this can be expressed as the total derivative of S . N with
respect to time, where S refers to mean size of the individuals and N refers to the number
of individuals, i.e.,
d(SN) -dN dS
--=S-+N~.
dt dt dt
The first change element, df:, is the net difference between new individuals entering the
system through birth or reproduction and existing individuals leaving the system through
death (mortality). The second change element, ~f, is normally referred to as survivor
growth, or, more precisely, growth on surviving trees. Total growth on surviving trees is
simply the sum of the accretions of individual trees. My objective here is to examine the
role of spontaneous and induced change in the survivor growth or accretion concept.
Because trees retain a partial record of their past states via their annual rings, forest
stands also contain collective partial records of their past states. In forest tree systems, as
opposed to forest tree aggregates, these partial records can be assessed to study induced
changes from two sources: tree~tree effects and "lower habitat" ~tree effects. A primitive
view of the accretion concept can be summarized in the following relation:

A : ST x OT --+ S.

That is, the accretion concept maps the Cartesian product of the set of subject trees, and the
set of other trees into statements, S. In a more realistic formation of the accretion concept,
the domain of the mapping should be expanded to include the "lower habitat" or site factor.
The modified accretion concept maps the Cartesian product of the set of subject trees, the
set of other trees, and the set of site factors into the set of statements, S, that make use of
the new accretion concept.
In recent years, apparently beginning with the work of Staebler (1951), the accretion
concept has been expanded by some persons to include spontaneous change of the subject
54 The Role of Change in Some Forestry Concepts

trees. Of course there remains induced change in the subject trees due to "lower habitat."
The importance and extensibility of the idea of introducing spontaneous change of the subject
trees is easily understated. Its effect is, in sum, to modify the argument list further, i.e.,

A" : ST l x ST 2 x OT x SF -> S

where
ST l designates the set of subject trees under other tree and site factor induced change,
ST2 designates the (hypothetical) set of subject trees under site factor induced change
only (I call this spontaneous change.),
OT designates the set of other trees,
SF designates the set of site factors, and
S designates the set of statements.
Armed with these specifications of domain elements, the reference classes of the three
different accretion concepts is clearly spelled out. Further, the reference class of propositions
making use of either of the three accretion concepts can be easily computed. For example,
determining the reference class of "Forest stands have time invariant mortality accretion
ratios," turns immediately to identifying the domain elements in the propositional function
form of the accretion and mortality concepts.
I show in Chapter 6 that the inclusion of spontaneous change in the accretion process
opens doors for synthesis that remain closed with out it, such as when one uses mean induced
change.

D. Assessing The Animal Producing Power Of Forests


A strong parallel exists between the historic problems of assessing the "forest producing
power" of a site and that of assessing the "animal producing power" of a forest. Each problem
has a school of thought advocating that attention be focused on the "field" - the soil complex
in the former case and the vegetation complex in the latter. Members of the former I have
been calling geocentrists. Those scientists who emphasize habitat quality assessments may
be referred to as phytocentrists at this level. They seek causes for observed characteristics
of the test body, the animal populations.
After many years of trying to unravel the workings of the soil system, in both its biotic
and abiotic aspects, to gain understanding of the processes that lead to a forest there appears
to be less confidence that it can be done to a satisfactory degree. Increasingly, attention is
being paid to the "test body" (tree or plant community, including herbaceous plants) as the
supreme device for "sensing" the "field."
Clearly, there must be knowledge of both test body and field in both kinds of assess-
ments. The framework for formalizing theories shows possible approaches that can be used
in this assessment task. Ingredients of a theory are:
primitives + definitions + axioms ----> theorems + corollaries,
with those on the left related to those on the right via logical deduction. The primary
claims to knowledge are in the axioms, while primitives and definitions playa clarifying
role. Recall from Chapter 2 that there should be an axiom for each major assertion in the
argument. Axioms are more than hypotheses or guesses, but they may not be full-fledged
law statements (corroborated hypotheses) either. Thus, one should be able to supply at
least some evidence supporting every proposition included as an axiom (Leary 1980). No
doubt, evidence for some will be better than for others. When the argument takes the form
of logical implication, that is
if A then B,
the antecedent A will normally be the conjunction of a set of axioms expressed as proposi-
tions.
The Role of Change in Some Forestry Concepts 55

The differences in the assessment tasks can best be seen by focusing on the referents
of the factual propositions that form antecedent axioms. In a geocentric theory of forest
productivity, most axioms deal with aspects of the soil-climate complex. Fewer deal with
the vegetation. The entailment relation produces consequents that deal with vegetation.
Since we test theorems, items in B, not axioms, the forest produced on a site must be
measured to test any predictions or deductions stated quantitatively.
A phytocentric theory of the animal producing power of a forest would have its focus on
the vegetation. Hence, most axioms in its theoretical formulation would deal with vegetation.
Not all, of course, but most. Other axioms would deal with relationships between animals
and plants. The theorems, in this view, would be deduced from the axioms and deal with
selected properties of the animal complex.
A zoocentric theory of the animal producing power of a forest would, on the other
hand, have its axioms and other presuppositions focused primarily on animal needs from a
physiological perspective. The theory's consequents expressed in its theorems would deal
with properties of the forest habitat based on the conditions of the animal complex. Tests
of the zoocentric theorems would involve measuring the habitat in various ways mandated
by the process of logical deduction.
Recall that an important aspect of the relative success of tree height as a phytocen-
tric expression of the forest producing power of a site is the relative invariance of tree height
growth to plant sociological factors. Whereas lateral stem increment is very sensitive to stand
packing, longitudinal stem increment (height growth) is very much less sensitive. Clearly,
this aspect of tree physiology is the key to the success of height as a measure of site produc-
tivity, and is the key antecedent axiom for the phytocentric school of assessing a site's forest
producing power.
Continuing the analogy, then, a zoocentric approach to assessing the animal producing
power of a forest must identify properties of an animal's physiology that function in the same
manner as the dominance of apical meristematic activity over lateral activity in plants. The
process must obey all the relationships given in the theoretical model in Table 3.2; it must be
influenced by its lower habitat (the vegetation and soil complex) yet not be affected by animal
sociological processes. This, it seems, would eliminate current population or reproductive
levels from consideration as a measure of the animal producing power of a forest.
In sum, the key individual in assessing the animal producing power of the forest may,
perhaps, not be the animal ecologist as much as it is the animal physiologist. Some evidence
in favor of this reasoning has been found in the use of size of horns in Dall sheep for assessing
habitat quality (Bunnell 1978; Heimer and Smith 1975).

E. Discussion

The concepts described in this chapter have been invented to account for aspects of
reality perceived by forest scientists. What seems to have gone unstated is that the concepts
are appropriate for a particular ontology - a possibilist ontology. Bunge (1977) argues that
since every thing changes, change must have been possible in the first place.
Each concept considered here reflects a possibilist ontology: the possibility of change
under the condition of no effect from "higher habitat" (see Table 3.2). In some cases, e.g.,
the concept of stand packing, an actualist ontology, what actually happens, is also used. The
combination of ontologies, possibilist and actualist, with notions of spontaneous and induced
change, appear to have wide application in inventing new forestry concepts.
In sum, there are no things in forests that do not change. Hence, many of the concepts
used to represent properties of forest systems have referents that are implicitly or explicitly
associated with a particular kind of change. Two classes have been considered: spontaneous
(possible) and induced (actual). The role played by these two classes is examined for a)
56 The Role of Change in Some Forestry Concepts

measures of the forest producing power of a site, b) measures of space occupancy by trees,
c) models of accretion in forest stands, and d) measures of the animal producing power of
a forest. Concepts that ignore spontaneous (possible) change are deemed less useful than
those making use of it.

F. Research Problems
1. Pursue Bentley's argument that it is the peculiar concept set that differentiates foresters
from other scientists. What concepts do you think constitute the minimum set needed
for this differentiation?
2. Analyze the logic in the statement" ... stocking is comparative to the stand while stand
density is expressed numerically" (Bickford et aL, 1957).
3. Examine the concept of density in other scientific disciplines and in the European
forestry literature. Compare your findings with the concept as described in this chapter.
How are they different? The same?
4. Survey the extensive literature on forest growth models. What fraction make use of
spontaneous change?
5. Assemble and review recent literature on forest stand stocking and density. How many
authors use the term density to designate the concept of packing? How many use density
to designate stocking?
6. Devise yourself, or search the literature for, other concepts that "measure" the wildlife
producing power of a forest. Develop an abstract representation, as in Table 3.2, to
support your idea. (See Seal (1977) and Anderson, et al., (1972) and more recent works
by these and other authors.)
7. Tourney (1924) states "the term site as used by foresters is the exact equivalent of
habitat as used by ecologists" Apply the ideas discussed in Chapter 1, and applied in
Chapter 2, to the 1924 version of forester's site and ecologist's habitat. How do you
plan to assess equivalence between the two? What is your assessment of the truth value
of Tourney's proposition?
8. In pursuit of a geocentric measure of site productivity, apply the phytocentrist's ap-
proach of observing a "test body" (a group of plants) as it reacts to a "field" (the
abiotic complex), but go down one level and consider the abiotic complex as the "test
body." What is the "field" at this new level? Refine the concept of an abiotic complex.
(See Jenny 1958,1961.)
9. Develop an abstract representation for the results of problem 8, complete with object,
schema, sketch, theoretical model, and generic theory.
10. In completing research problem 8, it may be necessary to identify a property of the
abiotic complex that is sensitive to its "lower habitat "but is not affected by its "higher
habitat. " What are some candidate properties?
4: Interaction
Geometry

"Interaction" is a class concept. It is also a relation concept of a particular kind, compar-


ative relational. This means that class membership can be assigned to a particular natural
object system only after comparisons have been made. The comparison is of an ordinal na-
ture; that is, I assert that under induced change the numerical values of variables designating
attributes that represent elements in the system's property set are 'greater than,' 'less than,'
or 'equal to' the values for the variables when the system is undergoing spontaneous change.
The purpose of this chapter is to show the transition of "interaction" from a mathematized
concept to a fully quantitative one taking the form of a mathematical coordinate system.
The transition will be accomplished in two stages, the first treating the fundamental parts
and the second treating the manner of part combination.

A. Fundamental Parts
There are two constructs that form what I call the fundamental parts of the interaction
coordinate system. One is based on a relational comparison, and the other is based on a
metric comparison.

1. The comparative relational form of "interaction."


The comparative relational form of "interaction" was first published III 1947 (Haskell
1947). Its derivation was explained two years later as follows:

°
In order to determine not only all observed, but all possible coactions ... , I substi-
tuted for the rate at which any given activity is carried on by each class in the
absence of the others, + for all accelerations of this activity in either class by the
other, and - for decelerations. Cross-tabulation of +, 0, and - gives all possible
coact ions as shown in [Figure 4.11 (Haskell 1949).
The orientation and contents of the tabulation in Figure 4.1 are identical to those specified
by Haskell, however labelling of the margins has been expanded and made, hopefully, more
clear. Each non-null interaction may have system conditions and mediary that cause it to be
classified in one of the corner cells of the coaction cross tabulation (Figure 4.1). For example,
if X and Yare different kinds of trees in a mixed stand, the "process" is nutrition, and the
property set elements are tree central stem dimensions, then the system interaction will in
most cases be (-, - ).
One step enroute to a more quantitative form of interaction is shown in Figure 4.2.
Published in 1949, this advance was ignored for 30 years if its use in published literature is
to be the judge (Lidicker 1979).
Perhaps a partial reason for the delayed recognition is that a key step in geometrizing
the coaction cross tabulation was revealed only recently. This sequence of steps is shown in
Figure 4.3. It is made possible by :> recognition that the body of the cross tabulation may be
58 Interaction Geometry

Effect of population of entity Y on


the normative (process) rate of a
population of entity X as measured
by (attribut es of earmark set)
o +
Effect of population
+ (-,+) (0,+) (+,+)
of entity X on the
normative (process)
rate of population
o (-,0) (0,0) (+,0)
of entity Y as
measured by (attri-
butes of earmark set)
(-,-) (0,-) (+,-)

Figure 4.1. The comparative relational "interaction" with place holders for process and
property set.

~70'

I
Figure 4.2. Geometrization of the coaction cross tabulation as it appeared on the cover of
Main Currents in Modern Thought summer 1949. Haskell calls this the coaction compass.
(With the permission of the Editors, Main Currents in Modern Thought.)

considered a region of two-dimensional space (Figure 4.3b). Once realized, the coordinate
system followed as described in the next paragraph.
The coordinate system in Figure 4.3c can be derived from the cross tabulation, Fig-
ure 4.3a, because two lines is all the space needed to represent interactions involving a single
Interaction Geometry 59

a
Effect of Y on X
0 +

>
c:: + (-,+) (0,+) (+,+)
0

..-
><
Y
0 0 (-,0) (0,0) (+,0) I
u (0,+)
=
CI

w (-,-) (0,-) (+,-)


(+,+)

Cl

---,
b -(-,0)+----(0,0)---+-(+,0)- X
Y

r---
(0,+)
I (-,+) (+,+) I
I I (-,-)

I (-,0) (0,0) (+,0) I X (0,_(-(0,0)


I I I
IL _
(-,-)
__ (0,-) _ _ _ -.JI
(+,-)

Figure 4.3. Evolution of the coaction cross tabulation to an interaction coordinate system
(Haskell 1973). The right-hand figure shows the (0,0) relation as a point (c), and as a circle
(d).

neutral effect (( -,0), ( +,0)' (0, - ), (0, + )). Also, the (0,0) interaction needs a dimensionless
space (a point) for its representation. This leaves the four interactions involving bi-ordinal
relations ((-,+),(+,-),(-,-),(+,+)) to occupy the four quadrants. In Figure 4.3c, the
"+" in (+,0) is the effect of Y on X and "0" the effect of X on Y. Another form of the
coordinate system, shown in Figure 4.3d, is possible if one considers a point as the limit of a
circle, i.e., a circle with zero radius. Henceforth, I will refer to Figure 4.3d as the relational
coodinate system.
Before leaving ordinal relations, I call attention to the symbol for reciprocal neutrality
(0,0). To avoid possible confusion of this symbol for relations between populations with
that for the numerical expression for the absence of both populations 0,0, I have in all
cases enclosed the relational zeros in parentheses. Nowhere are numerical zeros presented in
parentheses.

2. The quantitative form of "interaction."


The quantitative approach to studying interactions may be considered the classical
approach. First used extensively by Gause (1934), it consists of studies of organisms under
spontaneous and induced change. Typically, some measurement or assessment of population
size is made for all three conditions as the experiment progresses. The population sizes
may then be expressed geometrically in the familiar time series graph. Too many studies of
interactions stop at this point even though several extremely simple extensions are possible.
One such extension is to eliminate time from one axis, replace it with standing crop of
one of the organisms, and include time parametrically. The result is a graph called the phase
60 Interaction Geometry

plot and is said to be located in a phase plane. In many cases just two points, equilibrium
points for spontaneous and induced change, are plotted in the phase plane. Because of the
artificial nature of equilibria, my approach is to consider, instead, all points in the time
series and let the progression of these two points trace out a trajectory in the phase space.
Figure 4.4 shows the trajectories produced by an example experiment analyzed elsewhere
(Leary 1976). Still another extension is to develop a method of comparing the position of
each trajectory at any time of interest. A simple measure of direction and distance from the
spontaneous change, (0,0), trajectory to the induced change, - (0,0), trajectory will suffice.
And this requires only simple trigonometry (Leary 1972). It is in this manner that the (0,0)
trajectory serves as the reference framework from which to measure induced change.
Three classes of (0,0) phase plane trajectories deserve attention. One class extends into
the two dimensional region of the phase space. Organisms with this type of (0,0) trajectory
can live independently of each other and are the basis for facultative relations. Another class
of trajectories extends only on one axis of the phase space. One of the organisms with this
type of (0,0) trajectory cannot live independent of the other. Collectively these constitute
facultative-obligatory relations. The third class of (0,0) trajectories quickly proceeds to
the origin of the phase space from nonzero initial conditions. Neither organism can exist
without the other: it is an obligate-obligate relation. This class of (0,0) trajectories is found
in mutualistic relations.

600

1"'1 500
...E
400
c
300
"'
...
G>

::I
• pu re (0,0)
"' • mixed (0;0)

100 200
P. eaudatum in .5 em 3
Figure 4.4 Phase plane trajectories for spontaneous and induced change. By comparing
positions 01 trajectories at a common time, it is possible to compute the angle and distance
from the (0,0) point to the - (0,0) point. The angle () gives the type of interaction, and the
distance z gives one measure of interaction intensity. Numbers in the figure indicate days
since the study was begun (Leary 1976).

B. Combining Fundamental Parts


The relational and metric constructs provide the parts to be combined. The sense of
"combination" I use here involves juxtaposition of parts. Specifically, I take juxtaposition
to mean embedding one framework into another. The latter I call the host, the former,
the guest. The embedded framework is considered, then, to be subordinated to the host
framewotk.
There are four possible combinations of the two parts taken two at a time. They are
shown in Figure 4.5. Each combination that includes a phase plane has the two trajectories
Interaction Geometry 61

included from Figure 4.4. Figure 4.5a shows phase plane in phase plane. Here, the host
framework is quadrant I of the Cartesian coordinate system. The guest is all four quadrants
of the Cartesian coordinate system with positive axes and negative axes in their normal
orientation. The latter represent decreases from the numerical values associated with (0,0)
trajectory. Figure 4.5b shows the relational coordinate system as a moving reference frame
riding the (0,0) trajectory. At each point in time a line connecting (0,0) and ~ (0,0)
trajectories intersects the (0,0) circle to mark the type of interaction. Figure 4.5c shows
the (0,0) phase plane trajectory riding the (0,0) circle of the relational coordinate system.
Figure 4.5d is the only one without a metric, so the guest form of the relational coordinate
system can be embedded only in one of the host's four quadrants or on one of its four axes.
In contrast with the previous three combinations, where a continuum of positions is possible,
this combination allows only eight different positions, which would be equivalent to asserting
that there are only eight important types of interaction. For this reason Figure 4.5d is not
considered further.

"HOST" FRAMEWORK
"GUEST"
phase plane relational coordinate system
FRAMEWORK

phase plane

v ..
relational
coordinate
system

Figure 4.5. Four possible combinations of relational coordinate system and the phase plane.
Combination a) shows the phase plane as both the host and guest framework. Combination
d) shows the relational coordinate system as both host and guest. In these two cases the
guest framework is a variation of the host framework.

Examination of the alternative combinations in Figure 4.5 shows that, combination d


excepted, each requires an assessment of direction and distance from one point to another,
62 Interaction Geometry

Ie., from one relation to another. This brings to mind polar coordinates where, for any
two points, direction is measured from an oriented line, and distance is measured from a
privileged point on the line, normally at its left-most end. For example, the coordinates of
the point X below are given by rand e. The grid marks on the oriented line specify the unit
of distance. Here, r = 3.75 units of distance.

In a sense the relational coordinate system resembles a polar coordinate system. Its
resemblance can be shown as follows:

-+--------'1-----'-----+-(+,0)

Here the (+, 0) axis becomes the oriented line from which angles are measured, and a
point on the (0,0) circle becomes the privileged point from which distances are measured. A
serious deficiency exists, however, because the lines in this figure are just that-lines. They
are not metric axes; they are relations. Hence, there are not the reference axes to scale off
r, assuming it is known quantatively.
The deficiency is overcome by embedding the phase plane in the relational coordinate
system as shown in combination c of Figure 4.5 and below with metric axes "graduated."
It is the combination of the relational coordinate system as host and the phase plane as
guest that will be used throughout this book. This method of combination was first described
by Haskell and Cassidy (1960) and developed extensively by Haskell (1972). The method
suggested here is somewhat more flexible in that the point representing spontaneous change
of both populations, the (0,0) point, "floats" on the (0,0) circle of the relational coordinate
system. In fact, it is only the relation of reciprocal neutrality that the two frameworks share.
It follows, then, that the shared property should be brought into coincidence when the two
coordinate systems are combined.
This combination was selected in part because my prime interest is interactions, hence,
the host framework is interaction-based. Had my prime interest been absolute quantities,
the host framework should have been the phase plane. I can not argue that one method
of'combination is better than another without clearly specifying a primary objective, which
Interaction Geometry 63

---...1...----+----(+,0)

in turn specifies what is to be dominant and subordinate. All aspects cannot be dominant
simultaneously.

c. Mapping Experimental Results


In order to establish precisely the location of an experimental result in the coordinate
system two items are needed: the type of interaction, given mathematically by the angle 0,
and the intensity of interaction, given by the distance from the (0,0) point to the ~ (0,0)
when both are located in the phase plane. In practice, both computations are made using
phase plane coordinates, so no embedding of the coordinate system is actually carried out
in a physical way. Of course, it could be if experimental results are located in a phase plane
drawn on transparent paper.

1. Direction from (0,0) to ~ (0,0).


To know the direction from point to another in a plane, the coordinates of both the
(0,0) and ~ (0,0) points need to be known. The length of the sides of a right triangle must
be computed and the appropriate angle determined from trigonometric functions, i.e.,

-(0,0)

then Q = tan1y/x, and


0= 180 0 - Q.

Caution must be exercised in selecting the scales for ordinant and abscissa because they
determine the angle o.

2. Distance from (0,0) to ~ (0,0).


The distance from one point to another in the phase plane is not so easily determined as
direction. To date, no one distance measurement has been found that is totally satisfactory.
Probably the most familiar setting of distance measurement questions is the 3-dimensional
64 Interaction Geometry

physical world. Distance traveled, or distance "as the crow flies," is often the primary
interest. The formula for computing such distances is

when I = 90. What z means in physical space can be intuited: If it takes X time units +Y
time units to traverse the legs of the triangle, it should take

time units to traverse the diagonal (hypotenuse) directly from (0,0) to ~ (0,0). The meaning
of z is not at all clear when this formula is used to compute distance in a more abstract
space such as the phase plane. For example, what is the interpretation of z in Figure 4.4?
Needed is a measure of distance that tells us something about the nature of the process
taking place in the system. This problem is particularly acute when x and yare of opposite
slgn. Consider, for example, the distance measure

z' = Ixl + Iyl,

the so-called Manhattan metric. Since both x and y express changes between (0,0) and
~ (0,0) positions, z' does convey some sense of the amount of "action" that has occurred
under ~ (0,0) conditions. Of course, an equal contribution to "action" is made by a decrease
or an increase in population levels.
A measure of net change in the system that I often use later is given by the simple
equation

If x and yare of opposite sign, the "net" may be near zero (z -> 0), yet there may have
been lots of "action" (Ixl + Iyl > 0). Of course, when the signs of x and yare the same,
z and z' are identical. Population "exchange," tradeoff, or substitution rates are given by
ratios of increase and decrease, e.g., x/yo With either of these measures, the result has
some biological interpretability, whereas the Pythagorean measure of distance seems to defy
a biological interpretation.
The base population levels for the (0,0) relation largely control the magnitude of the
increases or decreases, especially the latter, so it is appropriate to normalize any distance
measure by expressing it as a fraction of the population levels for the (0,0) relation. The
(0,0) population level is given simply by summing the coordinates of the (0,0) point in
Figure 4.4. Of course, the coordinates must be conformable for addition so that one does
not try to add apples and oranges. A brief summary of these ideas is given in Table 4.1.
The (0,0) and ~ (0,0) coordinates have been specified in the table heading. The values
Rand R' express "distance" as a fraction of the summed coordinates of the (0,0) point.
Notice that the normalized interaction intensity values differ considerably depending on the
formula used for computing vector length. The resulting vector norm differences can be large
in quadrants II and IV (90 0 < 8 < 180° and 270° < 8 < 360°). However, in quadrants I and
III both norms give identical results.
The convention I use in plotting experimental results in Haskell's Coordinate system can
be stated as follows: (a) if 8> 315° and 8 < 135°, plot the ~ (0,0) point at (8, R), scaling a
distance outside the (0,0) circle equivalent to R times its radius, and (b) if 135° < 8 < 315°,
plot the ~ (0,0) point inside the circle at (8, R'), scaling a distance inward equivalent to the
R fraction of its radius (Cassidy 1972). The feasible regions of ~ (0,0) points is shown in
Figure 4.6.
Interaction Geometry 65

Table 4.1. Sample computation of type of interaction (II) and two measures of interaction
intensity (R, R'). The computations are based on the together (induced change) and separate
(spontaneous change) conditions: ~ (0,0) = 50,50 ,and (0,0) = 94,21 using both the net
(z) and total action (z') metrics.
~ (0,0) - (0,0)
,
II = 180 0 z= z' = R= ~-
II (0,0) II R'= II (0,0)
z
II
x y -tan-1y/x Ix+ yl Ixl + Iyl
-44 +29 146 0 15 73 .63 .13

Figure 4.6. Location of two regions of coordinate space where ~ (0,0) interactions are
plotted, and their relationship to the type of interaction. Feasible region is hatched.

D. A Wrinkle in Interaction Space

By using the methods of the previous section to map experimental results into the
Haskellian coordinate system, a series of points for different times in the same experiment
can be located. A particularly interesting case is when interaction intensity decreases with
time toward a zero level. In this condition, induced change is nearly the same as spontaneous
change. Assume that type of interaction (II values) does not change as intensity approaches
zero. So long as intensity is not exactly zero, there is a type of interaction, but the moment
intensity becomes zero, type of interaction changes to (0,0). Since the (0,0) "interaction"
type is represented as a circle, each such type may be considered to be everywhere on the
circle. When an interaction intensity develops again, it may be of a different type than had
existed prior to the decrease to a zero level of intensity. The result of all this is that the
(0,0) interaction functions as a wrinkle in interaction space. It allows trajectories to jump
instantaneously from one side of Haskell's coordinate system to the other. Figure 4.7 gives
66 Interaction Geometry

I
(0,+)
(_,+)76~ (+,+)
I 4
I
I 4
7. I
population 6. :
B 5.1
B' --------¥---- -(-,0) (+,0)-
1·3
: .2
I .1
I
\2
+-z...(O,O)
I
I 4 4
I
(-,-) (+,-)
A'
(0,-)
population A I

Figure 4.7. Idealized change in interaction intensity for a constant interaction type character-
ized in the phase plane and Haskell coordinate system. Because intensity initially decreases
to zero, the trajectory moves toward the (0,0) circle. When induced and spontaneous change
are identical (point 4 coincides with A', B' in the phase plane) the system is everywhere on
the (0,0) circle. As intensity begins to increase again after passing through the point of
mutual spontaneous change (A', B ' ), the trajectory moves away from the (0,0) circle.

a diagrammatic representation of this phenomenon in quadrants II and IV. Such jumps are
not limited to these two quadrants, however. They provide an alternative way for evolution
in type of interaction to occur.

E. Haskellian Coordinates as a Fourth Order Predicate


A predicate is a property of some thing or object. To this point I have limited the objects
to natural objects. I now let them be conceptual objects as well. The function of a predicate
is to map objects into statements containing the word that designates the predicate. Recall
that predicates can have different orders. First order predicates are properties of individual
natural objects. Second order predicates are properties of populations of natural objects.
Third order predicates are properties of properties of populations of natural objects. Fourth
order predicates are properties of properties of properties of populations of natural objects.
I submit that Haskellian coordinates have fourth order predicate structure. The manner in
which this is determined is given below:

predicate designator predicate formula predicate order


net action N A: SG A x IGAB ----> S second
net reaction N R : SG B x IGBA ----> S second
net interaction NI:NAxNR---->S third
Haskellian coordinates HG: N I X RJ X Rz ----> 5 fourth

where
SG A designates the set of spontaneous changes in population A,
IG A B designates the set of changes in population A induced by population B,
5G B designates the set of spontaneous changes in population B,
IGBA designates the set of changes in population B induced by population A,
NA designates the set of net actions of population B on population A,
Interaction Geometry 67

N R designates the set of net actions of population A on population B,


N I designates the set of net interactions between populations A and B,
He designates the set of Haskellian coordinates,
Rl designates the set of real numbers (angles) [0,360], and
R2 designates the set of real numbers (intensities) [0, (Xl).
Haskellian coordinates geometrically order the set of possible interactions. Interaction
orders the set of possible net actions and net reactions. Net action and net reaction order
the set of possible spontaneous and induced changes.
Haskellian coordinates geometrically organize the interactions in a very special way,
allowing one to speak about 'neighboring' relations, and about the evolution of interaction
type from one neighbor to another, or via the wrinkle, from any type to any other type.
The evolution of real object populations is represented as the changing location of a point
or series of points (trajectory) in an abstract space. Haskell first published an early version
of the coordinate system in 1949, nearly one-third century ago. Only a handful of persons
have shown interest in it since. On the other hand, many have used Haskell's coaction cross
tabulation of effects to produce a matrix of interactions, but many have changed the order
of the effects on the table margin, thereby scrambling the interactions. The most frequent
rearrangement seems to be

+
°
+ +,+ 0,+ -,+

° +,0
+,-
0,0
0,-
-,0
-
, -

(e.g., May 1973)' although Schwemmler (1980) suggested:

+
°
+ +,+ -,+ 0,+
+,- - , - 0,-

° +,0 -,0 0,0

The former rearrangement puts the (+, +) interaction in the upper left and so prevents
combining the phase plane (where tradition has positive axes directed rightward and upward)
with a relational coordinate system based on this cross tabulation. The latter rearrangement
puts the (0,0) interaction on the margin of the cross tabulation and prevents forming a
relational coordinate system at all.

F. Research Problems
1. Find other formulas used to characterize a distance between two points or trajectories.
2. What mathematical requirements must all distance measures meet? Do either of the
two suggested distance measures meet any of these requirements?
3. What is the relation between a valid distance measure and the geometry of the space
in which it is used?
4. Invent a measure of distance that has biological significance.
5. Test the proposition: Advanced sciences make greater use of higher order predicates
that less adv;;tnced ones. How do you plan to conduct your test?
6. Some research articles in scientific journals claim to report studies of interacting popula-
tions, but they report only on ~ (0,0) relations. Find several such reports and examine
their findings. Would anything be gained by investigating the (0,0) conditions as wei]?
68 Interaction Geometry

7. Examine the benefits authors gained by disregarding the ordering of signs on Haskell's
coaction cross tabulation.
8. Invent your own measure of distance from the - (0,0) to the (0,0) interaction. Consider
the possibility that interaction space is not Euclidean. What distance measures are
appropriate if interaction space is Lobachevskian or Riemannian? How does one find
an arbiter for the geometry of an abstract space?
9. In a phase plane construct a series of pairs of points-some near to one another, and some
at a distance. Designate one of the points in each pair as the (0,0) relation and the other
as - (0,0). Map all the point pairs into Haskell's coordinate system. What contributes
to 'closeness' in the phase plane? What contributes to 'closeness' in Haskell's coordinate
system?
10. Find a forestry concept that expresses a relation between relations. If you can't find
one, invent one.
Part II
Analysis
5: Haskellian Coordinates in
Analysis: Searching for Pattern

The concept of pattern has an important role in many aspects of human endeavor.
There is an important distinction to be made, however, between the purposeful generation
of pattern and the discovery of pattern in nature. Art, music, and design exemplify the
former; science, and scientific research exemplify the latter. The following brief discussion
of pattern is organized around three questions: Where does "pattern" fit in the research
process? What is the concept of pattern about? How does one look for patterns?
The place of pattern in the research process has been viewed differently by various
authors. Judson (1980) emphasizes the visual discernment of pattern as a first step in the
process of knowing. Pattern is treated early in his book The Search for Solutions as shown
by listing the chapter headings in abbreviated form: investigation, pattern, change, chance,
feedback, modeling, strong predictions, evidence, and theory.
Ziman (1978) places his discussion of pattern recognition in a chapter entitled 'common
observation.' Special emphasis is given to the relation between human capabilities, indeed
preference, for processing information in pictures, and maps, rather than prose. Bunge
(1967a) says that the chief goal of scientific research is the discovery of patterns and introduces
pattern when discussing law in The Search for System sequence: problem, hypothesis, law,
and theory. Laws, says Bunge, are supposed to reconstruct objective patterns and law
statements are supposed to be corroborated or confirmed hypotheses, perhaps ready for
inclusion in a theory. Kaplan (1964) does not treat pattern until his discussion of explanation,
i.e., his pattern method of explanation. At this stage of the scientific research process,
explanation is sought and Kaplan uses pattern merely as a tool to that end.
In sum, the place of pattern in the scientific research enterprise can be seen to vary from
providing incentive to be curious (to investigate), to being the object of common observation,
to being the definiendum of science itself, to being a tool for applying scientific ideas for
explanations. There seems little pattern to pattern's place in science.
The pattern concept itself is not nearly so indefinite as its place in science. However,
there appear to be three different concepts, evidenced by looking at their predicate formulas:
1. pattern form (shape) ---> statements
2. pattern: form x repetition --t statements
3. pattern: form x repetition x regularity ---> statements,
where statements make use of the term 'pattern' designating the concept "pattern."
The first pattern concept is due to Webster and is probably not the concept designated
by 'pattern' in "the chief goal of scientific research is the discovery of patterns" (Bunge 1967a).
The second and third concepts are due to Margenau (1972). When regularity supplements
repetition, symmetry, and asymmetry, broken symmetry and other keys to unobservable
properties can result. Of course, the repetitions and regularities may be in time, physical
space, or state space.
72 Haskellian Coordinates in Analysis: Searching for Pattern

Perhaps the major use by scientists of pattern concept 3 is in generating expecta-


tions. Judson (1980) claims "[t]he most tremendous perception of pattern in the history
of science ... was Dmitri Mendeleyev's creation of the periodic table of the elements." The
full power of patterns to lead scientific research is again seen in Mendeleyev's work, wherein
he insisted that if the known facts of chemistry did not fit his pattern the facts were at fault
(Judson 1980).
How does one seek patterns? Ziman (1978) gives several scientific applications of the
maxim that 'seeing is believing.' Undoubtedly true, that maxim nonetheless gives the im-
pression that patterns should be directly perceivable, i.e., visually apprehendable. Although
science may end with the production of perceivable patterns, it does not begin with them:
science is more than keen observation. Most patterns of interest to science are not immedi-
ately apparent, hence must be discovered, or unearthed, so-to-speak, after much effort. This
in turn implies a great amount of believing before seeing.
If, in fact, most patterns of interest to science are "unearthed," what tools are there
for digging? Bunge (1967a) claims experiment is a "torch" to be used in seeking patterns.
In general, if no pattern is directly evident in a bit of nature under study, a scientist at-
tempts to measure some aspect of that bit of interest and run the numerical values (of the
quantitative variables (symbols) designating concepts that represent properties of the con-
crete thing) through especially selected conceptual "filters." Typically, different "filters" are
tested until a semblance of pattern is evident. If appropriately powerful, the pattern will
generate expectations and prognoses, and may some day lead to explanations.
Heuristically stated, the quest for pattern requires undertaking the following ordered
tasks:

Yes Yes Ye.:,....... focus on regularity


form repetition repetition ~
~ . ---+-.IS regu I ar
t
exists eXists ~

No foc~r~ity
No No broken facts
pattern wrong

Epistemologically stated, the quest for pattern involves three stages: 1) leave the plane of
perceptions via measurements, 2) use formal operations to associate those concepts linked
directly to the P-plane thereby forming new concepts that organize information in patterns,
and 3) return to a different bit of nature (the P-plane) with an expectation (Figure 5.1).
Stages 1 and 2 are typically executed several times before confidence is built to the point
where public disclosure of expectations seem "safe." Mendeleyev's boldness in publishing his
expectations (later shown to be essentially correct) remains an inspiration to many, judging
by the numerous claims to classification systems embodied in periodic tables (Muir 1962.
1969; Haskell 1972; Pianka 1974; Schwemmler 1980).
In the tradition of first things first, I attempt to gain some experience with the concepts
of type and intensity of interaction, and Haskellian coordinates. Later in this chapter there
will be time for anticipations. For experience, I examine some examples of competition, prey-
predator, parasite-host, and plant-herbivore interactions. Gause's studies of competition are
a good place to start.

A. Competition (-, -)
Studies of competition (-, -) are frequently reported in the ecology literature. Because
of the fairly strict assumptions made about the conditions under which studies of this kind
are conducted, I briefly review the experimental framework for a famous experiment reported
Haskellian Coordinates in Analysis: Searching for Pattern 73

plane of
perception construct field

measurements

NATURE

NATURE' c

Figure 5.1. Three stages in the search for objective pattern characterized using Margenau's
format. At stage a leave the plane of perceptions via. measurements. At stage b use formal
operations on the concepts linked directly to measurement, to produce new concepts. At
stage c expectations are checked against a different bit of nature.

by Gause (1934). I then examine his experimental results. When conducted in the laboratory
most competition studies are designed to keep the mediary (growing medium and space in an
experimental container) at a constant level and configuration. Thus, the change equations
governing mediary dynamics are

ds/dt = 0, (space)
dm/dt = 0, (nutrients)

starting from some initial level of each s(t = 0) = k z and m(t = 0) = k 1 . Three separate
experiments are needed to assess the type and intensity of a (-, -) competitive interaction:
1) species 1 growing by itself in k, and k z amounts of mediary, 2) species 2 growing by itself
in k, and k z amounts of mediary, and 3) both species 1 and 2 growing together in k, and
k z amounts of mediary.
The population levels from experiments 1 and 2 provide observations on the solution to
the two separate logistic- based growing equations:

dYJ/dt = b,Y,(K, - Yd/K 1 (1)

dYz/dt = bzYz(K z - Y z )/ K2 (2)

with Y1 (t = 0) = y, and Yzlt = 0) = yz. It can be assumed that the upper limit of population
levels K 1 and K 2 are functions of the mediary levels kl and k 2 .
74 Haskellian Coordinates in Analysis: Searching for Pattern

The population levels that result from starting YI individuals of population 1 and Y2
individuals of population 2 in the same mediary quantities ki and k2 provide observations
on the results of the interaction, hence on solutions to the competition equations

(3)

with YI(t = 0) = YI, Y 2 (t = 0) = Y2, KI and K2 determined from equation 1 and 2.


Equations 1, 2, and 3 can be used to synthesize the results of Gause's study. Once the
numerical constants have been estimated, trajectory comparisons will provide a new view of
the dyanmic nature of interactions by clearly separating type of interaction from intensity
of interaction. Other interaction studies reported later in this chapter will compare direct
observations and conjectures. In the process of comparing these studies they will show that
the Haskellian coordinate system can assemble information produced from good guesses or
solutions to simultaneous nonlinear difference equations with variable coefficents.
Gause (1934) was interested in examining the probable course of interaction between
several protozoan species. Considerable effcrt was exerted to develop
1. a constant food source for the protozoan (a bacterium Bacillus pyocyaneus cultivated in
petri dishes and a fixed quantity thereof placed in 10 em 3 of Osterhout's salt solution
where the bacteria could no longer reproduce),
2. a fixed physical environment (graduate tubes of 10 ee capacity with wool stoppers),
3. a fixed quality of the medium (the medium was changed daily by centrifuging the proto-
zoa into the bottom of the test tube, the medium above drawn off, discarded, and new
growing medium added in its place).
4. Precautions were also taken to prevent an accumulation of toxic materials in the test
tube when the medium was changed.
The course of development of two protozoan species, P. aurelia and P. caudaturn, was
assessed daily (except after the first day) by stirring the growing medium and removing a
.5 ee sample of liquid. The number of infusoria of each species in the sample was determined
by counting, after which the sample was destroyed. The growth experiments were replicated
as follows:

number of microcosms
microcosim contents (replications)

P. caudaturn separately 4
P. aurila separately 3
P. aurelia + P. caudaturn together 3

The population levels for the three conditions above are given in detail in Gause (1934)
Appendix I, Table 3. Because of certain fluctuations in the observations that appear to reflect
experimental and/or sampling errors, I have "smoothed" the observations by summarizing
them into mathematical equations.
The coefficients in the difference equation form of the logistic equation (1) were de-
termined from the P. caudaturn observations and equation (2) from the observations on P.
aurelia reported as noted above. The method of estimating numerical constants in the equa-
tions requi)"es solution of nonlinear multipoint boundary value problems (Leary 1970; Leary
and Skog 1972). Equations summarizing the pattern of development of P. caudaturn and
aurelia separately are estimated to be (respectively):
Haskellian Coordinates in Analysis: Searching for Pattern 75

6.Yd 6.t = .97OYI (203.65 - YI )/203.65 (4)


6.Y2/ 6.t = 1.188Yz(535.2 - Y z )/535.2 (5)

In estimating the coefficients in equations (4) and (5), the effect of sampling was included
by reducing the population level to .9 of its original level at the start of each day (except
the first day).
With equations (4) and (5) specifying how each protozoan develops by itself in the
growing medium, their mutual development can be approximated by equation (3), which
makes an adjustment to the righthandside of each change equation to account for the presence
of the other species. Clearly, if Q = 1 in equation (3), one P. aurelia has as much effect on
growth of P. caudaturn as one individual of the latter has on its own population. An identical
statement can be made for the coefficient 13 in equation (3).
An examination of the Q and 13 values necessary to permit equation (3) to characterize
the mixed culture development of the two infusaria species showed that Q and 13 cannot be
constant over the entire time interval of 25 days. By assuming Q and 13 relatively constant
over a five day period, the 25 days of the experiment were divided into 11 overlapping five
day subintervals 0 - 5,2 - 6,4 - 8, etc. Coefficients of struggle, Q and 13 were estimated for
each subinterval. A graph of the resulting trends is shown in Figure 5.2. Visual examination
of these points suggested that Q and 13 were, themselves, state variables changing signifi-
cantly during the course of the experiment. Because of the unique powers of the method of
estimating the numerical constants used here, it was possible to reformulate the equation
(3) into the larger system of equations:

6.Yd 6.t =.970YI (203.65 - Y1 - QYz )/203.65


6.Yd 6.t =1.188Yz (535.20 - Y z - j3Yd/535.20
(6)
Q = - (clexp(cZYz ) - .5)
13 =d 1 + dzt,
with CI, Cz and d l , d z , numerical constants to be estimated in the coefficient of struggle
change equation. System of equation (6) was calibrated using the mixed culture data from
Gause's experiment by using the observations on P. caudaturn and P. aurelia numbers as
boundary conditions on the first two equations. The estimates of Q and 13 shown graphically
in Figure 5.2 were not used in estimating Cl, Cz, d 1 , d z : they were used only for judging
appropriate algebraic form of the change equation for Q and 13. Final estimates of the
numerical constants are

Cl = 4.54, Cz = -.02872, d 1 = .1844, d z = -.26507.

The agreement between predicted and observed infusoria numbers for the mixed culture
are shown in Figure 5.3. The results of this analysis reveal two points:
1. The fact that P. caudaturn numbers get progressively smaller in the mixed culture while
P. aurelia does not increase up to its pure culture level suggests either a) an effect toxic
to P. aurelia has been introduced into or allowed to develop in the mixed culture, or
b) there has been over the 25 days natural selection of protozoans such that the P.
caudaturn has a much larger effect on P. aurelia per individual of P. caudaturn than at
the beginning of the experiment.
2. Given the manner in which the coefficients of struggle are changing, the ultimate out-
come of the experiment, as extrapolated using system of equations 6, is that P. caudaturn
supplants P. aurelia. This is contrary to statements by Gause (1934, pages 102, 103),
although Gause presents no evidence that he actually ran this particular P. aurelia vs.
P. caudaturn experiment to exclusion of caudaturn.
76 Haskellian Coordinates in Analysis: Searching for Pattern

• • •••••••••••••
o ••
-1

-2

10

6

4 • •
2
• •
(3 0

-2

-4

-6

-8

-10
0 2 4 6 8 10 12 14 16 18 20 22 24 26

tim., days

Figure 5.2. Estimates of coefficients of struggle, Q and /3, when they are assumed to be
constant over a five-day interval. Points are plotted over midpoints of the time intervals.

With equations 4, 5, and 6 calibrated, all information is available to do an interaction


analysis and plot the experimental results in the Haskellian coordinate system. Table. 5.1
shows population levels computed with above equations under pure and mixed culture, in-
teraction type (0) and interaction intensity (R).
Figure 5.4 shows the experimental results plotted in Haskell's coordinate system. Note
that net "action" intensity coordinates are used with interaction type to locate each point.
Table 5.1 shows that the intensity coordinates for days 5,6, and 7 are slightly different when
using absolute "action".
Are any observations forthcoming from Figure 5.4 that are not forthcoming from a time
series analysis of the same experiment? Gause, based on a time series, observed:
It is easy to see that the growth of the mixed population consists of two periods:
(a) during the first period (till the eighth day), the species grow and compete
for the seizing of the still unutilized energy (food resource). But the moment
approaches ... when all. .. taken hold of, and the total biomasses of the two species tend
to reach the maximum possible biomass under given conditions. (This happens on
the eight day; the total biomass is equal to about 210.) . . . (b) After this there can
Haskellian Coordinates in Analysis: Searching for Pattern 77

120 •
100 •

c: 80 • •
• •
60
• • •••
40 • •
20
• • •
Or-----------------------------------------~

only arise the redistribution of the already seized energy between the two species,
i.e., the displacement of one species by another (Gause 1934).
Based on Figure 5.4, several observations are in order:
1. The growth of the mixed population consists of five periods:
a) days 1-4: Until day 5 there is little interaction evident although some enhance-
ment is apparent for P. caudaturn over its pure culture development
level.
b) days 5-6: Following day 4, aurelia begins to be detrimentally affected although
the net interaction is still positive (gain to caudaturn exceeds loss to
aurelia) .
c) days 7-8: There continues to be a rapid change in interaction type but little
change in intensity. Between days 6 and 7, net effect became negative.
d) days 8-9: Change in interaction type slowed considerably but interaction intensity
increased significantly.
78 Haskellian Coordinates in Analysis: Searching for Pattern

Table 5.1. Example of interaction type and intensity computation based on estimated pure
and mixed cultures of P. caudaturn and P. aurelia. Population levels in columns 2-5 are
computed using the appropriate difference equation and represent number of individuals in
a .5 cc sample of the groWlllg medium. Two measures of interaction intensity, net action R
and absolute or total action R', are given. These values differ only in quadrant IV near 315 0
(see astericks at days 5,6,7). Theta values are based on tan-1y/x but adjusted for proper
quadrant.
pure mixed
culture culture z= R= z' == R'=
days rj r2 r2 - rj Ix + yl Nz Ixl + Iyl N
z

P.c. P.a. P.c. P.a x y ()

0 2 2 2 2 0 0 0 0 0 0 0
4 4 4 4 0 0 0 0 0 0 0
2 8 10 8 10 0 0 0 0 0 0 0
3 13 19 15 19 2 0 0 2 .06 2 .06
4 23 36 28 36 5 0 0 5 .08 5 .08
5 39 69 52 66 13 -3 351 11 .10 15 .14*
6 63 126 86 115 23 -11 334 12 .06 24 .13*
7 96 220 114 178 18 -42 293 24 .08 60 .19*
8 135 347 116 233 -19 -114 261 133 .28 133 .28
9 169 466 103 267 -66 -199 252 265 .42 265 .42
10 189 527 89 289 -100 -238 247 338 .47 338 .47
12 200 539 69 317 -131 -222 240 353 .48 353 .48
14 200 539 56 332 -144 -207 235 351 .47 351 .47
24 200 539 33 341 -167 -198 230 365 .49 365 .49

P. a.

2 4 P. c.
----_-,-t;> (+'0)
13/ -5
-6

(-,-) (+,-)
(0,0)
(0'-)

Figure 5.4. Gause competition experiment results plotted in Haskell's coordinate system
(from Leary 1976).

e) days 10-24: Neither interaction intensity nor type changed significantly for the sec-
Haskellian Coordinates in Analysis: Searching for Pattern 79

ond one-half of the experiment.


2. Multiplication of the P. aurelia population level in Table 5.1 by .39, and adding to the
P. caudatum population level, gives total biomass in P. caudatum units. At day eight,
the biomass is 210 units, the same as Gause determined. Further, maximum biomass
comes at an interaction intensity level of .27, not 0.0 as one might first expect.
3. The separation of type from intensity of interaction allows their covariation to be studied.
In the above figure, interaction was more intensive in (-, -) than (+, - ).
Additional observations on the course of interactions will be possible when examining systems
in other quadrants of Haskell's coordinate system.
Another widely studied interaction is prey-predator, to which I now turn.

B. Prey-Predator (-,+)
According to the interaction event analysis described in Chapter 2, (-, -) biological
competition involves a mediary such as physical space, light, moisture or another part of the
abiotic flux. Prey-predator interactions do not involve a mediary; contact is direct between
predator and prey. However, since the prey is often a herbivore, the configuration of the plant
material can significantly affect the outcome of prey-predator relations. For example, Gause,
et aI., (1936) found that the predator (Cheyletus eriditus) was able to rapidly diminish a prey
(Aleuroglyphus agilis) when the environment in which the two infusoria were placed consisted
of millet grains. The prey was easily locatable on the surface of the grains. When wheat
flour was used as food for the prey species, the predator had to exert greater effort to locate
prey. Hence, the predator needed 43 days to eliminate the prey instead of the 19 days needed
in the millet environment. A one-to-one mixture of wheat flour and millet was intermediate
in providing both food and cover to the prey, and the predator took 36 days before the
prey was eliminated. I conclude, then, that specifying the particular predator and prey
species does not determine the pattern of type and intensity of interaction. Characteristics
of a prey-predator system can be seen by examining the theoretically possible locations of
trajectories in quadrant II of Haskell's coordinate system. The drawing below shows four
idealized regions of quadrant II.

c
( -,O)-+_ _ _ _~D_~

Each region can be described as follows:

A small reduction in prey numbers/biomass due to predator, and low num-


bers/biomass of predators
B small reduction in prey numbers/biomass due to predator, and large num-
bers/biomass of predators
C large reduction in prey numbers/biomass due to predator, and large num-
bers/biomass of predators
D large reduction in prey numbers/biomass due to predator, and low num-
bers/biomass of predators.
80 Haskellian Coordinates in Analysis: Searching for Pattern

Prey-predator systems typically oscillate between regions. Region B systems show high
biomass conversion efficiencies; a small reduction in prey biomass converts to a relatively
large predator biomass. Region D systems show low biomass conversion efficiencies; a large
reduction in prey biomass converts to a small predator biomass.
Predator efficiency are of two types. One is typified by region B systems. Here predators
appear to take only what prey they need to survive and reproduce. There can be little waste
in a region B system. Region D systems show high killing efficiencies because large reductions
in prey numbers or biomass are accomplished with small numbers or biomass of predators.
Of course, considerable interest exists in the evolution of systems from one region to
another. This question will be left as a research problem for prey-predator systems, and
taken up again in Chapter 9 for parasite-host systems.
A study of wolf predation in the natural ecosystem of Isle Royale national park (Peterson
and Page 1984) and of a predator in a laboratory setting (Gause, et al., 1935) will illustrate
characteristics of prey-predator systems. For brevity, I dispense with the development of
mathematical equations characterizing "separate" and "together" development, using instead
the unsmoothed observations.
The study of moose-wolf dynamics on Isle Royale National Park, Michigan, provides
an example of natural prey-predator oscillations. Unfortunately, not every example in this
Chapter is of such a natural system. Begun over 20 years ago, the Isle Royale moose-wolf
studies have annually estimated the moose population using both summer and winter ground
and aerial survey procedures and have completely enumerated the wolf population via. aerial
survey in February of each year (Peterson 1977, Peterson and Page 1983). Best estimate
population levels for the moose and wolf (the 'together' condition) are given in Table 5.2
(Peterson 1984).
Somewhat at issue are the moose and wolf population levels in the separate condition.
Peterson and Page (1983) describe the moose dynamics prior to wolf colonization of the
island:
Moose colonized the island in the early 1900's and, in a predator free environment
previously untouched by large ungulate browsers, increased by 1930 to a peak level
of 2000 - 5000. During the mid-1930's winter dieoff reduced the population to
a relatively low level, but a large fire in 1936 renewed the forage base for moose
on about 20% of the island. The moose population subsequently increased, and
substantial numbers of moose were again dying of malnutrition in the late 1940's
(Krefting 1974). At that time wolves colonized the island.
Moose numbers in the absence of wolves are estimated to be around 2000. Moose weights
are estimated to average about 859 pounds (Peterson 1984), and wolf weights 80 pounds
(Mech 1984). Wolf numbers in the absence of the moose are assumed to be zero (Peterson
1984).
The moose-wolf interactions over the past 15 years are plotted in Figure 5.5. Notice
that although the wolf is a very efficient predator, there still exists a definite cycle in the
interaction. The type and intensity of interaction in 1983 and 1984 is nearly the same as
the type and intensity of interaction in 1972 and 1973. There is a clear zone of interaction
type and intensity (around 175° and .64) where moose and wolf seem not to interact. With
the exception of a two year retrogression in years 1978 and 1979, the interaction showed a
nicely consistent clockwise cycle.
More information on this outstanding long term study of natural prey-predator rela-
tions can be found in the literature (Peterson 1977, Mech 1966), and annual reports on the
continuing studies.
A different habit of predation is evident in a study of the predator Cheyletus eriditus
and its prey Aleuroglyphus agilis (Gause 1935). The prey was reared under both separate
and together conditions in a flour medium that was not renewed as the experiment pro-
Haskellian Coordinates in Analysis: Searching for Pattern 81

Table 5.2. Estimated population levels for moose and wolf system on Isle Royale National
Park, Michitan. Biomass values are used in the interaction intensity computations. The
biomass in t e (0,0) condition is estimated to be a constant 1,718,000 pounds.
~ (0,0) ~ (0,0) - (0,0) B = 180 0 r=
year moose wolf x y -tan-1y/x Ixl + Iyl R = II(O~O)II
1969 1150 17 -850 17 178.85 731,510 .43
1970 966 18 -1034 18 179.00 889,646 .52
1971 674 20 -1326 20 179.14 1,140,634 .66
1972 836 23 -1164 23 178.87 1,001,716 .58
1973 802 24 -1198 24 178.85 1,031,002 .58
1974 815 31 -1185 31 178.50 1,020,395 .59
1975 778 41 -1222 41 178.08 1,052,978 .61
1976 641 44 -1359 44 178.15 1,170,901 .68
1977 507 34 -1493 34 178.70 1,285,207 .75
1978 543 40 -1457 40 178.43 1,254,763 .73
1979 675 43 -1325 43 178.14 1,141,615 .66
1980 577 50 -1423 50 177.98 1,226,357 .71
1981 570 30 -1430 30 178.80 1,230,770 .72
1982 590 14 -1410 14 179.43 1,212,310 .71
1983 811 23 -1189 23 178.89 1,023,191 .60
1984 781 24 -1219 24 178.87 1,049,041 .61

Figure 5.5. Haskell coordinate system plot of Isle Royale moose-wolf prey predator interac-
tions. The wolf is a very efficient predator, as indicated by the nearness of the interaction
to 180 0 .
82 Haskellian Coordinates in Analysis: Searching for Pattern

gressed. Contrary to most other prey-predator studies, Gause correctly attempted to grow
the predator in the flour medium. They lived only about 15 days before dying off. The pop-
ulation estimates were made periodically by a complete enumeration of individuals in the
experimental containers (see columns 2-5 in Table 5.3). Samples of mediary, and predator
and prey were seldom taken, but when they were taken they were returned directly to the
containers. This procedure contrasts with the example given for (-, -) biological competi-
tion where the 10 percent of the population sampled was discarded after counting and the
nutrient medium replaced each day.

Table 5.3. Prey-predator experiment of Gause (1935). Two kinds of intensity are estimated:
total action using the Manhattan metric z' = Ixl + Iyl and net action metric z = Ix + y!.
Both z and z' are expressed relative to the total production in the separate condition Ih I .
Note in some cases z' is greater than IITlll, hence R' is greater than 1. When this happened
all R' values were determined and then normalized to the largest value.
pure mixed
culture culture z= R= z' == R'=,
days Tl T2 T2 - Tl Ix+ yl M
z
Ixl + Iyl z
hll
A.a. C.e. A.a. C.e x y ()

0 86 7 86 7 0 0 0 0 0 0 0
4 76 .5 44 4 -32 3.5 173 28.5 .37 35.5 ( .35)
15 110 .5 84 4 -26 3.5 172 22.5 .20 29.5 (.20)
21 304 0 128 160 -176 160. 137 16 .05 336 (.84)
27 494 0 32 192 -462 192. 157 270 .55 654 (1.00)
33 1141 0 96 32 -1045 32. 178 1013 .89 1077 (.71)
39 1658 0 206 24 -1452 24. 179 1428 .86 1476 (.67)
47 1584 0 352 64 -1232 64. 177 1168 .74 12~6 (.62)
52 1360 0 512 128 -848 128. 171 720 .53 976 ( .55)
58 1664 0 96 64 -1568 64. 177 1540 .90 1632 (.74)
64 1664* 0 32 64 -1632 64. 177 1568 .94* 1696 (.77)
70 1664* 0 0 56 -1664 56. 178 1608 .97* 1720 (.78)
76 1664* 0 0 8 -1664 8. 179 1656 .99* 1672 (.78)

Since the quality and quantity of the medium changed significantly over time, Gause (1935)
estimated that the maximum mixed population level was reached at age 40 after which prey
population growth became negative and its level declined. In the numerical example here,
the prey was excluded from the system and the predator followed closely behind. Although
it could be argued that prey decline resulted from either a diminished nutrient supply or
from predation by C. Eriditus, it should be noted that prey numbers reached a level of about
1600 individuals per 0.2 gram of flour when grown separately in unreplenished flour medium.
Thus, it is probably safe to assume that up to the time of the prey exclusion by predator,
mediary dynamics equations are
dn/dt = 0,
ds/dt = 0,
where n designates nutrients available and s designates space.
Interation type ((}) and intensity (R') are computed (columns 8 and 12 in Table 5.3) and
plotted in Haskell's coordinate system (see Figure 5.6). The total action metric, z = Ixl + Iyl,
is used as a measure of interaction intensity. The first cycle (upper figure) took 52 days with
the greatest total action occurring at day 27. Also, note that the cycle "opens" toward
the (-,0) interaction (180°), and it is near this relation that exclusion occurs. The second
Haskellian Coordinates in Analysis: Searching for Pattern 83

cycle (lower figure) began at day 58, and showed little variation in interaction type, moving
immediately to near 180 0 where exclusion occurred. Note the distance between points at
days 4 and 52, both starting points for a cycle.
We can see differences between the interaction patterns generated by the two predators
treated in this section. Briefly high-lighted, they are:
1. C. eriditue is somewhat less efficient as a predator than the wolf studied on Isle Royale.
It takes more of the former to reduce the respective prey population than it does of the
latter to reduce its prey population level. The values correspondingly deviate signifi-
cantly from the 178 0 -179 0 typical of the Isle Royale moose-wolf system.
2. C. eriditus drove its prey to exclusion, something the wolf has not yet done. Perhaps
in a natural environment A. agilis would have evolved behavioral traits allowing it to
avoid C. eriditus.

prey (-.oLJ~~Ha~~il~;lillllll

Figure 5.6. Haskellian coordinate plot of two cycles of interaction type and intensity that
leads to exclusion of the prey (A. agilis) by predator (C. eriditus) at day 70 of the experiment.
84 Haskellian Coordinates in Analysis: Searching for Pattern

In sum, the pattern of interactions found in prey-predator relations is peculiar to


a) the parties to the interaction,
b) mediary quantity, quality and dynamics,
c) relative initial proportions of prey to predator, and
d) experimental procedures (sampling, etc.) used in the course of study.
I now look at the pattern of interactions in parasite-host systems.

c. Parasite-Host (+,-)
A famous series of experiments exammmg parasite-host relations was conducted by
Utida (1957). Parties to the interaction were the host, azuki bean weevil (Callosobruchus
chinensis), and parasite, the braconid wasp (H eterospilus prosopidis). Utida reported on
three experiments run for 112 (experiment A), 47 (C), and 84 (E) generations. Only the latter
experiment ended "naturally" by exclusion of the host. The other two sets of experimental
populations were accidentally destroyed. A major purpose of the study was to continue an
earlier experiment lasting 25 generations to determine if there was a continuing fluctuation
in host and parasite numbers. Further, Utida wished to check for possible changes in "wave
length and amplitude" of the fluctuations after many generations of association. A host
generation spanned about 20 days, so the longest experiment analyzed here endured for
about 6 years.
Parasite-host systems, as with prey-predator, have no mediary. But, the host is, like the
prey, sometimes an herbivore, and the configuration of the plant material can have significant
influence on the type and intensity of interaction. Utida's was a laboratory experiment with
the following basic structure:
a) host food was 10 grams of azuki beans,
b) container was a petri dish 1.8 cm high and 8.5 cm diameter,
c) temperature and relative humidity were maintained at 30 and 75 percent, respectively,
d) both host and parasite were counted at 10 day intervals,
e) new azuki beans (10 grams) were added to the petri dish every 10 days to serve as host
food. Old or apparently ruined food was removed when counts were made, and
f) experiment A began with 64 weevils and 8 wasps, and experiment E with 512 weevils
and 128 wasps.
Utida's experiments are not complete because he did not examine the "separate" devel-
opment or spontaneous change of either the host or the parasite. That is, he established no
reference frame from which to judge induced change, or the "together" development. Nev-
ertheless, I have attempted to use his results. I felt safe assuming that the parasites placed
alone in the azuki bean environment would soon die. I thus assumed a zero population level
for the parasite in the "separate" condition. Perhaps less safe is my assumption that ·the
host population, growing in the absence of the parasite, exhibits a monomolecular pattern of
development: it rapidly increases, approaches a maximum and stays at the maximum level.
I selected a maximum host population of 1200 individuals per petri dish, about 15 percent
above the host population level under a very low parasite population level. Copies of the
original time series for the two experiments analyzed here are shown in Figure 5.7. Approxi-
mate numbers of hosts and parasites under induced change were determined by enlargement
of these figures and hand scaling of numbers of hosts and parasites at each generation. These
numbers then formed the basis for computing type and intensity of interaction. The host
has been referenced to the vertical axes and the parasite to the horizontal axes. This puts
the interaction between 270 0 and 315 0 .
Figure 5.7 contains ample illustration of the cyclic effect of parasite-host interactions.
When analyzed using Haskellian coordinates, one period of oscillation in the time series
converts to a polygon-like figure of interaction type and intensity. The polygon is more
clearly evident here than in the prey-predator examples. A natural initial condition and
Haskellian Coordinates in Analysis: Searching for Pattern 85

GENE.RAT!ON

Figure 5.7. Time series for Utida's parasite-host experiments. Host population values are
denoted with open circles, parasite with solid circles. Populations in A were accidently
destroyed. Host in E was excluded. Population density refers to numbers of individuals in
10 grams of azuki beans (Utida 1957).

trend in numbers is for the host population to increase, followed by the parasite population.
One full period of oscillation of host and parasite numbers between generations 13 and 18
in experiment A can be seen in Figure 5.7. Figure 5.8 shows its conversion to Haskellian
coordinates. In total, experiment A contains 22 distinct polygon-like figures of varying sizes
and lengths. Progression with time is counter-clockwise for nearly every segment of each of
the 22 cycles. A typical cycle has two parts, one when host numbers are large, parasite low
(loop toward the (0,0) circle in Figure 5.8)' the other when host numbers are low and parasite
numbers high (loop away from the (0,0) circle in Figure 5.8). As time series oscillations
become damped, s"Omething Utida sought to discover, the area "enclosed" by the polygon-
like figure is decreased (open circles in Figure 5.8). Large oscillations convert to polygons
with long perimeters. Oscillations with a reduced period length take fewer generations to
86 Haskellian Coordinates in Analysis: Searching for Pattern

·. .---- - - - - - - - -- - ----(+, 0) parasite

(0,-)
host

Figure 5.8. Polygon of interaction type and intensity for generations 13 to 23 from un-
smoothed data of Utida's experiment A.

enclose an area. Cycle 13 to 18 shows wide fluctuations in interaction intensity. Cycle for
generations 18 to 23 is much smaller. Both cycles have the same length, 6 generations, but
the latter encloses much less area. This means the amplitude of oscillations is decreasing.
The polygon for generations 23 through 30 (not shown in Figure 5.8) has a longer generation
length yet encloses an even smaller area.
Unfortunately there was no "natural" conclusion to experiment A; the populations were
accidently destroyed. Experiment E (lower part of Figure 5.7) ended when the host was
excluded by the parasite. Might this result have been anticipated prior to its occurrence?
Figure 5.9a shows the interaction polygon prior to the one during which exclusion occurred.
It has a typical form, two loops, counterclockwise orientation of generations, and near perfect
closure (points for generations 64 and 72 nearly coincide). This general pattern is repeated 11
times from generations one through 72 in this experiment. Starting at generation 72, progress
is "normal" to generation 73 but then extends way outside the previous interaction polygon
at generation 74 (Figure 5.9b). Following generation 74 the population interaction behaved
very erratically. In short, the pattern of repeated regular form was broken. Specifically,
a) for four generations (74-78) rotation was clockwise rather than counterclockwise,
b) a dramatic host population increase caused interaction intensity to drop markedly at
generation 75 as shown by the large "move" toward the (0,0) circle,
c) the parasite population level responded gradually to the dramatically increased host
population until generation 81 when it greatly increased,
d) following generation 80 the mixed population had a net positive interaction, hence oc-
Haskellian Coordinates in Analysis: Searching for Pattern 87

r-----------------(+,O) parasite

r----- - - - - - - - - - - - - - (+,0) parasite

Figure 5.9a. Interaction polygon for the cycle immediately prior to exclusion of the host in
Utida's experiment E.
Figure 5.9b. Polygon during which exclusion of the host occurred. Arrow indicates expected
interaction change after generation 74. Point for generation 81 (9 = 321°, R = .81) is, by
convention, plotted outside the circle and is omitted here. Note that exclusion occurred
between generations 82 and 83 and about 7° from the (0, -) axis.
88 Haskellian Coordinates in Analysis: Searching for Pattern

curred at () = 321 0 and R = .81 (total action). (Note: The progression beyond 315 0
may have been caused by my selection of too small a number for the host population in
the absence of a parasite. Recall Utida did not establish this level.),
e) total action was very high following the 80-81 period that was favorable to the parasite,
f) one generation following the high level of total action the host was excluded at () = 277 0
and R = .65,
g) the parasite population level followed closely the decrease in host level and presumably
disappeared at generation 85.
In retrospect it appears that the important event occurred at or just before generation
74 - nine generations before the exclusion event. One speculation is that bean weevils,
without a history of 74 generations coexistence with the parasite, were somehow allowed
into the experimental container. Although initially successful in escaping parasitism, the
braconid wasp needed only a few generations to become successful at parasitism and quickly
decimated the non-co evolved host population. There will be further discussion on the co-
evolution of parasite-host systems in Chapter 9.

D. Plant-Herbivore (-,+)
Three of the four example interaction studies discussed to this point differ from what
would seem the "normal" situation for interaction studies in forest ecology because:
a) the environment for each study was strictly controlled, either being totally closed to
food, party emigration, immigration or slightly open to food and emigration in the form
of harvesting, and
b) the parties to the interaction could be established in theoretically proper numbers for
both the "separate" and "together" conditions at the start of each interaction experi-
ment.
Of course interactions of various and varying types and intensities are a constant oc-
currence in forest communities. Scientists studying them face a dilemma. Both conditions
a) and b) above are true, yet they may wish to use Haskellian coordinates to organize and
display information about their findings. This section is devoted to a forest ecology example
wherein the investigators purposely viewed the system, and organized information about the
system, in such a way that their findings could be mapped into Haskell's coordinate system.
No great adjustments were needed; it suffices to pay attention to the needs of an interaction
analysis, i.e., both induced and spontaneous change information, something understood by
Gause but seemingly forgotten or considered unimportant by some present-day ecologists.
Mattson and Addy (1975) simulated the development of two herbivore - forest tree
systems: forest tent caterpillar (FTC) (Malacosoma disstria) - quaking aspen (Populus
tremuloides), and spruce budworm (Choristoneura fumiferana) - balsam fir (Abies balsam~a).
I shall confine my remarks here to their FTC - aspen study. Since it was not possible to
control the physical system as in laboratory settings, actually observing and measuring
development of each component under spontaneous and induced change, it was necessary to
synthesize specific information on aspen growth in defoliated stands with generally accepted
information about aspen growth in undefoliated stands. This had to be reconciled with
specific information on FTC biomass development when defoliating aspen stands. FTC was
assumed aspen - specific in this part of its range, so was assigned a zero level of annual
biomass production for the "without aspen" condition. The results of this simulation are
shown in Table 5.4.
Note that the numbers in columns 2-6 refer to annual biomass production growth, not
accumulated standing crop as in the previous examples. The Haskell coordinate system
repres~ntation of the FTC - aspen system is shown in Figure 5.10. (The figure differs from
the original because I have referenced aspen to the horizontal axes instead of vertical and
FTC to vertical instead of horizontal.)
Haskellian Coordinates in Analysis: Searching for Pattern 89

Table 5.4. Annual production of biomass in a quaking aspen stand with and without the
forest tent caterpillar. The difference in annual quaking aspen total production (stemwood
-"- foliage) is gIven in column headed "x", difference in FTC biomass in column headed "y".
Total action z' = :x; -r 'yi is used as the measure of interaction intensity, so z' and R values
differ from Mattson and Addy (1975).
Annual biomass production
age (91m 2 , dry weight)
yrs.
Forest without FTC Forest with FTC
--.--.-
stem- foli- : stem- foli- :
wood age: FTC wood age: FTC x y z' R ()

26 163 125 0 163 125 1 0 1 1 .004 0


27 165 125 0 165 125 2 0 2 2 .007 0
28 167 125 0 144 100 4 -48 4 52 .18 175
29 170 126 0 30 175 18 -91 18 109 .37 168
30 172 127 0 14 165 10 -120 10 130 .43 175
31 173 128 0 14 150 15 -137 15 152 .51 173
32 175 128 0 103 100 2 -100 2 102 .34 178
33 176 129 0 161 129 -15 16 .05 176
34 178 129 0 163 129 <1 -15 <1 16 .05 176
35 179 129 0 165 129 <1 -14 <1 15 .05 175
36 180 129 0 166 129 <1 -14 <1 14 .05 175
37 181 129 0 168 129 <1 -13 <1 14 .05 175
38 182 129 0 169 129 <1 -13 <1 14 .05 175
39 183 129 0 171 130 <1 -12 <1 13 .04 175
40 184 130 0 172 130 <1 -12 <1 13 .04 175

Judging by the type of interaction given in Table 5.4 and shown in Figure 5.10, FTC
could be rated a moderately efficient herbivore because once the outbreak occurred the
kind of interaction stayed close to 175°. However, the intensity was not great, reaching a
maximum near R=.51 at age 31, four years following the FTC population outbreak.
The information available spans a single 14 year period (stand age 26 to 40)' and during
these years the information was limited to a single interaction cycle of increasing intensity,
clockwise change, and then decreasing intensity. Previous examples, based on laboratory
experiments, showed cycles of interactions in this quadrant and quadrant IV. Should we
expect the same? Are there repeated FTC defoliations of the same stand, or is it a one-time
event in the development of a particular stand? Mattson (1980) stated that the aspen - FTC
interaction is a one-time thing; not all phytophagous insect-forest tree associations are this
way, but FTC - aspen is. Thus, cycles of interaction type and intensity are not expected.
Can FTC - defoliated aspen stands "catch up," in terms of total standing crop, to stands
not defoliated? To date this question remains unresolved. Clearly, the mathematics of the
matter requires that previously defoliated aspen stands grow faster than ones not previously
defoliated. The mechanism for this possibility is discussed below:
Wood production may increase after severe defoliations because the circulation of
important growth elements such as nitrogen, phosphorus, and potassium is en-
hanced or the-distribution of light and moisture is more equitable. Moderate to
severe defoliations can increase normal nitrogen, phosphorus, and potassium con-
tributions in the litter fall by 20 to 200 percent. This occurs because litter fall is
90 Haskellian Coordinates in Analysis: Searching for Pattern

forest tent
caterpillar
(0 +)

quaking
aspen

Figure 5.10. Change in type and intensity of interaction between quaking aspen and the
forest tent caterpillar for stand ages 26 through 40. Length of a significant interaction is
only about 5 years. Note the absence of cycles. Mean for years 28~40 is 174 0 indicating that
FTC is a fairly efficient herbivore.

not only greater but also richer than normal due to the exceptionally high con-
centrations of nutrients in dead insect bodies, insect excrement, and wasted food
parts. This enhances the soil organisms. The net results, perhaps, is increased
plant growth (Mattson and Addy 1975).
However, it seems a mistake to attribute excess in growth, previously defoliated stands over
nondefoliated stands, to the current direct interaction with an herbivore. Direct interaction
will typically have occurred several years prior to the advent of enhanced growth. Set up by
the direct interaction was a change in stand environment, for simplicity called site, so the
normal assumption of forest growth

ds/dt = 0,

where s designates site, in particular, nutrients, is invalid. More realistic is the assumption

ds/dt = f(lJ(t - k), R(t - k)) > 0,

where
IJ and R designate herbivore~plant interaction type and intensity, and
k designates a time delay constant.
Thus, any "catching up" done by defoliated stands is due to an indirect, rather than direct,
action of the phytophagous insect.
These examples are sufficient to establish the potential validity of the following propo-
sitions:
Haskellian Coordinates in Analysis: Searching for Pattern 91

1. Interaction trajectories in Haskellian coordinates take predictable forms, depending


upon the quadrant in which they occur.
2. The forms are typically repeated, if not in time, then with change in parties to the
interaction.
3. When the forms are repeated, they show a high degree of regularity.

E. Pattern And Expectation


Bunge (I976b) argues that the concepts of prediction and retrodiction belong to the
family of anticipations and retrospections. Included in the family is, in order of scientificness:
expectation, guessing, prophecy, prognosis, and full scientific prediction. Recall from Figure
2.1, answers to What if? questions are suggested as being more difficult to obtain than
answers to What is the character of? questions. The examples in this chapter suggest how
Haskellian coordinates can be used as expectation facilitators, helping us to inch our way
out into the column headed "prediction" in Figure 2.1.
Having gained some experience with Haskellian coordinates by analyzing form, repe-
tition of form and regularity of the form repetitions in five experiments, I focus briefly on
the notion of anticipation as a consequence of pattern. The referent of our anticipation is
another bit of nature, the same system at a later time or under different environment, a
system comprised of different parties but similar environment, and so on. Simply put, our
anticipation is that a regularly repeated form will be repeated. For example, we have seen
regularly repeated polygon-like interaction trajectories in quadrants II and IV, and simple
segmented line trajectories in quadrant III. My anticipation is that I will not see closed
polygon-like trajectories in quadrant III. Because I am using empirical generalizations, it
appears warranted to call my anticipations prognoses (Bunge 1967b).
An aid to prognosis is the analysis of attributes of the interaction trajectory. Is it
polygon-like? In what way might the polygons lack closure (for beginning and ending points
to coincide)? Does it lack type closure (agreement of (J values), intensity closure (agree-
ment in R or R' values), or both? Are successive polygon-like figures encompassing less or
a constant area (approaching a stable or dynamic euilibrium), or more area? Is the num-
ber of generations needed to complete a trajectory polygon, especially in quadrants II and
IV, increasing (period increasing), decreasing (period decreasing) or not changing? Is the
progression clockwise in quadrant II and counterclockwise in quadrant III? What are the
coordinates of the centroid (e.g., (J,R') for each polygon-like trajectory? Is there regular
change in location of these coordinates for successive trajectories? These questions simply
hint at formal attributes of trajectories that could be developed.
In two of the five experiments analyzed one party to the interaction was excluded from
the system. Both exclusions occurred near the axis of decrease for the party detrimentally
affected. I would expect exclusions to occur near, if not on, the (-,0) and (0, -) axes.
Further, I would expect the intensity of interaction at exclusion to be reasonable high,
certainly in excess of .5. Further, trajectories in quadrant III can also lead to exclusion at
or near the (-,0) and (0, -) axes. However, not every trajectory heading toward or crossing
one of these two axes includes an exclusion in its future.
Figure 5.11 summarizes the process of developing attributes of interaction trajectories
by examining the ordered triple < (J, R', t >, where the trajectories are polygon-like or simple
segmented lines, and synthesizing the attributes into prognoses about nature at another place
or time. I have used the same methodology for moving away from the plane of perception,
P, into the construct field, C, from five different locations in nature. As more experience is
gained, more repetitions manifested, confidence is fostered that Haskellian coordinates are
not only useful organizers, but are more useful prognosis facilitators than current methods
rooted in the Cartesian coordinate system.
92 Haskellian Coordinates in Analysis: Searching for Pattern

plane of
perception

measurement

NATURE

trajectory length
area enclosed
closure of "polygon"
trajectory centroid
NATURE' orientation of change
(
etc.

Figure 5.11. Summary of the process of searching for pattern and generating expectations
of another bit of nature using Haskellian coordinates.

F. Research Problems
1. Plot the observed values obtained by Gause in the P. caudatum - P.aurelia study in
Haskell's coordinate system. What has the mathematical smoothing done to interaction
type variation?
2. Rerun Gause's experiment in problem 1 above, but use a range of different harvesting
levels. Plot the results. What is the pattern? Can interaction type and intensity be
controlled?
3. Analyze all cycles in Utida's famous experiment and plot the results in Haskell's coor-
dinate system. For each cycle, compute a centroid. Is there a pattern to the centroid
locations as time of association increases?
4. For each cycle in problem 3, compute period length (number of generations to close
a polygon), and amplitude (area enclosed by one polygon), and regress them against
mean generation number. Is there a trend in either relation?
5. An important assumption in the moose-wolf analysis is that the (0,0) moose population
level is 2000 animals throughout the analysis period. Assume a slight decrease in moose
ha9itat quality, because of a maturing of the vegetation on Isle Royale, hence a reduction
of the (0,0) moose level with time. What effect does the assumption have on the
interaction trajectory?
Haskellian Coordinates in Analysis: Searching for Pattern 93

6. Develop the necessary information on biomass production for other important forest
insects, e.g., the gypsy moth, plot the results, and compare the trajectory with that for
forest tent caterpillar - quaking aspen.
7. Select an experimental study of a prey-predator or parasite-host system that has many
cycles. Plot the first third of the cycles in a phase plane. Extrapolate the population
levels to later cycles using the phase plane.
8. Plot the same first few cycles used in problem 7 in Haskell's coordinate system. Ex-
trapolate the pattern of e and R values to later cycles as done in problem 7. Solve the
equations

e= 180 0 - tan-1y/x

R' = Ixl + Iyl


11(0,0)11
for x and y (Use 360 0 if a parasite-host system). Compute the population levels at the
extrapolated times.
9. Compare the extrapolations based on the phase plane and those based on e and R' with
the actual population levels observed during the later cycles. Is one method significantly
better than the other at anticipating the future?
10. Are predators more efficient at reducing prey numbers than parasites at reducing host
numbers? What is your criterion of efficiency?
6: Patterns of Interaction
in Mixed Forest Stand Dynamics

Foresters have long been interested in mixed forest stand dynamics. Traditionally, their
view of the problem has reflected a utilitarian philosophy: at the time of stand establishment,
which combination of two species will give the greatest total yield over a fixed time interval?
Nelson (1964) reviewed some of the controversy surrounding this question in the writings
of early European foresters, described a mathematical model for plant competition studies
developed by Wit (1960) and others, and suggested it as an aid to answering the above
question. Swindel (1970) elaborated on this model.
Forest scientists have not, as a rule, viewed the assessment of mixed stand dynamics
as sharing anything in common with other forest interactions, e.g., prey-predator, or plant-
herbivore (see Chapter 5). Thus, Gause's inspiration from the forest ecologist V. Sukachev
has, with the .exception of Turnbull (1963), not been reciprocated. In general and sum,
studies of mixed stands may be faulted on two counts:
a) Often the analyses deal only with the species growing together in the mixed condition
and with the "important" species in a pure stand. The "unimportant" species are
seldom grown in a pure stand. Examples are the studies of Douglas fir and red alder
(Miller and Murray 1978) and black walnut and autumn olive (Funk, et aI., 1979)
b) Because the traditional assumption has been that the interaction is competitive and be-
cause the philosophical view has been utilitarian, the model used to design and evaluate
mixed stand studies makes use of an "incorrect" set of initial conditions for the second
of two sets of experiments: each species growing separately, and the two species growing
together.
An argument for a new approach to studying mixed stand dynamics may add to these
faults the as-yet-unmet need to assess the likelihood of tree species being excluded from the
stand, and the resultant likelihood that this results in forest diversity changes. Anticipat-
ing forest diversity changes has arisen as a special need only recently with the passage of
legislation and development of directives concerning this property of forest stands:
For each planning alternative, the interdisciplinary team will consider how diversity
will be affected by various mixes of resource outputs and uses, including proposed
management practices. To the extent consistent with the requirement to provide for
diversity, management prescriptions ... will preserve and enhance the diversity of
plant and animal communities, including endemic and desirable naturalized plant
and animal species, so that it is at least as great as that which would be expected
in a natural forest and the diversity of tree species similar to that existing in the
planning area. Reductions in existing diversity of plant and animal communities
and tree species will be prescribed only where needed to meet overall multiple-use
objectives .:.. (Federal Register, Vol. 40:181, p. 53998,1979).
Since exclusion of an important plant species from a forest community may significantly
affect community species diversity, a theory of diversity dynamics must include all cases
96 Patterns of Interaction in Mixed Forest Stand Dynamics

that can give rise to an exclusion. In the last chapter I examined cases of prey-predator and
parasite-host exclusions. To this I must add the traditional competitive (-, -) interaction-
based exclusion. This will complete my treatment of interactions that neighbor the valleys
of exclusion, the (0, -) and (-,0) axes.
In this chapter, I attempt five things:
(a) to compare and contrast the methods of studying plant-plant interactions due to Wit
(1960) and Gause (1934),
(b) to introduce a mathematical model of forest dynamics and use it to simulate the devel-
opment of pure and mixed stands of balsam fir and quaking aspen,
(c) to examine a few initial stand structures, compositions, and elapsed times for which
quaking aspen may be excluded from the stand by balsam fir,
(d) to represent these dynamics in Haskell's coordinate system, and
(e) to examine the degree to which balsam fir-quaking aspen interaction type and intensity
can be controlled by thinning the stand.

A. Historical Review
Harper's (1977) splendid synthesis of information on plant population biology gives
background on methods commonly used to study mixed plant populations and also those
suggested by Lotka, Volterra, and Gause (L- V-G) that have been applied mainly to ani-
mal populations. He identifies two experimental models for studying the growth of mixed
populations of plants: additive and substitutive. In studies of two-population mixtures it
is possible to vary either density of the total standing crop or proportion of each species,
or both. The difficulties in analyzing these variations are fewer in the substitutive type of
experimental design. Wit (1960) devised a substitutive design, called a replacement series,
for studying competition in two-species mixtures, that requires sowing two species in varying
proportions while maintaining overall density constant. Proportions include the zero level
for each species. The four basic forms that experimental results can take from substitutive
experiments are explained by Harper (1977, pp. 255-257).
Harper later illustrates a variety of models of two-species interactions for plants and for
animals (pp. 302-304). Omitted from his discussion is the fact that the replacement series
model makes fundamentally different presuppositions about the natural system than the well
known approach described by L- V-G. Harper uses the latter approach in a way not incorrect,
but not typical either. By associating the replacement series with the L-V-G approach, he
is giving the impression that the two approaches share presuppositions, but this impression
is incorrect.
The presuppositions of each method are reflected in the initial conditions (initial pop-
ulation levels) necessary for the parties to interact under conditions of induced change (the
mixed condition). In the replacement series approach, the presupposition is that the two
species are competing. In the 1-V-G approach, there is no specific anticipation made of the
kind of interaction between parties. A concrete example will illustrate the difference.
Because Wit's replacement series model has been suggested as a framework for the
study of mixed forest stands (Nelson 1964, Swindell 1970), I will examine it in forestry
terms. It has design and experiment stages that involve "establishing" (in analogy with
Harper's "sowing") m trees of species A per unit area in a pure stand, n trees of species B
per unit in a pure stand, and, for example, m/2 trees of species A per unit area and n/2
trees of species B per unit area in a mixed stand. Other proportions need to be established,
of course, to more completely examine the mixed stand standing crop. Henceforth I refer
to these as Wit-type initial conditions. The experimental stage then involves measuring the
standing crop of both species in all proportions established for the intended time interval.
By comparison, a scrutiny of the experiments reported by Gause shows that he stu-
diously adhered to the use of a particular design equivalent to letting species A have m trees
Patterns of Interaction in Mixed Forest Stand Dynamics 97

per unit area in a pure stand, species B have n trees per unit in a pure stand, and then
letting species A have m trees and species B have n trees on the same unit area for the
mixed stand. Henceforth I refer to this approach as having Gause-type initial conditions.

B. Role of Presuppositions in Study Design, and Analysis


The analysis of experiments with Wit-type initial conditions makes use of the math-
ematical devices shown in Figure 6.1. A comparison is made between the species in their
pure condition and the sum of the species in the mixed condition. The same species cannot
be compared in their pure and mixed condition because of unequal initial conditions. Fig-
ure 6.1b shows the analysis framework suggested by Nelson (1964) and Swindel (1970) to
evaluate a mixture (assumed to be in the ratio 1:1) at time to. The a-line connecting the
points Pa and Pb shows the expected mixed stand production if it is directly proportional
to the representation of each species in the mixture. In this schematic example, the point
M is slightly below the line, indicating a slight mutual depression effect.
The replacement series model and Wit-type initial conditions presupposes that:
(1) trees of the two species are competing, and
(2) the question of interest is the effect of species composition on total stand productivity.

BASAL AREA
STANDING CROP
PER UNIT AREA

A pure stand - species A


_ - mixed stand - (A+B)
/././- pure stand - species B
/'
~~~~------------------~~~

/r------------~p~br--a
M

0% A 100 % A
TIME 100 % B O%B
50 % A
50 % B

Figure 6.1. Analysis framework for the method of Wit (1960). A. Time series showing
standing crop of species A and B in pure stand and the mixed stand. B. Framework for
analyzing the effect of species on mixed stand production.

The analysis of experiments with Gause-type initial conditions call for comparisons
between standing crop of species A in the mixed and pure stand, and comparisons between
standing crop of species B in the mixed and pure stand. The sum of the standing crop in
the two pure stands gives total standing crop on two unit areas. If trees of species A and B
do not interact, the sum of standing crop of A and B in the mixed stand also gives standing
crop on two unit areas. That is, if A and B are neutral toward each other, there is one unit
of abiotic flux available to trees of species A and another unit available to trees of species B
in the mixed stands. If trees of species A and B compete, one would expect less than two
units worth of production in the mixture.
Figure 6.2a shows the time series framework which characterized the Gause method of
analysis. Each species may be compared in its pure and mixed condition because the initial
density is the same for each condition. Figure 6.2b shows the analysis framework, a phase
98 Patterns of Interaction in Mixed Forest Stand Dynamics

plane, that permits the representation of the theoretical totality of possible interactions.
Naturally most mixed forest stand interactions will likely fall in the hatched quadrant. Points
are located in Figure 6.2b, for example to, with stand age represented parametrically. A point
on either axis is a pure stand and off the axes is a mixed condition.
The phase plane approach and Gause-type initial conditions presuppose that:
(l) the trees of the two species are neutral toward each other, and
(2) the purpose is to estimate the type and intensity of interaction between the two popu-
lations.
No explicit consideration is given to determining the species composition that gives maximum
or minimum productivity. The hatched area is the quadrant where competitive interactions
occur.

STANDING CROP, BASAL AREA


BASAL AREA PER UNIT AREA
PER UNIT AREA IN SPECIES A
I
A I
B I
I
pure stand - species A~ I
I
I
pure stand - species BI7=;::2-:::::_ _ __ Pa~""",,"",,,~~~l~o_____________ _
A Pa

mixed stand - species B

TIME
BASAL AREA PER UNIT AREA IN SPECIES B

Figure 6.2. Analysis framework for method of Gause (1934). A. Time series showing standing
crop in pure and mixed stands for each species. B. Phase plane representation of A with
time parametrically included.

C. Example Mixed Stand Dynamics


I know of no field experiments that meet the design requirements for mixed stand
studies, that is, have Gause-type initial conditions, so it is necessary to simulate growth and
development of two-species in order to estimate the type and intensity of interaction between
the populations. The simulation requires three "runs":
• species A in a pure stand,
• species B in a pure stand, and
• species A with species B in a mixture.
Three simulations over time using these conditions as starting points will produce standing
crop values necessary for a Haskellian analysis.
The simulation model I use has been described elsewhere (USDA 1979). Additional
background on the approach can be found in Leary (1980). Basically, I view a stand of
trees as an interacting aggregate. The aggregate amount (total standing crop) at any time
is estimated by
AA = nS, (1)
where:
AA denotes the aggregate amount,
n denotes number of trees in the stand, and
Patterns of Interaction in Mixed Forest Stand Dynamics 99

5denotes mean size of tree in the stand.


Aggregate dynamics is given by the equation
6AA n6S S6n
- - -"-
--- (2)
6t 6t ' 6t'
where:
';,; denotes change in mean size, and
';,~ denotes change in number of live trees.
For the purposes of this example, ';,~ will be considered synonymous with mortality, and
always negative. Forms of representation for the two parts of total change are, for change in
mean size (represented by stem thickness at breast height),

6d
=
[potential diameter
growth of the tree
1 resistance to the
[ potential being
1 (3)
6t of mean size realized
and for change in number of live trees,
6 n
~
6t
= '"
~
Pr(i) (4)
1-=1

In equation (6.4), Pr(i) denotes the probability that the ith tree (of n trees) will "die" in
the current year and is presented as

Pr(i) = [
1 + exp (Q1 + Q2.exp - 1 .
((d - (3)/Q4)2 + Q56d"c)
1 (5)

given that the tree has grown 6d in the last year to attain a current size of d inches of stem
thickness at breast height (Buchman 1980).
The equation representing change in mean size has a part representing potential growth
and a part representing resistance to the potential being realized. The former takes the form
of an additive metabolic balance equation:
constructive - destructive = net metabolism.
The latter takes the form of an exponential extinction of photosynthetic active radiation
(PAR) as stand density increases. The potential and resistance combine in a multiplicative
manner to give a realized (actualized) growth for the mean tree.

1. Model form and calibration.


Not all models of forest dynamics have the form required to study interactions in mixed
stands using the theory of interactions developed here. Typically the models are unable to
grow each species in a pure stand. The models developed elsewhere (USDA 1979, Leary
1979) were explicitly designed to enable us to represent both pure stand growth and mixed
stand growth. The model for change in standing crop is:

(6b)
100 Patterns of Interaction in Mixed Forest Stand Dynamics

where:
Cr designates mean crown ratio for the respective species populations,
s designates site index for the respective species population,
Y. designates sum of stem diameters at breast height for ith species population,
n. designates number of trees in the ith population
d. = (Y./n.) designates mean stem diameter at breast height for ith population,
Ct') and (3.) designate numerical constants,
CtI2(dd"13 represents live cell volume (respiring phytomass),
CtI4SCr(dd"" represents the volume of phytomass capable of constructive metabolism,
and
(1 - exp (I(d., n.)) represents extinction of PAR in a stand of trees.
The first major term in brackets on the right of equation (6.6) represents the potential
of trees to increase in stem thickness if the tree
a. is not interacting with other trees,
b. is growing in a physical environment given a site index value of s,
c. has a crown ratio (percent of total height in full live crown) of C r , and
d. has a diameter d.
The equation form shows my bias toward a possibilist ontology (Bunge 1977). The
product of the first two major bracket terms represents the potential of the stand of n trees
to increase the cumulative stem thicknesses of all its trees assuming there is no interaction
among the n trees. The product of the three terms in brackets gives the realized growth in
stem thickness in one year given that there is interaction both between trees of the same
species and trees of different species.
The calibration phase of model development attempted to separate the intraspecies
interactions from the interspecies interactions. This was done in three steps:
l. Pure or nearly pure stands of balsam fir were located in a large data base of permanent
growth plot information (Christensen, et a!., 1979), and with (313 set equal to zero, a
series of (311 values were determined. These values determined how balsam fir interacts
with balsam fir. Sixteen plots were found suitable for the pure balsam fir analysis and
the 16 estimated (311 values were found to be related to mean stand diameter as follows:

(311 = 86(1 - exp (- .4923 Y1)). (7)


D1

2. Pure or nearly pure stands of quaking aspen were identified in the data base, and
methods of analysis identical to those for balsam fir were used to identify twenty-one
stands. With (323 in equation (6.6) set equal to zero, a (321 value was determined for each
stand. These numerical constants were found to be a function of mean stand diameter
as follows:
(8)
3. Mixed stands of balsam fir and quaking aspen were located in the calibration data
base. Sixteen plots were found that met the requirements on species composition and
number of measurements made on the same trees (three were required). At this point
the (311 and (321 values in equation (6.6) were considered known, and the purpose was to
estimate numerical values for (313 and (323. Each of the 16 plots was treated separately so
that the two unknown numerical constants were estimated simultaneously for each plot
using summaries of repeated tree measurements as boundary conditions in a multipoint
boundary value problem approach to system identification (Leary 1970, Leary and Skog
1972). A constraint on the numerical constants for each plot was (313 = 1/(323. Sixteen
sets- of values were derived and the average determined to form the interaction terms
in the right- hand-side of equation (6.6). The values were (313 = .581 and (323 = 2.327,
where the first value represents the contribution of one square foot of quaking aspen to
Patterns of Interaction in Mixed Forest Stand Dynamics 101

the reduction of balsam fir growth (approximately one half as much as another square
foot of balsam fir of the same diameter), and the second represents the contribution of
one square foot of balsam fir to the reduction of quaking aspen growth (over twice as
much as an additional square foot of aspen of the same diameter).
Since the model was calibrated on pure and mixed stands, it can be used to simulate the
growth of pure stands of aspen and balsam fir as well as mixed stands of these two species.

2. Density, structure, and composition of stands projected.


The mathematical representation of mixed stand change given in equation 6.6 is a key
axiom in a theory of mixed forest stand dynamics. It is my intent here to examine some
deductions from the theory using the key axiom (the equation) to make projections of as-
sumed initial conditions and to examine the stand densities, compositions, and structures
for possible exclusion of quaking aspen from the stand. Also, I wish to see when, if ever,
exclusion occurs and to scrutinize the interaction trajectory for clues to action by man that
could prevent the exclusion.
Because I had no records of stands that met exactly the requirements for Gause-type
initial conditions, it was necessary to fabricate a list of trees (species and sizes) that, if
summarized, would produce the desired stand characteristics (mean size, number of trees,
and basal area). The two primary stand characteristics of interest are stand structure,
especially the relative sizes of the two species, and stand density. The tree list fabricator
accepts three constraints: mean size of tree for each species (the mean dbh of trees of each
species), the standard deviation of sizes (it assumes a normal distribution of sizes around
the mean), and the basal area density for the species. When these three constraints are
specified, the number of trees is allowed to vary as needed to meet the other constraints.
Quaking aspen and balsam fir commonly grow in mixed stands in the Lake States. Due
to its sprouting ability, aspen typically will reoccupy an area after a mother stand has been
clear-cut. Adequate regeneration of the cut stand with aspen will occur only if nearly all
the mother stand trees are felled and if there are at least 50 trees or 20 square feet per acre
of quaking aspen in the mother stand (Perala 1977). The sprouting character of aspen gives
it an initial advantage over the more tolerant balsam fir that must regenerate from seed
originating on felled trees or blown in from trees adjacent to the clear-cut. Examination of
records from a long-term study of this mixture in Wisconsin (Cooley and Lord 1958) showed
that at the time the study was established the ratio of average diameters (aspen/balsam fir)
was about 2.0. The study was located on a site with a rating of 69 for aspen site index. An
attempt was made to fabricate tree lists for pure and mixed stands that had characteristics
similar to those of the Wisconsin study area (Table 6.1). Pure stands of aspen and balsam
fir were simulated together with the appropriate mixed stands, e.g., stands 1,4, and 7 were
simulated in the same projection (Figure 6.3). Simulating both pure and mixed stands at
the same time makes it possible to continuously determine type and intensity of interaction,
the subject of the next section.

3. Interaction analysis using Haskell's Coordinate system.


The numerical values for pure and mixed stands of balsam fir and aspen, such as those
used to construct the time series graph in Figure 6.3, are easily converted to Haskell coor-
dinates and plotted in the coordinate system. Figure 6.4 shows points 10 years apart on
the trajectories of interaction type and intensity for three sets of stand structures shown
in Table 6.1. I shall refer to the structures as A4B4 (aspen mean dbh 4 inches, balsam fir
mean dbh 4 inches)' A4B3 and A4B2. Because the type of interaction is (-, -), the choice
of metric for interaction intensity is not at issue because Ixl + Iyl = Ix + yl when x < 0 and
y < o.
Within 10 years of the start of projection, the A4B4 trajectory was at 253 and intensity
0

nearly .3. Another 20 years showed interaction intensity had increased to .475 and type
102 Patterns of Interaction in Mixed Forest Stand Dynamics

Table 6.1. Initial conditions used to fabricate pure and mixed stands of quaking aspen and
balsam fir that are needed as initial conditions for trial projections of mixed stand dynamics.
(fixed) (variable)
stand basal area number of
number specIes per acre mean dbh trees per acre
pure quaking aspen 41 4.0 460
stands 2 balsam fir 41 4.0 460
3 balsam fir 41 3.0 820
4 balsam fir 41 2.0 1840
mixed 5 quaking aspen 41 4.0 460
stands & balsam fir 41 4.0 460
6 quaking aspen 41 4.0 460
& balsam fir 41 3.0 820
7 quaking aspen 41 4.0 460
& balsam fir 41 2.0 1840

pure balsam fir

:: 160

...
Q)

u
~, pure aspen
CIJ
~
Gi 120 '~
0.
CIJ .~
...CIJ
Q)

..
CIJ
CIJ
.c
80
"".
Q)
> 40

°0~----~2~0~--~4~0~--~6~0~--~8~0~--~1~0~0----~1~2-0-----1~40
elapsed time, years

Figure 6.3. Time series of projected A4B2 live standing crop basal area for a pure stand of
each and one mixed stand of balsam fir and quaking aspen. Projection was made with the
generalized forest growth projection system (USDA 1979) with modified mortality equations
given in equation 6.5. Site was set at site index 60 for both aspen and balsam fir. Experience
shows both species' populations decline more rapidly than given by the model.

to 261 o. In theory, intensity should rem am at about .5 because there is in reality only
about one acre's worth of abiotic flux available. The A4B4 system rapidly passes through
this region toward greater intensity and toward the (0, -) axis near which exclusion of the
aspen should occur. There are two categories of "exclusion" that can operate here. First,
there is the event wherein the top and root system of the last aspen tree in the stand dies so
Patterns of Interaction in Mixed Forest Stand Dynamics 103

that nc form of propagule is available for a future generation. This is the event of "total"
exclusion. Also, there is the event of aspen trees dying from competition until there are so
few trees left that after a clearcut the species will not regenerate with a sufficient number
of stems to dominate the site. I call this event "effective" exclusion. Effective exclusion in
one rotation increases the probability of total exclusion in the next rotation. Perala (1977)
describes the conditions under which this can occur.
Estimates of either type of exclusion can be made with the help of equation 6.6, a gen-
eralized computer program, and the idealized stands summarized in Table 6.1. Accordingly,
I estimate that "effective" exclusion occurs in the A4B4 system only after 60 years elapsed
time. This corresponds to a stand age of 90 to 100 years. "Total" exclusion does not occur
in A4B4.
The A4B3 system shows an interaction type near 244 after 10 years (Figure 6.4),
0

approximately 10 degrees less detrimental for the aspen than the A4B4 condition. The
trajectory is not unlike that for A4B4, although they cross when the trajectory for A4B3 takes
a sharp turn toward the (0, -) axis. Neither "effective" nor "total" exclusion is estimated to
take place in the A4B3 system during a projection of over 100 years.
Projection of the A4B2 system showed an initial interaction type near 220 0
Near the

center of quadrant III, aspen and balsam fir are affecting each other in approximately equal
amounts. This is, of course, not accidental. The /313 and /323 coefficients in equation 6.6 give
weight to the effect of aspen on balsam fir and vice versa. When the (Y2/n2)/(YJ/nd (ratio
of mean diameters) values are approximately 2:1, as is the case in A4B2, the ratio cancels the
differential weight given by the f313 and /323 numerical constants. In theory, these numerical
constants are thought to be different due in part to different shade tolerances of the two
species. The projection of this condition shows that aspen affects balsam fir approximately
the same as balsam fir affects itself even though there is a two inch difference in mean sizes.
After 30 years of elapsed time the projection shows a marked change or "corner" to where
the intensity is not changing but type of interaction is rapidly becoming more detrimental
to the aspen.

D. Controlling Interactions by Thinning


A primary reason for locating mixed stand interaction trajectories in Haskell's coordinate
system is that the resulting pattern may show relationships not evident when the same
information is plotted in a different coordinate system or left in tabular form. Further, the
relationship may provide a cue to action.
As a case in point, compare the Haskell coordinate system plot of the A4B2 trajectory
(Figure 6.4) and the time series plot of the separate (pure stand) and together (mixed stand)
development (Figure 6.3). From the latter figure it is difficult to see that a major shift in
the nature of interaction occurred in the decade between 30 and 40 years of elapsed time.
On the other hand, the shift is very evident in Figure 6.4. Interest is, of course, in cues
that lead to "effective" action. The "corner" in the A4B2 trajectory indicates quaking aspen
is rapidly losing the struggle for space in the forest to balsam fir trees. Knowledge of the
"corner's" existence may, if it is desired, be used to develop an action that prevents the
corner, and what follows, from occurring. In order to test the usefulness of knowledge about
the interaction trajectory for prescribing thinning, I attempted to maintain a fairly constant
type of interaction, say between 210 and 230
0
0
so that neither population is projected to
,

have been either "effectively" or "totally" excluded (Figure 6.5).


Clearly, the interaction trajectory can be maintained in the vicinity of 215 to 225 0 0

by thinning. The more traditional time series form of the standing crop development for
the mixed stand shows that balsam fir was thinned at elapsed times of 10, 30, and 60 years
in amounts of 20, 30, and 30 square feet, respectively (Figure 6.6). Aspen was thinned 30
square feet at time 90 (from 97 to 67 ft). Of course, I do not advocate the last thinning unless
104 Patterns of Interaction in Mixed Forest Stand Dynamics

I I,

10.-1
quaking .I p a nl

Figure 6.4. Mixed stand trajectories for A4B4, A4B3, and A4B2 stand conditions. Exclusion
of aspen should occur at or near 270 0 The interaction trajectories for A4B3 and A4B2 are

markedly different. "Effective" exclusion of aspen is estimated to occur only in the A4B4
system projection.

1800 ( - ,0) -n:-rIL:r::::t==E3~


balsam 9
fir
.7

.5

.3

quaking
aspen
(0 ,-)
270°

Figure v.5. Pattern of interaction type and intensity when balsam fir is thinned from the
mixed stand at 10,30, and 60 years following start of the projection. Type of interaction is
easily maintained between 215 and 225
0 0

Patterns of Interaction in Mixed Forest Stand Dynamics 105

one wishes to maintain an interaction near 220 Interaction theory dictated that the same
0 •

amount of the mixed stand basal area removed in a thinning was also to be removed from the
pure stand of the species being thinned. Just as a thinning causes a discontinuity (increase)
in the standing crop over time graph (Figure 6.6), it causes a discontinuity (increase) in the
interaction intensity. This short term jump in intensity does not persist. The intensity never
exceeds .55 and remains quite close to the expected .50 level, as can be seen in Figure 6.5.
The implication of the pattern of points in Figure 6.5 is that type and intensity of
interaction between populations of natural objects may to some extent be controlled by
man.

100
~ ----~:
~_~' aspen:

Q.CII
-...
N

o.. ..CJ
80
1/'
,~'
i
L/
I
CJ eu I
I I

til ..
60 I I
I
I
c CII

~
I
.- Q.

-.
"1:1
C eu
eu CII
!/leu 40 balsam fir
CII-
> !/l
._ eu
-eu
..a
20

0
16


III

.c 14
CJ
.c c
..c- 12
"1:1 -
"1:1

--
'"
II

CJ '"
-"1:1
Q..
ceu 10

8
"c .-
><
eu E 6
E.c
• CII
4
c
2

00 20 40 60 80 100

elapsed time, years

Figure 6.6. Traditional time series graph of thinning trial showing standing crop of mixed
stand as a function of elapsed time. From a production perspective, the thinning at time
60 is not advisable. However, it significantly altered the trend in interactions as shown in
Figure 6.5.
106 Patterns of Interaction in Mixed Forest Stand Dynamics

E. Geometrizing Silviculture
Let me immediately reduce 'silviculture' to those operations undertaken to control stand
occupancy by trees. (See Chapter 3 for a discussion of measures of occupancy: "stocking,"
"density," "packing.") Further, in this brief section I do not limit myself to mixed stands.
The normal course of development for a forest stand is for density to increase as the tree
stems increase in thickness. It is well-known that if density is too great the trees may stagnate
or many of them die for lack of abiotic flux that a larger growing space would provide.
Thinning operations have as a goal the control of occupancy so that desired objectives are
reached, e.g., a merchantable size is reached by a specified age. Tree classes, e.g., dominant,
codominant, intermediate, suppressed, are often used to categorize the social environment
of trees in close proximity and supply a basis for a partial thinning typology: thinning from
above, thinning from below, high and low thinning.
Several different stand and tree properties can be used as cues to the need for thinning.
Basal area density provides such a cue. In other cases the live crown ratio (percent of total
tree height in full live crown) in high density stands can be used as a cue to action. These
criteria, basal area density and live crown ratio, appear to me to be proxies for intensity
of interaction. They are used apparently because they are either directly perceivable or
estimable with a small effort given appropriate tools. Of course, usefulness of live crown ratio
is limited to stands in hydric and mesic conditions. In arid regions intensive interactions
may be taking place between trees widely separated. Also, in so-called agroforestry, widely
spaced trees may be competing intensively with interplanted food crops grown under a
tanguya system.
It is well known that the basal area densities when thinning should be done, for example
in red pine, depends on the site productivity of the growing area. More productive sites can
carry higher densities. Thus, there is not a one-to-one mapping between basal area density
levels and thinning needs, either between sites for the same species or between different
species on the same site. I suspect that type and intensity of interaction (0 and rare more
consistently related to thinning need. Thus, rather than saying, for a hypothetical example,
that one needs to thin a 40-year-old red pine plantation established at 1200 trees per acre
as follows:
site index 90, thin to 130 square feet per acre
site index 70, thin to 110 square feet per acre
site index 50, thin to 90 square feet per acre,
I would say for a 40-year-old red pine plantation
maintain 215 <:; 0 <:;225 and .45<:; r <:;.55,
0 0

when dominant + codominant trees are assigned to the vertical axes and intermediate +
suppressed trees are assigned to the horizontal axes of Haskell's coordinate system. Thus,
it is anticipated that these coordinates would apply to all sites, but the ranges for other
conifers and hardwoods would probably be different.
The interaction between tree overstory and shrub-grass-forb understory vegetation is
also of critical importance in the silvicultural measures undertaken before harvest cuts for
natural regeneration. Contrary to tree-tree interactions where moderate to low interaction
intensity is preferable, one seeks high intensity to the detriment of the understory vegetation
immediately prior to either a harvest or reproduction cut. New tree seedlings become estab-
lished more easily and have less competition when there is a nearly excluded shrub-grass-forb
component.
In sum, several important silvicultural operations are designed to control type and
intensity of interaction between plant components of forest stands. A recognition of this
fact may help to develop cues more directly related to interaction type and intensity. At
a minimum, Haskellian coordinates can be used to assemble in one framework results of
current prescriptions based on observable and estimable cues, pending the development of
Patterns of Interaction in Mixed Forest Stand Dynamics 101

cues more directly related to type and intensity.

F. Research Problems
1. Recent legislation has directed foresters to manage forests in such a way that "diversity
of plant and animal communities ... is at least as great ... " Analyze the predicate
"diversity" for its degree, order, and reference class using methods of Chapter 1.
2. Argue the merits of each set of presuppositions for mixed stand dynamics studies devel-
oped in this chapter.
3. Is there an experimental design in which information for both Wit and Gause types of
analyses can be obtained? If there exists no experimental design, invent one.
4. Develop an analog of the stocking guide methodology developed by Gingrich (1967) for
the (-, -) quadrant in Haskell's coordinate system. Begin by letting the B line in the
stocking guides coincide with the (0,0) circle.
5. Criticize the argument that since foresters are now legislated to be more concerned
about diversity dynamics, hence species exclusion, they need an analytical framework
for studying mixed stands in addition to the traditional utilitarian one due to Wit.
6. Develop a set of thinning prescriptions for mixed stands of your choice expressed as (!
and r targets.
7. Make a historical review of "unimportant" tree species and their financial standing
today.
8. Link the concepts of diversity and species extinction/exclusion. Begin by developing an
equation for diversity dynamics.
9. The growth model described in USDA (1979) was purposefully designed to account
for mixed stand dynamics. It has been revised (Belcher 1982). Test the mixed stand
dynamics projected by this revised model called STEMS.
10. Test the projections made by an individual tree growth model of your choice on a mixed
stand that has Gause-type initial conditions. Compute interaction type and intensity
that the model estimates. Do the results agree with your experiences?
Part III
Synthesis
7: Haskell's Coordinate System
in Synthesis

Nearly three hundred years ago Gottfried Wilhelm Leibniz observed:


But as things are, man's knowledge of nature seems to me like a shop, well stocked
with goods of all kinds but lacking any order or inventory. (Leibniz (ca 1683)'
Parkinson 1973).
Has the situation changed in the nearly 300 years since Leibniz made these remarks? The
general argument is no, if anything, it has become much worse. Leibniz's "shop" has become
at least a warehouse, if not a maze of them. The continuous accumulation of scientific knowl-
edge, accelerating in recent years, has forced an organization to evolve. That organization
is the set of scientific disciplines. Ackoff (1973) says that scientific disciplines are nothing
more than categories that facilitate filing the content of science. Others contend that the
accumulation of knowledge by scientific disciplines contributes to the fragmentation of our
knowledge of nature. This, as a result, places man in a poor position when he is expected
to intelligently "govern" nature. Nature simply is not organized like our knowledge of it is,
nor like our institutions where we study nature.
One of the major thrusts to deal with the discrepancy between man's filing system on
nature and the reality of it has been called inter-, multi-, or trans-disciplinarity of teaching
and research (Jantsch 1972). Ackoff (1973) explains that this approach did not prove fruitful
for long because the interdisciplines sought recognition and status just as the disciplines had.
Their focus became the same as the discipline, and their role the same, discovering and filing
new knowledge, but on slightly different parts of nature.
For many years the application of fragmented knowledge has had unexpected results,
second order consequences, when applied to the natural environment. When the magnitude
of the consequences rose to unacceptable levels, Congress responded with the National En-
vironmental Policy Act of 1969 wherein it is mandated among other things that planning be
done by interdisciplinary teams. If research scientists do not integrate the new knowledge
they discover, with both old and other new knowledge, the burden falls primarily to the
users of knowledge. Normally, this puts users in a very difficult position. Scientific research
should be organized to produce knowledge in ordered wholes that planners and managers
can use with a minimum need for alteration on their part.
Forestry and ecology are scientific disciplines that are integrative in their basic approach;
the former is more narrowly defined, but they both deal with chunks of nature as integrated
wholes. The "systems approach" was not exactly a revelation to many in these disciplines.
Bakuzis (1978) has aptly defined forestry as a special purpose systems science. But, a
distinction must be made between what a discipline claims to be and the method of operation
of scientists in the discipline. Forest ecology can be synthetic in nature, concerned about
the whole, or it can masquerade as synthetic but actually be analytic. The latter, it seems,
is becoming more the rule than exception. The objects of research are often described
initially in holistic terms, but soon the scene changes to a small facet and never again is
110 Haskell's Coordinate System in Synthesis

the indivisible whole, the system, mentioned. Ackoff (1973) describes analytic thinking as
outside-in thinking. The problem, the "outside," is explained by the "inside," the parts, or
sub-systems. Contrarily, he describes synthetic thinking as inside-out. The "problem" is
now the inside, and it is explained by the outside, by the system or system of systems of
which it is a part.
The purpose of this chapter is to contribute to the integrative thrust of forest ecology
by showing how knowledge of Haskell's coordinate system can contribute to our ability to
do "inside-out," that is, synthetic, thinking.

A. Synthesis in General
Synthesis is the process of producing a structured whole. A synthesis is the structured
whole. Eliminated are heaps of knowledge or information bits as well as neatly stacked bits
that are not based on a structural principle. A synthesis need not be a theory, a web of
propositions related by the relation of entailment, but it may. Theorizing is, after Charlotte,
spinning webs, but the strands are of law stature, whereas theory-less synthesizing is spinning
with lesser quality stuff.
If synthesis is the assembling of parts into a structured whole, the key concepts are
"assemble" and "structure." What are the connotations of "assemble?" Of "structure?" To
assemble implies some form of aggregating of parts into part groups. "Structured" connotes
"ordered." Structured assembly implies grouping in an ordered manner which translates
roughly as classification. Thus, a synthetic approach to science will have its roots in the
principles of classification, at least in its early stages.
Of particular concern is the availability of tools and procedures to accomplish the syn-
thesis task. Are there tools and procedures that are more helpful than others? If so, what are
they? But, before choosing tools, the nature of the probable obstacles must be understood.
Of course, the tools selected must work on the obstacles encountered. In my opinion a major
obstacle to synthesis in forest ecology is diversity or variety in the parts or elements that
are candidates for assembly. Variety may enhance or spice our lives, but it inhibits synthetic
science. The situation is capsulized in: "The diversity of nature will persist as long as things
(entities) are focused on" (Margenau 1972). I am speaking, then, of persisting diversity.
A tool traditionally used to deal with variety and diversity is abstraction, the process
of leaving out of consideration certain properties in other things. To better understand
how abstraction functions consider a continuum of possible relations between parts that are
candidates for assembly. I have identified four hypothetical categories in the continuum:

Relation Between
Candidate Parts Description of Parts

strict identity are the same in all respects

partial identity are the same in most respects

partial diversity are the same in a few respects

total diversity are the same in no respect

At the bottom are parts totally diverse; they are the same in no respect. If true, the
synthesis effort is misguided because none is possible. On the other extreme, strict identity
is possible but with zero probability. No two things are ever identical in all respects. Reality
is somewhere between these extremes of strict identity and total diversity. The process of
abstracting suggests that diversity be dealt with by focusing not on what is different about
Haskell's Coordinate System in Synthesis 111

the objects, rather, on what is similar about them. The process of abstraction plays a large
role in Ackoff's conception of science in the systems age: the search for similarities among
things that are apparently different (Ackoff 1973). But what does one do with partial identity
and partial diversity?
Partial identity is the basis for classifications, generalizations and laws expressing
the patterns or invariants of things and events amidst variety and change (Bunge
1967a).
The goal need not be to eliminate part diversity, only to control it. One means of
doing this is by subordinating it to something else, to an abstraction. An important set of
abstractions are called integrative principles (Margenau 1972). Before discussing them let
me state completely the strategy:
Because the diversity of the parts of nature will persist as long as things (entities)
are focused on, they must be made to vanish or be subordinated to an integrative
principle before a synthesis is possible.

B. Integrative Principles
An integrative principle is a concept that has power to organize information and knowl-
edge. Concepts suggested as having this power are: invariance, symmetry, quantization,
feedback, homeostasis, isomorphism, and interaction (Margenau 1972, Andrews 1972, Tay-
lor 1972, Thornton 1972, Laszlo 1972). Laszlo and Margenau (1972) examined the concepts
of invariance and symmetry, and argued that symmetry is a special case of invariance with
respect to movement (rotation, reflection, translation) about a point or axis. Rhythm in
music as invariance with respect to time displacement, and ornamental patterns on tapestry
as invariance with respect to space displacement were also suggested. Their major emphasis
was, however, on the character of scientific law statements. In order for laws to be universal
they must hold in every reference frame; i.e., they must be invariant with respect "to a
change of coordinates corresponding to passage from one inertial system to another" (Laszlo
and Margenau 1972).
Invariance is a relation concept with predicate degree three: J(x, y, z), where x is the
object prior to the "change" operation z, and y is the object following z. If invariance
is a relation concept, are not the other integrative principles also relation concepts? Cer-
tainly homeostatis is, as are isomorphism, feedback, and interaction. Only quantization
(continuity-discontinuity) is in doubt.
Of the four kinds of concepts, individual, class, relation, and quantitative, relation
concepts have the power to deal with diversity in a way that counteracts what Margenau
calls the "dissolution of learning." The implication is that class concepts will not prove as
powerful for synthesis as relation concepts. On the other hand, quatitative concepts are
more powerful than relation concepts when the need is for synthesis.

c. Forms of Scientific Synthesis


Without claiming to be exhaustive, there appear to be three general forms of synthesis
In use today: encyclopedic, mathematical, and quantitative. The three are not mutually
exclusive, so what I am suggesting are forms based on dominance. The encyclopedist's
strategy is to bring together existing information from a variety of sources, perhaps filling
in crucial bits of missing information through analysis, and organizing it in such a way as
to create a holistic picture of some chunk of reality. Often the information takes the form of
verbal statements, tabular data, simple graphs, and mathematical equations. The challenge
in this form of synt·hesis is to bring together and structure information from a wide range of
the sciences, perhaps even the arts. Encyclopedic synthesis is sometimes criticized by those
engaged in other forms, but it should be viewed as a necessary first step, as the laying of a
112 Haskell's Coordinate System in Synthesis

foundation for other forms. Large scale efforts at quantitative synthesis are likely to be of
limited success if a thorough encyclopedic-type synthesis has not been pursued to a fruitful
concl usion.
In the mathematical form of synthesis I include the combination of parts using all
mathematical procedures short of mathematical equations (algebraic, differential, difference,
etc.). Included are ordinal and cardinal numbering, points, lines, and curves in coordinate
systems of various sorts, and the use of advanced mathematical topics such as group algebra,
and category theory. The mathematical form is more oriented to geometry than, say, to
mathematical analysis.
A particularly fruitful procedure in the mathematical form is the construction of a graph
of the parts thought to be assembleable into a whole. A graph is simply a diagram showing
relationships between parts, wherein the graph node (usually shown as a point) represents a
part, and an arc connecting two nodes denotes the existence of the relationship between the
connected nodes. Rashevsky (1954, 1960) was one of the first biologists to employ arc-node
graphs of biological systems. Others have followed with the result that a very large literature
currently exists on applications of graphs toward the solution of many complex problems as
well as the relationship between graphs and nonquantitative areas of abstract mathematics
(topology, group theory, coloring problems, etc.) (Deo 1969, 1971; Levins 1975; Henry 1980).
I shall return to an application of the mathematical form of synthesis after a brief discussion
of quantitative synthesis.
By quantitative synthesis, I mean the use of tools commonly studied in that branch
of mathematics called analysis, especially differential and difference equations. Certainly a
very powerful tool for synthesis, and widely used with success in the physical sciences, it
has been applied repeatedly in the biological sciences. Its success ratio has been lower in
ecology and biology than in the physical sciences, judging by the relative scarcity in biology
of law stature statements formulated as differential equations. Perhaps the best example
of near-law statements are the differential equation form of the allometric law, and of the
general exponential growth equation characterization of "That which results from biological
growth is itself typically capable of growing" (Medawar 1940). A very bold effort at this
synthesis form was made by the grassland biome study group of the International Biological
Program (Innis 1978). Other biome study groups used the quantitative approach but placed
less reliance on it for successful synthesis.

D. Mathematical Synthesis Using Haskell's Coordinate


System
The mathematical form of synthesis represents a sort of middle ground between the
mostly verbal encyclopedic form and the quantitative form where computer modeling has
become widespread. The recent tendency has been to give a small effort to the encycl9pedist
form, skip the mathematical entirely, and become deeply involved in quantitative synthesis.
Part of the reason for skipping the mathematical form of synthesis is that it may be phe-
nomenological in its approach (Bunge 1967a). To use a phenomenological approach is not
a sin; to use a phenomenological approach forever may be. I argue that something can be
gained by investing time in the mathematical form of synthesis. This section is devoted to
elucidating that something.

1. Example one.
A hypothetical example suffices to exhibit some aspects of mathematical synthesis that I
want to emphasize. Imagine a forest system wherein we have identified four component pop-
ulations that are of particular importance: conifer trees, hardwood trees, a phytophagous
insect, and white-tailed deer. Developing a representation of the system begins with the
components abstracted to form the nodes in a system graph. Each node represents a pop-
ulation of the system element. On the basis of prior knowledge I know that some system
Haskell's Coordinate System in Synthesis 113

elements (populations) interact with one another with respect to a particular process, e.g.,
nutrition. Combinations of system elements that interact directly are identified by drawing
an arc between their respective nodes. The insect and deer populations do not interact with
respect to nutrition, hence do not have a connecting arc. In place of the arc between nodes
I could have strung the words "interact directly with one another" The arc represents the
relationship stated in words. It is crucial, of course, that the arc always represents the same
verbal statement throughout the graph.
A refinement of a graph is called the directed graph. A directed graph makes explicit
which components are affecting other components. This is done by replacing the single arc
by arrows that represent the direction of effect. No arrow from deer to hardwood tree,
indicates that the deer has no effect on hardwood tree nutrition, but hardwood tree does
have an affect on deer nutrition. Thus, in going from a graph to a directed graph, arcs are
replaced by one or two arrows indicating the direction of effect.
A signed digraph is produced by placing a sign (+ or -) adjacent to each arrow head in
a directed graph. Sign placement rules are simply that a + sign denotes enhancement of the
process due to the affecting population, and a - sign denotes a diminished process rate due
to the affecting population. A signed digraph can be simplified by replacing the two signed
arrows with a sign combination from the coaction cross tabulation of Chapter 2 (Figure
7.1a). In the case of the deer-hardwood tree interaction the positive effect of hardwood on
deer is paired with a 0 (symbol for no affect) for the effect of deer on hardwood tree. Since
the deer and insect populations do not interact directly, there is no arc connecting them. It
makes little sense to insert a (0,0) sign combination, so it is omitted.
The introduction of the sign combination brings to mind the evolution of the coaction
cross tabulation and its combination with the phase plane to form the Haskell coordinate
system. It is a small step in pursuit of this evolution to replace each sign combination in
Figure 7.1a with a small Cartesian coordinate system (quadrant I only) (Figure 7.1b). Each
coordinate system contains a (0,0) point and a ~ (0,0) point in the quadrant bearing the
appropriate sign combination. The axes of each coordinate system are labeled specifically
for the elements interacting. Thus, in looking over the five sets of Cartesian quadrant I's, it
is clear that one way to assemble the information in the graph is by locating one point in
four dimensional space for each interacting pair of populations. Each of the four dimensions
is used to represent one system element.
The information in Figure 7.1c can be assembled by following a different logic. In
Chapter 4 it was shown that any phase plane containing a (0,0) and ~ (0,0) point could
be mapped into Haskell's coordinate system. If each of the Cartesian coordinate systems in
Figure 7.1 b is replaced by a Haskell coordinate system, the ~ (0,0) point can be located as
a point in the coordinate system, thus showing the type and intensity of interaction between
system elements (Figure 7.1c). What started out as a graph of a hypothetical forest system
is characterized as the original set of system elements and several occurrences of a coordinate
system. It seems logical that we should attempt to reduce the number of coordinate systems
to unity.
The intriguing last step in going from graph to map is to imagine that the symbols in
Figure 7.1 c are portrayed on something like a store window. Each node has as many copies
of the symbol laid one on top of the other as there are arcs leading to and from it. For
example, the insect node has two copies and the conifer tree node three copies. Consider a
bystander on the sidewalk observing the window. Working from inside the store I take one
copy of each of the deer and conifer symbols and place them adjacent to one another (deer
to the right from the bystander's vantage point) at the spot marked "." in their connecting
coordinate system. Repeating the step for hardwoods and deer places the symbol pair near
the (0, +) axis ana shows that there are no more copies of the deer symbol left. Moving on
to all the arcs between system elements and placing the element symbol pair at the position
of the "." in the intervening Haskell coordinate system we find, upon completion that the
114 Haskell's Coordinate System in Synthesis

Figure 7.Ia. Simplification and


completion of signed digraph
using symbol combinations of
Chapter 2. The no effect sym-
bol, "0", is used to complete a
sign combination only when the
other sign is not a "0".

Figure 7.1 b. Reminder that sign


combinations in above figure are
based on comparisons of points in
a phase plane. All point positions
are schematic. The axes are la-
beled with the right-most popu-
lation on the abscissa.

Figure 7.Ic. Graph of system af-


ter each phase plane has been em-
bedded in the Haskell coordinate
system. Populations on the right
go with the horizontal axes and
those on the left wi th the vertical
axes. Each point in Haskell's co-
ordinate system stands for a par-
ticular point in time. I do not
assume it to be an equilibrium
point.
Haskell's Coordinate System in Synthesis 115

store window now contains five identical coordinate systems each containing a single symbol
pair, and all the nodes are gone. It is a simple matter to bring the coordinate system axes
into coincidence thus forming a single coordinate system. All system element interactions
have been mapped into one set of axes, and both type and intensity are represented (Figure
7.2).

.
(0,+)

~r ~(
(- . +)

r 'if)
~r~V1
(-,0) -+--------- (+,0)

(-,-)

(0,-)

Figure 7.2. End result of transformation begun with Figure 7.1c. Convention is that the
entity on the right has the effect on the left entity contained in the left sign. For example,
the insect has the effect "-" on hardwood tree and hardwood tree the "+" effect on insect,
because symbols for hardwood tree and insect occur in the (-,+) quadrant. Positional
locations are for a particular time. The symbol pairs may be assumed to move in the
coordinate system.

To better understand what has just been done and its relation to synthetic thinking
in science, compare the contents of Figures 7.1b and 7.2. In 7.1b the system elements are
dominant. The interaction is something that exists between elements and is implicit in the
relative positions of the (0,0) and ~ (0,0) points. Any number of new elements can be added
to the scheme as long as the necessary arcs and phase planes are also added. In Figure 7.2
the interaction types are dominant. The element pairs are simply particular instances of
a type and intensity of interaction. No new type of interaction is possible because nine is
the theoretical totality of possible interaction types between elements. Hence, new elements
may be added to the scheme simply by inserting the appropriate symbol pair in the proper
116 Haskell's Coordinate System in Synthesis

location.
The process in going from Figure 7.1b to 7.2 is somewhat akin to turning 7.1b inside-out.
What is particularly satisfying in doing so is that the potentially large number of system
elements, the diversity, has been neatly harnessed by subordinating the things to a maximum
of nine possible relations between things. Because of the power of the Haskell coordinate
system to harness variety through subordination of things to interaction between things, it
is an important new synthesis tool.

2. Example two.
Another synthetic bit of science is provided by Dindal's expansion of Haskell's coaction
cross tabulation (Dindal 1975). This predominately encylopedic synthesis can be mathema-
tized.
Dindal expanded the coaction cross tabulation by differentiating between facultative
and obligate symbiont relations and by identifying three degrees of negative effect intensities
(Figure 7.3). These categories established the effects forming the margins of the enhanced
cross tabulation. Combinations of symbols for appropriate rows and columns produced the
contents of each cell. Most of the interactions identified had been reported in the literature.
The references supporting the factual occurrence of each interaction are presented by Dindal
in a large table. The act of putting symbols together revealed two interactions that had not
been reported previously (+!) and (! +). These were included and named hemi-mutualism
by Dindal. Other interactions were named as shown in Figure 7.3. Upon seeing this figure I
was certain that it could be mapped into Haskell's coordinate system. What follows are the
results of this effort.
The symbols in each cell of Figure 7.3 are assumed to summarize the relationship be-
tween two points in a phase plane. When a symbol pair includes a (!) sign, e.g., Figure 7.4a,
it simply means that the "separate" point is located on one of the coordinate axes (point
marked "x" in Figure 7.4b). This is because A cannot exist without B, hence the amount of
A under the separate condition is zero. If neither A nor B can exist without the other, the
"separate" point corresponds with the origin of the coordinate system in Figure 7.4b. When
a symbol pair contains a :: or =, it is assumed that they connote increasing intensity of an
inhibiting effect. Figures 7.4c and 7.4d show the symbol meanings and how the symbols are
converted to points in a phase plane. The procedures shown in 7.4b and 7.4d were repeated
for each cell or group of cells in Figure 7.3 except those containing a ~ symbol.
Figure 7.5 shows the result of mapping 44 phase planes into a single Haskell coordinate
system, dropping the unneeded axes and placing symbols at the "together" equilibrium point.
Twelve of the symbols are due to Haskell (those involving =, < or > signs), and the rest are
Dindal's.
Interactions having one or more obligatory relations, !, are placed farther from the (0,0)
circle than their nonobligatory counterparts because the base for percentage calculation
will likely be smaller when one symbiont produces nothing in the "separate" condition.
These interactions also have a slightly different angular position than their nonobligatory
counterparts because of the probable difference in direction from "separate" equilibrium to
"together" equilibrium. Examples are interactions 14 and 15, 13 and 16,40 and 37, and 38
and 39.
A qualification is needed for the names in quadrants II and IV because several scientists
have argued there is need for either II or IV, but not for both. It is possible to have just
one quadrant if criteria used to assign symbionts to the axes are interaction independent.
Ordinarily, one establishes a criterion, for example, 'member of the animal kingdom' for the
vertical axis, and 'member of the plant kingdom' for the horizontal axis, and adheres to them
throughout. Doing so puts plant-phytophagous insect interactions in quadrant II (Mattson
and Addy 1975) and insectivorous plant-insect interactions in IV. If I insist on using only II
or IV, I may have to relabel the axes for every pair of organisms.
Haskell's Coordinate System in Synthesis 117

SYMBIONT A
j:"'Ul.TAnvt OBL'C;ATOIIY
o + +
+

+
+
~--~~~~~~~~--~~~~--~~~
M
8 ±
I

¥~~~~~~~~~~--~--~~
8

~ANTI.WTIC. C'...~~

~ OPPOSITE C0t.481WA,T'0N 01= .... "' ....TIOH$H.P .... I...ACAOV L ...... UO "'Bov.

Figure 7 .3. Dindal's coaction cross-tabulation formed by differentiating between facultative


and obligatory relations, three degrees of negative effects, and the usual +, -, and 0 relations.
The ~ sign denotes a relation where individual or population is alternately stimulated and
inhibited by the relationship (from Dindal1975).

Figure 7.5 shows a concentration of interactions in quadrant III. If Dindal's survey is


representative of the studies of biologists and ecologists, and if scientific study in the analytic
mode usually leads to partitioning of a general category, then ecologist's bias toward studies
of conflict and competition is confirmed. Risch and Boucher (1976) surveyed 12 ecolqgy texts
published in the period 1971 - 1976 to test their claim that practically the entire discussion
of organismic interactions in ecology texts has centered on predation and competition. They
found a total of 718 pages devoted to interspecific interactions. The breakdown is as follows:
321 pages on predator-prey interactions (quadrants II and IV), 362 pages on competitive
interactions (quadrant III plus adjacent axes), and only 35 pages on any kind of mutualistic
relationship (quadrant I plus adjacent axes). Their claim is that ecologists find what they
look for. If mutualisms are not looked for they will not be discovered and they will remain
"interesting but eccentric exceptions to the general rule" (Risch and Boucher 1976). Actually
the situation in quadrant I is worse than indicated by the number of symbols in Figure
7.5, because only two types are generally recognized + + (proto-cooperation) and tt
(mutualism). But, Figure 7.6 is like any map; known "territory" is dotted with the locations
of familiar objects'. Sparse regions on a map may be due to the region's vacuous nature or
because it is relatively unknown. Sparseness for the latter reason has traditionally led to
exploration.
118 Haskell's Coordinate System in Synthesis

Figure 7.4a. Symbols for facultative-


+ obligatory interaction Dindal called

/ +,
+ hemi-mutualism. (+) means relation
is enhancing, and that B is faculta-
Effect of symbiont A Effect of symbiont B tive. ~ means the relation is enhanc-
on symbiont B on symbiont A ing, and that A is obligatory.

+t/
Figure 7.4b. Phase plane interpreta-
Symbiont A
tion of hemi-mutualism. A is obliga-
tory on B, hence the (0,0) point oc-
curs on the B axis.
L -_________ I-----. ~

Symbiont B +

Figure 7.4c. Symbols used to de-


note intensity of negative effects. (-)

I \
means the individual or population is
smaller because of subnormal growth.
Effect of symbiont A Effect of symbiont B (=) means individual or population is
on symbiont B on symbiont A gradually decimated or forced to mi-
grate. (==) relation results in immedi-
ate death of individual or population.

Symbiont A

Figure 7.4d. Phase plane interpreta-


tion of three degrees of negative ef-
fects.
Symbiont B

The interactions containing the-+: symbol are for nonequilibrium conditions, interaction
trajectories, and cannot be associated with a single quadrant or axis in Figure 7.5 .. For
example, the + -+: interaction includes quadrants I and IV. The -+: == interaction includes
numbers 14 and 26; -+:~ includes numbers 41 and 4; 0: includes 42 and 17, and so on.

E. Discussion
Figure 7.5 shows 44 different interactions synthesized into Haskell's coordinate system.
Critics may suggest they are overcrowded into one coordinate system, and although a syn-
thesis, it is not an elegant one. There are opportunities to relieve what appears to be clutter,
especially inside the (0,0) circle. Most importantly, parties to the interaction have been left
unspecified. Some types of interaction may typically occur between plants, others between
plants .and animals, and still others between animals. By making this simple differentiation
between parties to the interactions we can improve Figure 7.5. How this can be done is the
subject of the next chapter.
Haskell's Coordinate System in Synthesis 119

SYMBIONT
A

SYMBIONT
8

Figure 7.5. Geometrization of expanded coaction cross tabulation in Figure 7.3. Placements
of the symbols are qualitative. Names are mostly from Dindal (1975).

F. Research Problems
1. An often used basic step in developing a simultaneous equation model to project commu-
nity population levels is to construct an interaction matrix. Find several examples in the
literature and perform the transformations that subordinates the things (populations)
to the interactions between populations.
2. What is your reply to someone who says that discussing concepts, constructs, planes of
perception, and integrative principles is philosophy, not forestry or ecology? Why?
3. Every few years the National Research Council of the National Academy of Sciences
inventories the scientific disciplines. Assemble these reports and prepare a graph of
numbers of recognized scientific disciplines over time. On the basis of your graph, what
will the number of disciplines be in the year 2000?
4. Assemble information on the number of forestry journa,ls and articles published over
the last half century. Verbalize your interpretations of a 30 year extrapolation into the
future.
5. Describe and interpret the differences between inter-, multi-, and transdisciplinarity.
(See Jantsch 1972.)
6. Theoretically each scientific discipline could have its own set of "integrative principles."
What are candidates for forestry? Do they have any common qualities with those in
other disciplines? If so, abstract out the common quality and form interdisciplinary
"integrative principles."
120 Haskell's Coordinate System in Synthesis

7. Gather a group of your colleagues together in a brainstorming session (no rebuttals


allowed) and develop a list of ways in which your research and teaching subjects are
similar to those of other members of the group. If none are evident, go to a higher level
of abstraction.
8. Study Figure 7.2b and respond to the question: "Why is the phase plane a poor tool
for synthesis?"
9. Is quadrant I in Figure 7.5 sparsely populated because it is little explored or because it
is in nature nearly empty?
10. Examine the extensive literature cited by Dindal and identify the parties to each relation.
Are there relations that are exclusively between plants and plants, plants and animals,
or animals and animals?
8: "Interaction" in a
System of Concepts

Something that hangs by itself may be said to "dangle". A concept that hangs by itself
call a dangling concept. Just as proper grammar has no place for dangling participles,
advanced science has no place for dangling concepts. A more complete description of such
concepts, as well as an examination of the degree to which this idea applies to "interaction",
are the subjects of this chapter. In particular I examine some scientific concepts that are
candidates to combine with "interaction" to form concept systems. But my first task is to
elucidate the idea of a dangling concept.
The world of the scientist may be divided roughly into two areas: that of nature and the
immediately given, and that of ideas. Margenau (1950) has characterized the relationship
between these areas as shown before in Figure 5.1 and simplified in Figure 8.1. The vertical
line, labeled P, denotes the aggregate of all experience that is immediately given, the so-
called plane of perceptions. The circles to the right of P denote concepts, the abstractions
of sensory experience arrived at through "imaginative supplementation of the perceptually
given." The horizontal lines connecting circles to the plane of perception are called rules of
correspondence. I skip correspondence rules and simply assume that somehow a translation
has been made from the P plane to the concept, C, field. The question I wish to address
concerns the nature of the concepts themselves.
Recall from Chapter 1 that representing concepts may be expected to meet what Mar-
genau calls metaphysical requirements: 1) logical fertility, 2) multiple connectivity, 3) per-
manence and stability, 4) extensibility, 5) causality, and 6) simplicity and elegance. Here I
focus only on requirement 2), multiple connectivity.
The character of the connections that concepts may enter are, according to Margenau,
two: epistemic and formal. Epistemic connections are between the concept and the plane
of perceptions: between an observed difference in, say, production in a "together" and a
"separate" condition and the idea of an interaction type and intensity. Formal connections
set concepts in purely logical relation with one another either through postulates or empirical
relations of a confirmed character. A dangling concept is of the character that few, if any,
connections of the formal kind are possible. Thus, even though an epistemic connection may
exist, the concept has no fruitful connections with others - it just hangs there, dangling, in
the C-field. Since it will not connect with other concepts, it is not found in concept systems,
the building blocks of scientific theories.
In a sense I wish to pursue the idea, rooted in Chapter 7, that "interaction" has "reined
in" diversity too much, e.g., forced every interaction to be referenced to a single (0,0) circle.
Further, my synthetic enterprise will be more fruitful of new insights if I explicitly recognize
as much diversity as possible. One manner of doing this is by combining "interaction" with
other concepts to form concept systems, but a system where relations will dominate the
things of nature. Specifically, I look at how, what, and where other concepts combine with
"interaction."
122 "Interaction" in a System of Concepts

P-plane
(plane of perception)

Nature

C-field
(construct field)

Figure 8.1. The relation between nature and concepts (Margenau 1950). The concept field
shows two types of concepts: "A" are those that allow formation of concept systems through
logical relations with one another, and "B", are those that bear no logical relation with other
concepts. Those of type B are dangling concepts.

A. Valence of "Interaction"
In chemistry, valence is the term used to designate the concept that represents the
capacity of an element to combine with another. In the study of protein molecules, "active
sites" are the locations where enzyme configurations allow bonding with other enzymes.
These two concepts are useful in evaluating the ability of "interaction" to participate in
concept systems.
The number of active sites is the valence. The valence of "interaction" is four, because
there are four sites where concepts can bond. When "interaction" is rooted in Haskell's
coordinate system, the active sites are: (a) the process referred to by the signs, (b) the
referent of the (0,0) circle and the number of circles, (c) the referents of the axes, and (d)
referents of the partitioned space. In some cases, the simpler ones especially, the concepts
at each site are, or can be, independent of those at other sites. In the more complex and
int.eresting, cases, the bonding concepts may be interrelat.ed.
Until the concepts bonding at. each site are specified, Haskell's coordinate system has
a form analogous to an open sentence or propositional function (Copi 1978). The form is
there and fixed, but the variables (objects/processes/events) have not been specified. In the
following discussion I use the symbol I Hcs to designate the core concept "interaction" as
embodied in Haskell's coordinate system. The ' ... ' symbols link the core concept with the
active sites where bonding, designated " ___ .," takes place. I begin with the signs.

Signs (+, -,0)


The signs refer to a process change, or probability of process change, that has resulted
because of the "together" condition. The concept bonding at this site may be independent
"'Interaction" in a System of Concepts 123

of concepts bonding at the other active sites or be interrelated with one or more of them.
Population processes of interest include, among others, growth (e.g., mortality, birth, form)
and regulation (fluctuations, dispersal, oscillations) (Kormondy 1965). Expressing the results
of an interaction experiment in terms of, say, dispersal is basically the same as we have
discussed. Needed is a measure of dispersal in both a "together" and "separate" condition.
A phase plane comparison then provides the type and intensity of interaction, with respect
to dispersal, of the two populations.
The broad groupings of processes - growth and regulation - apply only to biological
populations. In order to expand the study of populations so that it has the same referent
range as the ecosystem concept, organism and abiotic environment, some additions to, or
fundamental rethinking in establishing the groups are imperative. For example, mortality,
as a process, is unlikely to apply to populations in the abiotic state.
Haskell (1970) interrelated the sign (process) concept and a level structure (hierarchy)
concept bonded at the (0,0) circle site as follows:
Each component can increase the other's capacity to participate in the emergence
or maintenance of higher members of the systems hierarchy [+], can decrease this
capacity [-I, or leave it unaffected [01·
Generality of the sign referent (growth, regulation, emergence or maintenance) is important
to wide use of the Haskell coordinate system because it must be uniform throughout a single
use.

(0,0) Circles

Signs (+, -,0)

Anyone (0,0) circle always represents an interaction with respect to a process, but the
numbers of circles can be expanded to accommodate variety in popuations undergoing the
process or, theoretically, the number of processes can be expanded while not recognizing thing
variety. Of course, (0,0) circles could be added to a coordinate system simply to improve
clarity in the mapping of a large number of population interactions. To do so would miss an
opportunity to not only improve clarity but also to recognize variety in the populations.
A concept that seems to bond naturally at the (0,0) site is level or hierarchy. Ecologists
appear to use the level concept, e.g., trophic level, where others use hierarchy. Much has
been written about both concepts (Whyte, Wilson and Wilson 1969; Weiss 1971; Pattee
1973). I employ the terms interchangeably although Bunge (1969a) argues that hierarchy
harbors a dominance relation that is absent from the concept of level.
I need to clarify two points before proceeding. These relate to levels and level structures.
A level is an assembly of things of a definite kind that have definite sets of properties' and
obey certain sets of laws. A level structure, L, is an ordered pair L = (S, E), where S is a
family of sets of individual systems, and E is a binary relation in the family of sets (Bunge
1969a). Thus, a collection of levels does not constitute a level structure unless there is a
binary relation that exists between the levels. Requirements on the binary relation E are
three: it must be (1) a one-many mapping, (2) reflexive, and (3) transitive in S. Requirement
three simply means that if relation E applies between levels a and c in S, i.e., E( a, c), and
also between c and e, i.e., E(c,e), then relation E also obtains between a and e, E(a,e).
Binary relation E is also supposed to represent emergence or coming into being of novelty
or qualitatively new systems in a process (Bunge 1969a).
Level structure is therefore a concept, itself of valence two, with active sites at Sand E.
At S the requirement is that every member of the family S be a natural class. This means
that a relatively homogeneous collection of things consititutes a level provided E does not
124 "Interaction" in a System of Concepts

hold among members of the collection. Concepts used in reducing heterogeneous piles to
relatively homogeneous groups bond at S. At active site E, bonds are formed with concepts
used in establishing relations between levels.
An example of relation E is the trophic level structure of ecosystems. It is probably
the most widely known basis for a level structure in forest ecology. S, the family of sets of
individual systems, shares the basic properties and laws common to all biological systems:
natality, growth, accretion, death, reproduction, anabolism, catabolism, etc. These apply
to both plant and animal populations. Individual levels are the referents of groups having
more similar properties and laws. A binary relation, E, that operates in S to produce a
level structure, instead of just a collection of levels, is that of "energy-availing" (Lindeman
1941). Since E must not operate within a level, yet must operate between levels, nearly
all autotrophs are placed in a single level. The abiotic component of ecosystems is not
considered in the trophic structure. While on the other hand heterotrophs - the consumers
and consumers of consumers - are partitioned into many separate levels. The specificity
of the "energy-availing" relation to the animal component of ecosystems reduces its general
utility to recognize and differentiate on variety. "Energy-availing" has a fairly narrow range
in the context of an entire ecosystem. A concept underlying a natural system's level structure
should be general enough to encompass the diversity in that natural system. Trophic level
structures slight the vegetation, ignore the abiotic system completely, and focus on the animal
component.
Establishing an ecosystem level structure, i.e., determining the respects in which the
members of S are equivalent and defining a relation E that exists between but not within
levels of S, is no easy task. Haskell (1970) has used a strategy that may simplify the task,
however. It is to interrelate the concepts bonding at the sign site and at Sand E in the
level structure.

Axis Sets I Hcs ........ (0,0) Circles

Signs (+, ~,O)

The Haskell coordinate system has two axis sets, horizontal and vertical. From the
explanation of the form of the coordinate system in Chapter 4, the relations the axes represent
((0,+),(~,0),(0,~),(+,0)) comes from the relational coordinate system. The axes names,
if any, come from the metric coordinate system via the phase plane. In the very simple cases
involving paramecia interactions, the axes were labeled just as one would label the axes in
a phase plane. The interactions are represented, then, with a simple sign, e.g., ".", at the
appropriate angle and distance for the type and intensity of interaction. I could continue
in this manner as long as the interactions are between populations of the same organism.
However, when I want to plot interactions between populations of other organisms, the axis
labels must be changed. An approach I have found useful is to locate symbol pairs depicting
the populations at the appropriate type and intensity of interaction (see Figure 7.2). Thus,
I put the ordered symbols tree, insect at the appropriate angle from (+,0) and distance in
or out from the (0,0) circle. The immediate question arises: Why put tree, insect at (--, +)
instead of insect, tree at (+, ~)? To resolve this we must discuss labeling in greater depth.
First, convention rules that the symbol on the left in tree, insect goes with the horizontal
axes and the symbol on the right with the vertical. To label the horizontal and vertical axes
to tha~ effect is redundant. It is convention. The fact that labeling is a convention causes no
problem: it creates an opportunity. The opportunity is to introduce additional descriptors
of the populations. Whereas in the analytic mode the axis labels were individual concepts,
the synthetic mode permits the referents of the axes to be broadened. I do this by using
"Interaction" in a System of Concepts 125

class and relation concepts. By itself this is not difficult. For example, given a single (0,0)
circle, the comparative relational concept "power" can be used to derive two categories, say,
"strong" for the vertical and "weak" for the horizontal (Haskell 1949). Or, two concepts
from cybernetics, "work component" and "controller component," can be used as axis labels
because they may be considered aspects of the concept of strategicity (Haskell 1970, 1972).
As long as there is a single (0,0) circle and the axis set concept is independent from
the sign concept, things remain straightforward. They become more interesting if a level
structure concept bonds at the (0,0) circles. An example of this is shown in Figure 8.2.
An extremely simple trophic level structure is given at the circles (autotroph-autotroph,
autotroph-heterotroph, and heterotroph-heterotroph), and reproduction is given as the pro-
cess (the referent of the signs). For this example I use the concept "adaptation" at the axis
site and express it in a comparative relational sense. Of course the referents of "adaptation"
must be based on prior information, the separate condition. To do otherwise would confound
the sign concept with the axis concept.

·better"
adapted
(0,+)

(-,+) (+,+)

(-,0) -+---+----+--- '*-----t.ifffifff~~f---__t- (+,0) ••••••


adapted

het.rotroph-
heterotroph
(+,-)

(0,-)

Figure 8.2. The Haskell coordinate system with three of its active sites occupied, (1) signs,
(2) (0,0) circles, and (3) axis sets. Because of the level structure, the referents of the axes
change in going from one circle to another. / / / / / labels the axis of increase for autotroph-
autotroph level elements less adapted to reproduce. := labels the axis of increase for the
elements better adapted to reproduce of the autotroph-heterotroph level elements. \ \ \ \ \
labels the axis of decrease for autotroph-autotroph level elements that are better adapted to
reproduce.
126 "Interaction" in a System of Concepts

Space 0° :S (J :S 360 0

o :S r :S 100

Axis Sets I Hcs ........ (0,0) Circles

Signs (+, -,0)


In the process of assigning concepts to the coordinate system's three active sites dis-
cussed so far, the fourth site, the two dimensional space, is somewhat constrained. It is
worth spending time on, though, because the names applied to the space may form conve-
nient "handles" and be widely used. If the name one scientist gives to, say, quantrant II
is "XYZ," while another scientist gives the name "WXL" to the same quadrant, the usual
reaction is that one must be incorrect, or at least not adhere to standard terminology. The
fact is, both names may be correct if the concepts bonded at the other sites are different.
The space I am talking about includes both dimensions, angular from the (+,0) axis for 360
degrees, as well as on, inside, and outside the (0,0) circles. Here is an example from the
literature:

Active Site Bonding Concept


signs nutrition
circles -none- (only one circle used)
axis sets power (expressed as weak and strong)
space the categories of relations given in
the cover figure of Main Currents
in Modern Thought (Haskell 1949).
Because of the convention that net negative interactions are plotted inside the (0,0)
circle and net positive outside, an idealized pattern of interactions forms a cardioid-like
figure in the coordinate system with the cusp in quadrant III (Haskell 1949, 1970, 1972).
Harold Cassidy calls the area between the (0,0) circle and the cardioid (135° > (J > 315°)
the cooperators surplus, and that area inside the (0,0) circle (135 0 < (J < 315°) he calls the
conflictor's deficit.
In his survey of the literature, Dindal (1975, Table 1) presents an extensive list of names
for types and intensities of interactions that can be translated into regions of space in the
coordinate system. Some of the names are included in his extension of the coaction cross
tablulation given in Figure 7.3 and my geometrization of it in Figure 7.5. Lidicker (1979)
modified the relational form of Haskell's coordinate system and introduced names for regions
of interaction space.

B. A Cumulative Level Structure


The foregoing concepts and procedures may be clarified by applying them to a specific
problem. In the process of this application I will introduce one additional concept, that of
cumulative level structure.
Most of the material from the example comes from a study of the activities and coac-
tions qf animals at sapsucker trees (Foster and Tate 1966). Their study, conducted in a
forested region of northern Lower Michigan, focused on the array of interactions between
various animals that result from yellow-bellied sapsuckers (Sphyrapicus varius) puncturing
the phloem of hardwood trees and releasing a flow of phloem sap in spring and early summer:
"Interaction" in a System of Concepts 127

The result is a complex situation wherein a species, through a specialized habit,


attracts to its feeding area a host of animals which vie with one another, establishing
a pronounced social order or hierarchy (Foster and Tate 1966).
I have added one element to the system, soil, and have limited the number of tree species to
sugar maple.
The system functions approximately as follows: In the spring sapsuckers select trees
near their nest trees and on them make their characteristic band of square noles around a
large branch or the trunk. The resulting sap flow is used directly by the sapsucker, and it also
attracts a large number of animal associates. Foster and Tate found insects of nine orders and
22 families, 20 species of birds, and five species of mammals associated with sapsucker feeding
trees. I have simplified the complex into the following elements: (a) soil system, (b) sugar
maple, (c) sapsucker, (d) fruit flies, (e) moths, (f) birds (hummingbirds, warblers, flycatchers,
Downy woodpeckers), and (g) mammals (bats, northern flying squirrels). In a simple arc--
node (graph) format the structure of the complex is given in Figure 8.3. The unstructured
graph in Figure 8.3 is missing explicit information about the nature of interactions and the
level structure (hierarchy) operative at sapsucker trees. I introduce the level structure first.
Clearly, three broad levels are involved: abiotic, autotroph, and heterotroph. These form
stable subassemblies as given in Table 8.1.

hummingbirds

warblers

fruit flies - - - - flycatchers

sugar maple
trees
/ downy woodpeckers

/
~moth,_bats
northern flying
yellow-bellied squirrels
sapsuckers

Figure 8.3. Graph of feeding complex at sapsucker trees in northern lower Michigan. The
soil element has been added. The connecting arc denotes "interact directly with".

'Fhe symbols [ and) are borrowed from mathematics. "[ .. X .. ]" denotes a closed, unreac-
tive level subassemble. "[lX]- y)" denotes a closed subassembly, [X], linked with a reactive
element, Y, capable of forming links with elements higher in the level structure. Thus, at
level 2, sugar maple is the element capable of forming links with elements at higher levels.
It does this at level 3, but sugar maple is again the element that links with higher level
elements. Of course, it is not the sugar maple at level 2 that links to fruit flies and moths
at level 4; it is sugar maple modified by sapsuckers. At level 5 there are still two reactive
elements~' birds and mammals. Level 5 shows a cumulative level structure characterization
of what Foster and Tate called "a pronounced social order or hierarchy."
Consider now the interactions between levels. In particular I am concerned with in-
teractions between two elements: the new element added at each level, one not present at
the next lower level, and the reactive element of the next lower level. Thus, level 2 has
sugar maple, not present at levell, and the reactive element at level 1 is "soil" Continuing,
level 3 has sapsuckers, not present at level 2, and the reactive element at level 2 is sugar
128 "Interaction" in a System of Concepts

Table 8.1 Cumulative level structure showing level element interactions at sapsucker trees
in northern Lower Michigan. It is a simplification of the findings of Foster and Tate (1966).
The structure has been extended to the abiotic levels. The column at the right gives a verbal
description of the system.
Level Level Element Structures Interpretation
[H+L ~ Soil) Energy (heat and light) flux on
bare soil.
2 [ [H+L ~ soil] ~ sugar maple) Sugar maple-dominated forest.

3 [r[H+L ~ soil] ~ sugar maple) Sugar maple forest with sap


I
sapsucker j flowing from sapsucker damaged
trees.
4 [[u H+L ~ soil] ~ sugar ~ fruit flies) Sugar maple forest with sap
flowing from sapsucker damaged
maple ~ moths) trees and with fruit flies and
moths consuming the sap
sap!ucker]

5 [ [[[ [H + L - soil] - sugar - fruit flies ]- birds) Sugar maple forest with sap
flowing from sapsucker damaged
trees and being consumed by
maple - moths] - mammals) fruit flies and moths that are

'~'"'k"l
preyed on by birds and mam-
mals.

maple; level 4 has fruit flies and moths as new elements, and level three's reactive element is
sapsucker-modified sugar maple trees, and so on. The process I use to characterize the signs
of relations is that of Haskell (1970):
Each component element can increase the other's capacity to participate in the
emergence or maintenance of higher members of the system's hierarchy [level struc-
ture]' can decrease this capacity or leave it unaffected.
The approximate relations between elements are shown in Table 8.2.
In Table 8.2 I have identified and ordered elements using concepts bonding at the Haskell
coordinate system's active sites as follows:

1. sIgns "contribution to the capacity of an element to participate


in the emergence or maintenance of higher members of the
systems hierarchy"
2. (0,0) circles cumulative level structure based on the binary relation of
"energy-availing"
3. axis sets new element at level L+ 1 (vertical axes) and element at level
L as modified by lower levels (horizontal axes).

Introduced here is the idea that the level structure is cumulative rather than separable as in
Figure 8.2. This again presents an opportunity to introduce new concepts in geometrically
assembling the sapsucker system. Each subassembly, corresponding to a given level plus all
the lower levels, is in itself stable. That is, independent existence in its own right can be
"Interaction" in a System of Concepts 129

Table 8.2. Cumulative level structure showing type of interaction between new element at
level L + 1 and the reactive element at level L. I\umbers near elements are used in Figures
8.4 and 8.5.
Level Elements and Relations
Between New Element at Level
Level Comment
L +-
1 and Reactive Element at L.
1
[H+L-(O,+ ) soil)
2
2 [lH+L soil] - (+ < +) - sugar maple) some deciduous forest species are
"soil builders"
3 [[ [H+L soilj sugar maple)
(- < +) Injury to population of sugar maple
3
decreases their capacity to. .less
sapsucker]
4 than it increases the sapsucker's ca-
fruit flies) pacity to ...
"fruit flies formed
(0,+)
4 soil] sugar "clouds around sapsucker holes.
"moths replaced butterflies at sap-
maple (0, +)
sucker trees at night ..
moths)
sapsucker] 5

6
birds)
"a few species consistently visited
(+ > -) the trees for insects ... "
5 [[[[H+L soil] sugar fruit flies 1 "moths were caught by northern fly-
ing squirrels and bats ... "
maple moths 1
(-, +)
sapsucker 1 mammals)
7

expected. To denote this character I inscribe each interaction in a triangle (Figure 8.4).
The "x" in the smallest triangle marks the qualitative location of the interaction between
sugar maple, the new element at level 2, and "soil" enhanced by heat and light, the reactive
element at level!. The result of this interaction produces, over time, a sugar maple forest.
The triangle, then, stands for a sugar maple forest with sugar maple trees as its reactive
element. This subassembly supplies a base of support for the next level subassembly which,
in turn, supplies support for the next higher level subassembly, and so on.
The smallest triangle in the lower left of Figure 8.4 may be thought as generating a kind of
"force field" into which element 3 is placed as a form of "test body." The interaction over
time of "field" and "test body" produce a new "field" manifested at the apex of the second
largest triangle in the lower left of Figure 8.4. Of course, the investigation of Foster and Tate
considered only those "test bodies" that had been successfully integrated with a "field" to
produce a new "field." All those test bodies that failed because of negative interactions with
130 "Interaction" in a System of Concepts

Level 2

Figure 8.4. A cumulative level structure as a set of triangular, stable subassemblies pro-
ceeding upward. The apex of each triangle represents a force field produced by the levels
and interactions it encompasses. The new elements at each level, given numbers, represent
"test bodies." Levell is omitted. At level 4 the structure splits with both elements 4 and 5
serving as "test bodies." Each helps generate a special "field" supportive of elements 6 and
7, respectively, at level 5.

the "field" had interactions below the horizontal axes and are no longer present. Occasionally
man will introduce an organism (read "test body") into a forest community (read "field"),
such as the eastern timber wolf into northern Michigan, where the "field" is unsuited for the
"test body." The result is the disappearence of the organisms (Wise, et a!., 1975). Other
"test bodies" may be introduced naturally through mutation and other processes.
Figure 8.4 characterizes the sapsucker-tree interactions in terms of three concepts
bonded to the Haskell coordinate system. As shown, the concept's order of dominance
IS
dominant -- interactions
su bdominant - level structure
subordinate - reactive element

The figure reveals prominently the cumulative level structure and the elements at each level
that are new and those that are considered reactive. It is analogous to Figure 7.2 in that there
"Interaction" in a System of Concepts 131

are Haskell coordinate systems at each level just as they are between individual elements
in the separable strucutre in Figure 7.2. As it was possible to assemble the elements in
Figure 7.2 into a single Haskell coordinate system, it is possible to assemble the hierarchy of
triangles in Figure 8.4 into a single coordinate system (Figure 8.5). In so doing I again have
subordinated structure and elements (things) to signs (relations). The pattern of "x's" is,
of course, only qualitative, but it shows a probable clustering around 0 = 90 0 for all levels.
More precise estimates of the interactions are certainly possible with additional field effort.

(0,+)

"test body·
x 7

(-,O)----~--~--~--~--~--

(0,-)

Figure 8.5. Geometric assembly of interactions at sapsucker-trees in northern Lower Michi-


gan. The innermost circle characterizes soil [I! and sugar maple [21 interaction. The result
of this interaction, a sugar maple forest, goes with the horizontal axis at the second circle.
Its active element, 2, interacts with 3 (sapsuckers) to produce the [2/1,31 assembly that is
associated with the horizontal axis at the third circle. One view of this figure is that it
represents "fields" within "fields."

One view of the structure of plant-animal communities is that they are characterizable
by one or more so-called nuclear interactions, interactions that are sufficient to set in train
another series of interactions. The sugar maple-sapsucker interaction is the nuclear inter-
action in this case. The possibility is thus raised of mapping each nuclear interaction into
132 "Interaction" in a System of Concepts

the same Haskell coordinate system. The result would be a geometric map showing both
the level structure and process-based function of the plant-animal community. One consid-
eration in favor of such an approach is that the map itself serves as an effective means of
holistic communication (Rhyne 1972). The cumulative level structure approach developed
here has an advantage when dealing with complex yet structured communities. It provides a
means for simplification, for reducing the number of interactions one has to deal with simul-
taneously. This simplification comes with the cost of assuming that the interactions being
considered, e.g., between birds and mammals and their respective prey, are not significantly
influencing interactions lower in the structure.

C. "Interaction" and Concept Maps


Throughout the previous chapters, and previous parts of this chapter, the assumption
or actual act itself, has been to map events into Haskell's coordinate system based on com-
parisons, actual or inferred, between population trajectories under spontaneous and induced
change. In this section I take a brief aside to show how the coordinate system can be used
in the preparation of what are called concept maps.
Stewart, et a!., (1979) describe a concept map as "a device for representing the concep-
tual structure of a discipline, or segment of a discipline, in two dimensions." The entries in
the two-dimensional space are the words that designate the concepts. Stewart and associates
observe that the linear, one-dimensional outline is the traditional way of representing infor-
mation about a subject. The two-dimensional framework has at least two advantages over
a linear array: 1) it allows some differentiation between concepts, e.g., on their generality
or their referents, and 2) it allows connecting arcs between concepts to show propositional
relationships between the concepts (Stewart, et a!., 1979).
Concept maps have been used primarily for instructional purposes. Their utility is, in
part, premised on notions of learning theory developed by Ausubel (1968), and Ausebel, et
a!., (1978). Novak (1966, 1977, 1979, 1980) and Stewart, et a!., (1979) have reported on the
use of concept maps in science instruction (including curriculum development, instructional
use, and evaluation use) at the primary and secondary school levels. It appears that these
notions have as much applicability for university instruction in forestry.
To illustrate the idea of a concept map for an aspect of forest biology (specifically, the
stand tending aspects of silviculture), consider the concepts arrayed in Figure 8.6. Of course,
these concepts are only a small fraction of all the silviculture concepts. The solid lines in
the figure connect concepts included in the propostion: "trees in class 1 achieved a higher
increment in relation to basal area [density I than the rest of the stand." Both the concepts
and the proposition are abstracted from Assmann (1970).
This concept map can itself be mapped into Haskell's coordinate system. Its range
is quadrant III, referenced to the plant-plant (0,0) circle (Figure 8.7). Further, simplified
concept maps for portions of forest pathology and forest entomology map into the coordinate
system in quadrant IV (plant-plant (0,0) circle) and quadrant II (plant-animal (0,0) circle)
(Figure 8.7). The propositions indicated by connecting arcs are given in Table 8.3. These
may be called disciplinary propositions. Interdisciplinary propositions are comprised from
concepts in two or more of the quadrants. The primary use of a concept map such as in
Figure 8.7 is for representing the conceptual and propositional structure of a discipline or
interdiscipline, something of use in both education and research.

D. Summary
When the interaction concept is represented by the uninterpreted form of Haskell's
coordinate system, "interaction" appears to be a concept with a valence of four. Four
other concepts 'bond' with "interaction," one each at the process, axis, (0,0) circle, and
concept space 'active sites'. An analysis of coactions at sapsucker trees shows that the
"Interaction" in a System of Concepts 133

dominated tree class

heavy thinning

production efficiency
light
growing space thinning
thinning
social advantage low thinning

relative height tree class

growth capacity
\
social structure

increment

stand dominance) dominating


tree
\ social position
class
"--- density

Figure 8.6. Partial concept map for that portion of silviculture dealing with stand develop-
ment and tending via thinnings. Connected concepts are used in the proposition given in
the text.

(0,+)

(+,+)

tree populatlon/

(-,O)t------'-"'-,.---------- ----------+-(+,0)
domfnal9{l tree class
\ heavy thinning
produCtion eHlClency - - - " ",

growing space
\hml1ll'lg '.112;,1
lhlnnlr\g \
epidemic
low thinning I c:

(0,0) relative helg/'11 tree class- .. i :


!?
fungus
do Co facullatlve
SOCial stnJcture \ gfowlh capacity i \llfl,llerx:e "--" obligate
plant - animal saprophyte

~;~~T d~:::,~
canker causal
hosl \ vector agent
............. dlsease - a pathogen

-,~ (+,-)
suscep!

(-,-) plant - plant


(0,-)

Figure 8.7. Mapping of partial concept maps for silviculture, forest pathology, and forest
entomology into Haskell's coordinate system. The latter has three reference zero circles (only
two are shown). Concepts are abstracted from Assmann (1970), Boyce (1948)' Hepting
(1971,1974), and Graham (1963). Lines connecting concepts denote factual propositions
abstracted from these literature sources.
134 "Interaction" in a System of Concepts

Table 8.3. Disciplinary proposItIOns abstracted from the literature on silviculture, forest
pathology, and forest entomology. Note that only factual concepts are identified in the map
(Figure 8.9). Formal concepts such as "greater", "higher", etc., Formal concepts such as
"greater", "higher", etc., are generally applicable and not specific to a discipline.

Quadrant Proposition
II a " ... the greater the di versification of tree species the less fre-
quent will be insect outbreaks" (Graham 1963).
b " ... [tree] [d]ensity has an important influence upon insects
that require herbs and shrubs as alternate hosts" (Graham
1963).
III a " ... tree class 1 achieved a higher increment in relation to basal
area than the rest of the stand" (Assmann 1970).
b " ... the difference in growth-capacities between the two pairs
of classes under very light thinning is greater than that under
very heavy thinning" (Assmann 1970).
c " ... heavy thinning increases the efficiency of production of the
lower social classes bringing it on a level of equality with that
of the dominating classes" (Assmann 1970).
IV a "The chestnut blight, a canker disease, is the only plant
pathogen to have virtually eliminated its host" (Hepting 1971).
b " ... there is no evidence that the virulence of the fungus has
decreased" (Hepting 1974).
c " ... the fungus grows as a saprophyte on many forest species in
the genera. "(Hepting 1974).
------------_.

coordinate system has the flexibility to represent a cumulative level structure as well as
non-cumulative ones. Although most usess of the coordinate system here are to represent
events/things/systems, concepts can also be mapped into the coordinate system and thereby
exhibit partially the conceptual and propositional structure of a discipline or interdiscipline.
"Interaction" in a System of Concepts 135

F. Research Problems
l. Study the references in Dindal's paper (Dindal 1975) and identify the concepts bonded
at the sign (process) site and at the axis site.
2. Improve the concept maps for silviculture, forest entomology, and forest pathology in
Figure 8.7. Identify major propositions that are believed true in each discipline complete
with evidence to support the belief. From the literature identify propositions that use
concepts from all three of these sub-disciplines of forest biology. Evaluate the idea that
propositions (not concepts) are interdisciplinary.
3. Analyze for valence the ecosystem level concepts "stability," "sensitivity," "resilience,"
"persistence," and "complexity," and identify the active sites on each. (See Levin 1974.)
4. Simon (1969) first introduced the notion of stable subassemblies when analyzing the
evolution of complex systems. Find evidence for and against the idea that a sugar
maple stand is a stable subassembly.
5. Overton (1974) poses the question: "Decomposability-A unifying concept?". What
is your personal assessment of "decomposability" as a unifying concept? What is its
valence? Should "unifying" concepts have a higher valence than non-unifying ones?
6. Cartwright (1951) cautions that in order to develop a satisfactory system of concepts,
scientists have to be particularly careful about the way in which they develop their
individual concepts. Search the forestry literature for claims to the development of new
concepts. Are the concepts developed shown to be members of concept systems?
7. From forest and range ecology literature identify candidate nuclear interactions, and
the interactors, that set in train a set of interactions that predominate in each of the
quadrants or on each of the axes of Haskell's coordinate system. For example, the
sapsucker-induced interactions clustered around () = 90 0 . Find systems that cluster in
quadrants III, I, etc.
8. One view of the most strategic elements in a forest community is that they are the
populations comprising a cut point of the graph of the system (Harary, et aI., 1965,
Chapter 7). Do the directed graphs of forest communities possess cut points? If they
do not, does this mean that no forest community populations are more strategic than
others?
9. Prepare a cumulative level structure graph of the cocoa-mistletoe system described in
Room (1975). Approximate the interaction types qualitatively and prepare a cumulative
hierarchy of triangles and circles showing the pattern of interactions. Alternate problem:
Execute the above on the complex described by Korford (1958).
10. Invent a formula that relates comparative unifying power (up) of a concept to the
concept's valence. For example, is it closer to up =.c QV or to up = v", where v designates
the concept's valence, and Q denotes a numerical constant.
Part IV
Natural Selection of
Community Interaction Structures
In chapters 9 and 10, I introduce and attempt to rationalize a theory of natural selec-
tion of community interaction structures. "Structure" means different things to persons from
different disciplines, so, I briefly examine the commonly accepted meanings to foresters, ecol-
ogists, systems analysts, and structuralists. I refer to 'structure' as the system of interactions
between things; a reference most commonly used by systems analysts and structuralists. An
important element in the theory is the notion of indirect effects. Both chapters make use
of a new operator (in the mathematical sense of "operator"), which I call the "followed-by"
operator. Using the new operator, a system of populations comprising a model animal-plant
community, and some formal reasoning from abstract mathematics, I develop in Chapter 9
a theory of indirect effects and deduce three theorems that have importance for the natural
selection of community interaction structures. In keeping with my objective of formalizing
whenever possible, the theory is axiomatized. Chapter 9 is limited to the special case of a
community with just two populations.
In Chapter 10, I develop a method to analyze the interaction structure of a community
with many populations. The method requires focusing on each population temporarily and
computing the direct and indirect effects it receives from other populations. Also computed
are the effects, both direct and indirect, that the subject population has on the other com-
munity populations. Using a theorem and corollary developed in Chapter 9, I argue that
a certain combination of direct and indirect effects received (effects on) and delivered (ef-
fects by) the subject population leads to an exclusion of the relations. Because the relations
are excluded, the population common to the relations is eliminated. Thes new exclusion
principle is the primary mechanism by which community interaction structure selection oc-
curs. Further, it is the mechanism which warrants a move from the prediction column in
Figure 2.1 toward the explanation column. It is conjectured that the exclusion principle has
wider applicability than just to plant-animal communities.
9: A Theory of Indirect Effects and Implications For
The Natural Selection of Community Structures

The theory of natural selection plays a key role in the theory of evolution (Lewis 1980).
Traditional evolution theory has dealt with the evolution of traits of individuals in some
general relation to their environment. Recently, evolution theory has been extended to deal
with change in the type of interaction between individuals of different species populations.
That is, it has been extended to the structure cf a community, not just to its compo-
sition. Axelrod and Hamilton (1981) call their extension reciprocity theory. Most current
interest seems devoted to developing an explanation for the evolution of cooperation. In the
context of all possible interaction types, the study of the evolution of community interaction
structures adds a significant new dimension to evolution theory. What is to be explained is
not a property (trait) of individuals, but a relation between individuals of different species
populations. My objective here is to present a formal theory of indirect effects and then to
interpret theorems deduced from the theory in terms of the natural selection of community
interaction structures.
Only recently has attention been directed at the role of indirect effects in the important
question of community structure and function (Vandermeer 1980; Levine 1976, 1977; Lawlor
1979). The recent interest has had a quantitative mathematical character. My perspective
is also mathematical, but employs methods of qualitative mathematics. By remaining at the
qualitative level, I sacrifice precision of prediction, but I am able to develop an axiomatized
theory of the natural selection of community structure without getting deeply involved in
mathematical analysis methods, and without forcing on the community populations pre-
suppositions about their manner of change, e.g., logistic growth, constant coefficients, etc.
Certainly my theorems may be too simple to be totally true, but useful, I maintain, they
are.
Attention at this point is on binary systems, on communities formed of just two popula-
tions in direct interaction. Further, the treatment will be confined to biotic system compo-
nents and their biotic mediaries. Abiotic me diaries will be ignored. I deal here with indirect
effects, not with indirect interactions as such. Only when the indirect effects analysis is
completed can the analyses be combined into statements about indirect interactions and the
natural selection of community structure. To begin, I examine the concepts "community"
and "structure" followed by a look at the general concept of natural selection.

A. Community, Structure, Natural Selection

1. Community.
A frequent view of "community" is that of a plant-animal population complex that is
dominated in a specified region of physical space by a particular population. The community
name typically comes from the dominant population's name. The symbolic form of this
version of the predicate "community" is, using materials from Chapter 1:
138 A Theory of Indirect Effects

where

D designates the set of dominant plant or animal populations,


P designates the set of regions of physical space (places),
D x P designates the Cartesian product of dominant plant or animal popula-
tions and places, and
S designates the set of statements.

view the concept of community with a different focus. My view shares with that above
the notion of place or limited region of physical space. But, it differs in that the boundaries
to a community are set by fairly distinct breaks in the web of direct interactions between
populations in a larger web of interacting populations. Of course, there are probably never
complete breaks in the web, except perhaps for some islands, but there are significant changes
in the kinds and intensities of interactions between populations. The community boundary
occurs, then, at the significant break. This view of a community is closely akin to the notion
of a system and a widely used approach to system boundary specification. The symbolic
form for this notion of community is

where

Ps designates the set of plant and animal populations undergoing spontaneous


change,
P, designates the set of plant and animal populations undergoing induced
change,
P designates the set of regions of physical space (places), and
S designates the set of statements that contain the term designating the pred-
icate C z .

My concept of community appears to be close to that advocated by MacMahon, et aI., (1978,


1981), and Wilson (1980).
The communities I examine in this chapter are the smallest permissible, a special case
of only three interacting populations. In Chapter 10, I use the methods developed here and
treat communities made up of many directly interacting populations.

2. Structure.
The term 'structure' is both vague and ambiguous. Ambiguity is reduced somewhat by
our reference to forest community structure. Because of the particular sense in which the
term is Ilsed here, I highlight the various senses of 'structure' to foresters, ecologists, systems
theorists, and structuralists. Progressing through the list in the order given shows that the
term becomes more and more abstract. To set the stage for this, a brief look at the general
notion of structure is in order.
"Structure is normally thought to refer to order in space; function to order in time
or space-time. They are generally considered together. Structure refers to the constituent
parts, function to the activities of these parts" (Bakuzis 1974). Thus, structure as a general
concept is in a class with "pattern," "form," and "order."
Foresters, as in Spurr and Barnes (1980), typically invoke the term 'structure' when re-
ferring to the vertical distribution of trees in a stand. Thus, 'structure' designates "structure"
that represents a property of a forest stand, along with such properties as stand composition
A Theory of Indirect Effects 139

and density. The method of de Liocourt (the q method) refers to a concept widely used
to characterize the frequency of trees over a size dimension, e.g., diameter at breast height,
when the trees are of many sizes (Meyer 1953). Suckachev and Dylis (1964) use "structure
of phytocoenoses" in much the same sense as 'form' in order to describe layering of natural
communities.
Ecologists have taken a somewhat broader interpretation of the concept of community
structure. Kormondy (1969) describes Danserau's scheme of forest community structure as
having six features: 1) life form (trees, shrubs, herbs, etc.), 2) size (tall, medium, low),
3) function (deciduous, evergreen, etc.), 4) leaf shape and size (needle, broad, compound,
etc.), 5) leaf texture (filmy, membranous, etc.), and 6) coverage (barren, discontinuous,
continuous, etc.). Pianka (1974) also examines the structure of communities from a number
of perspectives, e.g., food webs and trophic levels, the interaction matrix for a community
nearing equilibrium (the community matrix (Levins 1968)), pyramids of energy, numbers
and biomass (Elton 1927), and species diversity. The ecologist's interpretation of structure
obviously makes use of many more general concepts than does the forester's interpretation.
Hence, it deviates considerably from merely equating structure with form in space. Also, see
Margalef (1963).
Systems theorists have an important role for structure:
"Whatever its kingdom, conceptual or concrete, a system may be said to have a
definite composition, a definite environment, and a definite structure. The com-
position [C] of a system is the set of its components; the environment [E], the set
of items to which it [the system] is connected; and the structure [S], the relations
among its components as well as among these and the environment" (Bunge 1979).
The symbolism used for specifying a system is, again

51 = {C,E,S},
where
C designates composition,
E designates environment,
S designates structure.
For the example that follows, forest community populations constitute the composition,
the environment consists of loosely related populations of plants and/or animals plus the
physical environment, and the structure consists of the enhancing (+), detrimental (-), and
neutral (0) effects that populations have on a specific process taking place in each community
population.
The last group whose view of structure I wish to examine is the structuralists, or more
specifically, relational structuralists. Piaget (1970a) puts into perspective the notion of
structure as used in the many different disciplines (mostly from the social sciences and
humanities) that employ the structuralist method. To Piaget "the notion of structure is
comprised of three key ideas: the idea of wholeness, the idea of transformation, and the idea
of self-regulation." A variant of structuralism called relational structuralism is described as:
positing systems of interactions ... as the primary reality and hence subordinating
elements from the outset to the relations surrounding them, and reciprocally, con-
ceiving the whole as the product of the composition of these formative interactions
(Piaget 1970b).
The view which resulted from our geometrization of the interaction web in Chapter 6 IS
similar to the relational structuralist's view: the populations were subordinated to the in-
teractions surrounding the coordinate system.
Both system theorists and structuralists seem to agree on the notion of what constitutes
a system (5), but if required to write the parts in order of importance, structuralists might
order them
140 A Theory of Indirect Effects

s= {structure, composition, environment} .


For 'structure' I read 'interaction' In this respect, this book presents a relational structural-
ist view of forest ecology and managment.
Bunge (1977) argues against the structuralist position by observing that except in pure
mathematics there are no structures in themselves, only structures of things (in the composi-
tion of some system) In other words, Eltonian structure may be a property of a community,
but one would not say that a particular community is an Eltonian structure. On the other
hand, it seems reasonable to view the problem as mathematicians might, any relation or op-
eration on a set of objects constitutes "putting structure" on that set (Lane 1970). Further,
it seems worthwhile to examine what the structure is. I shall not go to the extreme and say
that a forest community is a structure because I agree with Bunge that change is rooted in
things (biotic populations in a forest community in this case). But being able to say that
the indirect interactive structure of forest communities is such and such is in keeping with
my goal of occasionally removing things from their dominant position in forest science and
allowing relations to be dominant for awhile.
Recapitulating:
1. Foresters have traditionally viewed forest structure as simply distribution of tree stems
in three dimensional physical space.
2. Descriptive ecologists have used structure in a broader sense than just distribution in
physical space.
3. Systems theorists, including mathematical ecologists, use 'structure' to indicate various
kinds of relations between system components.
4. Relational structuralists suggest that structure (relations) is the primary, of course not
the total, reality, making things, populations, incidental to relations.
I use both the third and fourth notions of structure. Because of the critical importance
of interactions to the structure concept, I shall use interchangably community interaction
structure or just community structure.

3. Natural selection.
Natural selection is recognized as one of the major processes at work in organic evolution.
Haldane (1958) spells out natural selection as follows:
Natural selection may be defined as the elimination from a genetically mixed pop-
ulation of the less fit genotype. The fitness of a genotype is measured by the mean
number of progeny left by its members, subject to certain conventions. Clearly if
natural selection is not counteracted by some other process it will lead to change in
the population. And since the difference between a number of related species has
been at least in part analyzed in terms of genes, this change may be equated with
evolutionary change (Haldane 1958).
I make use of the gist of these ideas. Let me paraphrase his first sentence with place-
holders at key positions:
-A- selection may be defined as the elimination,
from a -B- mixed population, of the less -C-.
The process of natural selection can operate at several levels. For example, Wilson
develops a theory of the natural selection of populations and communities. This covers
placeholders A and B. But what about C in Wilson's case'? What are the qualities of B that
contribute significantly to its elimination? He emphasizes as follows:
Individual selection models predict that to be favored by selection, an organism
must have the highest fitness, relative to others in its trait group. In structured
demes, natural selection becomes increasingly sensitive to the differential productiv-
ity of trait groups as the amount of genetic variation among trait groups increases
(Wilson 1980).
A Theory of Indirect Effects 141

My application of the concept of selection will involve indirect effects and interactions in
place holders A and B and a relation between changes in type of interaction in placeholder
C.

B. The Set; The Operator; The Structure


Borrowing the mathematician's term 'set' should not distract us from remembering that
I am dealing with forest population systems. The composition of a system is typically viewed
as evolving because of the structure and environment of the components. But now I want to
address the question of structure evolution. To do so I proceed to a a higher level of system
concept. The level I leave is

populations direct effects }


strongly (+,0,-) of
weakly interacting
51 = { interacting populations on a
with populations
populations process in other
strongly interacting
populations

The level I proceed to can be represented

populations strongly populations weakly


the "followed-by" }
5 = { interacting directly interacting with
operator
2 in ways formed from populations strongly ,
(symbolized "0")
(+,0, -) effects interacting

or 5 2 ={C 2,E 2,S2}'


The operator "0" has the form given in Table 9.1.

Table 9.1. Specification of the rules of the "followed-by" operator. The signs in the body
of the table give the net result of an effect of population i on population j and the effect of
population j on population k.
the effect of population j on population k
o +
-------------- --------~

net effect of population I on population k


effect of + o +
population i
on population j
o o o o
+ o

The indirect interactive structure of forest communities is determined by examining the


properties of '0' using methods from abstract algebra, specifically, the theory of groups
(Kramer 1970). Group theory relates a set and an operation, in our case (C 2 ,0), with
criteria established by mathematicians for labeling the two as possessing a particular type
of structure. A label 'group' requires uniqueness, closure, associativity, identity, and inverse.
Other mathematical structures result from making fewer or more requirements. For example,
abelian groups are produced by also requiring commutativity. On the other hand, one may
relax the requirements and have mathematical structures called monoid, semigroup, and
groupoid.
142 A Theory of Indirect Effects

I examine the system (C z , 0) formed by, for now, arbitrary forest community populations
and their direct effects (+,0, -) (C z ) and the operator "0" (S2). I look at the requirements
in the order given above with a framework that calls for a definition of each requirement and
a test of its being met by (Cz,o):

1. Uniqueness.
Definition:The operator "0" returns, for any given input element of C 2 , a single element.
Test: Since the result of each application of "0" returns a single element of C z , the
operator "0" has the uniqueness property.
2. Closure.
Definition: The operator "0" applied to the set of elements, C z , "returns" an element of
the set C 2 .
Test: Clearly, the requirement of closure is met because the set of signs in the body
of Table 9.1 defining "0" does not contain signs absent from the margins.
Because closure and ulllqueness properties hold, the system (C 2 ,0) has at least groupoid
structure.
3. Associativity.
Definition: If a, b, c are elements in the set C z , then the following equation holds: (aob) oc =
ao(boc).
Test: Possession of the associativity property may be easily checked. If any a, b or c E
C 2 is the neutral element (effect '0'), it ensures that the order of composition
is unimportant. The cases where a, band c are either + or - effects need be
checked. It suffices to simply enumerate them here.
(+0-)0- +0(-0-)
(+0-)0+ +0(-0+)
(+0+)0+ +0( ...... 0+)
(+0+)0- +0(+0-)
(-0+)0+ -0(+0+)
(-0+)0- -0(+0-)
(-0-)0+ -0(-0+)
(-0-)0- -0(-0-)
In each case the equivalence relation holds, hence associativity is a property of
the operator.
Because the system (C 2, 0) has uniqueness, closure, and associativity, it has at least semi-
group structure. In order to have monoid structure, the system must also have an identity.
And to have full group structure, it must have the identity property and a unique inverse
for each element of C z .

4. Identity.
Definition: The set of elements contains an element that, when "combined" with any other
element, "returns" the other element.
Test: The set C z contains the identity element "+". It is easily spotted because "+"
is the only element that returns each element composed with it.
The system (C z , 0) therefore has at least monoid structure.

5. Inverse.
Defihition: For any element in the set C 2 , call it a, there is a unique element, call it a-I, also
in the set C 2 , such that when combined according to "0" returns the identity
element,!.
A Theory of Indirect Effects 143

Test: Existence of an inverse presupposes existence of an identity because inverse is


defined using identity. The following inverses are found when examining each
element:
+ is the inverse of +, because + 0 + = +, and
- is the inverse of -, because - 0 - = +.
The neutral effect (0) has no inverse because there exists no a-I such that
a-1oO=+.

Because of the absence of an inverse for the neutral effect (0) in the set C 2 , the system (C 2 , 0)
fails to meet all requirements for mathematical group status. It does, as noted above, fully
meet all requirements for monoid structure. Thus, the indirect interactive structure of a
forest community is a monoid. On this basis, then, structure is based on the propagation of
interaction type through a forest community.
The grist on which the operator operates is three populations in a community, i.e.,
p. ----> p) ----t P k . In Chapter 10, I will treat the case where i i' k, but for now I confine
attention to the special case where i = k.

c. A Theory of Indirect Effects

I assume the existence of two items; call them primitives:


a) a set of biological populations, S = PI, ... , Pi, p), ... , P n , and
b) a set of direct effects operative between the populations of the set S wherein one pop-
ulation may affect another population of S in one of three ways (+,0,-). The sign
refers to a particular process throughout the system, taken here to be survival of the
population.
The axioms of the theory are the nine rules of the operator "0":

Ai (P, ----> + P)IIP) P k )=* p. >---> Pk


A2. (P. ----> + p)IIP) ----> 0 Pk )=* P, >---> 0 P k

A3. (P, ----; + p)IIP] ----t + P k )=* p. >---> + Pk


A4 (P, 0 p] II p) Pk ) =* P, 0 Pk

.
----t ----t - ----t

A5. (P . ---> 0p] "'P] ----t °Pk)~


----.'
P.. ----t °Pk
A6. (P, ----t - p) II p] ----t - Pk ) =* p. ----t + Pk
A7 (P. ----t - p) liP] ----t + P k )=* P, ----t - Pk
- 0 ) 0
A8. ( P, ----t p]IIP] ----t Pk =* Pi >---> Pk
0 + 0
A9. (Pi ----t PJ II p] ----t Pk ) =* p. ----t Pk

Clarifying definitions are as follows:


144 A Theory of Indirect Effects

def
P, - - - (T, +) - - - PJ ,
def with T a place holder for the unspecified effect
of jon i.
def
"and" (logical conjunction)
def
v "or" (logical disjunction)

Df4 : (x) =} (y) def if x the y (logical implication)

def
(::lPJ)(P, -+ PJ 1\ PJ -+ Pd
(There exists a population PJ such that popu-
lation P, affects it, and it in turn affects pop-
ulation Pk .)
Three theorems may be derived from the axiom set:
Theorelll 1:
From A3, setting k = i, and from A6, setting k = i,

Proof: Let

B=P,f--++P,

Argument:
A =} B(from axiom 3)
C =} B(from axiom 6)

... (A V C) =} B

Althought clearly true, let me give a brief proof. The proof rests on being able to show that
the conjoined premises do in fact imply what the theorem states, that is,

if (A =} B) 1\ (C= B) then (A V C) =} B.
Working the antecedent,

(~AV B) 1\( ~C V B) (by the definition of material implication)


(~AI\ ~C) VB (by axiom of distribution)
~ ( A V C) VB (by OeM organ 's axioms)
( Av C) =} B (by definition of material implication)

or if (( A V C) = } B) then (( A V C) = } B),
or if p then p,
which is ·a tautology. (See any basic logic text for further information, e.g., Rescher 1964.)
Theorelll 2:
From axiom 7 and axiom 1, setting k = i,
A Theory of Indirect Effects 145

(P, --> - PJ /\ PJ ----> + P,) v (P, ----> + PJ /\ PJ ----> - P,) ==* P, ----> - P,

Proof: See proof of Theorem l.


Theorem 3:
From axioms 2, 9, 4, 8, and 5, setting k -= i,

(Pi ----> 0 PJ /\ PJ P,

Proof: See proof of Theorem 1.

D. Discussion of Theorems
Before I get into a detailed discussion of each theorem, I will give an overview of the
form each discussion will take, as well as my underlying assumptions.
Although many factors affect a population's survival in a community, I shall make the
simple assumption that survival is more likely if a population indirectly enhances itself.
Thus, I assert that best interest for a population is survival, and survival is more probable
if a population is a self-enhancer. Best interest is assigned, no consciousness is attributed
to populations involved. In the discussion of each theorem I ask this question: What is
the actual indirect effect of a population on itself as given by the followed-by operator? If
the population is an indirect self-enhancor, good! Should natural selection work to increase
the intensity of the indirect enhancing effect, things would be even better. On the other
hand, should natural selection work to reduce the intensity of indirect self-enchancement,
the population's situation is not so favorable. If a population in an interaction is not self-
enhancing, such as (-, +) and (+, -), then I ask, does natural selection work to reduce the
intensity of the indirect detrimental effect? If yes, natural selection helps to offset the indirect
self-annihilating tendencies. If no, natural selection exacerbates an already bad situation.
The above steps are taken in an examination of each party to each of the eight non-null
interaction types.

Theorem 1.
I call Theorem 1 the self-help theorem. It states that any population indirectly affecting
itself in either a - 0 - or + 0 + manner is indirectly "helping" itself. The + 0 + case is clear:
a population enhancing an enhancer is enhancing itself.
In the - · 0 - case it takes only a moment to see that a population inhibiting an inhibiter
is a self-enhancer. This logic says that for a population, call it the subject population, A, to
continue to be a self-enhancer it should influence the "other" population, B, in such a way
that population B continues to exist. And, of course, the theory of natural selection states
that the less "fit" individuals will be eliminated from a population because of a detrimental
effect, so the more fit will remain. Population B will continue to exist, and if it does,
population A will continue to be a self-enhancer based on its indirect effect through one
intermediate population.
Since the relation is symmetric, that is it holds whether we focus on population A or
B, one might suspect that the parent (-, -) interaction would be a very stable relation. It
may not be a sta~le relation because the fitness of surviving individuals of both population
A and population B is increasing. But, the rate of fitness increase may be very different. If
one population of individuals should increase in fitness at a more rapid rate than another
population of individuals for a long enough time, the parent (-, -) interaction may become
146 A Theory of Indirect Effects

quite unstable. There are two important classes of conditions that affect the differential rate
of fitness change in a (-, -) system: time and space.
Temporal relationships that affect differential fitness change are well recognized (Odum
1971). Two dimensions of the temporal relation are of particular importance and will be
briefly discussed: length of time the two populations have affected one another, and com-
parative generation lengths of the two populations.
The first dimension, length of time populations have affected one another, can be viewed
as having two categories of possibilities (of course, a continuum is more realistic):
a) The subject and "other" population are "new" associates in the community. Time of as-
sociation may not have been sufficient for reciprocal selection to have taken place. Thus,
there will have been insufficient time for niche shifts, resource partitioning (Schoener
1974), and character displacements to take place, all of which tend to reduce the inten-
sity of detrimental effects, hence the likelihood of one population being excluded from
the community. Recall, reduced effect intensity heads the interaction trajectory toward
the (0,0) circle. Depending on the comparative generation lengths interactions of newly
associated populations may result in a quick exclusion.
b) The subject and "other" population have existed together long enough in this or another
community for reciprocal natural selection to have occurred. Implicit here is, of course,
the set of conditions in both populations necessary for natural selection to be possible:
1. varation in the genotype of the "other" population through recombination and
mutation,
2. differential survival of the genotype,
3. heritability of traits in each population that reduce the detrimental effect resulting
from the interaction with another population.
The second dimension is, also for simplicity, given two categories of possibilities:
a) approximately equal generation lengths. (I assume absolute length is not crucial.), and
b) decidedly unequal generation lengths. Other characteristics of the populations are no
doubt useful, e.g., similarity of growth form, and ontogenetic stage of individuals in the
two populations at the time of association, but I will ignore these here for the sake of
brevity. A cross tabulation shows four possible combinations (Table 9.2).

Table 9.2. Simplified set of combinations between time of prior association


and comparative generation lengths for (-, -) competitors.
I-~-· -~--~
comparative generation lengths
I approximately equal decidedly unequal j

I newly
I prior associated a b
I association
previously
L__ ~_ associated c
------ d_ _ ~

The first row in Table 9.2, combinations a and b, characterizes newly associated popula-
tions, for example, the introduction of exotic plant species into a forest or range ecosystem,
or the more gradual introduction in primary succession (Spurr and Barnes 1980). In the for-
mer case the exotic may become a pest or be excluded. Unequal generation length will likely
increase the rate at which one of these two possibilities is realized. Unequal generation length
charact(!rizes the early stages of primary succession: the transition from mosses and annuals
to perennial forbs and grasses to mixed herbaceous to shrubs and to intolerant trees. Equal
or nearly equal generation lengths are more typical of the latter stages of primary succession.
A Theory of Indirect Effects 147

The second row, combination c and d, characterizes populations that at a previous


time affected one another sufficiently to have exerted selection pressure on each other. In
the context of forest succession, the second row applies to secondary succession or to the
recovery process following disturbances (Mclntosh 1980).
The likelihood that large differential fitness increases in surviving populations will even-
tually result in the displacement of one population by another is greatest for combination
of conditions b in Table 9.2: newly associated individuals of decidedly unequal generation
lengths.
Spatial variation of the individuals of the two populations is the second major char-
acteristic thought to affect the rate of differential fitness change in a (-, -) interaction. I
shall not elaborate the various proposition under scrutiny in this area of ecology. Interested
persons can find ample information in the works of Pielou (1969), Tilman (1982), Thompson
(1982), and the references cited by Thompson.
The inclusion of - 0 - and + 0 + indirect effects, together under the self-help theorem,
is dictated by the mathematical properties of the "0" operator. Their inclusion together
says nothing about the relative value of either. Most effect types in forest communities are
"wired in" by the nature of the parties involved, but the intensities may be changed by their
environment, which includes human actions.

Theorem 2.
I call Theorem 2 the self-annihilation theorem. It states that populations engaged in
either + 0 - or - 0 + indirect effects are "harming" themselves. Further, it matters neither
whether a population is enhancing another, detrimentally affecting another, or whether it is
being enhanced or detrimentally affected by another. Each is engaged in a self-annihilation
process. A tentative proposition may be offered:
A population's contribution to the likelihood of its own exclusion from a community
is proportional to the number of indirect - 0 + and + 0 - effects it has with other
populations in the community.
Because these populations are indirectly "harming" themselves or their future generations
one would expect to find numerous mechanisms that tend to minimize the effect intensity,
or change effect type so that it is not indirectly detrimental. These mechanisms are there.
Indeed, the dominance of discussions of the mechanisms of parent (-, +) and (+, -) inter-
actions in ecology texts has been criticized (Risch and Boucher 1976). In the discussion of
this theorem I present evidence for and against it, trying not to get lost in details. Most
general ecology texts, e.g., Ricklefs (1973), have good discussions of detailed mechanisms.
There is ample evidence that forest trees have not annihilated themselves by making
available to herbivores and phytophagous insects a large and continuous food supply. The
evidence? Landscapes are still predominately green. Also, there still are herbivores that
are prey for large carnivores. And there still are forest tree populations that are hosts to
plant disease orgamisms. Further, tree disease organisms have not completely eliminated
themselves from the forest. For example, Endothia parasitica, causal agent of the chestnut
blight, is still extant in eastern North America. In sum, giving the name 'self-annihilation'
to this theorem may be overstating the situation somewhat.
The - 0+ and + 0- indirect effects differ from the - 0- and + 0+ indirect effects because
one party may rightly be considered the initiator of the action and the other the reactor to
the action: wolves attack deer that flee; fisher attack porcupines that protect their vulnerable
head and underbody; microorganisms enter wounds in tree stems and colonize an area that
the tree isolates by compartmentalizing it. It is convenient to view these interactions as
follows:
a) the subject population is not the initiator of the action, and
b) the subject population is the initiator of the action.
148 A Theory of Indirect Effects

Of course, this is artificial. Action initiation and reaction to it occur concurrently. It is only
for discussion purposes that I separate them this way.
Schematically, the two cases are represented as:

a) subject -...-- other populations (initiator)


-~
b) subject +.- ~ -
other population
(initiator)

In case a) the subject population could be a plant (SUbject to grazing or colonization by a


disease organism), an herbivore (subject to predation) or one carnivore subject to predation
by another, to pick three examples. In case b) 'subject' could be a grazer, a predator, or a
parasite.
In case a) the subject population's best interest is unilateral avoidance. Natural selec-
tion in the subject population favors individuals with superior avoidance capabilities. For
example, survivors of predation in a prey population may be somewhat better adapted to
survive because of traits that helped in avoidance. Mechanisms of avoidance can be said
to produce a 'separation' between the subject population and 'other' population. Three
mechanisms have received much attention:
i) physical separation, e.g., deer herds occupying areas at the boundaries of wolf pack
territories (Mech 1977; Rogers, et aI., 1980, Nelson and Mech 1981),
ii) biochemical separation, e.g., plants producing secondary compounds harmful to grazing
insects (for a start on the vast and rapidly increasing literature in this area see Rosenthal
and Janzen 1979)'
iii) appearance separation, e.g., mimicry (see Robinson (1981) and other articles in the same
journal).
Thus, in case a) selection pressure on the subject population by the action initiators com-
plements best interest for the subject population.
In case b) above, the subject population is most likely to survive in a community if it
is moderate in its detrimental effects on the 'other' population. Overwhelming detrimental
effects may lead to exclusion of the 'other' population, and then to exclusion of itself. Mod-
eration is, then, in the initiator population's best interest. However, individualistic natural
selection favors "cheaters" in any population of individuals with a range of detrimental ef-
fects. With time, "cheater's" genes constitute a greater and greater proportion of the gene
pool of the action initiators. Mechanisms by which moderation can evolve under such con-
ditions have received wide attention. By no means is there agreement on these mechanisms.
Two examples are commonly cited as evidence that moderation can evolve. To organize the
following discussion, I separate the subject population, the action initiator, into two tradi-
tional classes: parasite and predator. I begin with a look at natural selection for moderation
in parasites.
Parasites: Ricklefs says that natural selection in parasites is against virulence:
The benign character of most parasites may be explained as follows. Any strain of
parasitic organisms that is so virulent as to kill its host also kills itself, therefore
natural selection acts strongly toward the evolution of benign levels of infection
(Ricklefs 1973, page 106).
Were this widely true, the direction of effect change for best interest of the subject popula-
tion arid that for natural selection in the subject population would be complementary. This
complementarity, when combined with the complementarity for best interest/natural selec-
tion in the host (case a above)' would make for a dual complementarity for parasite-host
A Theory of Indirect Effects 149

relations.
Since the benign nature of infection suggested by Ricklefs takes a long time to evolve, few
laboratory experiments have been of sufficient length to assess a trend toward moderation.
Complicating the assessment is the attribution of interaction change observed in the course
of experiments to evolved traits enhancing avoidance mechanisms and to evolved traits that
enhance exploiter moderation. Conceivably, these two contributions could be identified in a
long-term study done by Pimentel and coworkers (Pimentel 1968; Pimentel and Stone 1968;
Pimentel and AI-Hafidh 1965). They studied the behavior of host and parasite under two
sets of conditions:
i) two experimental systems were started with offspring of a host and parasite that had
coexisted for 1004 days, and
ii) one system was started from wild stock of host and parasite.
They use a wasp parasite (Nasonia vitripennis Walker) of the house fly (Musca domestica).
Unfortunately they studied only two of the four combinations needed to estimate the relative
contributions of these two kinds of evolution to interaction change. In sum, there are four
conditions required:
Experiment Host Stock From: Parasite Stock From:
A wild populations wild populations
B wild populations coevolved populations
C coevolved populations wild populations
D coevolved populations coevolved populations

Only experiments A and D were done by Pimentel and coworkers. However, it is instructional
to examine them even though the two components of change can not be identified. Figure 9.1
shows the time series graph of the experimental results. Experiment 1, the top graph, shows
the fluctuations for wild stock of both host and parasite.
The dramatic increases and decreases of host and parasite numbers are not found in
graphs of experiments 2 and 3. The genetic stock for these two experiments carne from pop-
ulations of hosts and parasites that had coexisted for 1004 days under controlled conditions
described by Pimentel and AI-Hafidy (1965). Pimentel and Stone were unable to explain
why the host population level for experiment 2 is considerably below that of experiment
3. Even so, these experimental results show fairly stable levels of both host and parasite.
Never does the parasite population become larger than the host population, and with one
exception in experiment 2, nowhere does a parasite population increase appear to induce a
substantial decrease in host numbers.
An interaction analysis is made difficult because the authors did not specify the "sep-
arate" numbers, those associated with what I have been calling spontaneous change. Since
there were no data produced, I made the following assumptions for the numbers in the (0,0)
relation: for experiments 1 and 2, I assumed equilibrium population numbers of 0 for the
parasite in the absence of the host and 500 for the host in the absence of the parasite; for
experiment 3, I assumed a zero level for the parasite and 700 for the host. The numbers
are parasite adults and pupae produced per cell per week in a 3D-cell population cage. The
interaction analysis was completed for each experiment with type and intensity computed
at five-week intervals. The total action metric z = Ixl + Iyl was used to measure interaction
intensity. A plot of the trajectories for each interaction was then made. The plot for experi-
ment 1 showed dramatic changes in interaction type and intensity. In one period from week
50 to 55, interaction type changed 38 0
while during another five-week period (week 30 to
,

35) intensity changed from .12 to .96. Fluctuations in intensity and type of interaction were
considerably reduced in experiments 2 and 3. There was also a progression of the trajectory
center toward the (0,0) circle and toward the (0, -) axis. To assess the extent of this change
in trajectory, a "centroid" of each trajectory was computed by averaging the intensities, R'
values, and types, 0 values, at five-week intervals with the following results:
150 A Theory of Indirect Effects

Figure 9.1. Time series graphs of three experiments reported by Pimentel and Stone (1968).
Solid lines depict host population levels and dotted lines the parasite population level. Ex-
periment 1 contained wild stock of both host and parasite, while experiments 2 and 3 were
started with offspring of a host and parasite system that had coexisted for 1004 days.

Experiment 1 Experiment 2 Experiment 3


mean type (0) 294 ' 282 ' 283 '
standard deviation
of type 16 ' 11' 8'

mean intensity (HI) .77 .55 .29


standard deviation
of RI .33 .23 .11

A graph of the interaction trajectory "centroids" is given in Figure 9.2.


Comparing the mean type of interaction of experiment 1 with those of experiments 2
and 3 shows a change in type of about 11 degrees in favor of the host. Mean intensity
in experiment 1 (.77) was more than double that in experiment 3 and about 50 percent
greater than in experiment 2. Figure 9.2 shows that the overall direction of change in type
of interaction is clockwise, to a position in the (+, ~) quadrant where the relationship is less
A Theory of Indirect Effects 151

wasp
, . . - - - - - - - - - - - - - - ( + , 0 ) parasite
(Nasonla vltrlpennls)

housefly host
(Musca domestlca)
Figure 9.2. Centroids of interaction trajectories for three experiments reported by Pimentel
(1968) and Pimentel and Stone (1968). Numeral 1 locates centroid for wild stock of both,
numerals 2 and 3 locate centroids for coevolved stock of both host and parasite. Change in
interaction type and intensity is due to what Pimentel calls genetic feedback.

advantageous to the parasite population and less detrimental to the host population. Had
experiments been run using stock of host and parasite as suggested for experiments Band C,
one could separately estimate the contribution of host avoidance (separation) and parasite
moderation to interaction intensity reduction and interaction type change.
Interaction change between a forest tree and a disease organism is subject to the same
components of change as Pimentel's organisms: 1) change in susceptibility of host to the
parasite, 2) virulence of the parasite, and after my argument here, 3) self-annihilation of the
parasite.
Take for example, Hepting's (1974) pessimistic views for the future of the American
chestnut:
1. The vast crops of sprouts from blight-killed adult trees are genetically identical to the
parent trees. Further, only one generation of trees has developed from seed since the
blight first struck in 1904. [Thus, no evolution of avoidance has had time to develop in
the chestnut.)
2. There is no evidence that the virulance of the fungus is any less now than in 1904.
Thus, no measurable host avoidance ability nor parasite virulence reduction has evolved. The
152 A Theory of Indirect Effects

major short-term hope for the chestnut appears to rest with the parasite self-annihilation in
many forest communities in the northeastern United States. Hepting (1974) also reports:
Observers in ... areas where the blight has worked the longest tell us that the
sprouts are now getting much larger. Often they grow old enough to produce
nuts before succumbing. Seedlings three or more inches through have been
reported.
These pockets formed by parasite self-annihilation now support American chestnut in areas
where it was thought unlikely to occur again, and they are bound to play an important role
in the future of this blight-tree interaction.
On the other hand, neither host avoidance nor parasite virulence moderation should have
been expected to develop in this important tree disease because: a) the generation length for
chestnut is long in an absolute sense (probably about 40 years), and b) the parasite can live as
a saprophyte on many forest species in the genera Acer, Carya, Quercus, and Rhus (Hepting
1971). In order for selection against virulence to operate most effectively, the parasite must
be obligate, i.e., require living tissue of the host to complete its life cycle. Saprophytic stages
and/or alternate hosts are not permitted. It follows, then, that in a population of obligate
plant parasites a most likely form of 'cheating' is the evolution of saprophytic stages and/or
alternate hosts.
Predators: The case for moderation on the part of predators is less advocated than
the case for parasites. Wynne-Edwards (1962) suggested a behavioral self-limitation on the
part of predator species that evolved by the process called group selection. Wilson (1975)
summarizes Wynne-Edwards initial conception as follows:
a cluster of small groups exists, completely isolated except for a trickle of dispersers;
within each group natural selection promotes increased resource utilization, even
to the point of overexploitation; groups that overexploit go extinct; given variation
in the composition of genotypes between groups differential extinction can create a
form of "group selection" promoting resource management.
Self-limitation supposedly produces "prudence" on the part of predators. However, Slobod-
kin (1974) claims that "prudence" can evolve without group selection. "Prudence" on the
part of predators, he suggests, refers to the kind of prey taken by the predator, not to the
quantity of prey taken. For example, a predator taking postreproductive prey can be as
greedy as desired, and it will not affect prey-predator stability. Given as an example is the
eastern timber wolf and its tendency to assess prey vulnerability with a preliminary test-
ing process, a brief chase, during which the wolf makes a decision to continue the chase or
abandon it.
In sum and in order of certainty:
1. Because populations having indirect effects of the types - 0 + and + 0 - are indirectly
harming themselves, numerous mechanisms have evolved and have been identified that,
in one way or another, reduce the extent of harm.
2. The direction of interaction trajectory change reflecting the best interest of the host /prey
population is the same as the direction resulting from individualistic natural selection
in the host/prey population.
3. The direction of interaction trajectory change reflecting best interest of the parasite/-
predator may very well not be the direction of change produced by individualistic natural
selection in the parasite/predator population. Much current research is focused on units
of selection other than the individual organism to account for stabilization of parasite-
host systems (Levin and Pimentel 1981). Evidence to date supports the idea that
fundamental differences exist in the processes leading to interaction change for parasite-
host systems and prey-predator systems. Carefully designed and excuted experiments
can, for certain classes of organisms, assess the relative contributions to interaction
change of a) avoiding the initiator, and b) moderation by the initiator.
A Theory of Indirect Effects 153

Theorem 3.
I call Theorem 3 the helpless theorem. It states that a population having indirect effects
of the type + 0 0, 0 0 +, 0 0 - , or - 0 0 has no opportuni ty to exert an influence on the party
it is interacting with so that it has an effect on itself. In short, the concept of best-interest,
operative in Theorems 1 and 2 is not operative in the indirect effects covered by Theorem 3.
Because there is considerable range in the net result of this "helpless" situation it is not
possible to make a single general statement of the likelihood of a population persisting in a
community of populations when the subject population is a party to these indirect effects.
Let me introduce the different situations as corollaries of Theorem 3:
Corollary 1. A population PI in a web of interacting populations will be favored to
persist the more indirect + 0 0 effects it has with other populations.
The arc-node characterization is:

Population PI has no way of affecting P 2 directly, but PI does receive some sort of desirable
effect from population P2 . This is the well studied interaction of commensalism.
Corollary 2. A population PI in a web of interating populations is a neutral or island
population the more indirect 0 0 - and 0 0 + effects it has with other web populations.
The arc-node characterizations are:

PI -

Recall, populations interact, but I examine only certain properties of individuals or the
entire population as being affected by the interacting party. Further, these properties are
linked to the process used in specifying the signs (see Chapter 7). Thus, the term "neutral
population" should be interpreted as neutral property of population PI. Neutral properties
of populations have for the most part attracted little attention.
Corollary 3. A population PI in a web of interacting populations will be less likely to
persist the more indirect - 0 0 effects it has with other populations.
The arc-node characterization is:

Population PI can neither enhance P2 , in which case its best interest is to reduce the amount
of enhancement, nor can it detrimentally affect P2 , in which case its best interest is to
continue its detrimental effect. Since population PI is detrimentally affected by population
P 2 , it is to its advantage to avoid P2 . But, the fact of avoidance in no way deprives population
P2 . In short, PI is in a helpless situation.
Not only is the - 0 0 indirect effect the worst possible situation for population PI,
according to the theory developed in this chapter, it is near the parent (-,0) interaction
where PI exclusion occurs. Recall that if population PI is associated with the horizontal
axes in Haskell's coordinate system and population P 2 with the vertical axes, exclusion of
PI occurs near 180 0
, i.e., near the (-,0) axis (Figure 9.3).

E. Natural Selection Levels in Community Interaction


Structure
I am now in a position to examine the various levels at which natural selection operates
in the process of community interaction structure change in the 51 system, symbolized by
154 A Theory of Indirect Effects

(-,+) (+,+)

(0,0)
------+-(+,0)
)~o~~)-?

---
I
(-,-)

(0,-)
Figure 9.3. Phase plane and Haskell coordinate system representation of the exclusion of
population PI. In the phase plane PI exclusion occurs on the E·" axis. In Haskellian coor-
dinates, exclusion of PI occurs on the axis of decrease for PI, the (-,0) aXIs. Arrows in
Haskell's coordinate system suggest that exclusion can originate from with (-, +) or (-, -)
interactions.

The previous arguments can be synthesized into two conjectures about interaction struc-
ture natural selection. The conjectures relate to levels at which the process of natural selec-
tion operates-indirect effects and direct interaction. The two conjectures are combined in the
next section into a proposition about overall evolution of community interaction structure.
There are two aspects to structure selection: intensity selection and type selection.
Treatments of interaction selection should consider both aspects of an interaction. In ecolog-
ical time and for populations with prior association, selection of interaction intensity from
high levels (far from the (0,0) circle) to lower levels (nearer the (0,0) circle) is both antici-
pated by the theorems of this chapter, observed in nature, and widely reported to occur in
forest communities. Pimentel and coworkers exhibited it experimentally, and Odum (1971)
treats it as a cardinal principle of population ecology. For these reasons, intensity selection
seems the less interesting of the two aspects of structure selection, and will be treated later,
where I show that it can play an important role in type change.

1. Indirect effect selection.


Recall from Haldane (1958) the general statement of what the concept of natural selec-
tion is about:

-A- selection may be defind as the elimination,


from a[n] -8- population, of [the less]C-

For indirect effects, placeholders are assigned as follows:


A - indirect effect,
8 - varied indirect effects, and
C ~ indirect effects with little or no complementarity between direction of interaction
type change for a population's best interest and the direction of interaction type
change due to natural selection at the individual/group level.
A Theory of Indirect Effects 155

For example, self-enhancing indirect effects come from both parent (+, +) and (-, - )
interactions, and in both cases individualistic natural selection works to enhance the capacity
for self-enhancement. So, complementarity exists for indirect effects from parent (-, -) and
(+, +) interactions. But parent (-, +) and (+, -) direct interactions show complementarity
of indirect effects only for the population being acted upon - prey or host. Action initiators
(predators or parasites) fail the test of complementarity because cheaters tend to increase in
the gene pool relative to moderates. My argument, then, is that natural selection operates at
the level of indirect effects and it selects against indirect effects starting with action initiators
in parent (-, +) and (+, -) interactions. The other indirect effects are all neutral, and so
are missing half of the basis for comparison included in placeholder C. The parties to these
parent interactions can not be self-enhancers, nor can they minimize the detrimental effects
they receive.

2. Interaction selection.
The second level at which structure selection operates is at the level of direct interactions,
or what I just referred to as the parent interaction. Recall again Haldane's definition of
natural selection that I paraphrased earlier. With regard to interaction selection, I make the
suggestion: A-Interaction selection may be defined as the elimination, from a population of
B-direct interactions of varying types and intensities, of the less C-dually complementary
interactions.
Dual complementarity is simply indirect effect complementarity both ways, i.e., Pl~)P2
and P{~P2' This assessment can be summarized as follows:
Interaction dual complementarity? (+, +) yes, (-, -) yes, (-, +) no, (+, -) no, (+,0)
no, (0, +) no, (-,0) no, (0, -,) no.
Clearly, only (+,+) and (-,-) have dual complementarity. Two others, (-,+) and
(+, -), have single complementarity. For the remaining interactions the concept does not
apply because of the neutral effect. It does appear, however, that (+,0) and (0, +) interac-
tions should rank above (-, +) and (+, -) because the former do not fail the test of dual
complementarity; they simply can't take it. It is tempting to rank the interactions according
to their "score" on this test of complementarity. Doing so produces the following result:
direct interaction rank - (+,+) 8, (-,-) 7, (+,0) 6*, (0,+) 5*, (+,-) [parasite/host] 4,
(-,+) [prey/predator] 3, (0,-) 2*, (-,0) 1*
Adjacent interactions with an asterisk should be considered interchangeable. Of course,
these are just ordinal rankings. Recall that parasites have traits that permit some moderation
in detrimental effect on their hosts, so (+,-) [p/h] should rank above (-,+) [pip]. The
lowest ranking goes to the interactions in which parties are helpless (see Theorem 3, corollary
3) .

F. Communmity Interaction Structure Evolution


Natural selection of direct interactions is the primary mechanism behind the evolution of
community interaction structure. My concern here is the relative frequency of types of direct
interactions among the community populations. A direction may be given to community
structure evolution as follows:
Community interaction structure evolves toward those structures having high de-
grees of dual complementarity in the constituent interactions.
Thus, community structures with many (-, +) and (+, -) interactions evolve toward com-
munities with relatively more (+, +) andlor (-, -) interactions. Of course, each community
probably has some of each interaction type, each having a different degree of dual comple-
mentarity. But, should natural selection of direct interactions operate as envisioned, then
community interaction structure should evolve to where (-, -) and (+, +), and (0, +) and
(+,0) predominate. There would be fewer (-, +) and (+, -) and still fewer (0, - ) and (-,0)
156 A Theory of Indirect Effects

direct interactions. This rarity should be more evident in communities with a long period
of close population association, for example, in communities nearing a climax state. On the
other hand, there should be relatively more (+, -) and (-, +) direct interactions in forest
communities made up of newly associated populations and early successional stages.
A schematic geometric picture of suggested directions of community interaction struc-
ture evolution is given in Figure 9.4. The thicker arrows indicate more probable avenues
of natural selection. Thus, evolution out of (+, -) (parasite-host) is more probable than
evolution out of (-, +) (prey-predator), giving an asymmetry to interaction evolution. Fur-
ther, either can evolve toward (-, -) or (+, +) quadrants, or the interaction intensity can
significantly reduce to near (D, D).

(0,+)

(-,0) - . . - - - - - - - - (+,0)

(0,-)

Figure 9.4. Directions of interaction type evolution due to natural selection that favors high
degree of dual complementarity of indirect effects. Evolution out of parasite-host (+, -) is
more likely than out of prey-predator (-, +). Arrow thickness is schematic to indicate more
likely probability of the various evolutionary paths.

At this point, several observations are in order:


1. Essentially two regions are identified,
i) quadrant I and adjoining axes of increase along with quadrant II!, and
ii) quadrants I! and IV and their neighboring axes of decrease.
2. The regions of high ordinal rank lie separated by regions of low rank.
3. There is an apparent asymmetry around a line passing through 45 0 and 225 0 .
4. The "high" ordinal rank of the (-, -) interaction is unexpected.
5. Evolution is away from quadrants II and IV and in the direction of either quadrant
or III.
How do these observations compare with other's ideas about t"he evolution of community
interactions? Odum (1971) argues:
In terms of the overall picture of the ecosystem the nine types of interactions can be
redU(,ed to two broad types, namely, the negative interactions I( -, -), (D, -), (-, D),
(+, -), (-, +)1 and the positive interactions I( +, D), (D, +), (+, + )1. Two principles
regarding these categories are especially worthy of emphasis: (1) In the evolution
A Theory of Indirect Effects 157

and development of ecosystems negative interactions tend to be minimized in favor


of positive symbiosis that enchances the survival of the interacting species ... (2)
Recent or new associations are more likely to develop severe negative coactions that
are older associations ... (Odum's emphasis).
Agreement or lack of it between my view of structure evolution and Odum's is made
difficult by the word 'minimized.' Since there may be a larger total number of species
populations in a forest community at an early stage of succession (more nodes in the graph of
the community), it stands to reason that there will be more total interactions (arcs connecting
nodes). So, if "minimized" means a decrease in absolute number of negative interactions, the
statement is true, no doubt; but it could also be said of positive interaction. But I assume
that Odum means the relative fraction of negative interactions to total interactions is less in
more evolved ecosystems. On the basis of the theory given above, I would argue that only
(-, +), (+, -), (0, -), (-,0) are less in an evolved ecosystems than in less evolved one.
Ricklefs (1973) argues, "Most mutualistic interactions probably evolved by way of host-
parasite interactions .... " There is no contradiction here. Further, by omission Ricklefs is
ascribing an asymmetry to interaction type evolution: interactions evolve out of quadrant IV
but not out of quadrant II. Ricklefs omits the route from the location of a typical host-
parasite interaction trajectory in quadrant IV (see Chapter 5) to that of, say, a hemi-
mutualistic relation (DindaI1975). Parasite-host evolution to hemi-mutualism can be viewed
geometrically in Figure 9.5. To get from point X to point Y there are at least two possible
routes. One is the gradualistic route, perhaps assumed by Ricklefs, as shown by the solid
line. The other is the route shown by the dashed line. It consists of three stages:
i) a decrease in intensity and move from point X to point A on the (0,0) circle,
ii) tessaring along the (0,0) circle to point B, and
iii) gradual increase in intensity in moving from point B to point Y.

G. Summary
In summary, let me review briefly the main points of this chapter:
1. community structure can be viewed as the array of direct interactions between commu-
nity populations;
2. community structure evolution can best be studied from a higher level of system concept
that makes use of net indirect effects as given by the monoidic "followed-by" operator;
3. the nine rules of the "followed-by" operator form the axioms for a theory of indirect
effects;
4. three theorems are deducible from the axioms (they relate to how a community pop-
ulation affects itself through another population, and are called the self-help, self-
annihiliation, and helpless theorems);
5. for each part to each interaction covered by each theorem, an analysis is made of
a) the indirect effect predicted by the "followed-by" operator,
b) the indirect effect more desirable for survival (self-enhancement),
c) the direction of effect change produced by individualistic or group selection,
d) the existence, or lack, of complementarity between directions given by b) and c)
above;
6. indirect effects for each party to a dyadic interaction are synthesized into an ordinal rank-
ing for each interaction (produced are two regions of interaction space using Haskell's
coordinate system); highest ranking is (+,+), followed closely by (+,0), (0,+), and
( -, - ); the other region of interactions with lower rank contains (+, - ), (-, + ), (0, - ),
and (-,0); and
7. conjecture is made that natural selection of community interaction structures is against
those with lower rank, favoring ones with higher rank.
158 A Theory of Indirect Effects

{O,+l

{-,Ol - - + - - - - - - - ------~I---+{+,Ol

{O,-l

Figure 9.5. Two courses for interaction change from parasite-host relation (point X) to a
hemi-mutalistic relation (point V). Solid line represents a gradualistic course of change.
Dashed line represents a postulated course wherein the interaction tessars from point A to
point B across the wrinkle in interaction space (L'Engle 1962).

This completes the basic elements of my theory of indirect effects for the simple case of
a community comprised of just two populations. Let us look now at indirect effects In
communities made up of many populations.

H. Research Problems
1. Relate the notion of a population being an indirect self-enhancer with the notion of
autocatalysis (Wicken 1984).
2. Identify interacting forest populations (either plant-plant, plant-animal or animal-
animal) that correspond to- each condition in Table 9.2.
3. Develop major dimensions of the spatial variation of populations that affects the likeli-
hood of interaction trajectories of various kinds. Develop a table analogous to 9.2 but
for spatial relations.
4. Catalog as many avoidance mechanisms as you can for prey-predator (-, +) and
parasite-host (+, -) relations.
5. Conduct an experminent similar to that conducted by Pimentel and colleagues, but test
the four combinations of wild and coevolved populations suggested in the text. What are
your assessments of the relative contribution of host avoidance and parasite moderation
to the change in interaction type and intensity?
6. Execute the experiment in Problem 5 for a prey-predator system. Assess any differences
in the results between your results and those for Problem 5.
7. Develop an experiment capable of falsifying Corollary 3 of Theorem 3.
8. Gather evidence to falsify my conjecture about the direction of community interaction
structure evolution. Analyze the relative frequency of direct interaction types in many
A Theory of Indirect Effects 159

forests at an early stage of succession and at a later stage.


9. Design an experiment that will drive a pairwise population interaction system to the
(0,0) circle. Where does it go after reaching the (0,0) circle?
10. Examine the notion that the (0,0) circle allows complex bifurcations in relations space.
Contrast with bifurcation in relata space.
10: Natural Selection of Interaction Structures
in Communities with Many Populations

I now expand to communities of many populations the method of examining community


interaction structure selection and evolution begun in Chapter 9. Required is the abstract
view of system used in the previous chapter:

populations
( and direct "0" , the "followed-by" ) .
System = , environment,
operator
effects (+,0, - )

The focus is the implications of the proposition that is corollary 3 of Theorem 3 in Chapter 9:
A population, Pi, in a web of interacting populations will be less likely to persist in
a community the more direct (-,0) interactions it has with other populations.
The context is the case where many populations are linked in an interaction web. (Of
course, the interactions must be with respect to the same process throughout the web.) My
goal is to suggest that the dominant mechanism behind interaction structure evolution is
summarized in a new exclusion principle. But first I must outline some community structure
analysis techniques.

A. Representing Results of The Structure Analysis


Our point of departure is the diagram accompanying corollary 3 (Chapter 9):

subject ~ other
population " - - - - - - - - - - - - - - : " population

I split the "other population" into two parts, those that affect the subject population and
those affected by the subject population:

• other populatins that


/ directly affect the
subject population
subject population.

~. other populations directly


affected by the subject
population.

Of course, the two sets are not disjoint. By placing the subject population at a privileged
point of a continuum, I can convert into a quantitative form the relationship between the
subject population and the "other population's" two parts. For now, the continuum divides
162 Natural Selection of Interaction Structures in Communities

the other population into that part directly affecting the subject population and that part
directly affected by the subject population.
The beginnings of a coordinate system are evident when the continuum is joined with
a perpendicular vertical axis, crossing the continuum at the privileged point, and graduated
in frequency of direct effects (enhancing and detrimental). A "band" is placed between the
horizontal continua that makes room for the number of direct neutral effects of the other
populations on the subject population (left) and of the subject population on the other
populations (right). For example:

4
+ +
effects of 2 effects of
other subject
populations population
neutral neutral
on subject on other
population
x populations
2

The simplest conceivable community is represented by the condition to which corollary 3


applies:
subject ~ other
population ° ---------;0 population
The interaction structure of this "community" is represented in the above coordinate system
as an "x" in the lower left quadrant and as a 1 in the neutral band on the right side. (To
simplify the language henceforth, immanating ("flowing in") refers to the effects of other
populations on the subject population and emanating ("flowing out") refers to the effects of
the subject population on the other populations.)
As the number of populations in a community increases, both indirect immanating and
indirect emanating effects can be examined. The continua in the coordinate system are easily
extended right and left to accommodate a number of degrees of indirectness. The number
of intervening populations in a tree of direct effects is taken as the degree of indirectness.
Thus, direct effects are zero-degree indirect, those with a single intervening population are
I-degree indirect, and so on.
The type of effect (enhancing, detrimental, or neutral) is determined for indirect effects
by repeated application of the "followed-by" operator. Further development of the interac-
tion structure analysis methods requires a brief aside at this point to spell out some of the
details and exhibit them with an example.

B. Interaction Structure Analysis Methods


The process of analyzing communities for emanating and immanating indirect effects can
be illust~ated using a hypothetical forest community represented schematically as a signed
digraph (Figure 10.1).
There are six steps to follow:
Natural Selection of Interaction Structures in Communities 163

LTH - large (den) tree population (hardwood) SH - shrubs


MT c - medium tree population (softwood) WLF - wolf
ST H - small tree population (hardwood) DR - deer
Dl - defoliating insect PO - porcupine
FR - fisher CN - cavity nesting birds

Figure 10.1. Graph of hypothetical forest community showing connectedness and type of
interaction between population pairs.

1. Focus in turn on each node in the graph and call the population at the node the subject
population or the head node (Wirth 1976).
2. Consider each subject population to be the head node of a tree of effects that "flow
in" toward the subject population. Construct a tree of these effects (populations and
their direct effects on each other). For example, the tree of immanating effects on the
medium size conifer tree in Figure 10.1 is:

MTc

SH

The signs indicate the type of direct effect the population at the arrow's base has on
the population at the arrow's head.
3. Repeat step 2, but construct a tree of emanating effects.
4. When completing steps 2 and 3, do not allow cycles in the trees where a cycle is defined
as a population's prior occurrence in the tree branch. Line number 1 (below) identifies a
164 Natural Selection of Interaction Structures in Communities

--- ~ ----
cycle in an emanating tree of non-neutral effects with population SH as the head node,
Lines two and three identify repetitions of populations

/1/
----~LT)
____ "DRJ
~
SrH
/' 2'...... ~SH
/.. --,.,--'-'----,
SH - - - - - - - . . . .~~ ST H
DR_~_~
--.. WLF

5. For each degree of indirectness use the "followed-by" operator to determine the type
of indirect effect on the subject population, For example, the net indirect effect of the
fisher population (FR) on the medium-sized conifer tree population (MTc) is enhancive
(a detrimental effect followed by a detrimental effect). This operation is repeated in
an immanating effect analysis for each population at each degree of indirectness, To
differentiate a net indirect effect sign from a direct effect sign, I place the indirect sign
at the base of each arrow in the tree. The immanating indirect effects tree for the
medium-sized conifer tree is, then

::~STH~
SH-
- - - - . . DR
CN~ . .-A"
LT
H ~ "(.
WLF~ :--..... ST ~ -0-..:..
, ---"'" SH~ DR o. MTc
----:?
H

DR~
. ~PO~
FR t,...---- / -
SH

6. For each degree of indirectness, count the number of populations having indirect enhanc-
ing, detrimental, and neutral effects on the subject population. A table of immanating
effects for the population of medium sized conifer tree is given below:

-degree of indirectness-
..,? 2 1 0
-frequency-
net 2 2
indirect
effect + 3

neutral 2 3
non-neutral
total 2 2 3 2

Steps two through six should be repeated for each node in the graph of the community
for both immanating effects and emanating effects.
Natural Selection of Interaction Structures in Communities 165

c. Interpreting The Analysis


On the basis of corollary 3, my interest is focused on the extent to which immanat-
ing effects are detrimental and emanating effects are neutral. But that interest is, in the
immanating case, not solely on the absolute number of detrimental effects. Rather, main
interest is on the relation between such actual and theoretically possible effects, given a
particular tree branching pattern for a subject population. Maximum possible frequency of
detrimental immanating effects is given by an actual tree structure where the direct effects
have been changed to those of collusive competitors. If the populations affecting node MT c
acted as collusive competitors, they would have direct effects as shown below. Note that the
signs have been changed so that all direct effects are enhancive except those that directly
affect node MTc. Because of the "followed-by" operator's rules, this pattern of direct effects
translates into all detrimental indirect effects for all degrees of indirectness .

. ~ .~

~.~~.~
. ~. ~.
. ~ ..

A table of worst possible indirect immanating effects is easily made using methods described
in the previous section. The profile of these potential detrimental immanating effects forms
the standard against which the profile of actual immanating effects can be compared. The
standard for comparison of the actual emanating effects is simply the neutral axis, no effects,
either enhancing or detrimental.
The two comparisons of actual effects with their standard for the worst possible condition
leads me to the following conjecture:
When the actual distribution of immanating and emanating effects "closely" ap-
proximates the worst possible case, it is likely that the subject populations will be
eliminated from the community.
The analysis methods of the previous section applied to the medium-sized conifer tree
give effect distributions shown in Figure 10.2. The immanating effect profile was determined
previously at step 6.
The properties of the effect profiles that are likely to lead to exclusion of the effects are
four: 1) a high frequency of detrimental immanating effects, 2) a small shaded area for the
detrimental immanating effects, 3) few emanating effects, and 4) small shaded areas for both
kinds of emanating effects. Based on these four properties, I submit that this community
population shows little likelihood of being eliminated. Not so with the white-tailed deer,
however (Figure 10.3).
Further discussion of these analyses can best be separated as follows:
Ill1ll1anating_ Clearly, there is an alternation of immanating effects between the effects
detrimental to and enhancing to the subject population. Alternation was also noted by
Vandermeer (1980), who expanded on ideas of Levine (1976) and showed that "as we go up
a trophic structure the dominant interactive forces switch from mutualism to competition
to mutualism." The alternation is caused by a predominance of detrimental direct effects
in the community graph (especially for deer) and the rules of the monoidic "followed-by"
operator. The pattern of immanating effects for the wolf population (not shown) is identical
to that of the deer, but shifted to the left by one degree of indirectness. Alternation prevents
166 Natural Selection of Interaction Structures in Communities

IMMINATING EMENATING EFFECTS


20,
+ +

D---D---O---O---~ --· 0 --- 0 --- 0 - -' 0


123 2 7 7 2
D --- D --- O --- O --- ~ ~ __ _ I __ _ I__ _ 0 -- 0 --- 0

,
5 1

,
10
I
I
,I
,
15
I
I
2'0
IMMINATING EMENATING
8 --- 7 --- 8 --- 5 --- 4 --- 3 --- 2 --- 1 --- 0 --- 0 --- 1--- 2 --- 3 --- 4 --- 5 --- 8 --- 7 - - 8
- degree of Indlrectne., -

Figure 10.2. Indirect effect distribution for the medium-sized conifer tree population of the
hypothetical forest community in Figure 10.1. Note the high frequency of emanating effects
and lack of alternation between enhancing and detrimental emanating effects.

the populations from having immanating effect distributions that coincide exactly with the
worst possible distribution. Lack of coincidence is highlighted by shading in the two figures.
The numbers in the neutral band indicate for each degree of indirectness the number of
branches that ended because a neutral effect was encountered.
Emanating. The frequency of enhancing and detrimental emanating effects in Figure 10.3
is quite low, but high in Figure 10.2. The former suggests that, based on frequency of
effects, deer populations are having little influence on the community. On the other hand,
the medium-sized conifer tree population seems to be exerting considerable effect, again
basing the judgment on frequency. Shading in the figures indicates the deviation of actual
distributions from theoretical distributions, with a small shaded area presumed positively
correlated with the likelihood of elimination. The neutral band numbers again indicate,
for each degree of indirectness, the number of branches ending because a neutral effect was
encountered. Figure 10.2 suggests that emanating effects lack the alternation character of
immanating effects.

D. Toward A New Exclusion Principle


Exclusion principles are laws of nature formed as vetoes, what nature will not be like,
Natural Selection of Interaction Structures in Communities 167

IMMINATING EMENATING EFFECTS


20
I
I

+
I
I
15
I
+
I
I
I
10
I
I
I

0 --- 0 --- 0 -- --- 0 --- 0 ---0- - -0,


5 11 9 3 2 2 2
0 --- 0 --- . . ...,0 ·--- 0 ---0 ---0

I
15
- I
I
I
I
20
IMMINATING EMENATING
8 --- 7 --- 6 ---5 ---4 ---3 --- 2 --- 1--- 0 ---0---1--- 2 --- 3 --- 4 --- 5 --- 6 --- 7 --- 8
-degree of Indlrec:tne •• -

Figure 10.3. Indirect effect distribution for white-tailed deer population of hypothetical forest
community shows a high frequency of immanating effects alternating between enhancing and
detrimental. Further, it shows a relatively low frequency of emanating effects.

while many, if not most, natural law statements take the affirmative, what nature will be
like. A general form of such a veto statement is:

A and B will not.. C.

Their form is dyadic: two items, A and B, will not do something, be somewhere, or
behave in some manner. In the following discussion I specify the placeholders A, B, and C,
examine and compare them for two well-known exclusion principles, and introduce a new
qualitative exclusion principle.
Pauli's exclusion principle is recognized as one of the most significant findings in the
physical sciences. It explained much empirical knowledge that had accumulated to that
date (1925) on chemical valence, the structure of the periodic table of chemical elements
(previously discovered empirically by Mendeleyev), spectroscopy, and magnetism (Margenau
1972).
In its most elemental form, and applying it only to electrons, Pauli's exclusion principle
states that no two electrons can be in the same state. Before quantum theory, the "same
state" meant the same position and velocity. With the advent of quantum theory, in partic-
ular the discovery of the fourth quantum number, the exclusion principle was stated more
168 Natural Selection of Interaction Structures in Communities

precisely: No two electrons can have the same four quantum numbers (n, a measure of the
electron's distance from the nucleus; I, a measure of the electron's angular momentum; m,
a number fixing the orientation of an electron orbit in space; s, a number (-1-1 or - 1) that
signifies the orientation of electron spin) (Margenau 1950, 1972). In terms of placeholders
in a veto format,

A- an electron with quantum numbers w, x, y, z


and B - an electron with quantum numbers w, x, y, z
will not C - be found in the same atom.

Margenau (1972) put into perspective the significance of Pauli's exclusion principle:
"Its success in solving problems was even greater than that achieved by relativity, but the
problems which the principle solved were very technical and therefore of interest to few,
and the peculiar integrative significance of the exclusion principle passed from view and has
even now been rarely recognized in its philosophical fullness." He states further that the
principle is a purely social law , simple in its basic formulation, yet immense in its collective
effect, and one that may lead to other symmetry principles, yet undiscovered, which will
unravel the mysteries of organization (Margenau 1944, 1972). Efforts to build connections
between biology and physics have emphasized the importance of Pauli's principle, and have
attempted to identify classes of biological entities that are analogs of elementary physical
particles called fermions that obey the principle and those called bosons that do not obey
the principle (Frazer 1955; Goldman 1971, 1973, 1980).
The competitive exclusion principle has been widely discussed, debated, criticized, rein-
terpreted and otherwise written about (Hardin 1960; Cole 1960; Gause 1935, 1970; Ayala
1969,1970; Grinnell 1904; Armstrong and McGehee 1980; Hutchinson 1961, 1978; Leak 1972;
Gilpin and Justice 1972; Antonovics and Ford 1972). Attribution is still unresolved. Nearly
every article about this principle contains a somewhat different form of its statement. In one
carefully ambiguous form, Hardin (1960) claims: "Complete competitors cannot coexist."
There is some difficulty in giving a precise interpretation to "complete competitors" and
"coexist." My interpretation of this statement, needed for the veto format given above, is:

A - a population with resource requirements x, y, z


and B - a population with resource requirements x, y, z
will not C - be sympatric.

By specifying that two populations have the same resource requirements, I am specifying
that they occupy the same niche.
Several questions have been raised about the competitive exclusion principle. Most have
been summarized elsewhere, so I only briefly mention them here:
l. It is tied to niche theory. For example, the claim of Ayala (1969) that he had experi-
mentally invalidated the competitive exclusion principle was rebutted by Gause (1970)
who claimed the populations used in the test occupied different niches.
2. When expressed mathematically, it is usually associated with logistic growth of popu-
lations and the well-known equations, often attributed to Volterra and Lotka, of mixed
population dynamics.
3. It is tied to the notion of equilibrium population levels in its formal expression. Hutchin-
son (1961) summarized this situation as follows:
Since the deduction of the principle implies an equilibrium system, if such
systems are rarely if ever approached, the principle though analytically
true, is at first sight of little theoretical interest.
At the present time, growing interest seems centered on systems far from or not at
equilibrium (Nicolis and Prigogine 1977, Pickett 1980).
Natural Selection of Interaction Structures in Communities 169

By themselves, niche theory, logistic growth relationships, Volterra-Lotka equations,


and the notion of stable equilibrium are useful constructs, but their union makes a tenuous
foundation for the competitive exclusion principle.
Both Pauli's and the competitive exclusion principles are thing-dyadic veto format law
statements. My work in the previous sections has involved two items, but they are distribu-
tions of effects, not things. The effects share a population much as Pauli's electrons share
effects (actions and reactions) and competitive populations share actions and reactions for
resources. Retaining the dyadic veto format, I suggest the following exclusion principle:

A - totally detrimental immanating effects


and B - totally neutral emanating effects
will not C - be properties of the same population in a system.

In other words, the effects exclude each other. In so doing, they cause to disappear from the
system any population they jointly characterize over a time period I suspect is specific to the
population. For biology, this principle extends the notions of Stern (1969), Stern and Roche
(1974), and Saki (1961) that competitive effects should be split into those exerted upon
other populations (Stern's competitive influence, my emanating effects) and those exerted
on the subject population by others (Stern's competitive ability, my immanating effects).
The principle also extends the notions of Patten (1979, 1982) on general systems "input"
environment and "output" environment into a predictive statement of association.
I claim that totally detrimental immanating and totally neutral emanating effects are
never found to characterize a population in a natural system. When the condition is ap-
proached, exclusion occurs between the effects and the shared population drops out of the
system. When this happens, the community structure is changed. Effect exclusion can be
avoided if the subject population forms new direct relations with other populations in such
a way as to move away from "total" detrimental immanating and "total" neutral emanating
effects.
The Pauli and competitive exclusion principles appear to have thing-based antecedents
of a very special nature, electrons (in fact, all fermions) and biological populations with
special resource requirements, respectively, hence they lack generality and extensibility. On
the other hand, the relation-dyadic antecedents of the suggested principle would appear to
make it applicable to a wide variety of populations of things occurring in systems. Should the
relation-dyadic principle prove as useful as hoped, it will require a more serious consideration
of the relational structuralist's perspective that relations (not things) are dominant in reality.
Further, the relational exclusion principle offers an alternative explanation for the exclusion
of a subject population from a community. Did the eliminated subject population occupy
the same niche as a single other population, or did the subject population's immanating and
emanating effects exclude each other?
With the formation of this exclusion principle, I claim to have taken route B in Figure 2.1
into the column for Why? questions.

E. Discussion
Before discussing practical uses and improvements needed, I consider presuppositions
of the suggested principle.
A crucial presupposition in the suggested principle is that the signed directed graph
is an adequate representation of the relationships in the community. This requirement is
significantly less ~han that of an apparently similar, yet much different, approach called
loop analysis (Levins 1975, Hutchinson 1978, Henry 1980). Loop analysis is based on the
community matrix concept which, in turn, presupposes equilibrium population levels in a
quantitative mathematical model of population dynamics. I presuppose only spontaneous
170 Natural Selection of Interaction Structures in Communities

change and induced change with respect to a process, constancy of effect type, and effect
transition according to the monoidic "followed-by" operator. Of course, the effects (enhanc-
ing, detrimental, or neutral) are specific to the process. Change the process, and the effects
will no doubt change as well. Changed signs and changed connections between populations
in a community will result in changed implications for indirect effect exclusion.
Important practical uses of the suggested exclusion principle, when used in conjunction
with the computational methods developed, appear to be:
1. It will help to identify populations in a community that are endangered by their so-
cial relationships, i.e., that place them in a position of receiving effects of change and
adaptation in the community's other populations, while at the same time preventing
the subject population from affecting its adapting neighbors.
2. It will aid in estimation of the likelihood of elimination of a population from a commu-
nity.
3. It will allow one to computationally estimate how effect exclusion can be prevented.
4. It will allow one to computationally estimate how effect exclusion can be made to hap-
pen.
5. It will allow one to estimate the effects of introducing a new population into a com-
munity, e.g., woodland caribou into northern Minnesota, or beaver populations into
landscapes populated with deer and wolf.
In more general terms, the suggested exclusion principle applies to relations between
relations; hence it seems at a higher level of abstraction than Pauli's and the competitive
principle. Thus, it is free of "thing bondage" and has general applicability to the fate of
populations in systems whether they be plant, plant-animal, or human. Further, the activi-
ties of humans to counter what the exclusion principle predicts has led to major structural
features in human society.
Because counting is the only kind of mathematics employed in the analysis of effect dis-
tributions, the current form is primitively mathematized. A means of weighting the indirect
effects should improve the resolving power of the emanating and immanating effect distri-
butions. For example, if a population has a detrimental effect on the subject population, it
makes a large difference whether the population has a short or long time to reproduction
relative to reproduction times of subject population individuals. An annually reproducing
insect population affecting a tree species population needing 40 years to reach reproductive
age has 40 more opportunities to adapt to the tree's traits than an insect population that
takes 40 years to reproduce (if one exists). A simple ratio of generation lengths is a place
to start to develop such a weighting factor. Further, skewness measures of the effect dis-
tribution should improve the method's resolving power. For example, subject populations
with immanating effect distributions skewed toward low degrees of indirectness probably
exert more pressure on the subject population than those skewed toward high degrees of
indirectness.

F. Summary
In sum, these are the major attributes of the suggested exclusion principle:
1. It combines ecological and evolutionary perspectives.
2. It has simple presuppositions.
3. It takes into account indirect as well as direct effects.
4. It incorporates ideas expressed by geneticists about a division of a population's role in
a community into those effects it exerts on other populations and those exerted on it by
other populations.
5. It summarizes the systemic notions of input and output environment into a proposition
about their combination and its effect on persistence of a population in a community.
Natural Selection of Interaction Structures in Communities 171

6. It is based on an axiomatized theory of indirect effects wherein the central axiom set
specifies the rules of an operator with defined mathematical properties (it is monoidic).
7. It is ontologically different than the thing-dyadic principles of Pauli and ecologists.
8. It applies to a population in a community of many populations.
9. It has application to populations of many kinds, not just plants or animals.
10. It is testable, at least in laboratory settings.
11. It helps to establish the importance of the relational structuralist's perspective that
relationships are the dominant reality.

G. Research Problems
1. Write a computer program that takes as input a signed direct effect matrix and returns
as output the indirect interaction trees; 1) emanating out from the subject population,
and (2) radiating in on the subject population.
2. Extend the computer programs in Problem 1 to generate indirect effect graphs for both
the actual and potential (worst possible) emanating effects and immanating effects.
3. Develop a rationale for indirect effect profiles for "constructive" and "destructive"
species populations.
4. Analyze the 31 food webs given in the appendix to Cohen (1978) and look for "con-
structive", "destructive", and "threatened" species populations.
5. Design and conduct an experiment that will show a population in the worst possible
situation (totally detrimental immanating effects and totally neutral emanating effects)'
yet it continues to exist in the community for many generations.
6. Develop a rationale for integrating indirect effect analyses for the different process es-
sential for survival of a population (e.g., nutrition, reproduction, protection).
7. Analyze the conjecture that too many enhancing relations in a community contributes
to community instability because:
a) the + effect is the identity element, and it simply "passes on" other effects,
b) many "passed on" enhancing immanating effects, if "captured" by a detrimental
effect, constitute the worst possible case - that of collusive competitors, and
c) enchancing emanating effects are unable to overcome neutral effects.
8. Find other natural systems containing monoidic or group operators. What is the relative
frequency of the identity element in these systems?
9. Adapt Patten's quantitative model of input and output environs to the presuppositions
and methods of this and the previous chapter (Patten 1982). What adjustments must
be made in the way Patten makes his analyses?
10. Because the suggested exclusion principle is relata independent, to what extent does it
have application in other disciplines as well. Human sociology, for example?
Part V
Decisi on-ma king
11: A Clarification and
Extension of Multiple Use

Nearly every discipline, scientific or not, has one or more concepts that are perenni-
ally troublesome. Ecologists have argued for years about "competition"; philosophers have
argued about "denotation" and "designation." Forestry's contribution is, without doubt,
"multiple use." But "multiple use" is different, for it is about power (Convery 1979). McAr-
dle (1953) put it bluntly: "this is no penny ante game; the stakes are tremendous." In
a more philosophical vein, Hall (1963) quotes Charles A. Reich: "the power of the Forest
Service is awesome, for the Service recognizes ... that its job is nothing less than the definition
of public good, a task once reserved for philosopher kings." Because multiple use plays such
an important role in forest management decision making, it bears further scrutiny.

A. 'Multiple Use,' "Multiple Use," Multiple Use


The pattern of expressed frustrations with troublesome concepts is repeated in the
literature on multiple use. They deal with meaning and definition of the term multiple use,
what the concept "is," and to what the concept refers. Some highlights from the literature
illustrate this frustration.

1. The term.
Stagner (1960) suggests that the term "is sometimes used so loosely that one wonders
if it has any meaning at all." Zivnuska (1961) asks, "Exactly what do we mean by multiple
use?" He suggests three uses (but recommends a fourth): a) use it as a "slogan term" for
multiple products, b) use it as an economic term for joint product economics applied to
forestry, c) use it as a philosophy and method of land management. He recommends that
multiple use be considered a concept of management. The Multiple Use Sustained Yield
Act of 1960 contains a multiple use means paragraph. Hall (1963) is not satisfied, claiming
that "the meaning of multiple use has not been established-either by a consensus among
natural resource experts or by legislative decree." Behan (1967) expresses dissatisfaction
with clarification efforts, suggesting that there has been only one concise definition and
exposition of multiple use (Waugh 1936). The others, Behan claims, have been sloganeering,
and that multiple use is really a doctrine "a) ill defined, b) vague, c) ambiguous, d) generally
fuzzy, and e) mostly meaningless." Convery's solution is to "bury both phrases [mUltiple
and dominant use] since they encourage decision making by slogan rather than by reason"
(Convery 1979). Teeguarden (1979) accommodates by introducing the term multiple services,
yet states explicitly that multiple use, multiple service, and integrated use or management
"are all intended to mean the same." His interpretation of multiple use is "a program of
managerial inputs rationally selected to produce a desired set of forest services." Snyder
summarizes the situation by saying, "The great difficulty with multiple use is the almost
complete disagreement on what it means." His remedy is strict honesty in the use of terms
114 A Clarification and Extension of Multiple Use

related to multiple use. Needed is honest adherence to terms that have agreed definitions
(Snyder 1979).

2. The concept.
The literature on the concept of multiple use has centered on the context in which it
does, or should, find application. Stagner (1960) argues that "it is a planning concept ... not
an operating method ... it is not a practice to be applied to each individual unit of land within
a land estate." Behan (1967) attributes to Evans (1938) the notion that multiple use is a
"conception of forest management." Behan further states that the Forest Service Manual
says multiple use is a "principle of management" rather than "a system or method of land
use." He argues that multiple use is a doctrine that became law in 1960 with enactment of the
Multiple Use Sustained Yield Act. Ciracy Wantrup (1938) refers to the principle of multiple
use and argues that the event of multiple use "is the actual or hypothetical result of economic
conditions ... and to regard multiple use as a general objective .. .for policy is not warranted
logically-it puts the cart before the horse" (Ciracy Wantrup 1961). Hall (1963) also says
multiple use is a principle, but he notes that the Forest Service converted the principle into a
slogan and produced a 5 part emblem to symbolize "the multiple contributions of the forests
to national well being." Starr (1961) claims the concept of multiple land use management
deals with 1) management systems (distributed tract system, single tract system) and 2)
types of land units (inventory units, management units, administrative units). Pearse (1969)
writes "toward a theory [not concept or principle] of multiple use." Gregory (1955) performs
an "economic analysis of the multiple use concept."

3. The event/thing.
Two historical perspectives on the event of multiple use have drawn heavily on farming
analogies. Pearson (1944) says that "effective multiple use is merely organized and coordinate
specialization .... When foresters have developed an effective program of specialized use, they
will realize that wild land management is nothing more than a form of agriculture." Behan
(1967) reviews the article by Waugh (1936) who wrote about intercropping beans and corn.
Claiming this an example of multiple use [of land], Waugh said another example can be
found on our National Forests. According to Pearson, good Midwestern farmers know there
is a place and a time for every activity (Pearson 1944).
Several important observations can be made in regard to this brief literature review.
First, consideration is directed mainly at two categories of concern in Chapter 1: term and
concept. The event of multiple use remains little examined. Second, there is the usual heavy
reliance on definition as a way to specify term meaning. Recall, Bunge argues that terms get
their meaning vicariously through the concepts they desig nate, and he suggests that terms
be said to have significance (Bunge 1974a). Concepts can have meaning, defined as the 2-
tuple <reference, sense>. Third, many attempts to explicate the multiple use concept take
the form of adding a label rather than digging into the concept itself. Clearly, multiple use
is a relation concept of at least ternary predicate structure. Applying the method suggested
by Bunge (197 4a) to consider a predicate as a propositional function that maps objects into
statements, I start with

M R x L x T --+ S,

where
M designates the concept (predicate) "multiple use,"
R designates the set of resources,
L designates the set of locations (tracts, places),
T designates the set of times,
S designates the set of statements that use the concept,
A Clarification and Extension of Multiple Use 175

x designates the mathematical symbol for Cartesian product of sets, and


designates mathematical mapping.

Fourth, there has been minimal effort to spell out an ontology for multiple use. What are the
referents of the concept? What properties of the referents are needed in representations of the
event of multiple use? Without a well formed ontology, unattainable tasks can arise that, for
example, call for analyses of multi resource use interactions. Fifth, the ontological confusions
are made worse because the biology of natural object populations and the sociology /biology
of user populations are almost totally ignored. Sixth, there is confusion about natural
objects and constructs, and the permissable relations among and between items in these
categories. For example, in Bunge's ontology natural objects can interact, but concepts can
only be interdependent. Sticking with this ontology I ask, to what category does timber
belong? Is it construct or natural object? If it is construct, timber growth is ontologically
inappropriate. If it is natural object, timber must be expected to interact with other natural
object populations, for example, trees. Seventh, the agricultural analogies of Waugh (1936)
and Pearson (1944) misrepresent the complexity of multiple use of forests. Intercropping of
beans and corn, wherein both are grown on the same piece of private land, harvested and
removed from the land, and consumed elsewhere or sold to be consumed elsewhere, has no
analog in forestry. For example, in multiple use of public forest land, the manager is not
the owner, the "crop" is "consumed" by only a small fraction of the owners, and much of
the "consumption" is done on site. These differences make multiple use of forests, especially
public forests, much more complex than mere intercropping of beans and corn.
In sum, the difficulty foresters have had with multiple use is evidence of what was argued
in Chapter 1; term definition alone is unable to clarify a complex notion. The examination
must be extended from terms to the concepts they designate, to the natural objects to which
the concepts refer. In short, multiple use will not be explicated without recourse to each
node in Figure 1.1. and the relations of designation and reference.

B. Analyzing "Multiple Use"


Bunge's formula for determining the meaning of a concept is to identify the concept's
reference and sense. I confine my efforts in this section to exposing the referents of the
concept. Needed is the concept expressed in propositional function form. I shall progress
through four steps in analyzing the referents, briefly discussing each step's adequacies and
inadequacies.
In its simplest form, "multiple use" can be said to map the Cartesian product of the set
of forest resources, the set of locations, and the set of times into statements. Symbolically,
the result is the expression given earlier in this chapter. The sets Land T allow inclusion of
rotating use and single uses across a landscape under the multiple use umbrella.
Elementary set theory analyzes the many kinds of sets and relations between them.
Three kinds and one relation are needed here:
a) kinds of sets
i) the empty set (designated 0)
ii) a set containing a single element (this kind of set is called a singleton)
iii) a set containing more than one element
b) relation between sets
iv) equality of sets (sets A and B are equal if and only if set A is contained in set B
and set B is contained in set A).
By varying membership number in sets L, T, and R, the special variations of "multiple
use" are identified. For example, if Land T are both singletons and R is, say, neither null
nor singleton, the result is the classic case of multiple use referring to the production of
two or more resources on the same area at the same time. Many object to this version of
the concept, e.g., Pearson (1944). The major inadequacy of this expression of the concept's
176 A Clarification and Extension of Multiple Use

reference is that the first set, resources, is of doubtful value as an ontological category. What
is a resource? A population of objects? A population of objects in a particular relation to
another population? Or is "resource" just a relation?

c. Extending the Concept of Multiple Use


Multiple use can be sharpened by considering "resource" to be a relation between pop-
ulations of natural objects (trees, animals, etc.)' and populations of prospective users of
the natural object populations. This leads to splitting the set of resources into two sets,
populations of objects and populations of users. Symbolically, this step gives:

Mz P x U x L x T -t S

where
M z designates "multiple use,"
P designates the set of natural object populations,
U designates the set of users of natural object populations, and
L, T, and S are sets as given above.
While resource has been rejected as an ontological category and removed from the
referents of "multiple use," it has been replaced with two less troublesome yet still nonspecific
categories: natural object populations and populations of users of natural object populations.
But, both of the latter are still too general to participate as members of the referents of
"multiple use."
Another refinement comes from partitioning both the natural object populations and
user populations into two subsets. I break the set of natural object populations, P, into
biotic populations, P b , and abiotic populations, P a . Also, I break the set of user populations
into biotic users (nonhuman), U b , and human users, U h . Symbolically, this version of the
multiple use concept is

where
P b designates the set of biotic populations, i.e., trees, not timber; animals, not wildlife;
shrubs, not brush, etc.,
Pa designates the set of abiotic populations, e.g., water bodies (streams and lakes,
etc.), natural "mineral licks," etc.,
U b designates the set of populations of biotic users of biotic and abiotic natural object
populations,
U h designates the set of populations of human users of biotic and abiotic natural object
populations, and
L, T, and S are sets as above.
This concept of multiple use has more acceptable referents, ontologically speaking. l\'ote
that the set of populations of biotic users contains members that are also included in the
set of biotic populations. Wolves prey on deer, deer browse on aspen shoots, etc. However,
no nonhuman population preys on wolves; but some may parasitize them. Further, biotic
populations are users of abiotic populations. Trees transpire groundwater, deer and moose
obtain nutrients at mineral licks, and so on.
A fourth, and for now last, step is to recognize a second group of humans, the set of
managers, as distinct from the set of human users. In the set are forest managers, wildlife
managers, range managers, etc. Let this set be designated M, giving:
A Clarification and Extension of Multiple Use 177

where
M designates the set of managers, and other sets are as above.
The concept of multiple use given by this propositional function has referents that are
ontologically satisfactory for my purposes. It includes sets normally omitted from consider-
ation in the many discussions of multiple use. Further, it provides an ontological supporting
frame work for an added dimension to "multiple use" decision making.
Several observations can be made about these four steps in identifying the referents
of "multiple use." First, by letting one or more of the sets be null or singleton, a whole
genus of multiple use concepts results. Consider the species "wilderness." It is that version
with sets U h and M null, and many element sets Land T. Or consider the concept of ex-
clusive use timber production. Here it can be argued that P b contains, but is not limited
to, trees; M is the singleton, forest (timber) manager; U b is the empty set (anything that
would have consumed trees has been eliminated or exterminated), and U h is the singleton,
timber harvester/processor/consumer. Second, depending on the category of locations, L,
the set of owners, call it 0, can be defined in two ways: 1) 0 1 = M = U h, e.g., a small
private landowner may well be the manager and the only human user; and 2) O 2 = M u U h,
e.g., state or federal lands are owned by the union of the set of human users and managers.
Normally it is assumed that managers of public forest lands are automatons implementing
policy, with no ownership interest. Third, depending on the owner set's attitudes or beliefs,
e.g., dominion over nature or stewardship of nature, attempts may be made to significantly
change the biotic user set, U b . Fourth, this form of the multiple use concept casts hu-
man users of natural object populations into a mold similar to the ecologist's treatment of
resource use and partitioning by biological populations. Fifth, biological populations use
natural object populations both biotic and abiotic. Trees use water in photosynthesis and
transpiration; insects consume tree foliage; deer obtain nutrients from tree seed (mast) and
get protection provided by tree canopies; and so on. Sixth, human populations use natural
object populations-both biotic and abiotic. Trees are used for lumber and plywood to build
houses; water for drinking, irrigating crops, watering livestock; places (space) for solitude,
risk, adventure and so on. Seventh, when two or more user populations require the same
object population, three classes of user conflicts can emerge: biotic-biotic, biotic-human, and
human-human. Eighth, the traditional multiple use concept has been confined to human
users and to resolving human human use conflicts. The approach has had adverse effects
on biotic populations (Lennartz 1979). Ninth, the resolution of biotic-biotic user group
conflicts (these may be indirect through an element from sets Pa or Pb ) has traditionally
been left unresolved or has been considered a biological/ecological problem far removed from
multiple use. Typical of the methods man uses for resolving biotic user conflicts is to place
financial values on individuals of one, both, or all three populations. Resolution is made to
favor the populations with the highest financial value. When carried to extremes, widespread
elimination of biotic users of biotic and abiotic populations can threaten a biotic population
with exclusion from a community. Exclusion from many locations may lead eventually to
extinction, i.e., exclusion from all L, hence for all T. Extinction is the last exclusion. Tenth,
two alternatives exist to the typical method of resolving conflicts described in comment nine.
One is to resolve biotic biotic conflicts "naturally," i.e., in the absence of man. Here both
man the manager (set M) and man the consumer (set U h ) are null, nature being allowed to
"take its course." The other alternative is to resolve the conflict in such a way that neither
natural object population is likely to be excluded from the community. This is accomplished
by members of set M ensuring that population interactions (both direct and indirect) do not
become predominantly (0, ~) or (~, 0). I call this pancentric valuation, because equal value
is put on all natural object populations. Eleventh, biotic user-human user conflicts have in
the past been resolved in the short term in favor of the human users. Lennartz (1979) de-
scribes how, after a series of setbacks for biological user populations, the endangered species
act was passed, legislating the conditions under which the biotic user population must be
178 A Clarification and Extension of Multiple Use

favored. This legislative restraint on the typical anthropocentric view of forestry decision-
makers has reportedly caused much consternation (Hutcherson 1976). Twelfth, Snyder
(1979) argues that the set of locations, L 1 , where M operates should be a subset of all the
locations, L, i.e., Ll C L. Some areas, he suggests, should be left uncommitted to any form
of management.
In sum, the salient features of the generic concept "multiple use" are
1. sets of biotic and abiotic populations, rather than resources,
2. sets of biotic populations as users of biotic and abiotic popula tions,
3. sets of human populations as prospective users of biotic and abiotic populations,
4. sets of locations,
5. sets of times, and
6. sets of managers.
The particular "species" of multiple use of concern can be made from these sets using set
theoretic notions of null sets, singletons, and many element sets.
This completes my examination of the reference part of the 2-tuple
<reference, sense>
needed to specify a precise meaning for the concept "multiple use." I leave the completion
of the sense for another time.

D. Research Problems
1. Find as many different definitions of multiple human use as you can. Evaluate the
definitional strategy used in each.
2. What is the referent of the concept of multiresource interactions?
3. What is the referent of the concept of multiresource use interactions?
4. Using the generic formula for multiple use, use set theory to identify as many species as
you can. Give a name to each if one does not exist already.
5. Express the concept of dominant use mathematically using set theory.
6. Trace the origins of the idea of using forests for several purposes simultaneously.
7. Is the implicit modifier "human" in multiple use also implicit in the attitude of Eastern
cultures towards forests?
8. What is the concept of "resource" about?
9. Develop an alternative set of extensions of the multiple use concept.
10. Evaluate Lennartz's argument that for many species, multiple use is a promise unful-
filled.
12: Conservation
Traditions

Conservation has been defined as having "to do with the proper rate of use of a resource"
(Jeffers 1943). Human conservation of renewable natural resources is, however, sometimes
limited to selecting the proper rate of human use of the natural objects currently beneficial
to man. Conservation should not be confused with preservation, although preservation could
be considered a special case of conservation: one where the "proper" rate is a zero rate of
human use. Throughout this chapter I adopt a conservationist's stance, "proper" can range
from zero to a positive value, depending upon the tradition to which one subscribes.
The conservation traditions I use are those outlined by Petulla (1980): economic, eco-
logic, biocentric. My treatment of the economic tradition is more narrowly focused than
Petulla's; I emphasize multiple use economics and multiple use decision-making methods.
Ecologic and biocentric tradition treatments will likewise be more narrowly focused, on liv-
ing components of forest ecosystems.

A. Economic Efficiency Tradition


Petulla (1980) describes the focus of the economic efficiency tradition:
The economic perspective, sometimes more broadly called the utilitarian approach
to conservation, focuses on the optimal use of natural resources for the longer period
of time, or more recently, the assignment of costs to those who take away public
environmental amenities. Almost a century's worth of urban and resource planners
fit into this tradition because of their interest in the highest level of efficiency for
the benefit of the tax paying public.
In sum, this tradition is concerned with human action, the costs of these actions (preferrably
expressed in financial terms), and the benefits to man (also expressed in financial terms)
that result from the human actions. Thus, the economic tradition is also an anthropocentric
tradition-man and man's near term welfare dominate decision making.
The anthropocentric character of forest decision-making has long dominated the forestry
profession. Pinchot (1917) made his view abundantly clear that "the central idea of the
forester, in handling the forest, is to promote and perpetuate its greatest use to men. His
[the forester's[ purpose is to make it serve the greatest good of the greatest number for the
longest time." The "greatest use to man" emphasis of the profession has been translated into
a commodity ontology-timber, range, wildlife, water, recreation-that some suggest currently
dominates planning and decision-making in the U. S. Forest Service.
The extent to which the economic tradition dominates Forest Service decision-making
is shown by examining reports prepared for the periodic assessments of forest resources.
Schweitzer, leader of the wildlife and fish assessment team, argued that the fish and wildlife
aspects of recent national forest resource assessment have had a commodity or anthropocen-
tric character because they are legislated to be that way:
180 Conservation Traditions

The Resources Planning Act and related legislation make clear that the fundamental
question to be addressed by a national assessment [of fish and wildlife resources] is
not what might be done to improve the condition of wildlife and fish; the central
question is What might be done to improve the welfare of people? (Schweitzer, et
aI., 1981).
But it is well known that forest economists were instrumental in writing the resources plan-
ning act and related legislation. The philosophy with which economists approached the task
of assessing the condition of a biological resource should clearly separate adherents of the
economic efficiency tradition from those discussed later in the chapter.
The discipline of economics is, like most, diverse in its subdisciplines and specialty areas.
My primary concern here is on the subdiscipline used for many years in studying problems
of multiple use economics, the economics of joint production. Joint production economics is
appropriate when there is more than one object of production and when the output of one
affects the output of another. Ontologically speaking, the commodities are interdependent
because the natural objects interact; timber and range are interdependent because trees and
grass interact in a biological system.
Behan (1967) gives a good historical review of the multiple use concept and its imple-
mentation in the management of national forest lands. The association of joint production
economics methodology to multiple use problems was apparently first recommended in 1938
(Ciracy Wantrup 1938). Gregory (1955) spelled out the association in detail, and expanded
on it later (Gregory 1972). A good overview of the economics of multiple use (to 1969) is
presented by Lloyd (1969). Zivnuska (1961, 1979) has written on the economics of multiple
use, as have Duerr (1963) and Teeguarden (1979a,b).
The principal constructs used by economists in their quest for joint production efficiency
are 1) product-product graphs, 2) isocost curves, and 3) isorevenue lines. Details of these
constructs are given by Gregory (1972) and additional background information by Heady
(1952). Skipping over difficulties of constructing product-product graphs, the task is basically
to establish the units of two "products" of output, from the same forest for given levels of
inputs. The inputs will be expressed as dollars of cost. For example, in a hypothetical
case, three different combinations of Y and X can be produced for 10 units of input costs
(Figure 12.1). Suggested is a continuous range of outputs Y and X from this stand, all
costing 10 units of input. For simplicity I assume that for 20 units of input the combination
of outputs labeled B, Figure 12.1, can be produced. Additional combinations may be shown
for reasonable increments of costs, but are omitted here.
The quest for economic efficiency requires the introduction of financial values of the
respective products Y and X. Again, for simplicity I assume a linear relationship with
slope: (units of Y /units of X) needed to produce a set revenue. The total revenue assigned
to a line is determined by the price per unit output and the quantities of output (Figure 12.2).
The slope of the isorevenue line here is unity; output of 40 units of Y produces the same
total revenue as output of 40 units of X. If total revenue is $80, each of Y and X are worth
$2 per unit.
The combination of isocost and isorevenue relations in the same coordinate system is
shown in Figure 12.2 for this hypothetical example. The point of tangency between isocost
and isorevenue curves, for various levels of revenue, locates the combination of products one
should produce to gain that level of revenue at least cost. For example, $76 of revenue is
produced from about 23 units of X and 14 units of Y at a cost of $20, giving a benefit/cost
ratio of $76/$20 = 3.8. No other combination of outputs yielding $76 can be produced with
as large a ratio. Thus, the C combination is the most economically efficient combination.
(I ignare the concept of expansion paths and concepts used in maximizing total revenue for
a firm.)
There are two well-recognized problems with the theory of joint production as a decision-
Conservation Traditions 181

50

output 40
of Y

30 Bt

20 At
S, A2
'[2]
10 ,,
\
B2
0
0 10 20 30 40 50

output of X

Figure 12.1. Idealized steps in development of product-product graph for the joint production
of X and Y. In theory, a continuous curve can be passed through combinations of output
with equal cost.

50

output 40
of Y 8
30

20

10

0
0 10 20 30 40 50
output of X
Figure 12.2. Hypothetical isorevenue curve in the product-product graph. Slope of isorevenue
curve is dependent on relative value of products X and Y. Most efficient combination of
products X and Y for $20 expenditure are coordinates of point C.

making method for economic efficiency:


1. Muhlenburg (1964) argues that "to operate [the theory of joint productionJ ... successfully
it would be necessary to obtain continuous expressions of product combinations on a
series of isocost curves. [T]his becomes almost impossible in view of the complexity of
the real world. Therefore, the pure concept, while of great intellectual merit, is of little
practical importance .... " He suggests that point estimates are feasible, and develops a
method of handling such estimates.
2. Whaley (1970) argues that "many of the commodities do not command a well established
market demand." Hence, for example, "lnJo market established prices can represent the
182 Conservation Traditions

values of recreation and water in investment analysis." The inability to establish market
values of forest outputs prevents reliable estimates of isorevenue line slope, which in
turn prevents reliable estimates of the point where marginal cost of an additional unit
of output exactly equals the marginal revenue from producing that additional unit.
Whaley goes on to describe the ways of overcoming this problem; basically it is to avoid
it by using a method that does not require value data.
Clearly, both constructs-isocost and isorevenue-have been identified as barriers to the full
application of joint production theory to economic efficiency questions in multiple use man-
agement.

B. Ecological Tradition

Community stability is an important concept to followers of the ecologic tradition.


Petulla (1980) emphasizes:

The ecologic emphasis is derived from a scientific understanding of interrelation-


ships and interdependence among the parts of natural communities. The important
ecological concept for the group is a model of a stable community made up of plants
and animals (preferably rich and diverse) and a traceable flow of energy which may
be disrupted by natural phenomena or, more commonly, by human activity.

But community stability is a very complex concept that is interrelated with other community
concepts, e.g., complexity, interaction structure, diversity, and connectance. It should be
noted that, as Petulla puts it, "the conceptual difference between biocentric and ecologic
viewpoints lies in the commitment of the viewer, whether to the rights of nature in itself
or to a model of scientific understanding of how nature orders itself" (Petulla 1980). Of
course, scientists of an ecological persuasion commit themselves to the models of scientific
understanding because they believe them to be valid representations of how nature operates.
My intention is to focus in this book on the interactions in natural systems, so it is appropriate
to identify and summarize the principals who committed themselves to a model of scientific
understanding of how nature orders itself with respect to interaction structures. Several
person and dominant interaction names come immediately to mind:

Charles Darwin - struggle


Sir Petr Kropotkin - mutual aid
G. F. Gause - struggle
Eugene Odum - cooperation
Aldo Leopold - mutual and interdependent cooperation
Garret Hardin - competition
Edward Haskell - cooperation

These scientist's names can be mapped m an approximate way into Haskell's coordinate
system (Figure 12.3).
Conservation Traditions 183

<-,+)
4 /
Y
/ /
/ /
<+,0) Leopold/ /
/ /
/

/
/ .

E. Odum',/
, / /, , //
<+,+) .
. , Kropotkin

Figure 12.3. Scientists names mapped into Haskell's coordinate system according to the
class of interaction generally associated with their names. Scientists associated with the
upper right do not, of course, deny existence of net negative interactions but they argue that
evolution proceeds in their direction. Darwin, Gause, and especially Hardin, are generally
assoicated with the view that net negative interactions are dominant.

c. Biocentric Tradition
Biocentric tradition adherents do not have a commodity ontology. Nor, do they explicitly
espOllse a systems perspectives which incorporates Type II (natural object) and Type III
(thing connectedness) ontology. Petulla's words illustrate:
Those who look at the natural environment from the biocentric point of view con-
centrate value in nature for and in itself, apart from human uses of it. They would
follow Aldo Leopold's land ethic, which assumes equal rights for humans and beings
of the natural world (Petulla 1980).
On the surface, at least, it appears that biocentrists are concerned primarily with natural
objects. For my purposes, I limit the natural objects to living creatures, plants and animals,
although Petulla's category includes a much broader array of abiotic and biotic objects.
Principals who have expressed support for this tradition include John Muir, Albert
Schweitzer, Rachel Carson, and forester Aldo Leopold. Muir devoted much of his energy to
wilderness conservation, but he also made important contributions toward development of
biocentrism through writings and organization of the Sierra Club (Fox 1981). Schweitzer is
well known for his call for reverence for all forms of life, the theme of a sermon delivered
February 16, 1919 in Strasbourg. Rachael Carson brought widespread attention to the fatal
effects on animal life of the widespread use of chemical pesticides. Forester Aldo Leopold
was sensitive to the dominance of economic considerations in land management questions
when he wrote:
One basic weakness in a conservation system based wholly on economic motives
is that most members of the land community have no economic [financial] value.
Wildflowers and songbirds are examples. Of the 22000 higher plants and animals
native to Wisconsin, it is doubtful whether more than 5% can be sold, fed, eaten,
or otherwise put to economic use. Yet these creatures are members of the biotic
community, and if (as I believe) its stability depends on its integrity, they are
184 Conservation Traditions

entitled to continuance .... When one of these non-economic categories is threatened,


and if we happen to love it, we invent subterfuges to give it economic importance ....
The evidence [of the importance of songbirds] had to be economic in order to
be valid.... It is painful to read these circumlocutions today. We have no land
ethic yet, but we have at least drawn nearer to the point of admitting that birds
should continue as a matter of biotic right, regardless of the presence or absence of
economic advantage to us .... Some species of trees have been read out of the party
by economics minded foresters because they grow too slowly, or have too Iowa sale
value to pay as timber crops: white cedar, tamarack, cypress, beech and hemlock
are examples (Leopold 1949).
A major problem that prevents the biocentric perspective from participating as an equal
with the ecologic and economic perspectives in forest management is the lack of scientific
constructs that summarize the perspective's most important aspects. Indeed, this tradition
has been linked with religious convictions. For example, Fox (1981) suggests that Leopold
was influenced by the writir,gs of a Russian philosopher, Peter Ouspensky. Muir, he sug-
gests, was greatly impressed by a book on Hinduism. Other biocentrists have also expressed
admiration for beliefs other than Judeo Christian traditions. Fox summarizes:
From these varied sources-panthEism, classical Greece, American Indians, Asian
religions, and Western interpreters of Asian view points-conservationists took their
philosophical affirmations that seemed absent from the more familiar Judeo Chris-
tian tradition. They may have still called themselves Christians. But their ideas
came from elsewhere (Fox 1981).
The economic tradition places heavy reliance on "efficiency," the ecologic tradition places
heavy reliance on "stability" and "interaction structure." What does the biocentric view
rely on? Perhaps the most helpful construct will turn out to be "extinction." Certainly, the
negation of extinction is "continuance." I show in the next chapter how extinction relates
to selected constructs from the ecologic tradition.

D. Summary
The three conservation traditions outlined by Petulla (1980) are interpreted in light of
problems of managing the biotic systems in forests. Joint production economics is suggested
as the most appropriate construct for relating multiple use forest management considerations
to the quest for economic efficiency. Ecosystem stability and community interaction structure
are suggested as constructs that can fill needs of ecologic tradition adherents. Extinction
(or absence of extinction) is suggested as a construct that expresses much of the concerns
of biocentric pioneers Muir, Schweitzer, Carson, and Leopold. A synthesis of beliefs into a
perspective that permits subscribers of each tradition to participate in multiple use decision-
making is the subject of the next, and final, chapter.

E. Research Problems
1. Trace the formulation of the Forest and Rangeland Renewable Resources Planning Act as
amended by the National Forest Management Act (Society of American Foresters 1976)
and identify the role (if any) played by persons from the three conservation traditions
in its wording and requirements of the u. S. Forest Service.
2. For several years an emblem identifying commodities (timber, range, wildlife, water,
recreation) had a prominent role in U. S. Forest Service buildings and literature. Trace
i!s history.
3. Research the split that developed between utilitarian conservationists (Pinchot) and
biocentric conservationists (Muir) in the early 20th century. Upon whom did each rely
for philosophical direction?
Conservation Traditions 185

4. In your view, are scientist's names appropriately placed in Figure 12.2? If not, why?
Add others to the group.
5. Study G. Hardin's writings on competition. Does he argue that competition is intrinsi-
cally good, or that it is inevitable, hence we must learn to live with it (Hardin 1980)?
6. Find examples of joint production relations between timber and a noncommodity animal
population, timber and forage, timber and water. Construct isocost relations if they do
not already exist.
7. Scrutinize A. Schweitzer's Strasbourg sermon of 1916. Has it any relevance in the 1980's
and beyond?
8. What has been the direction of joint production research because of the problems de-
scribed by Whaley (1970) and others?
9. Is the economic status of white cedar, tamarack, cypress, beech, or hemlock changed
since Leopold's observation?
10. Find examples of economic analyses of forest production alternatives where economic
values have been assigned to noncommodity populations of plants or animals.
13: Toward a Synthesis of
Conservation Traditions

There is an urgent need for greater dialogue between subscribers to the three traditions:
economic, ecologic, and biocentric. One problem preventing dialogue is the institutional
residency of the traditions. Residency is most clear cut for the biocentric tradition. It
is resident in so-called special interest groups - Friends of the Earth, National Wildlife
Federation, World Wildlife Fund and other conservation organizations. Economic efficiency
permeates forest industry and, it is often claimed, the u.s. Forest Service. To some extent
it also is dominant in forestry schools. The ecologic tradition resides primarily in university
forestry, wildlife, and ecology departments, and to some extent in research branches of public
agencies. The extremely large sums of money spent in litigation between, for example,
biocentric organizations and public economic efficiency organizations is testimony to the
need for more dialogue between them.
Another problem preventing dialogue is the apparent absence of constructs that link
subscribers to all three traditions. The economic and ecologic traditions reside in scientific
disciplines. In theory, since both are scientific, there should be found some linking construct
or constructs. The biocentric has for the most part resided outside the area of science.
Where? Leopold (1966) argues it resides in ethics, a branch of philosophy. Others claim
biocentrism reflects religious views, beyond both science and philosophy. If true, linking the
biocentric and ecologic traditions would be extra difficult because of the traditional schism
between science and religion. In this chapter I first attempt to link biocentric and ecologic
traditions by making biocentric more scientific, and then I attempt to link the resultant
construct with a construct from economics.

A. Biocentric - Ecologic
When Aldo Leopold wrote "Yet these creatures are members of the biotic community,
and if (as I believe) its stability depends on its integrity, they are entitled to a continuance"
(Leopold 1966), I think he intended continuance to mean to not become extinct. To become
extinct is at the end of a continuum of continuances. To not become extinct is the rest of the
continuum. In the Conservation Ethic, Leopold asks "Why do species become extinct?" His
suggestion is "[b]ecause they first become rare. [and] Why do they become rare? Because of
shrinkage in the particular environments which their particular adaptations enable them to
inhabit" (Leopold 1933). Viewing this schematically, I suggest the progression in Figure 13.1.
The large circle encloses physical space; the dots represent individuals of a species population.
Species population individuals may, in periods of abundance, be distributed fairly uniformly
in space (A). As habitat disruptions occur, the population becomes divided into several
geographically isolated subpopulations (B). If further pressure is exerted on a population
or its habitat, some local populations may be eliminated (C). Using the theory of indirect
interactions and the new exclusion principle from Chapters 9 and 10, I suggest that in
C the immanating and emanating effects exclude themselves and the local population is
188 Toward a Synthesis of Conservation Traditions

eliminated. As pressure on localized populations continues, there may eventually be only one
subpopulation remaining. Its elimination is also the population's extinction. So, extinction
of a species is simply the last exclusion of a population's immanating and emanating effects.

A B c

+ +

o , - I_ _ 1---10

Figure 13.1. Species become extinct because they first become localized in portions of their
former range (B). If habitat suitability continues to decline, some of the localized populations
will be eliminated becasue their immanating and emanating effects exclude each other (C).
The last exclusion of effects is extinction of the species.

B. Ecologic - Economic
A principal construct for organizing natural object population interaction information
IS the phase plane, quadrant I of the Cartesian coordinate system with time represented
parametrically. The phase plane data come from four time series graphs of standing crop
levels plotted against time. They are represented geometrically in the phase plane as two
trajectories, one for populations developing "together" and one for populations developing
"separately" (see Chapter 4, Figure 4.4).
Typically, the phase plane is used to represent either trajectories at several points in
time or a series of values for one time, such as equilibrium population levels. The phase plane
can also be used to represent periodic amounts of standing crop removed during a rotation,
again with time represented parametrically. Or, the phase plane could be used to plot only
the total cumulative amount of each population removed in a specified period of time, say
one rotation for even aged forest management.
The removals phase plane represents an intermediate step between an ecologic construct
used to show natural object population interaction and the economic construct used to
show commodity interdependence. Commodity interdependence can be shown with a slight
modification to the construct described above. The cumulative removals need simply be
"produced" for every cell in a grid superimposed in quadrant I in a given time period, e.g.,
one rotation. The costs must be determined for each combination (below). Except for some
type of conversion from standing crop units, e.g., basal area, to commodity units, e.g., cords
or cubic feet of merchantable stem volume, this figure is identical to the economist's joint
production framework used in multiple use decision-making.
Toward a Synthesis of Conservation Traditions 189

60

quaking
aspen
40
11 etc.
harvested
10 11
20 9 ~O
12
8 9 11 13
7 910 12 14
5 7 10 11 13 15
20 40 60
balsam fir harvested
Figure 13.2. Costs of producing amounts of harvested balsam fir and quaking aspen III one
rotation (schematic).

Thus, I propose that the ecologic and economic traditions can be partially integrated through
their constructs for ordering information about natural object population interactions (the
phase plane) and commodity production interdependence (the product-product graph).

C. Biocentric - Ecologic - Economic


The three traditions can be partially integrated because
1. extinction is the last exclusion of effects,
2. population elimination occurs at particular interaction types ((0, ~) and (~, 0)),
3. management actions needed to prevent development of (0, ~) and (~, 0) interactions
can be identified using Haskell's coordinate system,
4. a record of standing crop removals required to accomplish 3. can be represented in a
cumulative removals phase plane,
5. the costs associated with necessary removals (to prevent population elimination) and
other removals can be estimated, and
6. the prices of products removed completes the information necessary to do a classical
economic efficiency evaluation.

D. Synthesis Illustration
To illustrate the integration of biocentric, ecologic, and economic efficiency action crite-
ria, I consider the population of mixed aspen balsam fir stands in the Lake States (Michigan,
Wisconsin, Minnesota). Quaking aspen and its associates, bigtooth aspen and paper birch,
are pioneer species. They are intolerant of shade and among the first tree species to invade
clearcut or burned areas. Understories of shade tolerant northern hardwoods and northern
conifers develop in many aspen stands and will eventually replace the aspen overstory if no
actions are taken to interrupt succession. Succession can be interrupted by reproduction
cuts - typically clearcuts. Partial cutting prior to the final reproduction cut can influence
the relative amounts of the two species in stands. The relative amount of the two species
will in turn affect the type of habitat supplied by the stands. How to manage mixed aspen-
balsam fir stands is a complex question. I now consider some of the ecologic, biocentric, and
economic facts which affect a final management choice. I begin with some broad ecologic
facts.
190 Toward a Synthesis of Conservation Traditions

1. Ecologic.
Heinselman (1954), in a study of the likely extent of natural conversion to other species
of the Lake States aspen-birch type, found that 39 percent of the commercial forest land
in the Lake States was in the aspen birch type as of about 1950. He attempted to esti-
mate the fraction of these lands that might convert naturally to other forest types by the
gradual replacement of quaking aspen. A major conclusion of his study was that in 10 to 40
years about 32 percent (representing 5,800,000 acres) of the aspen-birch type would, without
human interference, completely convert to either northern hardwoods or spruce-fir (largely
balsam fir). The latter, he estimated, would cover approximately 1,430,000 acres.
An indication of the extent of tolerant conifer invasion of aspen forests is given by taking
stands typed as aspen and looking at the fraction of growing stock volume in tolerant conifers
compared to the fraction in quaking aspen. Fractions at two points in time help indicate if
stands given an aspen type designation are now, more than previously, stocked with tolerant
conifers. (It should be recognized that balsam fir percent of total volume should never be
greater than 50 percent, or the stand would by typed spruce-fir.) The table below gives
these relation ships for Michigan's western Upper Peninsula (Spencer 1982; Spencer and
Pfeifer 1966):

1966 1980
million cubic feet million cubic feet
balsam fir 24.99 54.98
% softwood volume 29.10 39.81
% total volume 4.17 7.25
quaking aspen 354.40 366.55
% hardwood volume 69.08 59.06
% total volume 59.17 48.31

The distribution of volume in the aspen forest type in the northeastern region of Minnesota
is as follows (Jakes and Raile 1980; Spencer 1983).

1962 1977
million cubic feet million cubic feet
balsam fir 92.61 163.17
% softwood volume 47.14 47.17
% total volume 6.06 8.92
quaking aspen 1,148.8 1,082.5
% hardwood volume 86.25 72.79
% total volume 75.17 59.17

In both regions of the Lake States there was an increase in the percentage of balsam fir
tree volume to the total volume in stands typed as aspen. More dramatic is the decrease in
the percentage of quaking aspen tree volume to the total volume in stands typed as aspen.
In Michigan the latter was only 48 percent and in Minnesota it was 59 percent, down from
75 percent fifteen years earlier. In both areas, however, aspen was the most prevalent of all
species. These converging trends will result in a significant change in the Lake States forest
area that is considered the aspen forest type.
Natural succession leads to the gradual replacement of aspen by balsam fir. Because
the financial value of balsam fir pulpwood can be as much as twice that of quaking aspen,
there has been considerable interest in speeding up the natural succession. In 1950, a study
was initiated to test several methods of cutting the overstory quaking aspen to release the
higher-valued understory balsam fir in northern Wisconsin (Cooley and Lord 1958)'. The
following was part of the study rationale: Removal of the overstory will be advantageous if it
Toward a Synthesis of Conservation Traditions 191

increases the growth and vigor of the understory trees, thus making them less susceptable to
decay and [spruce] budworm [Choristoneura fumiferana (Clem.)] attack. In 1975 the study
area was part of a natural spruce bud worm outbreak area. The cutting methods that gave
balsam fir the greatest release from over-topping aspen, e.g., commercial clear cut of aspen,
also had the greatest balsam fir mortality from budworm as estimated from a remeasurement
of research plots five years later (Popp, undated manuscript in preparation). Balsam fir in
stands that did not have the aspen overstory removed were little affected by the budworm.
These findings agree with those in a report by Batzer (1969).
It is not so clear whether the presence of a balsam fir understory in any way mitigates the
effects of aspen "pests" such as the forest tent caterpillar (Malacosoma disstria Hubner) or
hypoxylcin canker caused by Hypoxylon mammatum Wahl. Mill. . In a study of hypoxylon
incidence, Anderson and Martin (1981) were unable to claim a significant relationship
between risk of infection and percent of other species present in the stand. It should be
noted, however, that their study was not designed to detect this relationship, so there
may in fact be one. There may be some evidence that understory species such as balsam
fir reduce the likelihood of severe forest tent caterpillar outbreaks. Severe outbreaks are
generally limited to pure aspen stands. Again, no studies are known that focused directly
on the contribution of understory to risk reduction.

2. Biocentric
Quaking aspen will be steadily excluded from mixed stands in the Lake States as the
aspen standing crop is reduced below that required for an aspen motherstand. Perala (1977)
estimates 20 ft 2 per acre basal area is needed to get a full stand of root suckers once the
overstory is removed completely. When deficient motherstands are harvested, some aspen
reproduction may appear in the next stand, but it will likely be a small amount. An even
smaller component will likely appear after the next reproduction cut. The replacement of
quaking aspen in the quaking aspen forest type has so far not been taking place rapidly, just
steadily, in the Lake States.
Probably occuring more rapidly is the elimination of habitat for the many herbivorous
wildlife species that are found in stands with a high proportion of aspen. Two of the most
studied herbivores are ruffed grouse (Bonasa umbellus) and white-tailed deer (Odocoileus
virginian us) . Pure balsam fir stands provide a fairly hostile habitat for most birds and mam-
mals (Gullion 1983). Exceptions are two herbivores, spruce grouse (Canachites canadensis)
and moose (Alces alces) and some insectivorous birds. Pure fir stands also provide thermal
cover for big game animals, but lowland white-cedar stands probably provide better quality
thermal cover.
Little is known about the exact habitat requirements of forest bird species. It has
been suggested that black throated green warbler (Dendroica virens), blackurnian warbler
(Dendroica fusca), red breasted nuthatch (Sitta canadensis)' and purple finch (Carpodacus
purpureus) frequent mixed stands of aspen and balsam fir (Probst 1983). Large scale con-
version of mixed to pure fir stands would likely eliminate much suitable habitat for these
species and essentially exclude them from these areas.
Figure 13.3 gives a conceptual summary of the information presented in the ecologic
and biocentric discussions. The hatched bands along each axis indicate the minimum moth-
erstand basal areas recommended to fully reproduce each species at the final harvest cut
(Perala 1977; USDA Forest Service 1967). Thus, the standing crop coordinates of the stand
must be beyond (above and to the right) the hatched area immediately prior to harvest cut-
ting. This condition applies no matter what has gone on before in the way of intermediate
removals.
192 Toward a Synthesis of Conservation Traditions

100

better
habitat
quaking 80 for
aspen
standing reduced risk to balsam fir
crop
/ from spruce budworm (conjecture)
/ and reduced risk to quaking aspen
---....
from forest tent caterpillar and
60
""-X hypoxylon
higher/. \
risk from
hypoxylon
& forest tent \
40 caterpillar
(conjecture) I

inade~uate I better I
aspen balsav habitat for \
mothermother_ moose, I high
stand ....... spruce grouse, /risk from
o / - st'd in~deq.
o
: & insect. ~irds
40
.
spruce budworm
,---
20 60 80 100
balsam fir standing crop

Figure 13.3. Summary of ecologic and biocentric facts about population interactions and
resulting habitat of selected wildlife species, in mixed quaking aspen balsam fir stands.

3. Economic.
Records show that in 1982 prices received by the State of Minnesota for standing volume
(stumpage) of balsam fir sold from state lands averaged about $4.00 per cord (79 cubic feet
of solid wood), with a range of $3.20 to $5.60 per cord. At the same time quaking aspen
on state lands was selling for an average $3.07 per cord, with a range of $1.80 to $5.27 per
cord (Blyth 1983). Thus, more income is produced for the state from mixed stands having
a large balsam fir component.
Forest managers, of course, recognize the implication of the aspen-balsam fir price dif-
ference for the profitability of their operations. The cutting methods study described briefly
above was designed to identify the proper timing of mixed to pure balsam fir conversion cuts
in relation to aspen age. They also hoped to identify the best way to accomplish the conver-
sion "whether by a single clear-cutting lof aspen] or a series of partial cuttings ... " (Cooley
and Lord 1958). The economic efficiency of converting mixed to pure balsam fir stands is,
however, questionable because of the man-hours required for cutting, peeling, and swamping
compared to that required for pure aspen stands (Cooley and Lord 1958).
Murphy (1981) examined alternative rules to follow in selecting mixed stands to be
harvest cut in such a way that higher prices paid for balsam fir are exploited. He examined
the present value of the physical yields obtained by managing groups of mixed stands. Of
particular concern were the timing and value of a stream of hardwood (mostly aspen) and
softwood (mostly balsam fir) products that could be produced during the next 40 years on
Toward a Synthesis of Conservation Traditions 193

state lands in northeastern Minnesota. Tested were five different rules for selecting the order
in which mixed stands should be harvested. Briefly they are:
1. harvest the oldest first
2. harvest first the stands with largest amount of softwood volume
3. harvest last the stands with largest amount of softwood volume
4. harvest the oldest stands first, then harvest the stands with largest softwood volume
5. partition the stands into those having a conifer understory and those not having a
conifer understory. Then partition each group of stands into a high site group and a
low site index group. The first priority is to harvest high site aspen at maturity for
fear of losing natural regeneration of aspen. Second priority is to stands with a conifer
understory. Overmature stands with a conifer understory are not harvested for 30 years
to accumulate growth on softwoods (primarily balsam fir). Assumed is that overmature
aspen is of poor quality, hence of little financial value. It is sacrificed to increase the gain
in higher valued softwoods. The increased return from more softwood volume at harvest
time occurs at the risk of the aspen component being reduced to less than required for
a mother stand.
The value flow (in millions of dollars) from implementing these five stand selection rules on
state owned lands in northeastern Minnesota is given below (Murphy 1981):

Planning Period
Discounted values assuming
Rule II III Total 10 percent annual return
(millions of dollars)
86.02 107.18 136.29 329.49 91.68
2 76.91 115.91 134.48 327.30 87.94
3 76.45 110.26 142.19 328.90 87.01
4 82.19 109.74 136.36 328.29 89.92
5 90.98 97.12 128.81 316.91 91.66

Rule 5, which deferred cutting on already overmature aspen that had a conifer understory,
gave the second highest return. It is probably not significantly different than Rule 1, which
gave the highest return.
Summarizing, economic efficiency criteria for action suggest that mixed aspen-balsam
fir stands be managed to increase the relative proportion of the more valuable conifer species.
Included is the complete conversion from mixed to pure balsam fir stands.
The combined biocentric - ecologic criteria for action can be integrated with economic
criteria because Figure 13.3 can be combined with Figure 13.2. The result, Figure 13.4,
is a construct that represents the biocentric and ecologic situation in the "future" (to the
right and above point C) and the economic (commodity) situation to date in the "past" (to
the left and below point C). The schematic example in Figure 13.4 shows a twice thinned
mixed aspen-balsam fir stand that has current standing crop 40 and 40 ft 2 /acre for aspen
and balsam (point D). The "past" (lower left) section shows that two intermediate cuts have
been made. The first cut moved removals point C from A to point B (2 cords of balsam fir
removed). The second cut moved the removals point from point B to C (3 cords of aspen
removed).
The "future" is coordinated in normally used control units, standing crop basal area,
whereas the "past" is coordinated in commodity units, e.g., cords of pulpwood. Cost figures
are entered in italics in the "past" section of Figure 13.4. In the "future" section discrete
cost figures have been converted to isocost curves. The "future" section of Figure 13.4
moves around on these cost values as periodic intermediate cuts are made in the stand. All
numbers and curves are, of course, schematic. If no intermediate cuts were made, point C
would correspond with point A. There would be no "past" actions.
194 Toward a Synthesis of Conservation Traditions

The difference between coordinates of point C and those of point D gives the possible
removals of standing crop at the present time. Normally, point D should not be located in
either hatched area. The stand component with an inadequate mother stand, should a final
harvest cut be made, will be excluded. However, it is biocentrically permissible for point D
to be in the hatched regions if in the near future the stand will grow in such a way as to
move the stand coordinates out of the hatched area (to the right or upward).

18

16

14
the future
\
Quaking aspen Quaking aspen
harvested "- standing crop
volume, basal area, ft2 /ac .

60~ ".~~'00
cords/ac.

10

""- ' \ $80


$70
40
8 ~

4 etc.

80 balsam fir
8 standing crop
{

: '--:......._:---L_e_t_c.L._.....L._....L._...l.-_.L---.J.L-_Jl....--.JL..l...--I....L.---L_......L.. bas a I are a, ft 21 a c.


(

O~ 2b
s 4 6 8 10 12 balsam fir
the past ~ ~ harvested volume,
c _____~--~' cords/acre
~.

Figure 13.4. Integration of biocentric - ecologic framework (Figure 13.3) with the economic
efficiency framework (Figure 13.2) to form a commodity-based "past" and a biotic-based
"future." Text contains a description of the figure parts.

The final step to fully integrate the biocentric - ecologic framework with the joint
production economics framework is to add the isorevenue curve (assumed to be a straight
line). Recall, the price of balsam fir pulpwood is about 1.3 times that of quaking aspen.
The slope of the isorevenue line is given by the ratio of physical units required to yield
one dollar in revenue. Let one unit of aspen yield one dollar. It takes 1/l.3 = 76 units
Toward a Synthesis of Conservation Traditions 195

of balsam fir to produce one dollar of revenue. The slope of the triangle's hypotenuse in
Figure 13.4 corresponds to the required slope of the isorevenue curve. Since the isocost
curves are schematic, there is little point in looking for an exact point of tangency. What
appears clear, however, is that the most efficient combination of aspen and balsam fir to
produce now or in the future will be stands with mostly balsam fir, but never less than
20 ft 2 /acre of aspen at the final harvest cut. The latter will, in theory, insure that aspen
appears in the next stand. Special habitat needs of noncommodity populations can easily
be incorporated into this framework by altering the hatched bands in Figure 13.4.

E. Observations
Several important observations should be made about the foregoing discussion:
1. In contrast to other scientists expressing preferences for interactions, especially cooper-
ative ones, I express preference against two interaction types, (0, -) and (-,0).
2. This approach is more likely to provide the basis for a bio-ethical perspective rather
than favoring cooperation, which two tree species can never do.
3. Every thing (natural object population) does not need a financial value to participate in
multiple use decision-making. Their "interests" can be represented in hatched regions
of the biotic "future" (Figure 13.4).
4. Biological populations need only exist for them to participate in multiple use decision-
making.
5. The method of combining decision-making constructs suggested here allows economic
efficiency to be considered in joint production decisions.
6. Economic efficiency is practicable only after care has been taken to insure against elim-
ination of natural object populations.
7. Attention to elimination reverses the usual practice where economic efficiency comes
first in a "trickle down" approach to habitat management for all biotic populations.
Populations persist if, by chance, actions based on economic efficiency are not too dif-
ferent from those needed to prevent interaction types and intensities that lead to effect
exclusion and population elimination.
8. There has, I suspect, always been the option to restrict the production possibilities
space. The specific restrictions derived in a biocentric view are not a matter of whim.
They should be based on available scientific evidence of what actions are necessary to
prevent effect exclusion.
9. Zivnuska (1979) argues that recent research has shown, contrary to conventional wisdom,
that many technical relationships on which production possibilities curves are developed
are competitive, not complementary. He uses this finding to reaffirm his preference for
so-called dominant use. In my simple example, dominant use is production of all quaking
aspen or all balsam fir.
lD. Integrated biocentric, ecologic and economic decision-making constructs will likely result
in fewer cases of dominant or exclusive use.
11. Dominant use is not prohibited by the integrated approach developed here. It is permit-
ted when no actions are needed, or are prohibited, to prevent a population from being
eliminated from a community.
12. The suggested approach to decision-making does not call for total concern for the fate
of natural object populations. It limits its concern to that needed to prevent population
elimination.
13. Because properties of the tree overs tory populations largely determine habitat charac-
teristics for other plant and animal populations, overstory characteristics may proxy for
habitat of other biotic populations.
14. Physical locations with many biotic populations may require quite different hatching
patterns than presented here.
196 Toward a Synthesis of Conservation Traditions

15. The permissible production areas (unhatched regions of production possibilities space)
mayor may not overlap.
16. Non-overlapping permissible production areas means the populations, or those proxied
for, cannot co-exist in the same location.
To conclude, biocentric, ecologic and economic efficiency criteria can be integrated into a
holistic approach to multiple use decision making. The possibility of integration is enhanced
when decision-makers include a Type IV ontology in their way of looking at the world.

Research Problems
1. Estimate the litigation costs to the people of the United States over the past decade
due to commodity centered organizations being sued by biocentric organizations.
2. Respond to the argument that much of the litigation referred to in problem 1 is because
of a difference in human values in each group. And, that the difference in values can not
be resolved unless or until all parties share the same ontology. In short, human values
are ontology dependent.
3. Assemble historical maps showing the distribution of a plant or animal species currently
threatened or extinct. To what extent do the maps show spatial distribution changes
as postulated in Figure 13.P
4. Evaluate the following proposition: Linear programming for multiple use planning is an
example of Kaplan's law of the instrument.
5. Reconcile the differences between the need for convexness of the set of constraints
(needed to allow linear programming methods to provide a solution) and the pattern
of constraints needed to ensure against direct (0, -) or (-,0) interactions as shown in
Figure 13.4.
6. In a forest ecosystem of your choice, develop phase plane constraints for tree populations
that will proxy for noncommodity populations found in the forests.
7. Develop the mathematical methods necessary to locate the point of tangency between
isocost and isorevenue lines in Figure 13.4. Ensure that the point located is always to
the right and/or above the hatched area of the phase plane.
8. Apply the methodology developed in this chapter to a multiple use management question
in an ecosystem of your choice.
9. In your view, can one claim to be bioethical and limit one's concern to preventing the
extinction of plant and animal species populations? Or, must each species population
be made to "flourish"?
10. Make an argument that what foresters need to do is to tackle their management problems
with a commodity view of the future and a biotic view of the past.
Literature Cited

Ackoff, R. L. 1973. Science in the systems age. Operations Research 21(3): 661-671.
Anderson, A. E., D. E. Medin, and D. C. Bowden. 1972. Indices of carcass fat in a Colorado
mule deer population. Journal of Wildlife Management 36(2): 579-594.
Anderson, G. W. and M. P. Martin. 1981. Factors related to incidence of Hypoxylon cankers
in aspen and survival of cankered trees. Forest Science 27(3): 461-476.
Andrewartha, H. 1961. Introduction to the study of animal populations. University of
Chicago Press, Chicago. 281 p.
Andrews, D. H. 1972. Quantization as an integrative concept. Pages 69-136. In: Integrative
principles of modern thought. H. Margenau, editor.
Antonovics, J. and H. Ford. 1972. Criteria for the validation of the competitive exclusion
principle. Nature 237: 406-408.
Armstrong, R. A. and R. McGehee. 1980. Competitive exclusion. The American Naturalist
115(2): 151-170.
Assmann, E. 1970. The principles of forest yield study. Sabine H. Gardner. English transla-
tion. Pergamon Press, Oxford. 506 p.
Ausubel, D. P. 1966. Meaningful reception learning and the acquisition of concepts. Pages
157-175. In: Analysis of concept learning. H. J. Klausmeir and C. W. Harris, editors.
Academic Press, New York. 272 p.
Ausubel, D. P. 1968. Educational psychology: A cognitive view. Holt, Rinehart, and Win-
ston, New York.
Ausubel, D. P., J. D. Novak, and H. Hanesian. 1978. Educational psychology: a cognitive
VJew. (2nd edition). Holt, Rinehart and Winston, New York. 733 p.
Axelrod, R. and W. D. Hamilton. 1981. The evolution of cooperation. Science 211: 1390-
1396.
Ayala J. 1969. Experimental invalidation of the principle of competitive exclusion. Nature
224(5224): 1076-1079.
Ayala, F. J. 1970. Invalidation of principle of competitive exclusion defended. Nature
227(5253): 89-90.
Bakuzis, E. V. 1978. Personal communication.
Bakuzis, E. V. 1974. Foundations of forest ecosystems lecture and research notes. Chapter 1.
Scientific .method. College of Forestry, University of Minnesota. 137 p.
198 Literature Cited

Batzer, H. O. 1969. Forest character and vulnerability of balsam fir to spruce budworm in
Minnesota. Forest Science 15(1): 17-25.
Behan, R. W. 1967. The succotash syndrome, or multiple use: A heart-felt approach to forest
land management. Natural Resources Journal 7(4): 473-482.
Bella, 1. E. 1971. A new competition model for individual trees. Forest Science 17(3): 364-
372.
Bellman, R., H. Kagiwada and R. Kalaba. 1967. Quasilinearization and the estimation of
time lags. Mathematical Biosciences 1: 39-44.
Bentley, W. R. 1968. Essential concepts in the professional forestry curriculum. Journal of
Forestry 66: 402-403.
Bentley, W. R. 1979. Personal communication.
Berlinski, D. 1976. On systems analysis. MIT Press, Cambridge. 186 p.
Beveridge, W. 1. B. 1957. The art of scientific investigation. W. W. Norton and Co. Inc.,
New York. 373 p.
Bickford; C. A., F. S. Baker, and F. G. Wilson. 1957. Stocking, normality, and measurements
of stand density. Journal of Forestry 55(2): 99-104.
Birch, L. 1957. The meanings of competition. The American Naturalist 91: 5-18.
Blyth, J. 1983. Personal communication.
Borsodi, R. 1967. The definition of definition. Porter Sargent Publisher, Boston. 121 p.
Boyce, J. S. 1948. Forest Pathology. (2nd edition). McGraw-Hill Book Co., New York. 550
p.
Boyce, S. G. 1978. Theory for new directions in forest management. USDA Forest Service
Research Paper SE-193. 19 p.
Boyce, S. G. and N. D. Cost. 1978. Forest diversity- new concepts and applications. USDA
Forest Service Research Paper SE-194. 36 p.
Buchman, R. G. 1980. Personal communication.
Bunge, M. 1963. The myth of simplicity: problems of scientific philosophy. Prentice-Hall,
Englewood Cliffs, NJ. 239 p.
Bunge, M. 1967a. Scientific Research I: The search for system. Springer-Verlag, New York.
536 p.
Bunge, M. 1967b. Scientific Research II: The search for truth. Springer-Verlag, New York.
374 p.
Bunge, M. 1969a. The metaphysics, epistemology, and methodology of levels. Pages 17-28.
In: Hierarchical Structures. (Whyte, et al., editors)
Bunge, M. 1969b. Analogy, simulation and representation. Revue internationale de philo-
sophic 23: 16-33.
Bunge, M. 1973a. Philosophy of Physics. D. Reidel, Boston. 248 p.
Bunge, M. 1973b. Method, model and matter. D. Reidel, Boston. 196 p.
Bunge, M. 1974a. Semantics I: Sense and Reference. D. Reidel, Boston. 185 p.
Bunge, M. 1974b. Semantics II: Interpretation and Truth. D. Reidel, Boston. 210 p.
Bunge, M. 1977. Ontology I: The furniture of the world. D. Reidel, Boston. 352 p.
Bunge, M. 1979. Ontology II: A world of systems. D. Reidel, Boston. 314 p.
Bunnell,. F. 1978. Horn growth and population quality in Dall sheep. Journal of Wildlife
Management 42(4): 764-775.
Literature Cited 199

Carmean, W. H. 1975. Forest site quality evaluation in the United States In: Advances in
Agronomy 27: 209-269.
Carmean, W. H., F. B. Clark, R. D. Williams, and P. R. Hannah. 1976. Hardwoods planted
in old fields favored by prior tree cover. USDA Forest Service Research Paper NC-134,
16 p.
Cartwright, D. 1951. Foreward In: Field theory in social science: selected theoretical paper
by Kurt Lewin. D. Cartwright, editor. Harper and Row, New York. 346 p.
Cassidy, H. G. 1969. Knowledge, experince and action- an essay on education. Teachers
College Press, Columbia University, New York. 205 p.
Cassidy, H. G. 1972. Summary of theoretical issues: What generalization of Mendeleyev's
periodic table means. Pages 3-19. In: Full Circle-The Moral Force of Unified Science.
E. F. Haskell, editor. Gordon and Breach, New York. 256 p.
Chamberlin, T. C. 1897. The method of multiple working hypotheses. Journal of Geology
V(VIII): 837-848.
Christensen, L. C., J. T. Hahn, and R. A. Leary. 1979. Data base. Pages 16-21. In: USDA
1979.
Christiansen, F. B. and T. M. Fenchel. 1977. Theories of populations in biological commu-
nities. Springer-Verlag, New York. 144 p.
Ciriacy- Wantrup, S. V. 1938. Multiple optimum use of wild land under different economic
conditions. Journal of Forestry 36(7): 665-674.
Ciriacy- Wantrup, S. V. 1961. Multiple-use as a concept for water and range policy. Economic
analysis of multiple-use, the Arizona watershed program - a case study of multiple-use.
Comm. on Econ. of Water and Econ. of Range Resour. Develop., West. Agri. Econ.
Res. Counc. Rep. 9: 1-11.
Clark, J. P. 1971. The second derivative and population modeling. Ecology 52(4): 606-613.
Cohen, J. 1978. Food webs and niche space. Monographs in population biology No. 11.
Princeton University Press, Princeton, NJ. 189 p.
Cole, L. C. 1960. Competitive exclusion. Science 131: 348-349.
Colinvaux, P. 1973. Introduction to ecology. John Wiley & Sons, Inc., New York. 621 p.
Convery, F. J. 1979. The case for multiple use. Pages 44-48. In: Proceedings of the Society
of American Foresters. 320 p.
Cooley, J. H., and W. B. Lord. 1958. A study of aspen-balsam fir cutting methods in
northern Wisconsin: five-year results. Journal of Forestry 56(10): 731-736.
Copi,1. M. 1978. Introduction to logic. (5th edition). Macmillan Publishing Co., New York.
590 p.
Cunningham, W. J. 1954. A nonlinear differential-difference equation of growth. Proceedings
of the National Academy of Sciences (USA) 40: 708-713.
Czarnowski, M. S. 1964. Productive capacity of locality as a function of soil and climate
with particular reference to forest land. Louisiana State University Studies, Biological
Science Series No.5. 174 p.
Darlington, P. 1972. Competition, competitive repulsion and coexistence. National Academy
of Science Proceedings 69(11): 3151-3155.
Davies, J. T. 1973. The scientific approach. Academic Press, New York. 185 p.
Deo, N. 1969, 1971. An extensive English language bibliography on graph theory and its
applications. Technical Report 32-1413 and supplement 1 to Technical Report 32-
1413. Jet Propulsion Laboratory. Pasadena, CA.
200 Literature Cited

Dobben, W. H. van and R. H. Lowe-McConnell, editors. 1975. Unifying concepts in ecology.


Dr. W. Junk B. V. Publishers, The Hague. 302 p.
Donald, C. 1963. Competition among crop and pasture plants. Advances in Agronomy 15:
1~118.

Duerr, W. A. 1963. Economic guides to multiple-use policy. Pages 163~166. In: Proceedings
of Society of American Foresters. 183 p.
Duerr, W. A., D. E. Teeguarden, N. B. Christiansen, and S. Guttenberg, authors and editors.
1979. Forest resource management: decision-making principles and cases. W. B.
Saunders Co., Philadelphia, PA. 612 p.
Edwards, P., editor-in-chief. 1967. The encyclopedia of philosophy. Macmillan Publishing
Co., Inc. and The Free Press. New York. 4206 p.
Einstein, A. and L. Infeld. 1938. The evolution of physics. Simon and Schuster, New York.
302 p.
Elton, C. S. 1927. Animal ecology. MacMillan Press, New York. 209 p.
L'Engle, M. 1962. A wrinkle in time. Farrar, Straus, and Giroux, Inc., New York. 210 p.
Evans, R. M. 1938. Multiple-use forest management. Journal of Forestry 36(10): 1028-1034.
Fontanilla, C. 1978. Writing about writing (or Theory of types). Canopy 4(3).
Ford-Robertson, F. C., editor. 1971. Terminology of forest science, technology practice and
products (English-language version). Society of American Foresters, Washington, D.C.
349 p.
Foster, W. and J. Tate. 1966. The activities and coactions of animals at sapsucker trees.
The Living Bird 5: 87-113.
Fox, S. R. 1981. John Muir and his legacy: The American conservation movement. Little,
Brown and Co., Boston. 436 p.
Frazer, W. R. 1955. Some indications of unity among the sciences. Philosophy of Science 22:
135-139.
Frothingham, E. H. 1921. Classifying forest stands by height growth. Journal of Forestry 19:
374~381.

Funk, D. T., R. C. Schlesinger, and F. Ponder, Jr. 1979. Autumn-olive as a nurse plant for
black walnut. Botanical Gazette 140 (supplement): s110-s114.
Furnival, G. M. 1964. More on the elusive formula of best fit. Pages 201-207. In: Proceedings
of the Society of American Foresters.
Furnival, G. M. 1971. All possible regressions with less computation. Technometrics 13:
403~408.

Gause, G. F. 1935. Verifications experimentales de la theories mathematique de la iutte pour


la vie. Actualites Scientifiques et Industrielles #277. Hermann et Cie, Paris. 62 p.
Gause, G. F. 1964. The struggle for existence. Hafner, New York. 163 p. [Originally
published in 1934.1
Gause, G. F. 1970. Criticism of invalidation of principle of competitive exclusion. Nature
227(5253): 89.
Gause, G. F., N. P. Smaragdova, and A. A. Witt. 1936. Further studies of interaction
between predators and prey. The Journal of Animal Ecology 5(1): 1-18.
Geach, P. T. 1968. Reference and generality. (Emended edition). Cornell University Press,
Ithaca. 202 p.
Literature Cited 201

Gerrard, D. J. 1969. Competition quotient: a new measure of competition affecting individual


trees. Res. Bull. 20, Agric. Exp. Stn., Mich. State Univ. 32 p.
Gilpin, M. E. and K. E. Justice. 1972. Reinterpretation of the invalidation of the principle
of competitive exclusion. Nature 236: 273-274,299-301.
Gingerich, 0., editor. 1975. The nature of scientific discovery: A symposium commemorating
the 500th anniversary of the birth of Nicolaus Copernicus. Smithsonian Institution
Press, City of Washington. 210 p.
Goldman, S. 1971. The mechanics of individuality in nature. Foundations of Physics 1:
395-408.
Goldman, S. 1973. The mechanics of individuality in nature. II. Barriers, cells, and individ-
uality. Foundations of Physics 3: 203-228.
Goldman, S. 1980. A unified theory of biology and physics. Journal of Social and Biological
Structures 3: 331-360.
Graham, K. 1963. Concepts of forest entomology. Reinhold Publishing Corp., New York.
388 p.
Gregory, G. R. 1955. An economic approach to multiple use. Forest Science 1(1): 6-13.
Gregory, G. R. 1972. Forest Resource Economics. Ronald Press Co., New York. 548 p.
Grinnell, J. 1904. The origin and distribution of the chestnut-backed chickadee. Auk 21 (July):
364-378.
Grosenbaugh, L. R. 1958. The elusive formula of best fit: a comprehensive new machine
program. U.S. Forest Service, Southern Forest Expt. Sta. Occas. Paper 158. 8 p.
Grosenbaugh, L. R. 1970. Design of growth studies in mixed stands. Pages 57-62. In:
Workshop for research on growth of mixed hardwood stands. U.S. Forest Service,
Athens, Georgia.
Gullion, G. 1983. Personal communication.
Hagglund, B. 1981. Evaluation offorest site productivity. Forestry Abstracts 42(11): 515-527
Haldane, J. B. S. 1958. The theory of evolution before and after Bateson. Journal of Genetics
56: 11-27.
Hall, G. R. 1963. The myth and reality of multiple use forestry. Natural Resources Journal
3(2) 276-290.
Harary, F., R. Norman, and D. Cartwright. 1965. Structural models: An introduction to the
theory of directed graphs. John Wiley & Sons, New York. 415 p.
Hardin, G. 1956. Meaningless of the word protoplasm. Scientific Monthly 82: 112-120.
Hardin, G. 1960. The competitive exclusion principle. Science 131: 1292-1297.
Harper, J. 1961. Approaches to the study of plant competition. Page 1 In: Mechanisms
in Biological Competition (Symposium of the Soc. for Exp. BioI. (Great Britain)).
Number XV.
Harper, J. L. 1977. Population biology of plants. Academic Press, New York. 892 p.
Haskell, E. F. 1947. A natural classification of Societies. New York Academy of Sciences
Transactions, Series II 9(5): 186-196.
Haskell, E. F. 1949. A clarification of social science. Main Currents in Modern Thought 7(2):
45 51 (also, cover design).
Haskell, E. F. 1970. Assembly of the sciences into a single discipline. The Science Teacher,
supplement to 37(8): 8-15.
Haskell, E. F. 1972. Full circle. Gordon and Breach, New York. 256 p.
202 Literature Cited

Haskell, E. F. 1973. Personal communication.


Haskell, E. F. and H. G. Cassidy. 1960. Plain truth. (Privately published and distributed,
New Haven, Conn.) 136 p.
Haukioja, E. and P. Niemela. 1979. Birch leaves as a resource for herbivores: seasonal
occurrence of increased resistance in foliage after mechanical damage of adjacent leaves.
Oecologia 39: 151-159.
Hayakawa, S. I. 1972. Language in thought and action. (3rd edition). Harcourt, Brace and
Jovanovich, Inc. New York. 289 p.
Heady, E. O. 1952. Economics of agricultural production and resource use. Prentice-Hall,
Inc., New York. 850 p.
Heiberg, S. o. and D. P. White. 1956. A site evaluation concept. Journal of Forestry 54:
7-10.
Heimer, W. E. and A. C. Smith III. 1975. Ram horn growth and population quality: their
significance to Dall sheep management in Alaska. Alaska Department of Fish and
Game. Wildlife Technical Bulletin 5. 41 p.
Heinselman, M. L. 1954. The extent of natural conversion to other species in the Lake States
Aspen-Birch type. Journal of Forestry 52: 737-738.
Hempel, C. G. 1965. Aspects of scientific explanation and other essays in the philosophy of
science. The Free Press. Princeton, NJ. 504 p.
Hendee, J. C. 1984. Public opinion, and what foresters should do about it. Journal of
Forestry 82(6): 340-344.
Henry, C. 1980. Competitive interaction studied by the mathematical method of loop anal-
ysis. Com pates Rendus, Series D-Sciences Naturelles 290(12): 787-79l.
Hepting, G. H. 1971. Diseases of forest and shade trees of the United States. USDA Forest
Service, Agricultural Handbook No. 386.
Hepting, G. H. 1974. Death of the American chestnut. Journal of Forest History 18(3):
60-67.
Heyerdahl, T. 1958. Aku-aku: The secret of Easter Island. Rand McNally and Company.
384 p.
Holt, R. D. 1984. Spatial hetergeneity, indirect interactions, and coexistence of prey species.
The American Naturalist 124(3): 377- 406.
Honer, T. G. 1972. A height-density concept and measure. Canadian Journal of Forest
Research 2: 441-447.
Hospers, J. 1967. Introduction to philosophical analysis, (2nd edition). Prentice-Hall. En-
glewood Cliffs, NJ. 629 p.
Hull, D. L. 1980. Individuality and selection. Pages 311-322. In: Annual Review of Ecology
and Systematics, Volume 11.
Humphreys, D. 1978. Teaching at the concept level. The American Biology Teacher 40(11):
506-507.
Hutcherson, K. 1976. Endangered species: the law and the land. Journal of Forestry 74(1):
31-34.
Hutchinson, G. E. 1954. Theoretical notes on oscillatory populations. Journal of Wildlife
Management 18(1): 107-109.
Hutchinson, G. E. 1961. The paradox of the plankton. The American Naturalist 95: 137-145.
Hutchinson, G. E. 1978. An introduction to population ecology. Yale University Press, New
Haven. 260 p.
Literature Cited 203

Innis, G. S. 1972. The second derivative and population modeling: another view. Ecology
53(4): 720-724.
Innis, G. S., editor. 1978. Grassland Simulation Model. (Ecological Studies 26) Springer-
Verlag, New York. 298 p.
Jakes, P. J. and G. K. Raile. 1980. Timber resources of Minnesota's northern pine unit,
1977. USDA Forest Service, North Central Forest Experiment Station, St. Paul, MN
Resources Bulletin NC-44. 54 p.
Jantsch, E. 1972. Towards interdisciplinarity and transdisciplinarity in education and inno-
vation. Pages 97-121. In: Interdisciplinarity-Problems of teaching and research in
universities. Organization for Economic Cooperation and Development. 319 p.
Jeffers, D. S. 1943. Multiple use of wild lands in the Rocky Mountains and intermountain
region. Journal of Forestry 41: 627-632.
Jenny, H. 1958. Role of the plant factor in the pedogenic functions. Ecology 39(1): 5-16.
Jenny, H. 1961. Derivation of state factor equations of soils and ecosystems. Soil Science
Society of America Proceedings 25(5): 385-388.
Judson, H. F. 1980. The search for solutions. Holt, Rinehart and Winston, New York. 211
p.
Kaplan, A. 1964. The conduct of inquiry. Methodology for behavioral science. Chandler
Publishing Company, San Francisco. 428 p.
Keister, T. D. 1972. Predicting individual tree mortality in simulated southern pine planta-
tions. Forest Science 18(3): 213-217.
Klausmeir, H. J. and C. W. Harris, editors. 1966. Analyses of concept learning. Academic
Press, New York. 272 p.
Korford, C. B. 1958. Praire dogs, white faces and blue gramma. Journal of Wildlife Man-
agement Monograph Number 3. 78 p.
Kormondy, E. J. 1969. Concepts of ecology. Prentice-Hall, Inc. Englewood Cliffs. NJ. 209
p.
Kormondy, E. J., editor. 1965. Readings in ecology. Prentice-Hall, Englewood Cliffs. NJ.
219 p.
Korzybski, A. 1958. Science and sanity. (4th edition. First published in 1933.) Institute of
General Semantics. Lakeville, CT. 806 p.
Krajicek, J. E., and K. A. Brinkman. 1957. Crown development: an index of stand density.
USDA Forest Service., Cent. States For. Exp. Stn., No. 108. 2 p.
Krajicek, J. E., K. A. Brinkman, and S. F. Gingrich. 1961. Crown competition, a measure
of density. Forest Science 7(1): 35-42.
Kramer, E. E. 1970. The nature and growth of modern mathematics. Hawthorn Books, Inc.,
New York. 758 p.
Krebs, C. J. 1972. Ecology: the experimental analysis of distribution and abundance. Harper
and Row, New York. 694 p.
Kuhn, T. W. 1970. The structure of scientific revolutions. (2nd edition). University of
Chicago Press, Chicago. 210 p.
Lane, M. 1970. Introduction; I. The structuralist method; II. Structure and Structuralism.
Pages 11-39. In: Introduction to Structuralism. M. Lane editor. Basic Books, New
York. 456 p.
Laszlo, E. 1972. ·Integrative principles in art and science. Pages 362- 391. In: Integrative
principles of modern thought. New York. H. Margenau, editor.
204 Literature Cited

Laszlo, E., and H. Margenau. 1972. The emergence of integrative concepts in contemporary
science. Philosophy of Science 39: 252-259.
Lawlor, L. R. 1979. Direct and indirect effects of n-species competition. Oecologia 43: 355-
364.
Leak, W. B. 1972. Competitive exclusion in forest trees. Nature 236: 461-463.
Leary, R. A. 1970. System identification principles in studies of forest dynamics. USDA
Forest Service. North Central Forest Experiment Station RP-NC-45. 38 p.
Leary, R. A. 1972. Estimating coaction from experimental data. Paper presented at the First
International Conference on Unified Science, November 23-26, New York.
Leary, R. A. 1976. Interaction geometry: an ecological perspective. USDA Forest Service.
General Technical Report NC-22. North Cen. For. Exp. Stn. St. Paul, MN. 8 p.
Leary, R. A. 1979. Design. In: A generalized forest growth projection system for the Lake
States. USDA Forest Service. General Technical Report NC-49. North Cen. For.
Exp. Stn. St. Paul, MN. 96 p.
Leary, R. A. 1980. A design for survivor growth models. Pages 62-81. In: Forecasting
forest stand dynamics. Proceedings of the Workshop held at the School of Forestry,
Lakehead University, Thunder Bay, Ontario. K. M. Brown and F. R. Clarke, editors.
261 p.
Leary, R. A. 1985. A framework for assessing and rewarding a scientist's research productiv-
ity. Scientometrics 7(1-2): 29-38.
Leary, R. A. and K. E. Skog. 1972. A computational strategy for system identification in
ecology. Ecology 53(5): 969-973.
Lennartz, M. R. 1979. Multiple-use management-For some, a promise unfulfilled. Pages
97-106. In: Multiple-use management of Forest Resources. Proceedings of the Sym-
posium. Clemson, SC. Donald D. Hook and B. Allen Dunn, editors. 267 p.
Leopold, A. 1933. The conservation ethic. Journal of Forestry 31 (6): 634-643.
Leopold, A. 1966. Sand county almanac: with other essays on conservation from Round
River. Oxford University Press. 295 p.
Levin, S. A., editor. 1974. Ecosystem analysis and prediction. Proceedings of a conference
on ecosystems. Alta, UT. 337 p.
Levin, S. and D. Pimentel. 1981. Selection of intermediate rates of increase in parasite-host
systems. The American Naturalist 117(3): 308-315.
Levine, S. H. 1976. Competitive interactions in ecosystems. The American Naturalist 110:
903-910.
Levine, S. H. 1977. Exploitation interactions and structure of ecosystems. Journal of Theo-
retical Biology 69: 345-355.
Levins, R. 1966. The strategy of model building in population biology. The American
Scientist 54(9): 421-431.
Levins, R. 1968. Evolution in changing environments. Princeton University Press, Princeton,
NJ.120p.
Levins, R. 1975. Evolution in communities near equilibrium. Pages 16-50. In: Ecology and
evolution of communities. M.L. Cody and J.M. Diamond, editors. Belknap Press of
Harvard University Press, Cambridge, MA. 545 p.
Lewin, R. A. and J. T. Enright. 1978. Browsers and their browse. Nature 276(5683): 10.
Lewis, R. W. 1980. Evolution: a system of theories. Perspectives in Biology and Medicine
23(4) 551-572.
Literature Cited 205

Lidicker, W. Z. 1979. A clarification of interactions in ecological systems. BioScience 29(8):


475-477.
Lin, J. Y. 1969. Growing space index and stand simulation of young western hemlock III

Oregon. (Unpublished Ph.D. thesis. Duke Univ.) 182 p.


Lindeman, R. 1942. The trophic-dynamic aspect of ecology. Ecology 23(4): 399-428.
Lloyd, R. D. 1969. Economics of multiple-use. Pages 45-54. In: Proceedings Conf. on
Multiple-use of Southern forests. Pine Mountain, GA, November, 1969.
Lotka, A. J. 1956. Elements of mathematical biology. (Dover reproduction of a 1928 printing.)
Dover Publications, New York. 465 p.
Lubcheno, J. 1978. Plant species diversity in a marine intertidual community: importance
of herbivore food preference and algal competitive abilities. The American Naturalist
112(983): 23-39.
Lubcheno, J. 1979. Comsumer terms and concepts. The American Naturalist 113(2): 315-
317.
Lyons, J. 1977. Semantics: Vol. I. Cambridge University Press, Cambridge. MA. 371 p.
MacLean, C. 1979. Relative density: the key to stocking assessment in regional analysis.
USDA Forest Service, General Technical Report PNW-78. 5 p.
MacMahon, J. A., D. L. Phillips, J. V. Robinson, and D. J. Schimpf. 1978. Levels of biological
organization: an organism-centered approach. Bioscience 28(11): 700-704.
MacMahon, J. A., D. J. Schimpf, D. C. Anderson, K. G. Smith, and R. L. Bayn, Jr. 1981.
An organism-centered approach to some community and ecosystem concepts. Journal
of Theoretical Biology 88: 287-307.
Malcolm, W. M. 1966. Biological interactions. Botanical Rev. 32(3): 243-254.
Margalef, R. 1963. On certain unifying principles in ecology. The American Naturalist 97:
357-374.
Margenau, H. 1944. The exclusion principle and its philosophical importance. Philosophy of
Science 11(4): 187-208.
Margenau, H. 1950. The nature of physical reality. McGraw-Hill, New York. 479 p.
Margenau, H., editor. 1972. Integrative principles of modern thought. Gordon and Breach.
New York. 522 p.
Margenau, H. 1972. Some integrative principles of modern physics. Pages 45-68. In: Inte-
grative principles of modern thought. H. Margenau, editor. Gordon and Breach, New
York. 522 p.
Martin, P. L. 1969. Conflict resolution through the multiple-use concept in Forest Service
decision making. Natural Resources Journal 9: 228-236.
Maruyama, M. 1966. Metaorganization of information. General Systems Yearbook 11: 55-
60.
Maslow, A. H. 1966. The psychology of science: A reconnaisance. Regnery-Gateway, Chicago.
168 p.
Mates, B. 1972. Elementary logic. Oxford University Press, New York. 237 p.
Mattson, W. J. and N. D. Addy. 1975. Phytophagous insects as regulators of forest primary
production. Science 190{ 4214): 515-522.
Mattson, W. J. 1980. Personal communication.
May, R. M. 1972. Will a large complex system be stable? Nature 238: 413-414.
206 Literature Cited

May, R. M. 1973. Stability and complexity in model ecosystems. Princeton University Press.
Princeton, NJ. 235 p.
May, R. M. 1976. Simple mathematical models with complicated dynamics. Nature 261:
459-467.
May, R. M., editor. 1976. Theoretical ecology: principles and applications. W.B. Saunders,
Philadelphia, PA. 317 p.
Maynard Smith, J. 1979. Game theory and the evolution of behavior. Proceedings of the
Royal Society of London, Series B: 205: 475-488.
McArdle, R. E. 1953. Multiple use-multiple benefits. Journal of Forestry 51(5): 323-325.
McClure, J. P., N. D. Cost, and H. A. Knight. 1979. Multiresource Inventories-a new concept
for forest survey. USDA Forest Service Research Paper SE-191. 68 p.
Mcintosh, R. P. 1980. The relationship between succession and the recovery process in
ecosystems. Pages 11-62. In: The recovery process in damaged ecosystems. J. Cairns,
Jr., editor. Ann Arbor Science Publishers, Inc., Ann Arbor, MI.
Mcquilkin, R. A. 1975. Errors in site index determination caused by tree age variation in
even-aged oak stands. USDA Forest Service Research Note NC-185. 4 p.
Mech. 1. D. 1966. The wolves of Isle Royale. U. S. National Park Service Fauna Series No.
7.210p.
Mech, L. D. 1977. Wolf-pack buffer zones as prey reservoirs. Science 198: 320-321.
Mech, L. D. 1984. Personal communication.
Medawar, P. B. 1940. The growth, growth energy, and ageing of the chicken's heart. Proc.
Royal Society. London. Series B, 129: 332-355.
Meyer, A. H. 1952. Forest Mensuration. Penns Valley Publishers, Inc., State College, PA.
357 p.
Miller, R. E., and M. D. Murray. 1978. The effects of red alder on the growth of Douglas
fir. Pages 283-306. In: Utilization and management of alder. Proc. Symposium held
at Ocean Shores, WA, April 25-27, 1977. David Briggs, Dean Debell, and William A.
Atkinson, compilers.
Milne, A. 1961. Definition of competition among animals. In: Mechanisms in Biological
Competition. Symposia of the Society for Experimental Biology. No. XV. Academic
Press, New York.
Muhlenburg, N. 1964. A method for approximating forest multiple-use optima. Forest Science
10(2): 209-212.
Muir, J. W. 1962. The general principles of classification with reference to soils. Journal of
Soil Science 13(1): 22-30.
Muir, J. W. 1969. A natural system of soil classification. Journal of Soil Science 20(1):
153-166.
Murphy, D. L. 1981. A decision support system approach to timber harvest scheduling.
(Ph.D. dissertation, University of Minnesota) 188 p.
Nahikian, H. M. 1964. A modern algebra for biologists. University of Chicago Press, Chicago.
236 p.
Nelson, M. E. and L. D. Mech 1981. Deer social organization and wolf predation in north-
eastern Minnesota. Wildlife Monograph No. 77. Journal of Wildlife Management
45(3): supplement. 53 p.
Nelson, T. C. 1964. Growth models for stands of mixed species composition. Pages 229-231.
In: Proceedings of the Society of American Foresters. 247 p.
Literature Cited 207

;'Iiewnham, R. M. 1964. The development of a stand model for Douglas fir. (Unpublished
Ph.D. Thesis, University of British Columbia, Vancouver) 201 p.
;'Iiicolis, G. and I. Prigogine. 1977. Self-organization in nonequilibrium systems. John Wiley
& Sons, New York. 491 p.
Niven, B. S. 1980. The formal definition of the environment of an animal. Australian Journal
of Ecology 5: 37-46.
Norreklit, L. 1973. Concepts: their nature and significance for metaphysics and epistomology.
Odense University Studies in Philosophy. Vol. 2. Odense University Press. 226 p.
Novak, J. D. 1966. The role of concepts in science teaching. Pages 239-254. In: Analyses of
concept learning. H. J. Klausmeir and C. W. Harris, editors. Academic Press, New
York 272 p.
Novak, J. D. 1977. A theory of education. Cornell University Press, Ithaca, NY. 295 p.
Novak, J. D. 1979. The application of psychology and philosophy to the improvement of
biology teaching. The American Biology Teacher 41(8): 466-470,474.
Novak, J. D. 1980. Learning theory applied to the biology classroom. The American Biology
Teacher 42(5): 280-285.
Odum, E. P. 1953. Fundamentals of ecology. W. B. Saunders Company, Philadelpha, PA.
384 p.
Odum, E. P. 1969. The strategy of ecosystem development. Science 164: 262-270.
Odum, E. P. 1971. Fundamentals of ecology. (3rd edition). W. B. Saunders Company,
Philadelphia, PA. 574 p.
Orenstein, A. 1977. Willard Van Orman Quine. Twayne Publishers, Boston. 180 p.
Opie, J. E. 1968. Predictability of individual tree growth using various definitions of com-
peting basal area. Forest Science 14(3): 314-323.
Overton, W. S. 1974. Decomposability: a unifying concept? In: Ecosystem analysis and
prediction. S. Levin, editor. Proceedings of a conference on ecosystems. Alta, UT,
July 1-5,1974. Sponsored by SIAM and NSF. 337 p.
Packard, J. M. and L. D. Mech. 1980. Population regulation in wolves. Pages 135-150. In:
Biosocial Mechanisms of Population Regulation. M. Cohn, R. Malpass, and H. Klein,
editors. Yale University Press, New Haven, CT.
Parkinson, G. H. R. 1973. Leibniz Philosophical Writings. G. H. R. Parkinson, editor; Mary
Morris and G. Parkinson, translators. J. M. Dent and Sons Ltd., London. 270 p.
Pattee, H. H.,editor. 1973. Hierarchy theory: the challenge of complex systems. George
Braziller, New York. 156 p.
Patten, B. C. 1979. Systems approach to the concept of environment. General Systems
Yearbook XXIV: 255-27l.
Patten, B. C. 1982. Environs: Relativistic elementary particles for ecology. The American
Naturalist 119(2): 179-219.
Pearse, P. H. 1969. Toward a theory of multiple use: the case of recreation versus agriculture.
Natural Resources Journal 9: 561-575.
Pearson, C. A. 1944. Multiple use forestry. Journal of Forestry 42(4): 243-249.
Pelz, D. C. and F. M. Andrews. 1966. Scientists in organizations. John Wiley & Sons, Inc.,
New York. 318 p.
Perala, D. A. 1977. Manager's handbook for aspen in the north central states. USDA Forest
Service, General Technical Report NC-36. 30 p.
208 Literature Cited

Peterson, R. 1977. Wolf ecology and prey relationships on Isle Royale. National Park Service
Scientific Monograph Series No. 11. 210 p.
Peterson, R. and R. E. Page. 1983. Wolf-moose fluctuations at Isle Royale National Park,
Michigan, USA. Acta Zoologica Fennica 174: 251-253.
Petulla, J. M. 1980. American environmentalism: values, tactics, priorities. Texas A and M
University Press, College Station, TX. 239 p.
Piaget, J. 1970a. Structuralism. Chaninah Maschler, editor and translator. Basic Books,
Inc. New York. 153 p.
Piaget, J. 1970b. General problems of interdisciplinary research and common mechanisms.
Pages 467-528. In: Main Trends of Research in the Social and Human Sciences. (Part
I-Social Sciences). Mouton/UNESCO, Paris/The Hague. 819 p.
Pianka, E. R. 1974. Evolutionary ecology. Harper and Row. New York. 356 p.
Pickett, S. T. A. 1980. Non-equilibrium coexistence of plants. Torrey Botanical Club Bulletin
107(2): 238-248.
Pielou, E. C. 1969. An introduction to mathematical ecology. Wiley-Intersceince, New York.
286 p.
Pienaar, L. V. and K. J. Turnbull. 1973. The Chapman-Richards generalization of von
Bertalanffy's growth model for basal area growth and yield in evenaged stands. Forest
Science 19: 2-22.
Pimentel, D. 1968. Population regulation and genetic feedback. Science 159: 1432-1437.
Pimentel, D. and R. Al-Hafidh. 1965. Ecological control of a parasite population by genetic
evolution in the parasite-host system. Annals of the Entomological Society of America
58: 1-6.
Pimentel, D. and F. A. Stone. 1968. Evolution and population ecology of parasite-host
systems. Canadian Entomologist 100: 655-662.
Pinchot, G. 1917. The training of a forester, (3rd edition). J. B. Lippencott Co., Philadel-
phia,PA. 157 p.
Platt, J. R. 1964. Strong inference. Science 146(3642): 347-353.
Popp, M. (undated manuscript draft) Effects of spruce budworm damage on a balsam fir
conversion cutting. Draft of final report for research studies FS-NC-2204 (81-01) and
FS-NC-ll03 (81-02). USDA Forest Service, North Central For. Exp. Stn., St. Paul,
MN.
Popper, K. 1963. Conjectures and refutations: The growth of scientific knowledge. Harper
and Row, Publishers. New York. 417 p.
Popper, K. 1965. The logic of scientific discovery. Harper and Row, Publishers. New York.
480 p.
Prigogine, I. and I. Stengers. 1984. Order out of chaos: Man's new dialogue with nature.
Bantam Books, Inc. Toronto. 349 p.
Probst, J. 1983. Personal communication.
Quine, W. V. 1960. Word and object. MIT and John Wiley & Sons, Inc., New York. 294 p.
Quine, W. V. 1972. Levels of Abstraction. Paper presented at the First International Con-
ference on Unified Science, November 23-26, New York.
Quine, W. V. 1975. Truth by Convention. (First published in 1936). Pages 77-106. In:
The ways of paradox and other essays, by W. V. Quine. Harvard University Press,
Cambridge. 335 p.
Rapaport, A. 1978. Various meanings of "theory." General System XXIII: 29-37.
Literature Cited 209

Rashevsky, N. 1954. Topology and Life: In search of general mathematical principles in


biology and sociology. Bulletin of Mathematical Biophysics 16(4): 317-348.
Rashevsky, N. 1960. Mathematical Biophysics: physio-mathematical foundations biology.
Volume 1,488 p., Volume 2, 462 p. Dover Publications, New York.
Rescher, N. 1962. The revolt against process. The Journal of Philosophy. LIX(15): 410-417.
Rescher, N. 1964. Introduction to logic. St. Martin's Press, New York. 360 p.
Rhyne, R. F. 1972. Communicating holistic insights. Fields within Fields within ... Fields
5(1) 93-104.
Ricklefs, R. E. Ecology. 1973. Chiran Press, Newton, MA. 861 p.
Risch, S. and D. Boucher. 1976. What ecologists look for. Ecoforum In: Bulletin of the
Ecolog. Soc. Amer. 57(3): 8--9.
Roberts, A. 1974. The stability of a feasible random ecosystem. Nature 251(5476): 607-608.
Robinson, M. H. 1981. A stick is a stick and not worth eating: on the definition of mimicry.
Biological Journal of the Linnean Society 16(1): 15-20.
Rogers, L. L., L. D. Mech, D. K. Dawson, J. M. Peek, and M. Korb. 1980. Deer distribution in
relation to wolf pack territory edges. Journal of Wildlife Management 44(1): 253-258.
Room, P. M. 1972. The fauna of the mistletoe Tapinanthus bangivensis (Engl. and K.
Krasue) growing on cocoa in Ghana: relationships between fauna and mistletoe. J.
Animal Ecology 41: 611-621.
Rosen, R. 1958. The representation of biological systems from the standpoint of the theory
of categories. Bulletin of Mathematical Biophysics 20: 317-341.
Rosenthal, G. A. and D. H. Janzen, editors. 1979. Herbivores-Their interaction with sec-
ondary plant metabolites. Academic Press, New York. 718 p.
Roth, F. 1916. Concerning site. Forestry Quarterly 15: 3-13.
Runes, D. D., editor. 1977. Dictionary of philosophy. Littlefield, Adams and Co., Totawa,
NJ.343p.
Saki, K. 1961. Competitive ability in plants: its inheritance and some related problems.
Pages 245-263. In: Mechanisms in Biological Competition. Symposia of the Society
for Experimental Biology. No. XV. Academic Press, New York.
Schoener, T. W. 1974. Resource partitioning in ecological communities. Science 185(4145):
27 -39.
Schweitzer, D., T. Hoekstra and C. Cushwa. 1981. Lessons from past national assessments of
wildlife and fish: information and coordination needs for the future. In: Transactions
of the 46th North American Wildlife and Natural Resources Conference.
Schwemmler, W. 1980. The triality principle as a possible cause of the periodicty of evolving
systems. Acta Biotheoretica 29( 198): 75 86.
Seal, U. S. 1977. Assessment of habitat conditions by measurement of biochemical and
endocrine indicators of the nutritional, reproductive and disease status of free-ranging
animal populations. Pages 305-329. In: Classification, inventory and analysis of fish
and wildlife habitat. Proceedings of a national symposium, Phoenix, AZ. 604 p.
Shigo, A. L. 1979. Tree decay-an expanded concept. USDA Forest Service, Agricultural
Information Bulletin Number 419. 73 p.
Shive, J. N., and R. L. Weber. 1982. Similarities in physics. John Wiley & Sons, New York.
277 p.
Simon, H. A. 1969. The sciences of the artificial. Massachusetts Institute of Technology
Press, Cambridge. 123 p.
210 Literature Cited

Slobodkin, L. B. 1974. Prudent predation does not require group selection. The American
Naturalist 108(963): 665-678.
Smith, J. H. G. 1963. Analysis of crown development can establish biological and economic
limits to growth of trees and stands. The Commonwealth For. Rev. 42(1): 27-33.
Snyder, T. A., Jr. 1979. User prospectives. Pages 20-26 In: Multiple management of forest
resources. Proceedings of a symposium. Clemson, SC. Donald B. Hook and B. Allen
Dunn, editors. 267 p.
Sober, E. 1975. Simplicity. Clarendon Press, Oxford. 189 p.
Society of American Foresters. 1976. The resources planning act as amended by the national
forest management act. Journal of Forestry 74(12).
Sparhawk, W. N., et al. 1923. Classification of forest sites. Journal of Forestry 21: 139-147.
Spencer, J. S. Jr. 1982. Timber resources of Michigan's Western Upper Peninsula Unit, 1980.
USDA Forest Service, North Central For. Exp. Stn., St. Paul, MN. 102 p.
Spencer J. S. Jr. 1983. Personal communication of timber resources reports published by the
Office of Iron Range Resources and Rehabilitation, St. Paul, MN
Spencer, J. S., Jr. and R. Pfeifer. 1966. The growing timber resources of Michigan 1966:
Unit 2-The Western Upper Peninsula. Michigan Department of Natural Resources.
Lansing, MI.
Spurr, S. H. and B. V. Barnes. 1980. Forest Ecology, (3rd edition). John Wiley & Sons, New
York. 687 p.
Staebler, G. 1951. Growth and spacing in an even-aged stand of Douglas-fir. Master of
forestry thesis submitted to the School of Natural Resources, University of Michigan.
54 p.
Stage, A. R. 1969. A growth definition of stocking: units, sampling, and interpretation.
Forest Science 15(3): 255-265.
Stagner, H. 1960. A second look at multiple use. American Forests 66: 24-25.
Starr, W. A. 1961. Multiple land use management. Natural Resources Journal 1: 288-301.
Stern, K. 1969. Einige Geitrage genetischer Forschung zum Problem de Konkurrenz in
Pflanzenbestanden. AUg. Forst- u. J. Ztg. 140: 253- 262.
Stern, K. and L. Roche. 1974. Genetics of forest ecosystems. Springer-Verlag, New York.
330 p.
Stewart, J., J. van Kirk, and R. Powell. 1979. Concept maps-a tool for use in biology
teaching. The American Biology Teacher 41(3): 171-176.
Sukachev, V. and N. Dylis. 1964. Fundamentals of forest biogeocoenology. J. M. Maclennan,
translator. Oliver and Boyd, Edinburgh and London. 672 p.
Swindel, B. F. 1970. On theoretical bases for mixed stand growth models. Pages 44-54.
In: Workshop for research on growth of mixed hardwood stands. Athens, GA, March
10-11. USDA Forest Service.
Tansley, A. G. 1935. The use and abuse of vegetational concepts and terms. Ecology 16(3):
284-307.
Taylor, A. M. 1972. Integrative principles in human society. Pages 211-289. In: Integrative
principles of modern thought. H. Margenau, editor.
Teeguarden, D. E. 1979a. Multiple services. Chapter 21, Pages 276-290. In: Duerr, et al.
Teeguarden, D. E. 1979b. The forest as a system. Chapter 14, Pages 159-172. In: Duerr, et
aL
Literature Cited 211

Thompson, J. R. 1982. Interaction and coevolution. John Wiley & Sons, New York. 179 p.
Thornton, R. M. 1972. Integrative principles of biology. Pages 137-210. In: Integrative
principles of modern thought. H. Margenau, editor.
Tilman, D. 1982. Resource competition and community structure. Monographs in Popultion
Biology Number 17. Princeton University Press, Princeton. 296 p.
Tourney, J. W. 1924. Foundations of silviculture upon an ecological basis. Part I-site factors.
Edward Brothers Publishers. Ann Arbor, MI. 171 p.
Tourney, J. W. and C. F. Korstian. 1947. Foundations of silviculture upon an ecological
basis. (2nd edition revised). John Wiley & Sons, New York. 468 p.
Trivers, R. L. 1971. The evolution of reciprocal altruism. Quarterly Review of Biology 46( 1):
35-57.
Thrnbull, K. J. 1963. Population dynamics in mixed forest stands: a system of mathematical
models of mixed stand growth and structure. Ph.D. thesis. University of Washington.
196 p.
USDA Forest Service. 1967. Silvicultural practices handbook-Chapter 100-Spruce-fir forest
type. USDA Forest Service Handbook 2409.26, R-9 Ammendment Number 2.
USDA Forest Service. 1977. An assessment of the forest and range land situation in the
United States (Review Draft).
USDA Forest Service. 1979. A generalized forest growth projection system applied to the
Lake States region. USDA Forest Service, North Cen. For. Exp. Stn., General
Technical Report NC-49. 96 p.
Utida, S. 1957. Cyclic fluctuations of population density intrinsic to the host-parasite system.
Ecology 38(3): 442-449.
Vandermeer, J. 1980. Indirect mutualism: variations on a theme by Stephen Levine. The
American Naturalist 116(3): 441-448.
Vincent, A. B. 1961. Is height/age a reliable index of site? Forestry Chronicle 37: l44-150.
Volterra, V. 1931. Lecons sur la theorie mathematique de la lutte pour la vie. Paris. 214 p.
Waugh, F. A. 1936. Reconciliation of land uses. Journal of Land and Public Utility Eco-
nomics. February pages 87-89.
Weiss, P. A., editor. 1971. Hierarchically organized systems in theory and practice. Hafner
Publishing Co., New York. 263 p.
Whaley, R. S. 1970. Multiple use decisionmaking-where do we go from here? Natural
Resources Journal 10: 557-565.
Whyte, 1. L., A. G. Wilson, and D. Wilson, editors. 1969. Hierarchical Structures. American
Elsevier Publishing Co., Inc. New York. 322 p.
Wicken, J. S. 1984. Autocatalytic cycling and self-organization in the ecology of evolution.
Nature and System 6(1984): 119-135.
Wickler, W. 1968. Mimicry in plants and animals. Translated from German by R. D. Martin.
World University Library. Weidenfield and Nicolson, London. 255 p.
Williamson, M. H. 1957. An elementary theory of interspecific competition. Nature 180:
422-425.
Williamson, M. H. 1972. The analysis of biological populations. Edward Arnold, London.
180 p.
Wilson, D. S. 1975, New model for group selection. (A book review of Group selection in
predator-prey communities by Michael E. Gilpin (1975)). Science 189(4206): 870.
212 Literature Cited

Wilson, D. S. 1980. The natural selection of populations and communities. The Ben-
jamin/Cummings Publishing Co., Inc. Menlo Park, CA. 186 p.
Wilson, E. B., Jr. 1952. An introduction to scientific research. McGraw-Hill Book Co. Inc.,
New York. 373 p.
Wilson, E. O. 1975. Sociobiology: the new synthesis. Harvard University Press, Cambridge,
MA. 697 p.
Wirth, M. 1976. Algorithms + Data Structures = Programs. Prentice-Hall, Inc. Englewood
Cliffs, NJ. 366 p.
Wise, T., W. Robinson, R. Hook, and 1. D. Mech. 1975. An experimental translocation of
the eastern timber wolf. Audubon Conservation Report No.5. National Audubon
Society, New York. 28 p.
Wit, C. T. de. 1960. On competition. Institute for Biological and Chemical Research on
Field Crops and Herbage. Vers!' Landbouwk. Onderz. 66.8 Wageningen. 82 p.
Woldenberg, M. J. 1979. A periodic table of spatial hierarchies. Pages 429-456. In: Philos-
ophy in Geography. S. Gale and G. Olsson, editors. D. Reidel, Boston, MA.
Wynne-Edwards, V. S. 1962. Animal dispersion in relation to social behavior. Oliver and
Boyd, Edinburgh. 653 p.
Ziman, J. 1978. Reliable knowledge: An exploration of the grounds for belief in science.
Cambridge University Press, Cambridge. 197 p.
Zivnuska, J. A. 1961. The multiple problems of multiple use. Journal of Forestry 59(8):
555-560.
Zivnuska, J. A. 1979. The case for dominant use. Pages 49-53. In: Proceedings of the Society
of American Foresters. 320 p.
Zon, R. 1913. Quality classes and forest types. Pages 100-104. In: Proceedings of the Society
of American Foresters.
Subject Index
A
accretion 53
action and reaction 33
active site of a concept 4, 122-126
analysis 4, 69-107
animal producing power of forests 54
anthropocentrism 1, 179
anticipation and retrospection 91
aspen-forest tent caterpillar interaction 88
autocatalysis 158
avoidance mechanisms 148
axiomatics 42
axioms 56, 143
B
bifurcation 40, 159
boundary conditions 39, 74, 100
C
Cartesian product 21
change
induced 4, 32, 52, 55, 60, 66
new theory of 44
spontaneous 4, 31, 32, 51, 55-57, 60, 66
classification III
coaction
cardioid 126
geometrization of cross-tabulation 119
compass 58
cross-tabulation 43, 57, 58, 67, 68, 116
theory 40
coefficient of struggle 75
coevolution in parasite-host systems 149-152
collusive competitors 165, 171
commodity ontology 179
community
interaction structure 5, 136, 137
as a concept 137, 138
stability 182, 184
competition 72-78
competitive ability 5, 169
competitive influence 5, 169
(see effects: immanating and emanating)
214 Subject Index

concept
as a kind of construct 12
dangling 121, 122
maps 4, 132-133
metaphysical requirements 121
logical structure of 13
relation to predicate 13, 14
set of a disipline 45, 56
and synthesis 111
system 4, 121
taxonomy of 13
unifying 135
valence 122, 134, 135
conservation traditions
biocentric 5, 183, 184
economic efficency 5, 179-182
eclogic 5, 182, 183
synthesis of 187-196
construct 3, 12
context 16
continuance 184, 187
cooperation, evolution of 137
coordinate system
Cartesian 61
Haskell's 4, 64-68
relational 59, 61, 62
copula 14
crown competition factor 51
D
definition 20, 40, 43
denotation 23, 25, 26
density 51, 56
description 36
designation 9, 24, 175
Dindal's expanded coaction cross tabulation 116, 119
dissipative structures 38
distance measures 63, 64, 68
diversity
and exclusion 107
and synthesis 110, 121
as a predicate 107
dynamics 107
of tree species 95
E
economic efficiency
conservation tradition 179-182
joint production basis 180, 181
effects:
degree of indirectness 162,164
emanating 162-165, 188
immmanating 162, 163, 164, 165, 188
environs, input and output 171
Subject Index 215

ethics in forest management 5, 187


evolution
community interaction structure 5, 155-158
exclusion 3, 5, 153
in mixed forest stands 102, 103
in parasite-host systems 83, 87. 88
in prey-predator systems 82
exclusion principle:
competitive 168, 169
Pauli 167-169
suggested 5, 169-171
expectation 91, 92
explanation 4, 36, 169
extincton 5, 184, 187
as the last exclusion 189
F
field (force) 54, 56
and test body 129, 130
followed-by operator 136, 141-157, 162-171
associati vi ty 142
identity 142, 171
inverse 142
uniqueness 142
forest growth
law 37
model 53
G
Gause
competition studies 72-82
prey-predator studies 80, 82-84
type initial conditions 96, 107
group, mathematical 141
groupoid 141, 142
monoid 141-143, 157
semi group 141, 142
H
habitat
higher 49, 55, 56
lower 49, 53-56
Haskellian coordinates
as fourth-order predicates 66
hypothesis 7, 8, 71
hypothetico-deduction 39

indirect effects 137


complementarity 155
theory of 143-158
integrative principles 111, 119
interaction
concept 9, 34
control by thinning 103
dually complementary 155
216 Subject Index

event 9, 31, 33
intensity 63-65
nuclear 131
parasite-host 84, 86
plant-herbivore 89, 90
polygon 85, 86, 92
positive and negative 156
prey-predator 79, 80
structures 137
term 9, 40
theory 37-40
type 63-65
valence of 122, 134
Isle Royale 80
isocost 180-182, 193
isorevenue 180-182, 194
J
joint production economics
(see economic efficiency)
K
L
land ethic 183
language 11, 19
law
of nature 37, 71
of the instrument 39
Leopold, Aldo 183
level 121
level structure
cumulative 126-131
separable 123, 126, 128
subassembly 127-129
trophic 123, 124
logic 8
loop analysis 39, 170
Lotka-Volterra equations 41
M
meaning of terms 40
measurement 8, 9
mediary 35,36, 73, 79, 82, 84, 137
mixed forest stands 95-107
Gause-type initial conditions 97-98, 101
Wit-type initial conditions 97
moose-wolf interaction 80
in Haskellian coordinates 81
multiple use
concept 1, 5, 173, 174
concept extension 173-178
decision-making 5,177,195
event 173, 174
term 173
N
Subject Index 217

natural selection
levels of 153-155
of community interaction structures 136
of indirect effects 154, 155
of interactions 155
of populations and communities 140
theory of 137, 140
o
object(s) 17
categories of 19
-object relations 22
juxtaposition of 23, 30
superposition of 23, 30
ontology
actualist 55
change 18
commodity 2, 17, 175
common sense 9
natural object 2,17,175
possibilist 18, 37, 55
relational 2, 17
repertoire 2
type IV 2, 3, 4, 18, 196
open sentence (see propositional function)
operationism 20, 39
P
packing 51-53, 56
parasite-host interactions 84
pattern 71, 72
and expectation 91, 92
Periodic coordinate system
(see coordinate system, Haskell's)
periodic tables 72
phase plane 38, 60, 61, 68,188
philosophy, exact 3
phytophagous insects 89, 90
plane of perception 72, 73, 121
plant-herbivore interaction 88
polar coordinates 62
predator efficiency 80
predicate-language item 11
predicate-logic item 13, 14
degree 21, 34
factual, formal 7, 8
order 8, 21, 34, 66, 68
reference class 50, 53
prediction and retrodiction 36, 91
prey exclusion 83
prey-predator interactions 79, 80
property of an object 19
proposition
and hypothesis 7, 8
218 Subject Index

in concept maps 132


kind of constuct 14
kinds of existential quantifiers in 14, 15
propositional function 15, 122
Q
question difficulty 36, 37
question type
what if? 37, 91
what is? 37
why? 36, 37, 169
R
reference class analysis 35
reference relation 26-28, 175
representation relation 9, 28-30
resources 35, 176, 178
s
science of science 7
science- protoscience 8
scientific method 8
semantics 7
silviculture
geometrized 106
partial concept maps for 133
site
forest producing power of 45
geocentric assessment 46
phytocentric assessment 46-49
site index 48,50
space effects on interactions 146, 147, 158
space occupancy measures 51, 52
stocking concept 51
stocking guides 107
structure
of a system 139
as a construct 138
structuralism (relational) 139, 140, 171
synthesis 4, 113-135
and abstraction 110
encyclopedic Ill, 112
mathematical 111,112
quantitative 111,112
using Haskell's coordinate system 107
symmetry III
system 9, 10, 19, 139, 161
system identification 38
T
terms 40, 41
test Dody 54, 56
theorems
self-help 145-147
self-annihilation 147-152
helpless 153, 155
Subject Index 219

theory
as nets 4
classification of 17
as kind of construct 16
relation to proposition 36
generic 49, 50, 56
relation to context 16
thinking
analytic (outside-in) 110
synthetic (inside-out) 110,115.116
trajectory comparisons 60, 63
U
Utida parasite-host studies 84, 85
V
valuation, ontological dependence 17
virulence, natural selection against 148-152
W
wrinkle
in interaction space 4, 65, 67
tessaring across 157, 158

You might also like