22.01.2014 Views

PE EIE[R-Rg RESEARCH ON - HJ Andrews Experimental Forest

PE EIE[R-Rg RESEARCH ON - HJ Andrews Experimental Forest

PE EIE[R-Rg RESEARCH ON - HJ Andrews Experimental Forest

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>PE</strong> <strong>EIE</strong>[R-<strong>Rg</strong><br />

RESEARC H<br />

O N<br />

C<strong>ON</strong>IFEROUS<br />

FOREST<br />

ECOSYSTEMS °<br />

a symposiu m<br />

I77t<br />

THIS PUBLICATI<strong>ON</strong> C<strong>ON</strong>STITUTES A C<strong>ON</strong>TRIBUTI<strong>ON</strong> TO TH E<br />

U.S./INTERNATI<strong>ON</strong>AL BIOME PROGRAM


The editors :<br />

JERRY F. FRANKLIN, Principal Plant Ecologist,<br />

Pacific Northwest <strong>Forest</strong><br />

and Range Experiment Statio n<br />

L . J. DEMPSTER, Consulting Editor ,<br />

Corvallis<br />

RICHARD H. WARING, Associate Professor ,<br />

Oregon State University ,<br />

Corvallis<br />

Proprietary or brand names are used only to document actua l<br />

experience ; their use does not imply approval of the product to th e<br />

exclusion of others which may also be suitable .


Research on<br />

Coniferous <strong>Forest</strong> Ecosystems :<br />

First Year Progress in th e<br />

Coniferous <strong>Forest</strong> Biome, USu RP<br />

Proceedings of a Symposiu m<br />

held at<br />

NORTHWEST SCIENTIFIC ASSOCIATIO N<br />

Forty-fifth Annual Meeting<br />

Bellingham, Washingto n<br />

March 23-24, 197 2<br />

Edited by<br />

Jerry F. Franklin<br />

L . J. Dempster<br />

Richard H . Waring<br />

Published in 1972 by<br />

Pacific Northwest <strong>Forest</strong> and Range Experiment Statio n<br />

<strong>Forest</strong> Service, U .S. Department of Agriculture<br />

Portland, Orego n<br />

For sale by the Superintendent of Documents, U.S . Government Printing Offic e<br />

Washington, D .C., 20402 - Price $2 .50 Stock Number 0101-0233


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Foreword<br />

The research program organized under th e<br />

Coniferous <strong>Forest</strong> Biome is probably th e<br />

largest and most comprehensive single effor t<br />

at ecosystem analysis being carried out in th e<br />

Western United States . After a long period o f<br />

planning and discussion, it is finally in its firs t<br />

year of full-scale activity . Despite this youthful<br />

state there is great interest among ecologists<br />

both within and beyond the Biom e<br />

"boundaries" in the conceptual basis for th e<br />

Biome's program and present and planne d<br />

research .<br />

Consequently, at the invitation of th e<br />

Northwest Scientific Association, members o f<br />

the Coniferous <strong>Forest</strong> Biome organized th e<br />

symposium, "Research on Coniferous Fores t<br />

Ecosystems : First Year Progress in the Coniferous<br />

<strong>Forest</strong> Biome, US/IBP." The symposium<br />

was presented at the 45th Annual<br />

Meeting of the Northwest Scientific Association<br />

on March 23 and 24, 1972 . It highlighted<br />

the concepts and plans underlying major segments<br />

of the Biome program and the numerous<br />

new insights, techniques, and data which<br />

are resulting from the varied research activities .<br />

This volume contains the proceedings of<br />

that symposium. Publication was made<br />

possible only by the full cooperation and considerable<br />

contributions of the Pacific North -<br />

west <strong>Forest</strong> and Range Experiment Station,<br />

USDA <strong>Forest</strong> Service . Station staff not only<br />

assisted significantly in the editorial phases<br />

but handled the production in its entiret y<br />

from edited copy to final printing . Furthermore,<br />

this was done under severe deadlines in<br />

order that the volume be available for th e<br />

Fifth Assembly of the International Biologica l<br />

Program in August of 1972 .<br />

The editors would also like to acknowledg e<br />

the numerous contributions others made t o<br />

the production of this volume : To Drs . Dal e<br />

W. Cole and Frieda B . Taub for their significant<br />

part in organizing and conducting the<br />

symposium ; to the majority of authors an d<br />

referees who cooperated magnificently i n<br />

producing the necessary materials unde r<br />

severe time limitations ; to George M . Hansen ,<br />

Betty J . Bell, Mildred I . Hoyt, and others i n<br />

Editorial Services of the Pacific Northwes t<br />

<strong>Forest</strong> and Range Experiment Station, wh o<br />

bore the burden of preparing the volume fo r<br />

printing ; and to Robert Romancier, Glend a<br />

Faxon, Virginia Hunt, James Overholser, Han s<br />

Riekerk, Charles Grier, and Laura Gregg fo r<br />

their part in obtaining the papers, typing, an d<br />

otherwise assisting in production of th e<br />

volume .<br />

Jerry F . Franklin<br />

L . J . Dempster<br />

Richard H . Waring


Contents<br />

SOME BROADER VIEWS OF BIOME ACTIVITIES 1<br />

Why a Coniferous <strong>Forest</strong> Biome ?<br />

Jerry F. Franklin 3<br />

Organization and research program of the Western Coniferous <strong>Forest</strong> Biome<br />

S . P. Gessel 7<br />

Findley Lake-the study of a terrestrial-aquatic interfac e<br />

P. R. Olson, D . W. Cole, and R. R. Whitney 1 5<br />

A comparative study of four lake s<br />

Frieda B . Taub, Robert L . Burgner, Eugene B . Welch, and Demetrios E . Spyridakis . . . . 2 1<br />

The modeling process relating to questions about coniferous lake ecosystem s<br />

Douglas M . Eggers and Larry M . Male 3 3<br />

Toward a general model structure for a forest ecosyste m<br />

W . Scott Overton 3 7<br />

Hydrologic modeling in the Coniferous <strong>Forest</strong> Biom e<br />

George W. Brown, Robert H . Burgy, R. Dennis Harr, and J. Paul Riley 4 9<br />

Preliminary considerations of the forest canopy consumer subsyste m<br />

M . A. Strand and W. P. Nagel 7 1<br />

An environmental grid for classifying coniferous forest ecosystems<br />

R. H. Waring, K . L. Reed, and W . H . Emmingham 7 9<br />

Pag e<br />

WATER AND NUTRIENT MOVEMENT THROUGH ECOSYSTEMS 9 3<br />

Modeling water movement within the upper rooting zone of a Cedar River soi l<br />

W. H. Hatheway, P . Machno, and E . Hamerly 9 5<br />

Elemental transport changes occurring during development<br />

of a second-growth Douglas-fir ecosyste m<br />

Charles C. Grier and Dale W. Cole 10 3<br />

Nutrient budget of a Douglas-fir forest on an experimental watershe d<br />

in western Orego n<br />

R. L. Fredriksen 11 5<br />

Nutrient cycling in throughfall and litterfall in 450-year-old Douglas-fir stand s<br />

Albert Abee and Denis Lavender 133


Page<br />

ESTIMATING BIOMASS AND OTHER STATE VARIABLES 14 5<br />

Direct, nondestructive measurement of biomass and structure i n<br />

living old-growth Douglas-fir<br />

William C . Denison, Diane M . Tracy, Frederick M . Rhoades, and Martha Sherwood . . . .14 7<br />

Estimation of biomass and transpiration in coniferous forests using tritiated wate r<br />

J. R. Kline, M . L . Stewart, and C. F. Jordan 15 9<br />

Theodolite surveying for nondestructive biomass samplin g<br />

Eugene E. Addor 16 7<br />

Estimates of biomass and fixed nitrogen of epiphytes from old-growth Douglas-fi r<br />

Lawrence H . Pike, Diane M . Tracy, Martha A. Sherwood, and Diane Nielsen 17 7<br />

Litter, foliage, branch, and stem production in contrasting lodgepole pin e<br />

habitats of the Colorado Front Range<br />

William H . Moir 18 9<br />

Small mammal and bird populations on Thompson Site, Cedar River :<br />

parameters for modelin g<br />

Sterling Miller, Curtis W . Erickson, Richard D . Taber, and Carl H . Nellis 19 9<br />

TERRESTRIAL PROCESS STUDIES 20 9<br />

Terrestrial process studies in conifers-a review<br />

R. B. Walker, D . R . M. Scott, D . J. Salo, and K . L. Reed 21 1<br />

Criteria for selecting an optimal model : terrestrial photosynthesi s<br />

Kenneth L. Reed and Warren L . Webb 22 7<br />

A model of light and temperature controlled net photosynthetic rate s<br />

for terrestrial plants<br />

Warren L. Webb 23 7<br />

Energy flux studies in a coniferous forest ecosyste m<br />

Lloyd W . Gay 24 3<br />

The lysimeter installation on the Cedar River Watershe d<br />

Leo J . Fritschen 25 5<br />

Initial steps in decomposition of Douglas-fir needles under forest conditions<br />

P. L. Minyard and C . H. Driver 26 1<br />

Seasonal and diurnal patterns of water status in Acer circinatu m<br />

James P. Lassoie and David R . M. Scott 26 5<br />

Development and testing of an inexpensive thermoelectrically cooled cuvett e<br />

David J . Salo, John A . Ringo, James H . Nishitani, and Richard B . Walker 273


Page<br />

AQUATIC PROCESS STUDIES 279<br />

Studying streams as a biological uni t<br />

James R . Sedell 28 1<br />

Exploring the aquatic carbon web<br />

Bruce Lighthart and Paul E . Tiegs 28 9<br />

Dynamics of nutrient supply and primary production i n<br />

Lake Sammamish, Washingto n<br />

Eugene B . Welch and Demetrios E. Spyridakis 30 1<br />

Hydroacoustic assessment of limnetic-feeding fishe s<br />

Richard E. Thorne 317


Some Broader View s<br />

of Biome Activities


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Why a Coniferous <strong>Forest</strong> Biome?<br />

Jerry F . Frankli n<br />

Principal Plant Ecologis t<br />

<strong>Forest</strong>ry Sciences Laboratory<br />

Pacific Northwest <strong>Forest</strong> an d<br />

Range Experiment Statio n<br />

USDA <strong>Forest</strong> Servic e<br />

Corvallis, Oregon,<br />

and Deputy Directo r<br />

Coniferous <strong>Forest</strong> Biom e<br />

US/I B P<br />

Abstract<br />

This introduction to the symposium outlines some of the reasons large integrated ecological researc h<br />

programs have been deemed necessary despite problems inherent in such "large science" programs .<br />

The Coniferous <strong>Forest</strong> Biome program is a n<br />

interdisciplinary research effort concerned<br />

with the structure and functioning of coniferous<br />

forest and associate aquatic ecosystems ,<br />

particularly as they occur in western Nort h<br />

America. Initiated as a part of the U .S. International<br />

Biological Program it is one of si x<br />

biome programs organized in major biotic -<br />

environmental divisions or regions of th e<br />

United States-Deciduous <strong>Forest</strong>, Grassland ,<br />

Desert, Tundra, Tropical, and Coniferous<br />

<strong>Forest</strong>. The first major funding of the Coniferous<br />

Biome occurred in September of 1970 ,<br />

and there are now over 100 scientists from a<br />

total of 15 universities, national laboratories ,<br />

and agencies involved in the program .<br />

Many aspects of this large and vital researc h<br />

program will be described in papers whic h<br />

follow. However, questions continue to arise ,<br />

such as, Is this program necessary? What differentiates<br />

the Biome efforts from the numerous<br />

existing research programs on coniferou s<br />

forests, both large and small? Why has thi s<br />

"big science" effort been mounted ?<br />

So, before looking in detail at the Biome' s<br />

research activities, I'd like to share with yo u<br />

some of my views on the need for and impor -<br />

tance of the Biome program . These views are<br />

not all inclusive nor necessarily shared by al l<br />

involved in the program .<br />

To get close to home, let's consider firs t<br />

the increasingly complex nature of problem s<br />

facing forest land managers. Questions used to<br />

be relatively simple : What cutting system will<br />

regenerate a desired forest type? What effec t<br />

do different thinning or fertilizing regime s<br />

have on production of merchantable timber?<br />

What herbicides will most effectively permi t<br />

reforestation and discourage brush on cutover<br />

areas? Such questions are still interesting an d<br />

important . However, many of the more critical<br />

land management questions reach far beyond<br />

the relatively narrow confines of maximizing<br />

production of goods from forest lands :<br />

How do different cutting methods influence<br />

the flow and quality of water and, further ,<br />

the characteristics of aquatic communities?<br />

What effect does fertilization have on th e<br />

nutrient content or fertility of the wate r<br />

draining from the treated forest land? What<br />

happens to various pesticides when applied to<br />

the land-how fast are they degraded, where<br />

do they accumulate, and what effect do the y<br />

have on nontarget organisms and the eco -<br />

3


system as a whole? And, if the biological an d<br />

physical questions were not sufficiently difficult<br />

(and they are), overriding economic an d<br />

social considerations further compound th e<br />

complexity of land management problems .<br />

The basic knowledge of natural science required<br />

for rational resolution of some of th e<br />

larger questions has thrust some new and<br />

difficult demarJds on the scientific community.<br />

Of coursel large amounts of information<br />

are required ; we often hear of an information<br />

explosion but, in fact, we are faced with a<br />

"demand" explosion-available data are to -<br />

tally inadequate to meet the information requirements<br />

of policymakers. More important ,<br />

however, is the need for new kinds of information,<br />

particularly information on linkage s<br />

between the parts of forest ecosystems, suc h<br />

as land and water, and between economics ,<br />

sociology, and natural sciences . In effect,<br />

information is required which involves th e<br />

links or relationships between the traditiona l<br />

units or disciplines of scientific attention an d<br />

organization .<br />

In addition to the study of linkages amon g<br />

different components of ecosystems, we ar e<br />

searching for properties that are unique to<br />

whole systems. We are asking how much tim e<br />

is required for an ecosystem to fully recover<br />

following disturbance, how many component s<br />

are redundant in the system, what are the<br />

major selection pressures to which ecosystems<br />

respond, and finally, are we forcing ecosystems<br />

to adapt to changes faster than is<br />

possible?<br />

It is also apparent that research program s<br />

must be planned so as to anticipate unforeseen<br />

questions and needs as well as answer<br />

immediate questions . In other words, scientists<br />

must work toward development o f<br />

general principles and models which will allo w<br />

us to predict from past research what wil l<br />

happen in a new situation which has not ye t<br />

been a subject of detailed study . Example s<br />

might be the use of basic models to anticipate<br />

the effect of a cutting system in a new forest<br />

environment or of a newly developed pesticide<br />

on the animal component of a forest<br />

stand. The importance of this capability for<br />

generalization is related to another recent<br />

phenomenon, the insistence that scientists<br />

provide immediate "best " answers to questions<br />

based on present available knowledge ;<br />

society is unwilling to wait for long periods<br />

for "final" solutions .<br />

These increased demands on the scientific<br />

community have to be met without the great<br />

increases in money and manpower experienced<br />

in the earlier postwar period . Since<br />

scientific resources are limited, they must be<br />

utilized more efficiently to meet society's<br />

needs for problem-solving information . Relevance<br />

and efficiency have become important<br />

considerations in planning and funding research<br />

programs .<br />

In many respects, the demands outlined<br />

above run counter to the traditional ways o f<br />

doing scientific research . Science has been<br />

strongly disciplinary in character with the<br />

greatest rewards going to the specialist who<br />

pursued his field in great depth . Even the applied<br />

scientist has tended to have a narro w<br />

focus looking, for example, at the effect of a<br />

cutting method on regeneration or thinning<br />

on wood yields, not at the overall effects of<br />

such treatments on ecosystems .<br />

From another viewpoint, science has been<br />

likened to an edifice of bricks gradually built<br />

up by the effort of many individual scientists .<br />

Unfortunately, there has been a strong tendency<br />

for each scientist to produce bricks of a<br />

dimension, shape, and material primarily of<br />

interest to him. Furthermore, at least in<br />

ecology, we have lacked overall blueprints for<br />

our edifices which would provide direction as<br />

to the kinds and number of bricks we need .<br />

There can be no questioning the need fo r<br />

strong basic research programs ; they are essential<br />

to the advancement of science and huma n<br />

knowledge. Traditional viewpoints and approaches<br />

do not adequately meet a great<br />

many of today's needs, however . .<br />

the scientific community is being<br />

forced to restructure a large part of its efforts ,<br />

partially from a sense of its own responsibilities<br />

and partially in response to society ' s<br />

pressures. Nowhere is this more evident than<br />

in the fields of natural resources and ecology .<br />

New programs are interdisciplinary effort s<br />

which try to overcome inadequacies in pas t<br />

research efforts . Resources are concentrated<br />

on critical areas related, either directly or in -<br />

4


directly, to solution of major natural resourc e<br />

problems .<br />

The Coniferous <strong>Forest</strong> Biome program i s<br />

one of these integrated, interdisciplinary efforts.<br />

It has as its overall objective an under -<br />

standing of how materials, such as water an d<br />

nutrients, and energy enter, move through ,<br />

and leave coniferous forest ecosystems, including<br />

both the terrestrial and aquatic components.<br />

The Biome program shares some<br />

common characteristics with many other larg e<br />

new ecological research programs : (1) It involves<br />

groups of scientists from many different<br />

fields working together toward some common<br />

goals. (2) It has a "systems" orientation ;<br />

it is concerned with all parts of the ecosystem,<br />

its total behavior, and the linkage s<br />

between the various components, not with a<br />

single piece . (3) Mathematical descriptions o r<br />

models of the various processes, subsystems ,<br />

and total ecosystem are a key to organizatio n<br />

and synthesis of the effort . Conceptua l<br />

models provide the framework for structuring<br />

the research effort and determining dat a<br />

needs. (4) Continuous communication between<br />

the scientist participants, including<br />

sharing of data, is an essential feature . The<br />

efforts progress via the constant interchange<br />

between scientists, between modeling and th e<br />

field and laboratory research .<br />

In our symposium, we will try to introduc e<br />

you to the kinds of research being conducte d<br />

under the auspices of the Coniferous <strong>Forest</strong><br />

Biome . The papers range widely in scope fro m<br />

general presentations on the conceptual basi s<br />

for major program segments to results of relatively<br />

narrow research projects . Unfortunately,<br />

it is too early to provide any major<br />

synthesis of activities ; this summer will be the<br />

first field season of essentially full funding .<br />

We have tried to emphasize the new concepts ,<br />

techniques, and data which are emerging fro m<br />

the Biome's efforts . We hope that the<br />

"samples " of activities and philosophy which<br />

follow will make clearer what the Biome i s<br />

trying to do and how we are going about it .<br />

Acknowledgments<br />

The work reported in this paper was sup -<br />

ported in part by the U .S. <strong>Forest</strong> Service in<br />

cooperation with the Coniferous Fores t<br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution<br />

No . 18 to the Coniferous <strong>Forest</strong> Biome .<br />

5


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Organization and research program<br />

of the Western Coniferous <strong>Forest</strong> Biom e<br />

An Integrated Research Component of th e<br />

International Biological Program<br />

S. P . Gessel, Directo r<br />

Western Coniferous <strong>Forest</strong> Biome<br />

College of <strong>Forest</strong> Resources<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

The International Biological Program was initiated in 1964. The Coniferous Biome research of the Analysis<br />

of Ecosystems component did not begin until October 1970 . Other Biome programs, with the exception of th e<br />

Tropical, preceded the Coniferous with the first research in the Grasslands Biome beginning in 1968 . The IBP<br />

research, originally planned to terminate in 1972, has been extended in the United States through 1974 . It is<br />

likely that the Analysis of Ecosystems program will be extended beyond that date in order to give the integrate d<br />

research approach an adequate test.<br />

This introductory paper describes the administrative and research organization of the Coniferous Biome and<br />

presents the broad objectives. Principal research sites and the plans for extension to other areas of the Biome are<br />

also described. As research is in the very early stages, progress can be best judged by reviewing other papers in thi s<br />

Symposium. Aside from accumulation ofdata and development ofmodels, we have made significant contributions<br />

to the education ofboth undergraduate and graduate students by introducing new modeling concepts .<br />

Introduction<br />

The International Biological Program (IBP )<br />

began in 1964 with the specific objective of developing<br />

a world study on the biological basi s<br />

of productivity and human welfare . The initiators<br />

were concerned with world-wide problems<br />

associated with increasing populations ,<br />

world food supplies, changes in environmenta l<br />

degradation, and increasing resource use .<br />

It was originally conceived as an 8-year pro -<br />

gram to be completed by 1972 . However, be -<br />

cause many national programs started late ,<br />

IBP has been extended through 1974. There<br />

are clear indications, however, that productiv e<br />

research programs such as the Biome effort s<br />

will continue beyond 1974 .<br />

The nature of IBP programs and rate of<br />

attainment of specific research objectives has<br />

been variable in the many participating countries.<br />

Rate of progress has been closely related<br />

to funds invested in the research . This discussion<br />

concerns the United States program and<br />

primarily those sections of the program under<br />

which Biome research falls. Our office will<br />

gladly supply interested scientists with references<br />

on the broader objectives and progra m<br />

of IBP .<br />

The United States IBP effort has evolved a s<br />

a series of large-scale, integrated research pro -<br />

grams coordinated by a national committee<br />

(sponsored by the National Academy of<br />

Sciences) and with the financial support of<br />

the National Science Foundation (NSF) .<br />

The Integrated Research Programs (IRP's )<br />

are in two broad groups :<br />

A. Human Adaptability<br />

1. International study of Eskimo s<br />

2. Population genetics of American Indian s<br />

3. Biology of human population at high<br />

altitudes<br />

7


4. Nutritional adaptation to environmen t<br />

5. Ecology of migrant peopl e<br />

B. Environmental Managemen t<br />

1 . Operating programs<br />

a. Convergent and divergent evolutio n<br />

in the Americas<br />

b. <strong>Experimental</strong> biography of the sea<br />

c. Physiology of colonizing specie s<br />

d. Atmospheric dispersal (aeriobiology<br />

program )<br />

e. Analysis of ecosystem s<br />

f. Conservation of ecosystems fo r<br />

scientific purposes<br />

g .<br />

Chemical and biological control o f<br />

organism s<br />

2 . Programs proposed but without funding<br />

a. Conservation of genetic material s<br />

b. Crop production under stress<br />

c. Productivity and conservation of<br />

marine mammals<br />

d. Nitrogen management<br />

e. Phenology<br />

U .S. IBP activities, particularly the Integrated<br />

Research Programs, provide a different<br />

focus for scientific inquiry ; i .e., integrated<br />

groups of scientists working toward solutio n<br />

of common problems instead of individual s<br />

working alone . The changes in focus have produced<br />

some trauma and a cadre of critics .<br />

Criticisms are still abundant but help to insur e<br />

better programs. Some critics say that integrated<br />

research efforts result in mediocre<br />

science, but program accomplishments mus t<br />

speak to this point now and in the future .<br />

Analysis of Ecosystems<br />

and th e<br />

Coniferous <strong>Forest</strong> Biome<br />

Having dealt with generalities of the total<br />

IBP, I now wish to explain briefly some specific<br />

parts of the total research effort. A s<br />

noted, the environmental management segment<br />

of U .S. IBP includes a subdivision calle d<br />

" Analysis of Ecosystems ." This is the largest<br />

and most active U .S . IBP program . It was originally<br />

conceived by Dr. Fred Smith of the<br />

University of Michigan, now at Harvard. The<br />

basic concepts involved are that world<br />

environments can be placed within broad<br />

units called Biomes and that integrated Biom e<br />

research efforts will provide the understanding<br />

for using and conserving the resources of<br />

that environment .<br />

Initially, the Analysis of Ecosystems pro -<br />

gram developed slowly . The Grassland Biome<br />

was organized first and has been a guide for<br />

other Biome programs . The total research<br />

effort will include the following Biomes :<br />

Grasslands Coniferous forest<br />

Desert<br />

Tundra<br />

Deciduous forest Tropical fores t<br />

Figure 1 illustrates the world environmenta l<br />

distribution of the Biomes .<br />

The first five Biomes are now functioning .<br />

The Tropical <strong>Forest</strong> Biome is still being organized;<br />

however, NSF is supporting tropical ecosystems<br />

research by such groups as the Organization<br />

for Tropical Studies .<br />

Figure 1 . Distribution of Biomes .<br />

The overall objectives of the Analysis of<br />

Ecosystems program are :<br />

1. To establish a scientific base for program s<br />

to maintain or improve environmenta l<br />

quality ;<br />

2. To derive broad principles of ecosyste m<br />

structure and function through an integration<br />

of the results of the six Biome studies ;<br />

3. To relate these principles to characteristic s<br />

of ecosystems such as persistence, stability ,<br />

maturity, and diversity ; and<br />

4. To develop and refine a generalized adapt -<br />

able simulation model suitable for use in<br />

8


planning studies for new developmen t<br />

projects .<br />

Generalized objectives of Biome researc h<br />

include :<br />

1.To determine the driving forces, the processes<br />

causing transfers of matter and energy<br />

among components, the nonconcentratio n<br />

characteristics, and the controlling variable s<br />

in each Biome ;<br />

2. To determine the ecosystem response t o<br />

the natural and man-induced stresses appropriate<br />

to each Biome ; for example, large<br />

herbivores in grasslands, extreme and rar e<br />

weather patterns in deserts, periodic fluctuations<br />

of rodent populations in tundra ,<br />

commercial use of timber in the coniferous<br />

forests, urbanization in the deciduou s<br />

forests, and nutrient retention in the wet<br />

tropical forests ;<br />

3. To understand the land-water interactio n<br />

characteristics of each Biome : Prairie pond s<br />

and reservoirs in the grasslands ; the abundance<br />

of shallow waters in wet tundra ;<br />

springs and temporary waters in deserts ;<br />

river and lake systems with anadromou s<br />

fish populations in coniferous forest ; pollution,<br />

and eutrophication in the deciduous<br />

forest; and large river systems in tropical<br />

forests; and<br />

4. To synthesize the results of these and previous<br />

studies into predictive models o f<br />

temporal and spatial variation, effects of<br />

pollutants and of exploitation, stability ,<br />

and other ecosystem characteristics necessary<br />

for resource management in eac h<br />

Biome .<br />

The Coniferous <strong>Forest</strong> Biome<br />

The history and research progress of ou r<br />

own Coniferous <strong>Forest</strong> Biome brings this discussion<br />

somewhat closer to home . Coincident<br />

with the development of an Analysis of Ecosystems<br />

program and the initial organizatio n<br />

and funding of the Grasslands Biome, a large<br />

body of scientists in forested areas of the<br />

West began discussing a Coniferous Fores t<br />

Biome. The University of Washington hoste d<br />

the group's first formal meeting at Pack<br />

<strong>Forest</strong> in February 1968 which produced a<br />

plan for development of a Biome research<br />

organization . The University of Montana<br />

helped to continue the discussions with a<br />

symposium on "Coniferous <strong>Forest</strong>s of the<br />

Northern Rocky Mountains " held at Missoula<br />

in September 1968 .<br />

In early 1969, I accepted an invitation to<br />

become the Director of the Coniferous <strong>Forest</strong><br />

Biome . An organizational proposal for establishment<br />

of a skeletal office and development<br />

of a Biome research program was prepared<br />

and submitted to NSF .<br />

With these very limited funds and som e<br />

financial assistance from the Central Ecosystem<br />

office and the University of Washing -<br />

ton, and with a great deal of volunteer labor ,<br />

a research proposal was prepared and submitted<br />

to NSF on December 31, 1969 . The<br />

initial $460,000 request was scaled to fund s<br />

NSF had available. The proposal was approved<br />

but not funded until September 15 ,<br />

1970. We are, therefore, still in the very earl y<br />

stages of research . Even before the first proposal<br />

was funded, a second-year proposal wa s<br />

prepared and this, in turn, was funded at 1 . 2<br />

million dollars on January 1, 1972 . We are<br />

now preparing a proposal for 1973 and 197 4<br />

with a submission date of July 1 .<br />

Coniferous <strong>Forest</strong> Biome<br />

Organization and Research<br />

The Central Biome office is at the University<br />

of Washington's College of <strong>Forest</strong> Re -<br />

sources . Intensive research sites have been<br />

established on the H. J. <strong>Andrews</strong> <strong>Experimental</strong><br />

<strong>Forest</strong> of the U .S. <strong>Forest</strong> Service in<br />

the Oregon Cascade Range and the Cedar<br />

River watershed of the City of Seattle in th e<br />

Washington Cascade Range . Although many<br />

other areas had been nominated as potentia l<br />

research sites, these two were chosen because<br />

of their concentrations of available scientists ,<br />

developed facilities, history of research, an d<br />

ability to provide the types of research area s<br />

needed to accomplish the research goals of<br />

the Biome .<br />

The total Biome research program originally<br />

called for additional coordinating site s<br />

9


where validation studies could be carried out .<br />

Funds initially restricted development outside<br />

the intensive sites ; however, plans for coordinating<br />

programs, sites, and activities are no w<br />

being developed by the Extrapolation an d<br />

Application Committee under the direction o f<br />

Dr. William Laycock of the U .S. <strong>Forest</strong> Service.<br />

It appears that a diverse group of Biome -<br />

wide or coordinating site activities including :<br />

(1) major efforts to coordinate with other on -<br />

going ecosystem research programs in th e<br />

Biome; (2) cooperative studies of ponderos a<br />

pine ecosystems with the Grasslands Biome ;<br />

(3) examinations of fire and disease influences<br />

on coniferous forests ; and (4) validation o f<br />

some initial modeling efforts will develop .<br />

The Biome 's organizational structure use d<br />

in developing, organizing, and administerin g<br />

the research is shown in figures 2 and 3 . The<br />

overall Biome research plan is developed b y<br />

the Scientific Directorate, but the individua l<br />

research proposals are generated by the re -<br />

search committees . Yearly programs develop<br />

about as follows . First, research tasks or<br />

objectives are defined. Funding guidelines fo r<br />

each of these are also outlined based on total<br />

funding expected from NSF . Research committees<br />

then define specific subprogra m<br />

objectives and after agreement by the Scientific<br />

Directorate, the chairman of a researc h<br />

committee calls for proposals from the scientific<br />

community. Proposals are reviewed b y<br />

research committees and by the Scientifi c<br />

Directorate for task relevance and adequacy .<br />

Following revisions, deletions, and additions<br />

the proposals are incorporated into the Biom e<br />

proposal. A group of scientists from othe r<br />

Biomes review the draft of the Biome proposal<br />

before its final preparation and sub -<br />

mission to NSF .<br />

Objectives of Coniferou s<br />

<strong>Forest</strong> Biome Research<br />

I will not attempt to detail the research i n<br />

this brief report, but I should comment o n<br />

the main objectives of all Biome research, i .e . ,<br />

development and validation of models of ecosystems<br />

function. The purpose of producing<br />

these models is to provide understanding o f<br />

coniferous ecosystems useful for management<br />

BIOME ADVISORY BOARDI DIRECTOR] -- ANALYSIS OF ECOSYSTEM S<br />

EXECUTIVE BOAR D<br />

{ CENTRAL OFFICE<br />

]<br />

[- TECHNICAL COMMITTEES -1<br />

Chemical Analyse s<br />

Remote Sensin g<br />

Genetic s<br />

Phenolog y<br />

Data Managemen t<br />

Internationa l<br />

DEPUTY DIRECTORS 1- EXECUTIVE DIRECTORATE<br />

SITE<br />

DIRECTOR S<br />

Terrestrial Division<br />

Aquatic Divisio n<br />

Ecosystem Integratio n<br />

Division<br />

SCIENTIFI C<br />

DIRECTORATE<br />

H . J . Andrew s<br />

Lake Washington Drainag e<br />

Validation Site s<br />

TRRESTRIAL DIVISIO N<br />

- ECOSYSTEM INTEGRATI<strong>ON</strong> DIVISIO N<br />

AQUATIC DIVISIO N<br />

Primary Production Modeling Management Lakes<br />

Food Chain Processes Interface Streams<br />

Biogeochemical Processes<br />

Extrapolation and Application<br />

Physical Processes<br />

Figure 2. Coniferous <strong>Forest</strong> Biome organizational chart .<br />

10


DIVISI<strong>ON</strong><br />

COMMITTEES AND SUBCOMMITTEE S<br />

PRIMARY PRODUCTI<strong>ON</strong> FOOD CHAIN PROCESSES BIOGEOCHEMICAL CYCLING PHYSICAL PROCESSE S<br />

TERRESTRIAL<br />

1 .<br />

2 .<br />

Processe s<br />

Biomass an d<br />

Structure<br />

1 .<br />

•2,<br />

Biologica l<br />

Geochemical<br />

1 .<br />

2 .<br />

Hydrolog y<br />

Meteorolog y<br />

LAKES<br />

STREAMS<br />

.-<br />

AQUATIC<br />

1 .<br />

2 .<br />

3 .<br />

Water Column Processe s<br />

Bottom Related Processe s<br />

Higher Consumer Processe s<br />

MODELING MANAGEMENT INTERFACES EXTRAPOLATI<strong>ON</strong> AND APPLICATIO N<br />

ECOSYSTE M<br />

INTEGRATI<strong>ON</strong><br />

Figure 3 . Research committees detail .<br />

and protection purposes . The following statement<br />

by Dr . Chapman, Deputy Director for<br />

Biome modeling activity, explains Coniferou s<br />

Biome modeling philosophy .<br />

The primary aim of this ongoing re -<br />

search is to increase our understanding<br />

of whole ecosystems within the Wester n<br />

Coniferous <strong>Forest</strong> Biome ; and to this<br />

aim an area of high priority is the development<br />

of models . . . with special emphasis<br />

on modeling of the aquatic-terrestrial<br />

interface .<br />

The analysis of an ecosystem shoul d<br />

begin by constructing models from existing<br />

data . The second step should be collection<br />

and analysis of data for testing an d<br />

refining the first models . This require s<br />

participating scientists to spend consider -<br />

able time thinking how their part of th e<br />

system relates to the rest and about general<br />

modeling processes. As the kinds of<br />

additional data that are needed are identified,<br />

scientists will be able to conduc t<br />

specific field studies designed to furthe r<br />

improve the models . This approach establishes<br />

a continual feedback loop .<br />

It is important, then, that participatin g<br />

scientists consider the understanding of<br />

systems modeling to be a major responsibility<br />

. The necessary modeling cannot be<br />

done by, say, a systems engineer, biometrician,<br />

or biomathematician . It will not<br />

be necessary for each participant to be -<br />

come a full-fledged system engineer or<br />

biomathematician (although it would be<br />

highly desirable if a few did so), but it will<br />

be essential that each participant re-orient<br />

his approach and philosophy somewhat ,<br />

for all should become systems modelers ,<br />

in one capacity or another .<br />

These points, made in our first proposal, can be<br />

reemphasized in other ways . In the first place ,<br />

every scientific study beyond simple observation<br />

involves some type of model-verbal ,<br />

graphical, or mathematical . The development<br />

of subsystem models is the responsibility of th e<br />

scientist doing the substantive research an d<br />

constitutes an essential part of their annual<br />

reports to the Biome . The modeling group provides<br />

support for those in need of modeling<br />

assistance and in fact has such a supportiv e<br />

function as one of its primary missions .<br />

The other primary mission of the modeling<br />

group is the development of the overall ecosystem<br />

model . This is one ultimate aim of th e<br />

Biome study, and while still at an early stage<br />

it will be possible to construct a total system<br />

computer model of low resolution . A better<br />

model will be achieved when reasonably<br />

sophisticated functional subsystem model s<br />

1 1


have been developed . Prior to their development<br />

the overall ecosystem model will be for<br />

the most part a translation of simple graphica l<br />

and verbal models into mathematical language<br />

with many aspects of the system interpolate d<br />

on the basis of empirical submodels or eve n<br />

crude estimates .<br />

Ultimately both total systems models an d<br />

submodels will require testing. This testing<br />

will indicate the further research needed to<br />

answer new questions or refine parts of existing<br />

models . The modeling team plays a role in<br />

assisting in the design of such new research. I n<br />

this way the model building plays an essentia l<br />

feedback role in the program .<br />

In addition, both subsystem models an d<br />

larger ecosystem models require estimates of<br />

parameters and available data often appear s<br />

inadequate. In these circumstances, it is neces -<br />

sary to test the sensitivity of the model to th e<br />

errors of estimation. Does it make a great deal<br />

of difference if the estimates are crude? If<br />

not, the model may proceed . If it does make a<br />

significant difference, additional observation s<br />

are required and further research in such area s<br />

is necessary. Obviously, such sensitivity tests<br />

cannot be made until models are fairly wel l<br />

developed ; some models will be at this stage<br />

this year.<br />

Specific Goals<br />

The goal of our Coniferous Biome research<br />

is development of a basic understanding of<br />

coniferous forest ecosystems, including both<br />

terrestrial and aquatic components, so tha t<br />

ecological constraints on and opportunitie s<br />

for increased production of fiber, food, water ,<br />

and wildlife can be recognized . The overall<br />

strategy includes identification of the majo r<br />

components and processes, both physical an d<br />

organic, within the ecosystem, and definition<br />

of their interrelationships . The definition of<br />

interrelationships will be accomplished<br />

through a systems analysis and modeling pro -<br />

gram . As an overall guide for Biome research ,<br />

we recognize the following general objectives :<br />

1. To determine the major factors, both components<br />

and processes, that control the<br />

productivity and distribution of organisms<br />

in coniferous forest ecosystems, includin g<br />

(a) an analysis of the structure and distribution<br />

of the principal resources, (b) definition<br />

of the functional relationships between<br />

biotic, decomposer, consumer, and<br />

producer components of the systems, an d<br />

(c) analysis of the forms and degrees of<br />

stability in these systems .<br />

2. To examine the linkage of terrestrial an d<br />

aquatic components in coniferous fores t<br />

ecosystems, including (a) water, energy ,<br />

and transport of chemicals (including pesticides),<br />

(b) direct transport of terrestrial<br />

products into the aquatic system, e .g. ,<br />

through litter fall and surface erosion, an d<br />

(c) return of organic and inorganic materia l<br />

from the aquatic to the terrestrial environment<br />

through movements of fish, birds ,<br />

and insects .<br />

3. To determine how various manipulation s<br />

influence the structure and function of<br />

coniferous forest ecosystems using bot h<br />

unit watersheds and plot studies . Special<br />

attention is directed to the influences of<br />

manipulations on (a) stability and productivity<br />

of these systems and (b) the linkage s<br />

between terrestrial and aquatic components<br />

of the systems.<br />

4. To understand population dynamics of<br />

those major components of each trophic<br />

level which appear to influence significantly<br />

the sustained productivity and stability<br />

of various coniferous forest ecosystems<br />

within the Biome .<br />

5. To produce models of temporal and spatia l<br />

variations in coniferous forest ecosystem s<br />

or system components. These models will<br />

include factors affecting productivity an d<br />

stability of the systems and the linkages between<br />

terrestrial and aquatic environments ,<br />

forecasting the behavior of these systems<br />

and their relationships to human manipulation<br />

.<br />

6. To apply to specific models in the solutio n<br />

of major use problems in the Biome are a<br />

and assist other groups or agencies to d o<br />

likewise .<br />

Since these long-term objectives provid e<br />

only general guidance, considerable time an d<br />

effort is spent in defining annual researc h<br />

tasks in order to focus the research an d<br />

12


modeling activities and provide a basis fo r<br />

phasing of the program .<br />

The specific research tasks planned for<br />

1973 and 1974 are to :<br />

1. Complete initial programs in terrestrial production<br />

process measurement and modeling,<br />

using both physiological and meteorological<br />

techniques, and link results to behavior<br />

(net productivity, transpiration ,<br />

etc.) of actual stands . Begin developing<br />

data for the production process models fo r<br />

other species and environments and for<br />

different age classes important in th e<br />

Biome, including linkages of individual tre e<br />

processes to behavior of stands .<br />

2. Develop elemental cycling models of forest<br />

stand level ecosystems based upon proces s<br />

and transfer functions such as decomposition,<br />

ionic leaching, weathering, organism<br />

uptake, storage, and return, and meteorological<br />

and biological additions . This will<br />

include stressed systems. This fine resolutio n<br />

work will serve as a basis for development o f<br />

unit watershed programs .<br />

3. Begin describing paths of energy and material<br />

flow through consumers (food webs) in<br />

coniferous ecosystems under study, wit h<br />

particular attention to movement and accumulation<br />

of toxic materials and consequen t<br />

effects of diversity, flow paths, and primar y<br />

producers .<br />

4. Complete initial nutrient, water, and energ y<br />

flow models for unit watersheds and begi n<br />

their refinement with particular attention to<br />

incorporation of process models . For selected<br />

compartments and transfers and i n<br />

simplified form, begin examining the abilit y<br />

of the watershed models to describe behavior<br />

of (a) larger drainages and (b) strongl y<br />

contrasting coniferous ecosystems across th e<br />

Biome, and to predict (c) system response s<br />

under various stresses and over long tim e<br />

spans (in connection with manipulation an d<br />

successional research) .<br />

5. Quantify and model paths and rates of material<br />

and energy transfers across terrestrial -<br />

aquatic interfaces, specifically includin g<br />

quantification of transfers in and comparisons<br />

between (a) undisturbed small stream s<br />

and small lakes, (b) lakes of various topolog y<br />

and climate, and (c) undisturbed and clear -<br />

cut watersheds with stream systems . In the<br />

comparisons of interface transfers in undisturbed<br />

and altered systems, study consequences<br />

of terrestrial disturbances on aquat -<br />

ic productivity and stability with initial<br />

emphasis on stream systems .<br />

6 . Evaluate and model the trophic dynamic s<br />

of four lakes of significantly different<br />

nutritional status, topology, and climate .<br />

Evaluation and modeling will be by way of<br />

three subdivisions of the aquatic system :<br />

(a) the water column, including the rapidly<br />

changing community of phytoplankton ,<br />

microbes, and zooplankton, (b) the<br />

benthos, including the moderately slowly<br />

changing community of larger plants, sediments<br />

and longer life span invertebrates ,<br />

and (c) higher consumers-long life spa n<br />

invertebrates and fish . Integrate these submodels<br />

into a total system productivity<br />

model that allows determination of importance<br />

of various internal and externa l<br />

nutrient sources to lake productivity an d<br />

suggests methods of management of productivity.<br />

Finally, explore opportunitie s<br />

for validation elsewhere .<br />

7. Examine responses of coniferous ecosystems<br />

or subsystems to selected stresse s<br />

(manipulations) in terms of material and<br />

energy flows and of productivity . This tas k<br />

is largely structured as validation of model s<br />

and will utilize existing as well as newl y<br />

acquired data. Manipulations may includ e<br />

(a) clearcutting, (b) fertilization, (c) addition<br />

of toxic materials (pesticides an d<br />

heavy metals), (d) addition to or remova l<br />

of nutrients from aquatic systems, and<br />

(e) burning .<br />

8. Examine long-term (successional) behavio r<br />

of coniferous ecosystems specifically including<br />

: (a) development of a successiona l<br />

model for prediction of forest compositio n<br />

and biomass changes, primarily by cooperation<br />

with the Deciduous <strong>Forest</strong> Biome an d<br />

acquisition of necessary data base for<br />

validation of the Oak Ridge model(s) ;<br />

(b) descriptions of selected subsyste m<br />

structures and processes across different<br />

forest age classes to test successional hypotheses<br />

regarding : (1) complexity and redundancy<br />

of detritus subsystems, (2) ratio<br />

13


of gross production and respiration, and<br />

(3) conservation of nutrients within the<br />

system ; (c) examination of eutrophication<br />

processes in lake systems using a series o f<br />

lakes of varying "ecological stage" (nutritional<br />

status) .<br />

Each of these tasks involves an integratio n<br />

of efforts by several discipline groups . Expanded<br />

and more rapid interchange betwee n<br />

discipline groups, specialists, and modelers ,<br />

and between interdisciplinary units workin g<br />

on different levels of resolution or systems, i s<br />

essential and will be pursued by all feasibl e<br />

methods. Expanded contact, including personnel<br />

exchange will be developed with othe r<br />

Biomes . Specific cooperative programs wit h<br />

the Deciduous Biome on succession and wit h<br />

the Grassland Biome on ponderosa pine -<br />

grassland systems are planned . Extension of<br />

program activities to a Biome-wide basis i s<br />

essential to the 1973-74 program .<br />

Progress<br />

The papers presented in this symposium<br />

will indicate to some extent the progress in<br />

Biome research . Something which perhap s<br />

will not be as apparent in the presentations i s<br />

our accomplishments in working together so<br />

that the total of our group research effort is<br />

greater than the sum of the individual parts . I<br />

believe some of our best progress has been i n<br />

getting terrestrial researchers to consider th e<br />

relationship of their work to aquatic components<br />

and how these systems interface. Th e<br />

same kind of crosslinks have been establishe d<br />

among different disciplines on each campu s<br />

involved in the study and perhaps even mor e<br />

important across state and county boundarie s<br />

between different institutions . The working<br />

relationships established between scientists at<br />

Oregon State University, the University o f<br />

Washington, and the U .S. <strong>Forest</strong> Service have<br />

been particularly outstanding. U .S. <strong>Forest</strong><br />

Service scientists, especially those in the<br />

Pacific Northwest <strong>Forest</strong> and Range Experiment<br />

Station, have made a major contributio n<br />

to the success of the program .<br />

For organizational and budget reasons w e<br />

have concentrated our research effort at tw o<br />

major sites, the Lake Washington-Cedar Rive r<br />

system and the H. J. <strong>Andrews</strong> Experimenta l<br />

<strong>Forest</strong>. We must depend upon research at<br />

these areas to provide the initial theory an d<br />

models. We do recognize the broad researc h<br />

responsibility embodied in the term Coniferous<br />

<strong>Forest</strong> Biome and that we must expan d<br />

the base by carefully choosing additional site s<br />

for study and data accumulation in order to<br />

test models and theory . We also hope that<br />

researchers will use the principal research site s<br />

to validate models they may have develope d<br />

elsewhere .<br />

We are currently engaged in a study of th e<br />

entire Biome area for the purpose of selecting<br />

areas and projects which will extend Biom e<br />

research in an orderly and efficient manner .<br />

We are also developing inter-Biome research<br />

efforts which will enable us to exchange<br />

models and validation studies with all othe r<br />

Biomes .<br />

These remarks should serve to illustrate<br />

that the Coniferous Biome is in the early<br />

stages of a research program . Much of our effort<br />

so far has been related to organizing<br />

working groups, setting up general and specific<br />

research goals and then writing the necessary<br />

proposals . We hope the next 2 years can<br />

be devoted to the actual research and that in a<br />

similar meeting in 1974 we can demonstrate<br />

the success of the program .<br />

Acknowledgments<br />

The work reported in this paper was sup -<br />

ported in part by National Science Foundatio n<br />

Grant No . GB-20963 to the Coniferous Fores t<br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution<br />

No . 19 to the Coniferous <strong>Forest</strong> Biome .<br />

14


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Findley Lake the study of a<br />

terrestrial-aquatic interface<br />

P. R . Olson and D . W. Col e<br />

College of <strong>Forest</strong> Resources<br />

an d<br />

R . R . Whitney<br />

College of Fisheries<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

The linkage of the aquatic properties of a small lake to the terrestrial landscape is under examination in th e<br />

Findley Lake Basin of the Cedar River watershed. This pristine, 10 ha, oligotrophic lake is situated at 1,128 m<br />

elevation in a 162 ha watershed. During the initial year of investigation, this research program has focuse d<br />

primarily on a description of the ecosystem components. This initial phase of the program and the long-term<br />

objectives are discussed .<br />

Introduction<br />

Findley Lake is located in the upper portion<br />

of the Cedar River watershed in Kin g<br />

County, State of Washington . Cedar River<br />

watershed is the municipal watershed of th e<br />

City of Seattle and is one segment of the Lak e<br />

Washington watershed which is one of the intensive<br />

study sites of the Coniferous Fores t<br />

Biome, a portion of the United States Inter -<br />

national Biological Program (Gessel 1972) .<br />

A mission of the IBP is to demonstrat e<br />

how sound biological information can be<br />

synthesized and used in assessing alternatives .<br />

More specifically the Coniferous Fores t<br />

Biome addresses its attention to specific<br />

resource problems within the Biome . In general<br />

terms the Findley Lake interface study is<br />

directed to: better understanding of productivity<br />

and its maintenance on forest lands an d<br />

adjacent surface waters ; to determine the role<br />

of land and vegetation management in water<br />

quality and quantity ; and to measure the<br />

interaction of animal populations upon these<br />

areas .<br />

The City of Seattle Water Department ha s<br />

agreed to dedicate this entire drainage basin a s<br />

a research area for the IBP, Coniferous <strong>Forest</strong><br />

Biome program. This cooperation includes<br />

securing the timber rights of the property and<br />

developing access into the area which at th e<br />

present time is by trail . The intent at Findley<br />

Basin is to maintain the pristine environmen t<br />

and minimize human influence while conducting<br />

the research .<br />

Watershed Description<br />

Findley Lake is located at 1,128 m elevation<br />

in a cirque with a maximum elevation of<br />

1,450 m and a total acreage of 162 ha . The<br />

aquatic portion of the Findley Lake basi n<br />

consists of Findley Lake proper with an are a<br />

of 10 ha and two smaller ponds of abou t<br />

0.4 ha each. Figure 1 defines the drainage<br />

basins of each of the ponds and lakes . Findley<br />

Lake proper intercepts 60 percent of th e<br />

watershed, and the one shallow pond intercepts<br />

14 percent. The lower, deep pond inter -<br />

15


Figure 1 . Aerial photo of Findley Lake basin indicating<br />

drainage basin of upper lake and two smalle r<br />

ponds.<br />

cepts each of the other two basins as well a s<br />

26 percent of the watershed directly .<br />

The soils of the watershed have bee n<br />

mapped by Bockheim and Ugolini (n .d .) .<br />

Many of the slopes range from 30 to 40 percent<br />

; widespread talus accounts for 16 .2 percent<br />

of the total basin. Soils of mixed materials<br />

are divided as forested, semiforested, an d<br />

nonforested ; they account for 56 .2, 4 .3, an d<br />

1.6 percent, respectively, of the total basin .<br />

The residual soils on the ridges are 17 .5 percent<br />

forested and 4 .2 percent nonforested .<br />

Del Moral (1972) has completed a vegetation<br />

survey of the watershed . Seven vegetation<br />

types were distinguished and their distributions<br />

mapped . Most of the area is timbere d<br />

by relatively homogenous old-growth Pacifi c<br />

silver fir (Abies amabilis) .<br />

Taber' has conducted a preliminary survey<br />

of the terrestrial vertebrate activity in the<br />

1 P. R. Olson, J . S . Bockheim, F . C. Ugolini, an d<br />

others . A terrestrial-lake interface program, Findle y<br />

Lake watershed. Coniferous <strong>Forest</strong> Biome Interna l<br />

Report No . 25 . (In press .)<br />

watershed. The elk (Germs canadensis) and<br />

pikas (Ochotona princeps) have a noticeable<br />

impact on the vegetation of the basin . Migratory<br />

,vertebrates such as deer (Odocoileus<br />

hemionus), elk, and birds could influence the<br />

movement of nutrients in or out of the basin .<br />

In addition, the elk activity in some of the<br />

meadow areas has obviously disturbed th e<br />

soil .<br />

Snow accumulations of 4 m often occur i n<br />

the Findley Lake basin . The open surface<br />

period on Findley Lake is approximately 4<br />

months. The lake frequently is not free of<br />

snow cover until late July and usually start s<br />

to close about mid or late November. A thin<br />

layer of ice will often form in early November<br />

and soon thereafter, snowfall will cover th e<br />

surface. The snowfall accumulates rapidly and<br />

insulates the surface from additional freezing .<br />

The latent heat of the lake water thaws th e<br />

ice layer originally formed and the snowpac k<br />

effectively supports itself . As snow accumulates,<br />

the light penetration is greatly reduced .<br />

In February of 1972 there were 2 .9 m (9 . 5<br />

feet) of snow ; the light readings beneath th e<br />

snowpack were approximately 1 .4 lux-0 .0 3<br />

percent of the surface illumination . Th e<br />

oxygen concentration of the water colum n<br />

was close to saturation except within 2 m o f<br />

the bottom where the oxygen decrease wa s<br />

quite abrupt. Temperatures at this tim e<br />

ranged from 1 0 C at the surface to 3° C a t<br />

25 m .<br />

Limnological surveys were performed b y<br />

Welch (see footnote 1) on a monthly interval<br />

during the open summer period . Findley Lake<br />

is extremely clear with visibility extending to<br />

depths of 15 m . Since the lake is relativel y<br />

deep (about 28 m) and unproductive, th e<br />

hypolimnion during the summer period remained<br />

well oxygenated . Maximum water<br />

temperature noted at the lake surface durin g<br />

the surveys was 18° C . The dominant zoo -<br />

plankton found was Diaptomus shoshon e<br />

with a modest population of Holopedium<br />

gibberum present early in the season. Th e<br />

Diap torn us population persisted after 3<br />

months of snow cover .<br />

The analysis of water samples b y<br />

Spyridakis (see footnote 1) is covered mor e<br />

extensively by Taub et aL(1972). Mean values<br />

16


of the most pertinent variables were : total<br />

phosphorus 5 µg/2, orthophosphorus 1 µg/2 ,<br />

nitrate-nitrogen 3 pg/2, silica 76 pg/Q, calciu m<br />

1 mg/2, magnesium 0 .6 mg/2, sodium 1 mg/2 ,<br />

and potassium 25 µg/2 . Chlorophyll determinations<br />

were 0.3 pg/2 and primary production<br />

was 370 mg C/m 2 /day .<br />

Other surveys which are relatively complet e<br />

at this time are the paleoecological analysis by<br />

Tsukada (see footnote 1) of numerous 1 m<br />

long sediment cores . These cores are bein g<br />

analyzed for concentration of pollen ,<br />

diatoms, Cladocera, inorganic elements, an d<br />

organic chemicals including chlorophyll . Th e<br />

1 m cores were not sufficiently deep to give a<br />

complete record ; however, they did disclos e<br />

the fact that the surrounding environmen t<br />

had been disturbed at least twice in the recen t<br />

past. There was evidence of a fire with subsequent<br />

pollen change and, in a deeper portion ,<br />

a 500 percent rise in sedimentary chlorophyll .<br />

Longer cores will be necessary to explain thi s<br />

phenomenon . Soil profiles disclosed evidence<br />

of two relatively recent fires as well as three<br />

volcanic ash depositions .<br />

A macrobenthic invertebrate survey wa s<br />

taken by Paulson (see footnote 1) of the lake ,<br />

the ponds, and the interconnecting streams .<br />

May flies and caddis flies appear to be th e<br />

dominant consumers in the lake . May flies<br />

were represented by Baetidae and caddis flie s<br />

by two species of Limnephilidae . Midge larvae<br />

were present, two species of stone fly and als o<br />

Simuliidae in the creek . In addition a sphaeri d<br />

clam and a planorbid snail were found in fai r<br />

abundance . Three species of amphibians wer e<br />

abundant in the lake and pond : rough-skinne d<br />

newt (Taricha granulosa), western toad (Bufo<br />

boreas), and Cascades frog (Rana cascadae) .<br />

Less common were the northwestern salamander<br />

(Ambystonza gracile) and the Pacifi c<br />

treefrog (Hyla regilla) . All evidence to dat e<br />

indicates there are no fish present in the lake .<br />

Present and Future<br />

Program s<br />

Year one study on Findley Lake was primarily<br />

a survey year . The effort at present is<br />

directed at a continuation of the previou s<br />

research with a higher resolution to supply<br />

the necessary input to modelers . Studies will<br />

be intensified on nutrient availability and<br />

quantification of primary, secondary, an d<br />

tertiary production in both the terrestrial and<br />

aquatic environments .<br />

The objective of the proposed pedological<br />

study by Ugolini is to understand soil -<br />

weathering processes and their relations to th e<br />

b i o geochemical cycle. Gravitational water<br />

interconnects the atmospheric, biological ,<br />

lithological, and hydrological components of<br />

the biosphere, bringing the reactants together<br />

and transporting the end products to ne w<br />

sites. The very essence of soil formation is th e<br />

migration of ions and particulate matter . Prior<br />

to establishing the kinds and rates of migration<br />

it is necessary to acquire a knowledge o f<br />

the soil system in both its chemical an d<br />

mineralogical composition . Thus estimations<br />

of nutrient capital in soils and forest floor o f<br />

the Findley Basin will be determined .<br />

The common denominator throughout the<br />

entire ecosystem is water . Water input as<br />

precipitation will be determined and followed<br />

through the aerial portion of the system . Thus<br />

crown wash and stemflow analyses, both<br />

qualitative and quantitative, are essential as<br />

well as determination of litterfall rates . The<br />

elements of prime concern will be initially N ,<br />

Ca, Mg, K, and P . This aspect will be investigated<br />

by Cole and Gessel who will follo w<br />

water through the soil system collectin g<br />

leachates beneath the forest floor, at th e<br />

boundary between A and B horizons, beneat h<br />

the rooting zone in the C horizon and withi n<br />

the ground water table. Techniques used here<br />

will be similar to those previously reporte d<br />

for the Thompson Site (Cole and Gessel<br />

1968) .<br />

The surface water flow within the basi n<br />

will be surveyed and chemically analyzed b y<br />

Spyridakis. Many of the surface flows ar e<br />

temporary in nature associated with snow -<br />

melt; hence, the quantity of surface flow i n<br />

the upper system will not be critically deter -<br />

mined, but Wooldridge will quantify the<br />

hydrologic flow from the entire basin. Th e<br />

nutrient accumulation within the snowpack<br />

will also be determined .<br />

17


Nutrient levels in stream and lake water a s<br />

a function of seasonal variations and samplin g<br />

location will be determined by Spyridakis .<br />

Determinations will include temperature, conductivity,<br />

suspended and dissolved solids ,<br />

light penetration, turbidity, color, dissolve d<br />

O 2 , pH, soluble and particulate C, chemical<br />

oxygen demand, dissolved SiO 2 , Na, K, Ca ,<br />

Mg, Fe, Al, Mn, chloride, sulfate, bicarbonate ,<br />

and carbonate . In addition, the lake sediment s<br />

will be characterized and their contribution t o<br />

the overlying waters will be determined .<br />

Welch will determine the annual and seasonal<br />

primary production in the lake an d<br />

relate it to the available nutrients . Special<br />

attention will be directed to regulation an d<br />

control dictated by the elements which may<br />

be limiting. Attention will be directed to th e<br />

phytoplankton cell size and their utilizatio n<br />

by zooplankton . The zooplankton composition<br />

and biomass will be determined on a<br />

regular sampling schedule. The predatio n<br />

impact upon zooplankton will probably b e<br />

quite minimal, because of the lack of fish i n<br />

the system ; however, observations will b e<br />

made to determine the potential consumptio n<br />

by insects and amphibians . A consideration<br />

for the fourth year of study may be the introduction<br />

of cutthroat trout into the lowes t<br />

pond in the lake basin to determine th e<br />

impact of predation upon the secondary production,<br />

after a comparison of the pond t o<br />

the lake for a period of time .<br />

The contrast existing between the tw o<br />

ponds and the lake presents an interestin g<br />

study by itself. Both ponds are 0.4 ha in area<br />

but have different depth profiles . The upper<br />

pond is only 1 .5 m deep and with the loss of<br />

the snowpack in midsummer will cease to<br />

overflow and decrease to 1 .0 m in depth . The<br />

lowest pond has not been accurately sounde d<br />

but appears to be at least 10 m deep and continually<br />

receives the overflow from the upper<br />

lake. The shallow pond opens earlier in th e<br />

summer, has a higher heat budget, and freezes<br />

earlier in the fall. This contrast will provide an<br />

interesting comparison in studies of decomposition,<br />

lake bottom water interface, an d<br />

growth and reproduction of amphibians an d<br />

invertebrates .<br />

One of the concerns in the aquatic area is<br />

the self-sustaining capabilities within that are a<br />

and the rapidity of recharge when the syste m<br />

is limited in some important essential . Questions<br />

that have been asked repeatedly in th e<br />

past deal with the role or importance of<br />

decomposition in the aquatic system. The<br />

Findley Lake basin presents a unique opportunity<br />

to measure the dynamics of decomposition<br />

when light is abundant and when ligh t<br />

is greatly reduced or eliminated for a 6 month<br />

period by snow cover. Findley Lake will be<br />

used primarily as a comparison to the othe r<br />

three lakes in the system as discussed by Tau b<br />

et al. (1972) . The flux and pool sizes of<br />

carbon in the water column will be deter -<br />

mined by Lighthart from a compartment<br />

analysis of (1) dissolved inorganic carbon ,<br />

(2) phytoplankton, (3) zooplankton ,<br />

(4) dissolved organic carbon, (5) detritus, and<br />

(6) heterotrophic bacteria using a C 14 technique.<br />

This technique requires the measurement<br />

of primary productivity as the initial<br />

step in measuring exchanges between compartments.<br />

The role of the aerobic, facultative<br />

anaerobes, and anaerobic bacteria will be<br />

determined by Matches .<br />

The oxidation of organic matter in th e<br />

water column to be done by Packard will be<br />

estimated by means of an oxygen uptak e<br />

method measuring enzyme activity (Packard<br />

1969). By this method he will compare th e<br />

seasonal changes in respiratory activity o f<br />

phytoplankton and zooplankton . Pamatmat<br />

will determine the annual deposition of<br />

organic matter on the lake basin and its rat e<br />

of decomposition . This will be determined by<br />

placing in the sediment a bell jar equippe d<br />

with oxygen probe, thermistor probe, and a<br />

stirring mechanism (Pamatmat and Bans e<br />

1969). The oxygen loss will be monitore d<br />

with regard to time so that consumption rate s<br />

may be determined. This will be related to th e<br />

oxygen budget of the overlying water column .<br />

Estimates of the contribution of detrital<br />

biomass to the aquatic food chain will b e<br />

determined by Taub . This will involve th e<br />

biomass estimate, and its organic carbon ,<br />

organic nitrogen, and caloric contribution t o<br />

zooplankton, benthic invertebrates, an d<br />

amphibians . In addition, Taub will investigat e<br />

the biological aspects of nitrogen transforma -<br />

18


tions (Klucas 1969) . An overall nitrogen<br />

balance estimate will be determined from estimates<br />

of nitrogen fixation and denitrification .<br />

Attempts will also be made to determine th e<br />

transfer from detrital material such as chiti n<br />

and protein to the dissolved plant nutrien t<br />

ammonia . The interchange between the<br />

various dissolved states of nitrogen, ammonia ,<br />

nitrate, and nitrite will also be investigated .<br />

Discussion<br />

The Findley Watershed relates to the overall<br />

objectives of the Coniferous Biome i n<br />

three specific ways :<br />

1. It provides an excellent opportunity to<br />

study linkage between terrestrial processes<br />

and the limnological properties of a lake<br />

system. The research is designed to assign<br />

values to the rates and quantities of elemental<br />

and organic exchange between these<br />

systems .<br />

2. Findley Lake represents an extreme i n<br />

oligotrophic conditions in lake systems and<br />

should provide a very interesting genera l<br />

lake model. It will serve as an excellen t<br />

comparison with the lower three lakes o f<br />

the Lake Washington drainage basin in<br />

studies of nutrient income vs . productivity<br />

relationships. Consequently, it is a critical<br />

component of the four lake program of th e<br />

Coniferous Biome .<br />

3. Because of the similarity in terrestrial community<br />

structure between Findley Lake<br />

and some of the reference stands at th e<br />

Biome's H. J: <strong>Andrews</strong> Intensive Sit e<br />

(Gessel 1972), it provides a linkage between<br />

the two biome study sites . Th e<br />

effort at the <strong>Andrews</strong> Site will concentrat e<br />

on unit watershed models, which will include<br />

transpiration, terrestrial primary production,<br />

and stream hydrologic models, al l<br />

of which should have excellent applicatio n<br />

in the Findley watershed .<br />

To study the interface phenomenon ,<br />

Findley Lake has been diagrammatically stratified<br />

into a series of discrete ecosystem components<br />

connected by transfer functions .<br />

Figure 2 represents a first approximation of<br />

how these components are interconnected<br />

Figure 2 . Terrestrial-lake interface diagram with transfer<br />

functions numerically listed : (1) atmospheri c<br />

exchange, heat, light, water and gas; (2) nutrient<br />

assimilation; (3) nutrient leaching ; (4) hydrologica l<br />

output ; (5) consumption; (6) detrital contribution ;<br />

(7) decomposition ; (8) return from decomposition ;<br />

(9) movement across interface; (10) migration i n<br />

and out of watershed .<br />

and the nature of the linkage between th e<br />

terrestrial and aquatic systems . Each of the<br />

component boxes will necessitate a quantitative<br />

estimate of the biomass or effective<br />

reservoir involved . The pathways between<br />

components will not only involve a quantitative<br />

estimate but also a measure of rate o f<br />

movement between the components . As the<br />

research progresses, the level of resolution wil l<br />

become extremely critical before a total systems<br />

analysis is possible . A few important<br />

variables which will be kept in mind are :<br />

(1) spatial dimensions of the system under<br />

study, (2) time base of the study, and (3) th e<br />

nutrient elements and/or caloric equivalent s<br />

of the components . It is very likely that th e<br />

modeling effort will help prescribe a desirabl e<br />

time base .<br />

19


Acknowledgments<br />

The work reported in this paper was sup -<br />

ported by National Science Foundation Grant<br />

No. GB-20963 to the Coniferous Fores t<br />

Biome, U .S. Analysis of Ecosystems, Inter -<br />

national Biological Program . This is Contribution<br />

No . 20 to the Coniferous <strong>Forest</strong> Biome ,<br />

IBP .<br />

Literature Cited<br />

Bockheim, J . G., and F . Ugolini .[n .d.] Soil s<br />

and parent materials of Findley Lake ,<br />

Snoqualmie National <strong>Forest</strong>, Washington .<br />

Northwest Sci . : (In press . )<br />

Cole, D . W., and S. P . Gessel. 1968. Cedar<br />

River research-a program for studyin g<br />

pathways, rates and processes of elemental<br />

cycling in a forest ecosystem. 53 p . <strong>Forest</strong><br />

Resour. Monogr ., Univ . Wash. Contrib . No .<br />

4. Seattle .<br />

Del Moral, R . 1972. Vegetation survey o f<br />

Findley Lake Basin . Am. Midland Nat . : (In<br />

press .)<br />

Gessel, S . P. 1972. Organization and researc h<br />

program of the Western Coniferous Fores t<br />

Biome . In Jerry F. Franklin, L . J .<br />

Dempster, and Richard H . Waring (eds .) ,<br />

Proceedings-research on coniferous forest<br />

ecosystems-a symposium, p . 7-14, illus .<br />

Pac. Northwest <strong>Forest</strong> & Range Exp . Stn . ,<br />

Portland, Oreg .<br />

Klucas, R. V. 1969. Nitrogen fixation assessment<br />

by acetylene reduction . Eutrophicat<br />

i o n -B i o s timulation Assessment Work -<br />

shop. Proc. 1969 : 109-116 . Univ. Calif . ,<br />

Berkeley .<br />

Packard, T. T. 1969. The estimation of th e<br />

oxygen utilization rate in sea water fro m<br />

the activity of the respiratory electro n<br />

transport system in plankton . 115 p . Ph .D .<br />

thesis on file, Univ . Wash .<br />

Pamatmat, M. M., and K. Banse . 1969 .<br />

Oxygen consumption by the seabed . II. In<br />

situ measurements to 180 m depth .<br />

Limnol. Oceanogr. 14 : 250-259 .<br />

Taub, Frieda B., Robert L . Burgner, Eugene<br />

B. Welch, and Demetrios E . Spyridakis .<br />

1972. A comparative study of four lakes .<br />

In Jerry F . Franklin, L . J. Dempster, an d<br />

Richard H . Waring (eds .), Proceedingsresearch<br />

on coniferous forest ecosystems- a<br />

symposium, p . 21-32, illus. Pac. Northwest<br />

<strong>Forest</strong> & Range Exp . Stn ., Portland, Oreg .<br />

20


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

A comparative study<br />

of four lakes<br />

Frieda B . Taub ,<br />

Robert L . Burgner,<br />

Eugene B . Welch,<br />

an d<br />

Demetrios E . Spyridakis<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

It is increasingly evident from many unfortunate case histories that lake communities are sensitive in a<br />

variety of inadequately understood ways to watershed management practices and abuses . If we are to conserv e<br />

or manage our array of lakes of the Coniferous Biome successfully, we must gain a sufficient understanding t o<br />

predict the impact of various perturbations on the general lake community structure and production . Initial<br />

effort in the Coniferous Biome is being directed toward an intensive comparative study of an array of four lake s<br />

in a single watershed that provide a spectrum of human disturbance and manipulation . These are lakes<br />

Washington, Sammamish, Chester Morse, and Findley in the Lake Washington-Cedar River Drainage. A study of<br />

Cedar River, the main drainage system and a link between two of the lakes, contributes to both the lakes' stud y<br />

and the smaller stream investigations at <strong>Andrews</strong> <strong>Forest</strong> .<br />

The Place of Lake Studies<br />

in the Biome<br />

It may well be asked why there should be<br />

an intensive study of lakes within a C<strong>ON</strong>IF-<br />

EROUS BIOME . Certainly, the biome is<br />

named for the most conspicuous organisms ,<br />

the terrestrial conifers . However, it can be<br />

shown that aquatic studies and the terrestrial<br />

studies each have much to gain by being<br />

combined .<br />

Unlike the European IBP, the US IBP chose<br />

to combine the initial working groups for<br />

Freshwater Productivity and Terrestrial Productivity.<br />

It was anticipated that this approach<br />

would best encourage truly biome -<br />

wide studies, not artificially divided b y<br />

discipline boundaries, such as the Findle y<br />

Lake study which is discussed in the previou s<br />

paper. This Biome also includes studies of th e<br />

marine contribution to freshwater and terrestrial<br />

communities by anadromous fish .<br />

Limnologists have long recognized that th e<br />

terrestrial influences on lakes are of utmos t<br />

importance. The physical characteristics of a<br />

lake are determined largely by its topology<br />

and climate which will determine the degree<br />

and frequency of mixing. It had long been<br />

thought that lakes progressed from oligotrophic<br />

to eutrophic as an inevitable<br />

sequence . Margalef (1968) and Odum (1969 )<br />

have suggested that lakes will progress towar d<br />

increased oligotrophy if the surrounding lan d<br />

does not contribute nutrients . Thus the majo r<br />

successional pattern of a lake may be determined<br />

by the terrestrial input .<br />

The nutrient inputs to a lake are very largely<br />

determined by the nature of the soil, the<br />

land use, and the amounts and patterns of<br />

water flow within the drainage basin . Since<br />

the major nutrient inputs may be made by the<br />

streams, rather than by shore drainage, events<br />

quite some distance from the lake can exert<br />

major effects. The Biome study provides<br />

greater opportunity for evaluating terrestria l<br />

21


inputs into the lakes . Recent studies have<br />

shown this kind of information to be necessary<br />

in considering effects of land use .<br />

The terrestrial scientists should not be unmindful<br />

of the effect that aquatic bodie s<br />

exert on the surrounding land mass . Man 's<br />

needs for water have resulted in virtually all<br />

civilizations developing in proximity to rive r<br />

drainage basins . Locally, the impact of water<br />

on man's land use is also obvious . Seattle<br />

developed into a major city because of it s<br />

location and port facilities on Puget Sound .<br />

Lake Washington has had an impact o n<br />

Seattle by preventing growth eastward in a<br />

contiguous fashion . A great desire for water -<br />

front property has prompted the developmen t<br />

of homes and parks all along its shore front ,<br />

and therefore brought about the necessity o f<br />

bridges and major road systems to permi t<br />

residents of the east shore to have convenien t<br />

access to Seattle . Similarly, the existence o f<br />

Lake Sammamish and its recreational an d<br />

esthetic assets have created the developmen t<br />

of parks, communities, and therefore, roads ,<br />

along its shore . Because Chester Morse Lak e<br />

serves as a water supply for Seattle, the entir e<br />

watershed of this lake has been fenced off an d<br />

been inaccessible to the public for 60 years .<br />

Land use has been restricted to controlle d<br />

logging designed so as to protect wate r<br />

quality. Within the Chester Morse watershed ,<br />

Findley Lake has been protected incidentally .<br />

Aquatic bodies also influence terrestrial<br />

communities by the frequency and extent o f<br />

floods. The temptation to build on level land<br />

has resulted in construction on flood plain s<br />

and the subsequent need for dams and othe r<br />

flood control measures . These in turn have<br />

altered the terrestrial environment and th e<br />

potential uses of the land .<br />

The distribution of terrestrial wildlife is<br />

also influenced to a very great extent by th e<br />

availability of the water supply . A host of<br />

water birds collect their food from the water<br />

but otherwise nest and conduct the nonfeeding<br />

parts of their lives on the land . In addition,<br />

there are animals such as otters (Lutra<br />

canadensis), racoons (Procyon lotor), and<br />

bears (Ursus spp.), which take a rather substantial<br />

amount of their food from aquatic<br />

bodies, especially after the salmon spawning .<br />

Over geological timespans, water has a<br />

major impact on the land . The streams an d<br />

rivers are major sculptors of the terrestrial<br />

topology. Lakes are not merely a sink for up -<br />

land nutrients but may become the source o f<br />

future land . Much of our prime agricultura l<br />

land is lake fill . The occurrence of peat in th e<br />

local drainage system gives further evidenc e<br />

that much of our present land was previousl y<br />

occupied by an aquatic community .<br />

Characteristics of the<br />

Four Lakes<br />

The four lakes in the Coniferous <strong>Forest</strong><br />

Biome 's Cedar River intensive site-Washington,<br />

Sammamish, Chester Morse, and<br />

Findley-provide contrasting conditions (fig .<br />

1) . The physical parameters of the lakes are<br />

shown in table 1 .<br />

1. Lake Washington (fig . 2), the lowermost<br />

lake, has a documented history of eutrophication<br />

and recent sewage diversion .<br />

The dominant fish is the anadromous<br />

sockeye salmon (Oncorhynchus nerka)<br />

which was introduced about 30 years<br />

ago . Other anadromous and resident fis h<br />

populations also occur here. The lake is<br />

used intensely for recreation, including<br />

sports fishing . The shores of the lake are<br />

largely urbanized .<br />

2. Lake Sammamish (fig . 3) represents an<br />

intermediate condition, as it was oligotrophic<br />

until very recently, underwent<br />

limited eutrophication as the forest wa s<br />

replaced by an increasing urban community,<br />

and is currently undergoing<br />

sewage diversion. The response of this<br />

lake to sewage diversion has been substantially<br />

less dramatic than that of Lake<br />

Washington . Surprisingly, the hypolimnion<br />

is anerobic much of the summer .<br />

Some of the same fish species, includin g<br />

sockeye salmon, occur here as in Lak e<br />

Washington, but in smaller numbers .<br />

This lake drains into Lake Washington<br />

from the north via Sammamish Slough .<br />

3. Chester Morse Lake (fig . 4) is the upper -<br />

most of the two large lakes in the Ceda r<br />

River watershed . Although the lake level<br />

22


Figure 1 . Map of Lake Washington drainage.<br />

Table 1.--Physical characteristics of lakes in Lake Washington drainage basi n<br />

Lake<br />

Maximum<br />

depth<br />

Mean<br />

depth<br />

Water<br />

Area Volume Length Elevation exchange<br />

tim e<br />

m m km 2 km 3 km m years<br />

Findley 30' 10 1 0.09 0.0009' 1,13 1<br />

Chester Morse 35 13 .0 6.54 .085 8 .1 473 0 . 3<br />

Sammamish 31 17 .7 19 .8 .350 12 .9 12 2. 2<br />

Washington 64 33.0 87 .6 2 .884 31 .5 8 .6 3 .0<br />

' Tentative values .<br />

2 3


was slightly raised in 1902, the lake in<br />

other respects is the least disturbe d<br />

major lake in the state . Access to th e<br />

lake has been rigidly controlled sinc e<br />

1911 and no fish have been planted o r<br />

harvested since that time. It contains no<br />

anadromous fish, but has resident populations<br />

of rainbow (Salmo gairdneri) ,<br />

Dolly Varden (Salvelinus malma), and<br />

whitefish (Prosopium sp .) . The shore is<br />

completely undeveloped .<br />

4. Findley Lake (fig . 5) is among several<br />

small lakes in a totally unmanipulate d<br />

watershed . It and its watershed are being<br />

used in an integrated terrestrial-aquati c<br />

study . The land area consists of undisturbed<br />

mature coniferous forest. Th e<br />

lake has no resident fish population bu t<br />

has zooplankton and salamanders i n<br />

moderate abundance .<br />

Specific tasks of the lake-stream study in<br />

the Lake Washington-Cedar River watershe d<br />

include assessment of (1) terrestrial influences,<br />

including forest and urban land use ,<br />

(2) nutrient budgets, (3) sources of energy t o<br />

support the food chain, (4) community structure<br />

and metabolism, (5) effects of food avail -<br />

ability on successive trophic levels, (6) effects<br />

of fish on lower trophic levels, (7) effects o f<br />

anadromous fish on other fish populations ,<br />

and (8) outputs to the marine community b y<br />

anadromous fish .<br />

Research Program<br />

In 1971, fieldwork was initiated on chemical<br />

budgets, primary and secondary productivity,<br />

and fish populations, in coordinatio n<br />

with other concurrent studies. A literature<br />

compilation of past aquatic studies in th e<br />

watershed was completed and the transfer o f<br />

pertinent data to the IBP data bank initiated .<br />

A major hydrological study of the lake-river<br />

system was initiated by another agency<br />

(R I B C O, River Basin Coordinating Committee)<br />

. Decomposer studies were organized<br />

in conjunction with the terrestrial program. A<br />

preliminary interface study on Findley Lake<br />

watershed was accomplished . A generalized<br />

trophic level model was developed and its<br />

shortcomings were analyzed . Modeling efforts<br />

on existing data from other lakes wer e<br />

initiated by the modeling group .<br />

In 1972 the fieldwork mentioned above ha s<br />

been expanded and fieldwork on decomposition<br />

and nutrient regeneration processes has<br />

been initiated . Field effort and modeling<br />

structures are reorganized into subsystems o f<br />

bottom related processes, water colum n<br />

processes, and higher consumer dynamics .<br />

Models of optimum strategies and density<br />

dependent relationships are being developed .<br />

Detailed descriptive models of the lakes wil l<br />

be completed . Data sets at potential extrapolation<br />

sites are being identified and a<br />

coordination framework was established .<br />

In 1973 the full field program will continu e<br />

with phasing and evaluation of specific<br />

studies, including elimination of those whic h<br />

have either met their objectives or which are<br />

not profitable lines of pursuit . Greater<br />

emphasis will be placed on using hydrologica l<br />

data and interfacing with terrestrial aspects of<br />

the program as well as on perfecting sub -<br />

system models and linking them together int o<br />

a more complete model . Coordination with<br />

extrapolation site programs will be increased ,<br />

including data set analysis .<br />

In 1974, successful and fruitful program s<br />

will be continued as necessary, with phasin g<br />

of research in many instances to accomplish<br />

tasks deferred for orderly sequential progression<br />

under limited funding . Major efforts wil l<br />

be directed to integration of all work and to<br />

use of extrapolation data for model validatio n<br />

and generalization of intensive site results .<br />

Focus of Research Program<br />

The stream-lake study will address a number<br />

of questions) which have relevance to<br />

both science and analysis of environmenta l<br />

impacts. These questions are concerned with :<br />

(1) the mechanisms by which purely physica l<br />

constraints, such as structure of lake basin,<br />

These questions and the subsystem models have<br />

been developed with the help of the modeling group .<br />

Mr. Larry Male of the Center for Quantitative<br />

Science, University of Washington, is particularl y<br />

acknowledged for his contributions to these ideas .<br />

24


OUTLE T<br />

PINS<br />

LANE<br />

LYIMIIMP<br />

MICOGS<br />

LANE<br />

LAKE IAMMAMIS H<br />

DEPTH IN EETEA 1<br />

LAKE WASHINGTO N<br />

DEPTH IN METERS<br />

o 2 ..<br />

0 I E km Figure 3 . Map of Lake Sammamish .<br />

Ip►OIA N<br />

<strong>ON</strong>EER<br />

Figure 2 . Map of Lake Washington .<br />

Figure 4. Map of Chester Morse Lake. Figure 5 . Map of Findley Lake .<br />

25


climate, surrounding terrestrial environment ,<br />

prevailing winds, lake circulation dynamics ,<br />

rates of water and inorganic nutrient inflow<br />

and outflow, and chemical precipitation of<br />

essential nutrients control the productivity ,<br />

water quality, and community structure of<br />

aquatic ecosystems ; (2) the dynamics of nutrient<br />

and energy cycling through both biological<br />

and physical processes ; (3) the<br />

mechanisms by which community structure<br />

may inhibit or enhance overall productivity<br />

and water quality ; (4) the response of complex<br />

aquatic ecosystems to a variety of<br />

natural and artificially created disturbances ;<br />

and (5) the mechanisms these systems have<br />

evolved for coping with these disturbances .<br />

The nature of these questions justifies th e<br />

coordinated program which is planned. Some<br />

of the questions this program will bear upo n<br />

and how they intend to be tackled will now<br />

be discussed .<br />

The phytoplankton and bacterial communities<br />

in lakes not only provide the basis fo r<br />

complex food webs, they essentially contro l<br />

the overall water quality and the apparent<br />

state of eutrophication . The sustained abundance<br />

and species composition of the bacterial<br />

and phytoplankton communities depend<br />

to a large extent upon the rate at whic h<br />

they are supplied with essential inorganic<br />

nutrients and energy sources. Most of these<br />

compounds are introduced from the surrounding<br />

terrestrial environment and the<br />

atmosphere through biological fixatio n<br />

processes . The rate at which a lake is supplied<br />

with allochthonous materials may determin e<br />

its productivity structure . Several mechanisms<br />

may be responsible for determining a lake' s<br />

response to allochthonous nutrient input .<br />

These mechanisms relate both to the suppl y<br />

of nutrients to the phytoplankton and bacteria<br />

communities and to the physical an d<br />

biological forces which influence the production<br />

dynamics of these communities .<br />

1 . Dynamics of release of inorganic nutrients<br />

from allochthonous organic material<br />

of varying nutrient richness : The first indication<br />

is that nutrients in nutrient<br />

poor organic material, e .g., wood, are released<br />

slowly. Since the four lakes in thi s<br />

study can be characterized by the nature<br />

and extent of allochthonous input, thi s<br />

mechanism can be studied .<br />

2. Nutrient availability as a function o f<br />

thermal distribution, circulation dynamics,<br />

and cycling dynamics of other chemical<br />

compounds : Vertical density gradients<br />

(determined by vertical therma l<br />

distribution gradients) can control th e<br />

availability of essential nutrients to autotrophic<br />

organisms . The density gradient s<br />

function as a barrier to the mixing o f<br />

nutrients from deeper waters into th e<br />

productive surface waters . Since the<br />

density barrier also inhibits the passag e<br />

of oxygen from the surface into th e<br />

deeper waters the deeper waters some -<br />

times become depleted of oxygen . Thi s<br />

anerobic condition enhances the solubility<br />

of compounds which bind phosphorus<br />

in forms unavailable to phytoplankton<br />

. Thus, the deep waters may be<br />

rich in nutrients which are relatively unavailable<br />

to autotrophic organisms . Th e<br />

four lakes in this study vary greatly i n<br />

the formation of their density gradient s<br />

and thus their capacity to trap nutrients .<br />

This theory is only partially applicabl e<br />

to any particular lake since there exis t<br />

other mechanisms for distributing nutrients.<br />

The shape of the lake basin, direction<br />

and strength of prevailing winds ,<br />

internal circulation currents, disturbanc e<br />

of the bottom sediments by mechanica l<br />

mixing, and dynamics of nutrient trans -<br />

port through the reduced layer of th e<br />

sediment are all mechanisms which ar e<br />

being studied to bear upon the proble m<br />

of nutrient cycling .<br />

3.Biological cycling of nutrients : Since the<br />

communities of phytoplankton and bacteria<br />

in lakes are often limited by the<br />

availability of essential nutrients they<br />

have tended to evolve elaborate mechanisms<br />

for the conservation and recyclin g<br />

of nutrients. In the absence of these<br />

mechanisms, production would mos t<br />

likely be nil. Nutrients cycle through<br />

communities of phytoplankton, zoo -<br />

plankton, bacteria, protozoans, and<br />

littoral aquatic plants. The nature of<br />

these cycles and how they operate with -<br />

26


in the restrictions of the physical environment<br />

are being tackled . Since the<br />

four lakes in this study have evolved to<br />

their present structure in the presence of<br />

different nutrient stresses we might<br />

expect them to exhibit various capacitie s<br />

to conserve and cycle nutrients .<br />

4. Effect of community structure on productivity<br />

and distribution of energy : Th e<br />

production of herbivores (zooplankton )<br />

cannot be explained entirely by production<br />

of primary producers (phytoplankton).<br />

The physical form of the production<br />

(i.e., size distribution or colonia l<br />

form of algal cells) may preclude consumption<br />

by zooplankton . However ,<br />

zooplankton may in fact contribute to<br />

changes in the physical form of primar y<br />

production by being size-selective<br />

feeders. Since changes in the form an d<br />

species composition of phytoplankto n<br />

communities are indices of pollution an d<br />

eutrophying environments this process i s<br />

being investigated .<br />

In the same way that changes in th e<br />

physical form of phytoplankton production<br />

may control the production of zoo -<br />

plankton, the community structure, siz e<br />

distribution, and spatial distribution o f<br />

zooplankton control the feeding behavior<br />

and subsequent production o f<br />

predators .<br />

The terminal link in the food web of three<br />

of the study lakes is the community of fis h<br />

species . Much literature has been devoted to<br />

models which try to explain the populatio n<br />

dynamics of these organisms without incorporating<br />

their interaction with the environment<br />

. The communities of fish can influenc e<br />

the nature of production by altering th e<br />

community structure of other vertebrate an d<br />

invertebrate communities upon which the y<br />

feed and indeed may alter the entire community<br />

(Hurlbert et al . 1972). Their own production<br />

dynamics are related to the structure<br />

and abundance of the prey communities an d<br />

also to mortality dynamics which are deter -<br />

mined by the state of the physical environment<br />

at critical life stages . Three projects wil l<br />

bear upon the questions related to the fis h<br />

community .<br />

Presently, a three submodel system (figs . 6<br />

and 7) is being used to develop the propose d<br />

field studies ; these are organized around water<br />

column, bottom related, and higher consumer<br />

processes . These subsystems were chose n<br />

because the couplings within each are mor e<br />

tightly linked and the span-of-time resolution s<br />

OVERALL SYSTE M<br />

Figure 6 . Overall system categories.<br />

LAKE SUBSYSTEM (C)<br />

NIGHE R<br />

C<strong>ON</strong>SUME R<br />

DYNAMIC S<br />

(C? ,<br />

Figure 7. Lake subsystem categories.<br />

27


more similar than between the different subsystems<br />

. For example, the processes in the<br />

water column involving exchanges betwee n<br />

nutrients, algae, shortlived zooplankton, and<br />

bacteria can better be handled by a singl e<br />

team than by several teams, each interested i n<br />

a specific trophic level. Further, this information<br />

may have to be handled on a daily o r<br />

hourly basis, whereas the higher consumer<br />

production, e .g., fish, will be handled on a<br />

seasonal basis .<br />

Studies Planned for 1973-74<br />

The individual studies tentatively planne d<br />

for 1973 and 1974 for the four lakes are<br />

shown in table 2 . These are described briefl y<br />

here. The interface study specific to th e<br />

Findley Lake watershed is described in th e<br />

previous paper .<br />

Water Column Process Studie s<br />

The categories and relationships are show n<br />

in figure 8. Much of the past work in th e<br />

Cedar River drainage lakes defined the annua l<br />

nutrient supply to Lake Sammamish an d<br />

Figure 8 . Categories and relationships of sub-syste m<br />

(C)a ; water column processes .<br />

documented its rate of recovery compared t o<br />

that of Lake Washington as a result of nutrient<br />

diversion. Recently, limnological conditions<br />

have been monitored in Chester Mors e<br />

and Findley Lakes to compare their trophi c<br />

character with that of Lakes Sammamish an d<br />

Washington . Ultimate objective is to develop a<br />

model with enough generality to encompas s<br />

the range in trophic character now observabl e<br />

among the lakes .<br />

Lake Washington responded rapidly to a<br />

diversion of over one-half its annual supply o f<br />

phosphorous . Edmondson (1970) has documented<br />

its recovery to a trophic state simila r<br />

to that recorded over 20 years ago in a matte r<br />

of only 3 years after the completion and 7<br />

years after the beginning of sewage diversion .<br />

Lake Sammamish, a mesotrophic lake wit h<br />

similar flushing time, located only 10 miles t o<br />

the east of Lake Washington, has no t<br />

responded noticeably in 3 years following<br />

nutrient diversion . Possible reasons for this<br />

difference in rate of response involve th e<br />

morphometry of the two lakes .<br />

The nutrient input to the lake and the<br />

levels of dissolved inorganic nutrients will b e<br />

related to the growth and production of<br />

phytoplankton and its consumption by zoo -<br />

plankton . These studies will be in sufficient<br />

detail to examine changes in species composition<br />

and the importance of size categories a s<br />

they may relate to optimum feeding strategies.<br />

More detailed, but less frequent estimates<br />

of the flow of carbon through all<br />

components of the food web will be made by<br />

the techniques of Saunders (1969) usin g<br />

radioactive bicarbonate, radioactive detritus ,<br />

and radioactive dissolved organics . Both of<br />

these field-oriented studies will be correlate d<br />

with the modeling on the dynamics of nutrient<br />

distribution and flow through aquatic ecosystems.<br />

A deferred study concerns the rol e<br />

of fungal parasites on algae to explore the<br />

hypothesis that algal blooms may cease<br />

through disease processes, rather than throug h<br />

the limitation of nutrients .<br />

The mathematical models which have been<br />

developed to explain the production dynamic s<br />

of phytoplankton production admit that the<br />

concentration of phytoplankton may vary continuously<br />

from the surface to the lake bottom .<br />

28


Table 2.-Proposed lake studies (by investigator and abbreviated title )<br />

Aquatic Director: R. L . Burgner<br />

Lake Studies : F. B. Taub, Chairman<br />

Department University Study<br />

Water Colum n<br />

(E. B. Welch, Chairman )<br />

1. E. B. Welch Civil Engineering Univ. Washington<br />

P. R. Olson <strong>Forest</strong> Resources Univ. Washington<br />

Zooplankton-productio n<br />

2 . E. B . Welch Civil Engineering Univ. Washington Phytoplankton-production<br />

3. B. Lighthart Inst. Freshwater Studies Western Washington Carbon web<br />

4. B. Lighthart Inst. Freshwater Studies Western Washington Bacteria<br />

5. T. Packard Oceanography Univ. Washington Oxidation-respiratio n<br />

6. F . B. Taub Fisheries Univ. Washington<br />

J. T. Staley Microbiology Univ. Washington<br />

Nitrogen transformation<br />

7.*D. E. Spyridakis Civil Engineering Univ. Washington<br />

R. F. Christman Civil Engineering Univ. Washington<br />

Nutrient budgets<br />

Bottom-Related Processes<br />

(D. E. Spyridakis, Chairman)<br />

8. D. E. Spyridakis Civil Engineering Univ. Washington<br />

R. F. Christman Civil Engineering Univ. Washington<br />

Nutrient budget s<br />

9. F . B. Taub Fisheries Univ. Washington Detrital inpu t<br />

10. J. R. Matches Fisheries Univ. Washington Bacterial decomposition<br />

11. M. M. Pamatmat Oceanography Univ. Washington Oxidatio n<br />

Higher Consumers<br />

(R. R. Whitney, Chairman )<br />

12. R. L. Burgner Fisheries Univ. Washington<br />

Limnetic fis h<br />

R. E. Thorne Fisheries Univ. Washington<br />

13. A. C. DeLacy Fisheries Univ. Washington Limnetic fish feed<br />

14. R. S. Wydoski Fisheries Univ. Washington<br />

Benthic-littoral fis h<br />

R. R. Whitney Fisheries Univ. Washington<br />

*See also Bottom-Related Processes.<br />

29


The net rate of primary production at any<br />

particular depth depends upon the temperature,<br />

light intensity, essential nutrient concentration,<br />

and zooplankton grazing pressure a t<br />

that depth. The vertical distribution an d<br />

abundance of phytoplankton is thus a functio n<br />

of varying net production, sinking rates of algae<br />

and mixing dynamics of the water column . Th e<br />

total production per unit area of lake surfac e<br />

may be obtained by integrating the concentration<br />

of phytoplankton from top to bottom .<br />

The next step is to incorporate the size selective<br />

feeding of zooplankton and determine it s<br />

effect upon production (since algal cells o f<br />

different sizes possess varying capacities t o<br />

absorb nutrients) .<br />

The effective grazing rate by zooplankto n<br />

is modeled as a function of the phytoplankton<br />

density (at high densities the zooplanktons<br />

feeding structures become saturated) .<br />

The efficiency of assimilation of phytoplankton<br />

by zooplankton is also allowed to depen d<br />

upon algae abundance (efficiency decrease s<br />

with increasing abundance) .<br />

The overall model relating the phytoplankton,<br />

zooplankton, and nutrients is a system o f<br />

partial differential equations (in time and<br />

depth) .<br />

The submodel describing the physical<br />

cycling of nutrients (developed in the botto m<br />

related processes group) is an integral part of<br />

the phytoplankton, zooplankton submodels .<br />

The number and kinds of heterotrophi c<br />

bacteria will be assessed to give some information<br />

on the constancy of community structure<br />

of the decomposers . The utilization an d<br />

ultimate destruction of fixed organic material<br />

for the entire system will be estimated by<br />

respiration studies and verified by substrat e<br />

disappearance . The respiration an d<br />

heterotroph abundance values will also be correlated<br />

with the regeneration of nutrients .<br />

The nutrient budget and effective concentrations<br />

will be assessed . All of the above studie s<br />

will be carried out in units of biomass or wil l<br />

be convertible to biomass and units of carbo n<br />

by calculation of size distribution. Determinations<br />

of the magnitude of nitrogen transformation<br />

within each of the above compartments,<br />

as well as the search for evidence nitrogen<br />

fixation and its opposite process, denitri -<br />

fication, will be made .<br />

The detailed carbon cycle through th e<br />

phytoplankton, zooplankton, bacteria and<br />

pools of dissolved inorganic carbon, dissolved<br />

organic carbon, and detritus are bein g<br />

modeled initially with a varying rate compartment<br />

system. The rates of flow between th e<br />

compartments are to be functions of light ,<br />

temperature, and P0 4 and NO 3 concentrations.<br />

Later the rates will also be modeled a s<br />

functions of 0 2 and Si concentration. Thi s<br />

model will allow an analysis of what paths o f<br />

the cycling process limit the rate of<br />

production .<br />

Bottom Related Process Studie s<br />

The compartments are shown in figure 9 .<br />

The necessary chemical analyses to assess thi s<br />

aspect of the nutrient budget will also be<br />

done. Inputs of detrital materials to the sediments<br />

and their potential incorporation int o<br />

higher food chains via bottom invertebrates t o<br />

the fish will be assessed . The oxidation of<br />

organic material will be assessed by respiration<br />

rates of the sediments. The disappearanc e<br />

of substrates such as cellulose and chitin an d<br />

information on the bacteria associated wit h<br />

Figure 9 . Categories and relationships of sub-syste m<br />

(C)b ; bottom related processes.<br />

30


these processes will be measured and correlated<br />

with nutrient regeneration and respiration<br />

rates. The production of littoral material s<br />

will be assessed and compared with limneti c<br />

production and terrestrial inputs . The combined<br />

information on limnetic, benthic, an d<br />

littoral photosynthesis and respiration wil l<br />

permit comparisons of P/R ratios for th e<br />

various lakes within this study as well as fo r<br />

wider comparisons .<br />

A sophisticated theoretical model describing<br />

the complex physical and biological cycle s<br />

of inorganic nutrients has been proposed . Thi s<br />

model incorporates the features of densit y<br />

gradients as a barrier to mixing, circulatio n<br />

currents, chemical precipitation of essentia l<br />

nutrients, transport and release of nutrient s<br />

from the bottom sediments as a function o f<br />

0 2 concentration, pH, and bacterial decomposition,<br />

nutrient input in organic and in -<br />

organic form from allochthonous sources, an d<br />

atmospheric input through biological fixatio n<br />

processes . This theoretical model will be mad e<br />

more precise and useful by formulating it as a<br />

collection of well-defined assumptions an d<br />

mathematical equations. This model wil l<br />

allow an analytic comparison of the cyclin g<br />

processes in the four lakes .<br />

Higher Consumer Process Studie s<br />

Compartments are shown in figure 10 . Th e<br />

measurements of population parameters o f<br />

young sockeye and other limnetic feeding fis h<br />

will provide comparative data on seasona l<br />

changes in numbers, biomass, growth, an d<br />

mortality rates relative to possible controllin g<br />

factors, including recruitment, physical an d<br />

chemical environment, food availability an d<br />

characteristics, behavior and competition ,<br />

predation and other removal . Characteristics ,<br />

habits, and interactions of the benthic an d<br />

littoral fish will be studied . Study of th e<br />

benthic food supply will be temporarily deferred.<br />

The interaction between the benthi c<br />

and littoral fish and the limnetic feeding fis h<br />

is receiving emphasis . In all cases, there will b e<br />

a search for understanding of feeding strategies<br />

and growth dynamics as they influenc e<br />

community structure of the zooplankton an d<br />

fish .<br />

Figure 10 . Categories and relationships of sub-syste m<br />

(C)c; higher consumer dynamics .<br />

The vast food webs, typical of lake ecosystems,<br />

seem to be so complicated as to defy<br />

description . The higher consumers not only<br />

respond physiologically to their surroundin g<br />

physical environment, they exhibit complex<br />

behavioral mechanisms to cope with changin g<br />

food supplies and varying physical conditions .<br />

A conceivable model for this system is a varying<br />

rate compartmental model where the rate s<br />

of transfer among various species depend o n<br />

the physical environment . Another approach<br />

which is finding favor among theoretical<br />

ecologists is the theory of optimal strategies .<br />

The essences of this theory is that selective<br />

processes and adaptive mechanisms tend t o<br />

produce populations of organisms which<br />

strive to maximize their energy intake subjec t<br />

to the physical restrictions imposed by the<br />

environment and their own physiological an d<br />

morphological limitations . If the community<br />

structure of the prey organisms available to a<br />

predator is known then a mathematical representation<br />

of an optimal strategy model wil l<br />

predict the amounts and kinds of prey a<br />

predator will ingest. Coupled with models<br />

which describe the mortality structure of<br />

populations the optimal strategy model will<br />

effectively handle the questions concernin g<br />

31


community structure and energy flo w<br />

dynamics .<br />

It is obvious the subsystems above are artificial<br />

and that coordination of techniques an d<br />

measurements is necessary. For example, th e<br />

nutrient budget of the lake cannot be developed<br />

without the water column and th e<br />

bottom processes information, nor without<br />

the hydrological studies being coordinated<br />

with RIBCO and assessment of nutrien t<br />

sources .<br />

The present (1972) studies are necessarily<br />

descriptive, to supply the material for the firs t<br />

rough models . As the inadequacies of the<br />

crude models become apparent during the<br />

course of 1972, there will be increasing inter -<br />

play between the modeling and the field<br />

investigations and programs correspondingl y<br />

modified .<br />

Acknowledgments<br />

The work reported in this paper was sup -<br />

ported in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferous Fores t<br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution<br />

No . 21 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Edmondson, W. T. 1970. Phosphorous nitrogen<br />

and algae in Lake Washington after<br />

diversion of sewage . Science 169 : 690-691 .<br />

Hurlbert, Stuart H., Joy Zedler, and Deborah<br />

Fairbanks . 1972. Ecosystem alteration b y<br />

mosquitofish (Gambusia affinis) predation .<br />

Science 175: 639-641 .<br />

Margalef, Ramon. 1968. Perspectives i n<br />

ecological theory. 111 p., illus. Chicago ,<br />

London: Univ . Chicago Press .<br />

Odum, Eugene P . 1969. The strategy of ecosystem<br />

development . Science 164 :<br />

262-270 .<br />

Saunders, George W . 1969. Some aspects of<br />

feeding in zooplankton . In Eutrophication ,<br />

p. 556-573. Nat. Acad. Sci., Washington,<br />

D.C .<br />

32


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

The modeling process relating to<br />

questions about coniferous<br />

lake ecosystems<br />

A bstract<br />

Douglas M . Eggers<br />

Fisheries Research Institute<br />

University of Washingto n<br />

an d<br />

Larry M . Mal e<br />

Center for Quantitative Scienc e<br />

University of Washingto n<br />

The role of sound conceptualization and meaningful questions in the modeling process is discussed . Th e<br />

salient features of lake communities are reviewed. Included are factors which must be considered in answering<br />

questions involving the higher consumers of lake ecosystems .<br />

Before attempting an analytical model on e<br />

must conceptualize the system being modeled .<br />

If one 's conceptualization is good, then one is<br />

able to ask relevant and meaningful questions .<br />

Since ecosystems are infinitely complex, on e<br />

must make assumptions and simplifications in<br />

constructing analytical models which are with -<br />

in our capabilities . Because of these assumptions<br />

and simplifications it is impossible t o<br />

construct general ecological models-a general<br />

model being one which will answer or anticipate<br />

all questions. Therefore models must be<br />

constructed relevant to specific questions .<br />

A set of meaningful questions should be th e<br />

foremost component in the strategy which a<br />

worker adopts in the construction of a model .<br />

These questions allow the modeler to make set s<br />

of assumptions and simplifications which mak e<br />

a model both meaningful and workable .<br />

Considering a coniferous lake ecosystem, if<br />

one is interested in questions of seasonal succession<br />

of phytoplankton or perhaps, more<br />

long-term successional changes along a gradient<br />

lake nutrient level, then one must construct<br />

models which differentiate the algae with<br />

respect to species . This model must incorporate<br />

algal adaptive mechanisms to critical resources .<br />

On the other hand, if the question is one of<br />

primary production, then it may be worthwhile<br />

to construct models where the algae are lumpe d<br />

together into a uniform mass of quasi -<br />

organisms . This is a fundamental assumption of<br />

the Riley et al. (1949) model . But without<br />

species differentiation a means of handling the<br />

interesting questions in ecology of competition,<br />

predation, selection, and adoption i s<br />

destroyed .<br />

The following are some of the factors w e<br />

have considered in developing our conceptua l<br />

framework of a coniferous lake ecosystem .<br />

The structure and dynamics of the salient<br />

features of aquatic communities are reviewed .<br />

Attention is given to processes which hav e<br />

traditionally been neglected in quantitative<br />

models .<br />

Generation times in the algae are short .<br />

Thus the selection process may be rapid . This<br />

can give rise to rapid shifts in the species composition<br />

of phytoplankton as the environment<br />

changes (Hutchinson 1967) . Because of seasonal,<br />

diurnal, temporal and other spatial<br />

patterns of environmental change, selectio n<br />

processes are in a continual state of flux . The<br />

patchiness in spatial distribution of the organisms<br />

plus interruption in selection processes<br />

by continually changing environmental<br />

conditions, allows many species to coexist i n<br />

the lake ecosystem . The state of being th e<br />

33


est adapted organism vacillates betwee n<br />

species with time as the environment changes .<br />

The level of nutrients in a lake ultimatel y<br />

determines the level of phytoplankton whic h<br />

the lake can support . Nutrients are utilized b y<br />

plants as components of structural molecules ,<br />

such as DNA, RNA and enzymes . If nutrien t<br />

levels are in excess of structural demands an d<br />

other conditions are favorable for photosynthesis,<br />

then reproduction can occur . Nutrient s<br />

are rare elements in lake ecosystems, very<br />

eutrophicated lakes being exceptions . Aquati c<br />

communities have evolved many mechanisms<br />

fOr their conservation (Pomeroy 1970) . Algae<br />

and littoral plants can absorb nutrients almos t<br />

instantaneously from the water . Aquatic plants<br />

can take up nutrients as they are available and<br />

store them until conditions become more suit -<br />

able for growth when the nutrients are utilized .<br />

Light and temperature conditions necessary for<br />

photosynthesis are much more predictable ,<br />

with definite seasonal and diurnal patterns ,<br />

than nutrient levels, which the algae cells en -<br />

counter more or less randomly . Then, from the<br />

viewpoint of algae, a good strategy to evolve<br />

would be a means of acquiring nutrient s<br />

whenever they occur and wa iting to make use<br />

of them until sufficient light and temperature<br />

conditions (these are predictable) prevail .<br />

In addition to being able to utilize nutrient s<br />

as they occur, the phytoplankton-zooplankto n<br />

community recycles nutrients rapidly . As th e<br />

algae die and sink to the bottom, nutrients are<br />

released in two ways . Autolytical release occurs<br />

by simple diffusion. Mechanical release occurs<br />

when the cell membrane ruptures . Betwee n<br />

25 and 70 percent of the nutrients containe d<br />

in a sinking, dead organism are released i n<br />

these ways (Johannes 1968) .<br />

Zooplankton eat substantial proportions o f<br />

the algae . Herbivores are large in relationship<br />

to their food supply . Since zooplankton are<br />

eating a rich source of essential nutrients ,<br />

they consume many more nutrients than they<br />

need for structural components. The excess is<br />

simply excreted into the water .<br />

The classical role of bacteria, as the principal<br />

agents of nutrient regeneration is questioned<br />

by Johannes (1968) and Pomeroy<br />

(1970). Bacteria accumulate nutrients ove r<br />

the level required in growth, as algae do .<br />

Johannes (1968) believes that protozoa, ofte n<br />

neglected as important organisms in lak e<br />

systems, are responsible for regeneratin g<br />

nutrients by eating bacteria and excreting th e<br />

excess nutrients .<br />

In addition to biological means of nutrient<br />

cycling, many purely physical processes affec t<br />

nutrient distribution in lake systems . Nutrients<br />

enter the lake through the inflow (both surface<br />

flow and seepage) and precipitation . Nutrients<br />

leave the lake through outflow (both surface<br />

flow and seepage) . Sediments are very important<br />

in the nutrient budgets of lakes . They may<br />

act as a trap for essential nutrients, such a s<br />

phosphorous. The phosphorous is tied up b y<br />

iron and precipitated out at high redox potentials.<br />

Decreasing redox potential usually ac -<br />

companies decreasing oxygen concentration .<br />

When the redox potential falls below a certain<br />

level the phosphorous goes into solution, there -<br />

by becoming available to photosynthesizers i n<br />

the waters above . In oligotropic lakes th e<br />

redox potential is always high and phosphorous<br />

bound in falling detritus is lost to th e<br />

system when it reaches the sediments . Th e<br />

cycling of nitrogen is more complicated an d<br />

even more intimately tied in with the redo x<br />

potential . The rate at which essential nutrients<br />

are recycled through the community, interchanged<br />

with the sediments, and lost and<br />

replenished through inflow and outflow deter -<br />

mines the productivity of the lake . The<br />

response of this system to perturbation, measured<br />

not only by the level of productivity but<br />

by changes in community structure, is th e<br />

central theme of our research effort .<br />

Production is usually defined as the total<br />

elaboration of organic matter by photosynthetic<br />

organisms in a specified time period ,<br />

while productivity is the production per unit<br />

time. Photosynthetic rate depends principall y<br />

upon light intensity, temperature, and essential<br />

nutrient concentration (Riley 1946 ,<br />

Rhyther 1956, Tailing 1961) . Photosynthetic<br />

rate is different for different species of alga e<br />

(Talling 1955) . The production under a uni t<br />

area of lake surface is the sum over species o f<br />

the integral with respect to time and depth o f<br />

the algae densities times their respectiv e<br />

photosynthetic rates . The above generalization<br />

may appear overly simple ; however, the<br />

34


complexity of lake productivity will become<br />

apparent.<br />

The exponential decay of light intensity is<br />

a function of depth . Temperature varies with<br />

depth . Although factors leading to lake stratification<br />

are well understood, the construction<br />

of predictive models for temperature distributions<br />

in lakes has only recently begun . Light<br />

intensity varies with time, because of seasonal<br />

and diurnal patterns plus changing weathe r<br />

conditions . There are vertical currents due t o<br />

advection and eddy diffusion in lakes (Rile y<br />

et al . 1949) . These currents cause temperature,<br />

nutrient concentrations, and the population<br />

densities to vary with time . We must als o<br />

consider feedback mechanisms such as selfshading<br />

(Tailing 1960), and local depletion of<br />

nutrients by algae .<br />

Another interesting feedback mechanis m<br />

involves the relation between phytoplankto n<br />

and zooplankton densities . The rate of grazing<br />

on phytoplankton is determined by zooplankton<br />

density and its relative grazing rate (Riley<br />

1947) . On the other hand, the relative grazing<br />

rate of zooplankton is a function of phytoplankton<br />

density ; a saturated grazing rate is<br />

reached at high algal population densities<br />

(Holling 1959, Riley 1947) . Zooplankton also<br />

influence nutrient availability by excretin g<br />

soluble nutrients into the water .<br />

In addition to the limnetic plankton communities,<br />

there are, in lake ecosystems, littoral<br />

communities and, in areas deeper tha n<br />

maximal light penetration, benthic communities.<br />

The processes that occur in these communities<br />

are somewhat different than in th e<br />

limnetic community . Littoral communities<br />

are similar to terrestrial communities in many<br />

respects . Emergent and floating forms have<br />

access to atmospheric carbon dioxide . Attached<br />

forms have direct access to nutrients<br />

held in the sediments ; whereas, limnetic algae<br />

are free floating and must depend upon larg e<br />

ratios of surface area to volume and upo n<br />

passive sinking for efficient nutrient absorption.<br />

In littoral areas, herbivores are small ,<br />

relative to plant size . They consume little of<br />

the plant biomass. Littoral plants show seasonal<br />

die back, with much material being de -<br />

composed . As in terrestrial systems, bacteri a<br />

and detritus feeders are the primary agents of<br />

nutrient regeneration .<br />

Hutchinson and Bowen (1950), in a radiophosphorous<br />

study, showed that littoral<br />

plants can soak up and hold a large quantity<br />

of nutrients . Thus limnetic and littoral plant s<br />

do compete for nutrients . The littoral plants<br />

are less efficient than algae in absorbing nutrients<br />

because of their smaller ratio of surfac e<br />

area to volume . The littoral system's ability to<br />

hold nutrients (because of slow nutrient re -<br />

generation by bacterial decomposition an d<br />

relatively small grazing pressure) offsets inefficient<br />

nutrient absorption. Phytoplankton ,<br />

because of their short life span and hig h<br />

predation rate, cannot hold nutrients as long<br />

as littoral plants, but they are very efficient in<br />

absorbing nutrients .<br />

Deep benthic communities are made up o f<br />

detritus feeders that feed upon seston rainin g<br />

down from the productive epilimnion . They<br />

may be important in nutrient regeneration<br />

(Johannes 1968) . These animals constitute<br />

important food sources for higher orde r<br />

consumers .<br />

A promising theory which may be of great<br />

help in generating and answering question s<br />

about higher consumers, is that of feeding<br />

strategies. For a review of a substantial literature<br />

on the subject, see Schoener (1971) . The<br />

essence of the theory of feeding strategies i s<br />

that selective processes tend to produce populations<br />

of organisms which strive to maximiz e<br />

their energy (essential nutriment) intake .<br />

Predators may accomplish this end in severa l<br />

ways; however, it appears reasonable that i t<br />

would pay to eat the largest prey or the easiest<br />

to capture whenever they are encountered ,<br />

chasing relatively small difficult-to-catch prey<br />

only as a last resort . It is currently being<br />

demonstrated that this theory applies to a<br />

large variety of animals. The consequence is<br />

that one is able not only to predict the foo d<br />

habits of individual species but also to estimate<br />

their growth dynamics .<br />

Acknowledgments<br />

The work reported in this paper was supported<br />

by National Science Foundation Grant<br />

GB-20963 to the Coniferous <strong>Forest</strong> Biome ,<br />

35


U .S . Analysis of Ecosystems, International<br />

Biological Program . This is Contribution No .<br />

22 to the Coniferous <strong>Forest</strong> Biome, IBP .<br />

Literature Cited<br />

Holling, C. S. 1959. Some characteristics of<br />

simple types of predation and parasitism .<br />

Can. Entomol . 91 : 385-398 .<br />

Hutchinson, G. E. 1967. A treatise on limnology.<br />

Vol. II. Introduction to Lake<br />

Biology and the limnoplankton . 1115 p . ,<br />

illus. New York : John Wiley & Sons .<br />

and V. T. Bowen. 1950 . Limnological<br />

studies in Connecticut . IX . A<br />

quantitative radiochemical study of th e<br />

phosphorous cycle in Linsley Pond . Ecology<br />

81 : 194-203 .<br />

Johannes, R . E . 1968 . Nutrient regeneration<br />

in lakes and oceans . In M . R . Droop and E .<br />

J. F. Wood (eds .), Advances in microbiology<br />

of the sea, Vol. I, p . 203-213 . New<br />

York : Academic Press .<br />

Pomeroy, L. R. 1970 . The strategy of mineral<br />

cycling . In Annu. rev. of ecol. and syst. ,<br />

Vol. I, p. 171-190. Palo Alto : Annu. Rev . ,<br />

Inc .<br />

Rhyther, J . H. 1956 . Photosynthesis in the<br />

ocean as a function of light intensity .<br />

Limnol. & Oceanogr . 1 : 61-70 .<br />

Riley, G . A. 1946. Factors controlling phytoplankton<br />

populations on Georges Bank . J .<br />

Mar . Res . 6 : 54-72 .<br />

1947 . A theoretical analysis o f<br />

the zooplankton population on George s<br />

Bank . J. Mar. Res . 6: 104-113 .<br />

, H. Stommel, and D . F. Bumpus .<br />

1949. Quantitative ecology of the plankto n<br />

of the western North Atlantic . Bull .<br />

Bingham Oceanogr. Coll. 12 : 1-169 .<br />

Schoener, T. W. 1971. Theory of feeding<br />

strategies . In Annu. rev. of ecol. and syst . ,<br />

Vol. II, p. 369-404 . Palo Alto : Annu. Rev . ,<br />

Inc .<br />

Talling, J. F. 1955. The relative growth rates<br />

of three plankton diatoms in relation t o<br />

underwater radiation and temperature .<br />

Ann. Bot. 14 : 329-341 .<br />

1960 . Self-shading effects in<br />

natural populations of a planktonic diatom .<br />

Wett. Leben 12 : 235-242 .<br />

1961 . Photosynthesis under<br />

natural conditions . Ann. Rev. Plant<br />

Physiol. 12 : 133-154 .<br />

36


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Toward a general model structur e<br />

for a forest ecosystem<br />

W. Scott Overto n<br />

Professor, Departments of<br />

Statistics an d<br />

<strong>Forest</strong> Management<br />

Oregon State University<br />

A primary objective of the Coniferous <strong>Forest</strong><br />

Biome is the development of a family o f<br />

models for a forest ecosystem as a whole . Thi s<br />

paper will discuss the general structures an d<br />

forms which represented our views toward th e<br />

end of 1971 . It is emphasized that these ar e<br />

evolutionary ; at least part of the value of thi s<br />

paper is as a record of the path we took i n<br />

model development . The value of these structures<br />

in furthering ecosystem theory is als o<br />

argued .<br />

Thus, a primary focus of this paper is wit h<br />

respect to ecosystem research at the "tota l<br />

system" level, that is, concerning the ecosystem<br />

as an object . It is another goal of ecosystem<br />

research to study the relations amon g<br />

component parts and further to study th e<br />

system properties of those parts . Populations ,<br />

communities, assemblages, food chains, productivity<br />

processes, and many other sub -<br />

system structures are of interest, in their ow n<br />

right, as are the functional and relationa l<br />

aspects of these structures . In fact, one of th e<br />

great difficulties in modeling the ecosystem as<br />

a (single) object is the strong tendency to<br />

perceive the ecosystem as an assemblage o f<br />

poorly defined subsystems such that th e<br />

couplings are vague, at best, and often entirely<br />

undefined . Model structures are developed<br />

which, it is hoped, will contribute t o<br />

elaboration of these subsystem concepts, a s<br />

well as the concept of the total system as<br />

object.<br />

A brief statement of my view of models ,<br />

and of their role in science seems desirable .<br />

Any particular model for a real world system<br />

(or object) can be decomposed into a set o f<br />

statements regarding that object . Some of<br />

these statements will be definitive, some wil l<br />

represent firm knowledge about the object ,<br />

and some will be in the form of infirm knowledge<br />

or simplifications (which statements w e<br />

usually call assumptions) . If we consider th e<br />

collection of all possible models for an object ,<br />

then the set of all statements of definition<br />

and all statements of firm knowledge represent<br />

all that is known about the object, s o<br />

that at this level the model, the canonical<br />

model, is identical to the state of knowledge<br />

about the object .<br />

Working models usually emphasize som e<br />

elements of knowledge and neglect others, s o<br />

that in practice the models we use are some -<br />

thing less than "complete ." However, most<br />

working models represent, in a very real way ,<br />

the knowledge and theory of a real world<br />

system . Some subset of the total knowledge i s<br />

coupled with some set of simplifying assumptions<br />

and the collection given a form an d<br />

structure by a mathematical or relational<br />

algorithm or convention . The process puts th e<br />

present knowledge into a new form . If th e<br />

new form serves to increase the understandin g<br />

of the system, then the modeling operation i s<br />

37


successful .<br />

Thus we can see modeling as the impositio n<br />

of structure on existing knowledge . As such ,<br />

it is a joint activity between those familia r<br />

with forms and structures and those familiar<br />

with a body of knowledge . In time, the structure<br />

imposed becomes integrated into th e<br />

scientific paradigm of the object, providin g<br />

new insight and a base for extending concepts<br />

. In this view, modeling is an integra l<br />

part of the theoretical advance of a subject<br />

science, not an external, nor peripheral ,<br />

activity engaged in by mathematicians o r<br />

other specialists .<br />

Now I don 't think that I am the only on e<br />

with this view, but there are sufficiently man y<br />

opposing views among modelers and sufficiently<br />

many misconceptions among subjec t<br />

specialists that verbalization of this perspective<br />

is essential in an introduction of th e<br />

present topic . I view our efforts to develop a<br />

total ecosystem model as a contribution to<br />

the scientific paradigm of ecosystems . Ou r<br />

primary activity as modelers is the conceptualization<br />

of theoretical ecosystem structure s<br />

and behavior. The present paper deals wit h<br />

the structural aspects of a developing family<br />

of models for the forest ecosystem at a date<br />

late in 1971, and with the strategy we have<br />

adopted to pursue our objectives .<br />

Approaches to System<br />

Theory and Definition<br />

Klir (1969, 1972) discusses in some detail a<br />

variety of definition forms and approaches t o<br />

developing a general theory of systems . This<br />

seems a productive direction and the following<br />

section is my own interpretation and<br />

elaboration of these passages in Klir . There<br />

may be inadvertent deviations from Klir 's<br />

intent .<br />

The State Variable (or State Space) approach<br />

seems to describe the model structur e<br />

currently most popular in several fields, including<br />

ecology . The essence of this approac h<br />

is that the system is defined as a set of variables,<br />

the state variables, and a set of variable s<br />

representing the environment, in a particular<br />

temporal resolution . A popular version de -<br />

fines compartments, with the state variable s<br />

describing the contents of the compartments .<br />

The definition is completed by specificatio n<br />

of algorithms for change of the state variables<br />

in time .<br />

Klir has proposed a general systems theory<br />

which, although not distinct from the stat e<br />

variable approach, has certain features whic h<br />

apply nicely to study of ecosystems . Specifically,<br />

he identifies five ways in which a<br />

system may be defined :<br />

1. By the set of external quantities and th e<br />

resolution level<br />

2. By the given activit y<br />

3. By the permanent behavio r<br />

4. By the Universe-Coupling structure<br />

5. By the State-Transition structure<br />

The models of the system follow definitio n<br />

forms 3, 4 and 5 . Definition form 1 is used in<br />

the planning stages of a modeling or data<br />

collection activity, and a collection of data i s<br />

a realization of definition form 2. It follows<br />

that definition form 1 is implied by form 2<br />

and by forms 3, 4, and 5. When applied to th e<br />

same system, the forms must be mutually<br />

consistent . In the Coniferous Biome, we hav e<br />

adopted a particular combination of the thre e<br />

model forms for our model of the forest ecosystem.<br />

We view each system (or subsystem )<br />

as modeled at two levels :<br />

1. Holistically, according to its Behavior (o r<br />

State-Transition structure) .<br />

2. Mechanistically, according to its Universe-Coupling<br />

structure .<br />

That is, we view Klir ' s Behavior and S-T<br />

structure as useful forms for characterizing<br />

the holistic behavior of a system "as a n<br />

object," and consider that such holistic characterization<br />

is necessary for each define d<br />

system or subsystem . Further, we consider<br />

that in most cases we will also wish to model<br />

the system according to the U-C Structure ,<br />

that is, as a collection of subsystems, eac h<br />

modeled according to its Behavior and with<br />

the collection coupled in a manner appropriate<br />

to the behavioral forms used .<br />

Some further elaboration of these term s<br />

seems necessary, but a formal definition (as i n<br />

Klir 1969 and Orchard 1972) is inappropriate .<br />

An attempt to informally define some of th e<br />

38


terms follows. The external quantities (or<br />

variables') are those relevant quantities associated<br />

with the attributes of the object. Part<br />

of them are inputs (produced by the environment)<br />

and part are outputs (produced by th e<br />

object). Specification of these quantitie s<br />

constitutes definition form 1 . By definitio n<br />

form 2, an activity is a particular set of value s<br />

representing the system over a particular se t<br />

of time instants .<br />

The permanent behavior is a time invarian t<br />

relation between the outputs of interest and<br />

other quantities. In order to develop this in<br />

general, it is necessary to augment the specified<br />

external quantities by past values of th e<br />

external quantities (i .e., by a memory), the<br />

augmented set of quantities constituting the<br />

principal quantities . The outputs of interest<br />

are then the dependent quantities (a subset of<br />

the principal quantities), and the relation i s<br />

defined between these and the rest of the<br />

principal quantities .<br />

It is necessary to emphasize that this definition<br />

specifies that the principal quantities ar e<br />

generated from the external quantities alone .<br />

If internal quantities (i .e ., quantities defined<br />

on the subsystems but not on the whole) were<br />

allowed, then the state variable approac h<br />

would be a special case of definition form 3 .<br />

As it is, the two approaches (S-V and Behavior)<br />

have special cases in common .<br />

The State-Transition Structure is defined<br />

by recognition of the State as the instantaneous<br />

value of the external quantities, of the<br />

stimulus as the instantaneous value of th e<br />

input quantities and by a time invariant relation<br />

which sends the system from a particular<br />

state and stimulus at time t into the set of<br />

possible states at time t + h . The S-T structure<br />

is a useful special case of the permanent<br />

behavior, and in subsequent treatment wil l<br />

not be explicitly identified .<br />

One last point regarding behavior . In either<br />

form (of behavior) the time invariant relatio n<br />

can be either deterministic (i .e ., the relation is<br />

a mapping) or probabilistic (i .e., the relatio n<br />

is one to many) and in the latter case eithe r<br />

definition (Behavior or S-T structure) must be<br />

' The quantities are defined on the object . Variables<br />

are model counterparts .<br />

augmented by an appropriate probability<br />

function .<br />

Our notation, then, for the holistic definition<br />

of the system will be :<br />

S={Z,M,Y ;R} (1 )<br />

where Z represents instantaneous input<br />

quantities (variables )<br />

Y represents instantaneous output<br />

quantitie s<br />

M represents memory quantities (past<br />

values of Z or Y quantities )<br />

and R is the appropriate time invariant relation<br />

between Y, on the one hand<br />

and Z and M on the other .<br />

It is the specification of the relation, R ,<br />

which is the goal of ecosystem research at thi s<br />

level. R is the time invariant relation whic h<br />

permits description of the behavior of th e<br />

system (definition form 3) in place of th e<br />

simple "data record" provided by definitio n<br />

form 2 . In fact, it is a fair statement that a<br />

field data collection relates to system definition<br />

by the second form and that it is the rol e<br />

of data analysis and theoretical synthesis t o<br />

progress to forms 3 and 4 .<br />

The behaviorial representation of the system<br />

S as an object, as by (1), is followed by<br />

its perception as a coupled collection of sub -<br />

systems (sub-objects )<br />

S = {So, S I , Sk ; C } (2)<br />

where C is the specification of the coupling s<br />

among the subsystems S i , . .., Sk. Typically ,<br />

the coupling between two subsystems, say S i<br />

and Sj will be the coupling variables defined<br />

by ZjnYj and ZjnYi . That is, some of the<br />

outputs of subsystem Sj become inputs o f<br />

subsystem Si and vice versa, and these coupling<br />

variables define the couplings among th e<br />

subsystems .<br />

In Klir's theory, it seems to be assumed<br />

that all of the external quantities of the system<br />

S appear explicitly as external quantities<br />

of the subsystems (along with many ne w<br />

quantities which are external to Si, i = 1, . . ., k ,<br />

but internal to S) . However, it appears to me<br />

that it is the nature of different levels o f<br />

organization that the external quantities<br />

differ not only in resolution and detail but<br />

possibly also qualitatively and that definition<br />

39


of the external quantities of S would be impossible<br />

in the context of isolated subsystems .<br />

What is needed is some translation of sub -<br />

system external variables into system externa l<br />

variables. Clymer and Bledsoe (1970) have<br />

addressed a similar question in consideratio n<br />

of "interfacing" two subsystems at the sam e<br />

level and use the term "slave model" to designate<br />

the set of equations for changing resolution.<br />

This term seems inappropriate in the<br />

context of composition of subsystem proper -<br />

ties into properties of the whole ; such a<br />

process is more master than slave . I have<br />

chosen the term "Ghost System i2 and notationally<br />

designated So as the integrating sub -<br />

system which derives the properties of th e<br />

whole from the properties of the parts . In its<br />

simplest form, So is a resolution expander of<br />

inputs and reducer of outputs . Whether it will<br />

need to assume a more complex form is yet t o<br />

be seen .<br />

In this manner we have, by statements (1 )<br />

and (2), defined the ecosystem at two organizational<br />

levels, and may elaborate lower level s<br />

by the device of decomposing subsystems i n<br />

the same manner . That is, if S may be defined<br />

at two levels, so may the elements of {Si}, an d<br />

the subsystems so defined, to any desire d<br />

degree of fineness . Ultimately, it is necessary<br />

to end with a behavioral model for all sub -<br />

systems, because it is this model that yield s<br />

explicit form to the relational expressions .<br />

Several points are of interest here . First, i t<br />

is not clear that a hierarchical decompositio n<br />

is always desirable . That is, a finer resolutio n<br />

may not best take the form of a decomposition<br />

of next most coarse form of interest .<br />

Hierarchical forms, however, have man y<br />

advantages and will be used wherever possible .<br />

Second, it is quite clear that a uniform resolution<br />

is unnecessary . That is, a very coarse<br />

resolution model of one subsystem may b e<br />

coupled with a fine resolution model of<br />

another, and such an arrangement will have<br />

many advantages . Not the least of these is th e<br />

advantage of manageability . If we want to<br />

take a closeup view of some parts of the system,<br />

there is no reason to, and many reason s<br />

not to, look at the rest of the system at th e<br />

2 After Koestler's "The Ghost in the Machine ."<br />

same degree of magnification . A hierarchical<br />

subsystem structure will be most conducive t o<br />

a variable resolution, but again, it is no t<br />

necessary .<br />

Application of the<br />

General Theory to the<br />

<strong>Forest</strong> Ecosystem<br />

With this introduction to the theoretica l<br />

background, we can now take a look at th e<br />

coarser ecosystem structures as we now perceive<br />

them . A diagrammatic convention i s<br />

established which will be recognized as simila r<br />

to Forrester's, but which differs in some<br />

important respects . Let boxes designate compartments<br />

(to be conceptualized as storage<br />

containers), let ovals or circles designate systems<br />

(or subsystems), and let diamonds designate<br />

a set of variables . Solid lines designate<br />

coupled flow variables and dotted lines represent<br />

control variables. As a convention ,<br />

diamonds are eliminated from the flow couplings,<br />

as these variables are apparent fro m<br />

the context, but identification of contro l<br />

variables is important . Occasionally these are<br />

also eliminated from a figure to reduc e<br />

clutter, and in the illustrations given here ,<br />

most control paths are eliminated for th e<br />

same reason .<br />

In figure 1 is depicted a representation of a<br />

system as a whole, with inputs and outputs . At<br />

this level, the entire system (the forest ecosystem)<br />

is defined according to (1) . In figure<br />

2, the same system is elaborated by identification<br />

of major subsystems, the terrestrial biota ,<br />

the aquatic biota and the hydrologic system .<br />

This involves definition of the system according<br />

to (2) and it should be emphasized at thi s<br />

point that there are very many ways in whic h<br />

this could be done .<br />

Implementation of this form (fig. 2) requires<br />

decisions regarding boundaries an d<br />

specifications of the nature of the coupling s<br />

between subsystems . An example will illustrate<br />

both points. Consider the uptake of<br />

water by higher plants . Question : Do we wish<br />

to consider this water as having transferred<br />

40


function, but this appears awkward .<br />

Once all the couplings are identified and<br />

specified, then it is no longer necessary to consider<br />

one subsystem in elaborating the interna l<br />

structure of another. The process of couplin g<br />

specification effectively uncouples the syste m<br />

into subsystems, each of which can be developed<br />

independently, provided that the identified<br />

external variables of the subsystems ar e<br />

maintained .<br />

I will use the word integrity in this context .<br />

Elaboration of the system at one level o f<br />

organization is unrestricted so long as th e<br />

integrity of the next higher level of organization<br />

is maintained. This does not appear at<br />

this writing to be a trivial problem, but th e<br />

problem exists under any formulation o f<br />

model structure and is well identified unde r<br />

the present structure, hence potentially<br />

manageable .<br />

Figure 1 . The system viewed as an object receivin g<br />

inputs from the environment and producin g<br />

outputs .<br />

from the hydrologic system to the terrestrial<br />

biotic system? Or, will we let stored water<br />

remain in the hydrologic system? The answer<br />

to this is arbitrary and should be made on th e<br />

basis of ultimate convenience . If the specification<br />

of couplings is simpler for one boundary<br />

definition than for another, then that boundary<br />

is preferred .<br />

The nature of the couplings (or couplin g<br />

variables) is also important . Suppose we le t<br />

water enter the biotic system at the momen t<br />

of uptake. Then the transfer of water fro m<br />

hydrologic to biotic depends on (1) the stat e<br />

of the hydrologic system (specifically, th e<br />

amount and distribution of water available i n<br />

the soil) and on (2) the biotic demand fo r<br />

water. These two variables must be identifie d<br />

as output variables of the respective system s<br />

in order to elaborate the transfer . It then i s<br />

immaterial whether the calculation of transfer<br />

takes place in the hydrologic or the bioti c<br />

system, but it seems appropriate that it b e<br />

included in the latter as an uptake function .<br />

In fact, one can even leave the function o f<br />

transfer outside both systems as a coupling<br />

Decomposition of the<br />

Terrestrial Biotic System<br />

With this background, I can develop further<br />

details of our present conceptualization .<br />

These results are presented in much greate r<br />

detail in a project document titled "Modeling ,<br />

Round One," prepared jointly by myself an d<br />

a number of others in our program . This<br />

document summarizes a series of seminar -<br />

workshops in which we examined the definition<br />

of subsystems and the nature of thei r<br />

couplings . Reports of details of several sub -<br />

systems are reported by others in this symposium.<br />

I will use several subsystems for illustration<br />

which are not treated elsewhere .<br />

In figure 3 is seen a specification of primary<br />

subsystems of the terrestrial biotic<br />

system. These are not even, in the sense o f<br />

evenness of importance, but rather are chose n<br />

to emphasize the relationships we wish t o<br />

emphasize at this level . Only two compartments<br />

are specifically identified, the plan t<br />

biomass (B. & S.) and the detritus (D .) compartments,<br />

with all others embedded in the<br />

respective systems . Identification of the biomass<br />

compartment forces explicit identification<br />

of a minor subsystem, litterfall. Th e<br />

41


Figure 2 . The major subsystems of a forest ecosystem . Specification of the coupling variables among the majo r<br />

subsystems allows independent development of their internal structure .<br />

42


Figure 3. The December 1971 version of the subsystems of the Terrestrial Biotic System . The external variable s<br />

of the whole system match those illustrated for this subsystem in figure 2 .<br />

traditional decomposition and fixation sub -<br />

system has been uncoupled into the three sub -<br />

systems represented as Decomposition, Fixation<br />

and Weathering, and Interchange and<br />

Uptake (I-U) .<br />

The value of this formulation is several<br />

fold. First, we can recognize that this is a<br />

formal expression of generally held ideas, s o<br />

that only the formality, and perhaps the<br />

identification of the I-U process, is in an y<br />

degree innovative. The formality gives sub -<br />

stance to the concepts, and the formal U-C<br />

structure identifies the tasks facing us as w e<br />

elaborate this structure . These tasks are :<br />

(1) specification of the coupling variables ,<br />

(2) elaboration of internal structure of the<br />

subsystems, and (3) analysis and synthesis o f<br />

behavior and properties of the systems and<br />

4 3<br />

t+


subsystems as objects . In addition, the formal<br />

structure effectively partitions our effort and<br />

ensures that all . bases are covered . Boundaries<br />

between subsystems and the appropriat e<br />

coupling variables must be identified jointl y<br />

by those responsible for the subsystems involved.<br />

After boundaries and coupling variables<br />

are agreed upon, then the activities o f<br />

elaboration of internal structure are independent<br />

among subsystems . This is another<br />

expression of the earlier observation tha t<br />

identification of the couplings effectivel y<br />

uncouples the subsystems .<br />

These values of the model formulation can<br />

be expressed as explication, conceptualization,<br />

and organization . Existing concepts are<br />

given explicit expression, new concepts ar e<br />

developed and research efforts are given structure<br />

and organization, all by the formal ecosystem<br />

model . The fact that the model structure<br />

is arbitrary in no , way obviates thes e<br />

values . On the contrary, the arbitrariness o f<br />

the model structure enhances these values ,<br />

because as we proceed with the process of<br />

elaboration, and bring existing knowledg e<br />

into sharper focus by the process of modeling ,<br />

we are constantly forced into adjustments, re -<br />

orientation, and reorganization .<br />

Structures, like figure 3, are not meant t o<br />

be permanent . They live and die like generations<br />

of insects . In fact, the insect analogy is<br />

quite appropriate for figure 3 . This structure<br />

was conceptualized in the spring and earl y<br />

summer of 1971, during Round One . I t<br />

developed underground, so to speak, all during<br />

the fall, undergoing several transitions, t o<br />

emerge in December in the form presented .<br />

By mid-January 1972 this form had lived ou t<br />

its life and given birth to a new form which is<br />

yet in the larval stages and not ready for th e<br />

light of day .<br />

At the next lower organizational level ,<br />

progress has been made in elaboration of the<br />

internal structure of the food chain subsyste m<br />

and the I-U subsystem . The first is reported<br />

here by Strand and Nagel, but the second i s<br />

not represented in this symposium and I<br />

would like to include a brief description of<br />

that subsystem as additional illustration o f<br />

the way in which we are using the Universe -<br />

Coupling structure .<br />

Figure 4 represents a model form date d<br />

back in 1971, which form has been replace d<br />

by a new but incompletely developed form, in<br />

accordance with the changes being made a t<br />

the next lower resolution (fig . 3). However ,<br />

figure 4 serves to illustrate the concept of th e<br />

I-U process and, again, the point is made tha t<br />

this is developed more fully in the document s<br />

of Round One .<br />

The I-U subsystem is postulated on the following<br />

statements :<br />

1. If nutrients remain in soil solution, they<br />

must be lost to the system by soil and<br />

groundwater transport .<br />

2. Nutrients must be in solution in order t o<br />

be available to uptake .<br />

3. Some microorganisms and some highe r<br />

plants are "leaky ."<br />

4. It is concluded that a successful syste m<br />

must have a tightly coupled, highly interactive<br />

and buffered subsystem of nutrient<br />

interchange and uptake .<br />

The modeling contribution here is th e<br />

recognition (4) that the three stated feature s<br />

of the traditional processes of nutrient inter -<br />

change are such that a successful terrestria l<br />

ecosystem (i .e., one which does not lose it s<br />

nutrients downstream) must have a tightly<br />

coupled I-U subsystem . That is, the traditional<br />

study by individual process canno t<br />

possibly answer the questions we want to ask ,<br />

unless we explicitly define the couplings .<br />

Coupling definition is difficult in a tightl y<br />

coupled system, and appears exceedingl y<br />

difficult in this one . The identified task is th e<br />

conceptualization of properties and behavior<br />

of the I-U subsystem as a whole .<br />

It is my very strong conclusion here that<br />

the Biome research effort should be re -<br />

oriented to accommodate the concept of th e<br />

I-U subsystem. Present conceptualization an d<br />

specification of that subsystem is yet very<br />

primitive and will receive considerable attention<br />

in the next year .<br />

From the model structure point of view ,<br />

the primary productivity subsystem is th e<br />

least well defined in our model . This is, in<br />

part, because most of our research effort ha s<br />

been at lower organizational levels than i s<br />

necessary for ecosystem models . It is a giant<br />

step from tree to community . The primary<br />

44


I<br />

I<br />

I<br />

t<br />

/ INTERCHANG E<br />

a UPTAKE BY SOI L<br />

HIGHER<br />

MICROORGANIS M<br />

PLANT S<br />

I<strong>ON</strong> I C<br />

B<strong>ON</strong>DI NG,<br />

HUMU S<br />

1<br />

SOI L<br />

SOLUTI<strong>ON</strong> ----~--<br />

1<br />

Figure 4 . The December 1971 version of the subsystems of the Interchange and Uptake System. The externa l<br />

variables of the whole system match the couplings of the I-U subsystem of figure 3 .<br />

effort has been at the tree level (and below )<br />

and the model structures for the ecosystem s<br />

are communities. The necessary effort here is<br />

apparent .<br />

Representation of Spatial<br />

and Environmental Variatio n<br />

The Universe Coupling structure is als o<br />

appropriate to the problem of modeling a<br />

heterogenous ecosystem . If the system i s<br />

stratified into homogenous subsystems ,<br />

identified by spatial or environmental criteria ,<br />

then the models appropriate for each of the<br />

strata can be coupled together to form a<br />

model for the whole . This is the approach we<br />

have taken in our watersheds . A watershed is<br />

sufficiently variable, with regard to physical<br />

and biological processes, that we feel it i s<br />

necessary to stratify so as to provide essential<br />

uniformity of processes within strata. Th e<br />

process of composition, of coupling th e<br />

4 5


several models together to form the model for<br />

the watershed as a whole, is the reverse of the<br />

decomposition process previously discussed.<br />

Now we are faced with the question of defining<br />

appropriate external variables of th e<br />

watershed as an object after having initiall y<br />

defined the watershed as a collection of sub -<br />

objects .<br />

These questions are obviously appropriate<br />

to consideration of a general model for, say ,<br />

the entire H . J. <strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong> ,<br />

or for the Coniferous <strong>Forest</strong> Biome, sob that<br />

we consider them central to our overall<br />

objectives. By concentrating on the proble m<br />

of stratifying a watershed, and on the problems<br />

of modeling the strata and the collective ,<br />

we are attempting to devise strategies which<br />

will be useful in later extension of the modeling<br />

process .<br />

It turns out that if we view the construction<br />

of strata in essentially the same perspective<br />

as the construction of subsystems in th e<br />

U-C structure, the same general criteria hold .<br />

Generally speaking, subsystems (strata )<br />

should be constructed in such a manner tha n<br />

couplings are minimized (or simplified), an d<br />

such that important processes are containe d<br />

within the subsystems . This criterion is no t<br />

always compatible with the second one, tha t<br />

strata should be environmentally homogenous,<br />

within, and occasional conflicts arise .<br />

In any event, we are attempting hierarchica l<br />

stratification of watersheds, with topographic<br />

features defining the primary strata and vegetation<br />

types the secondary, the two levels giving<br />

uncoupling and homogeneity, respectively .<br />

Summary<br />

Some of the ways in which the Universe-<br />

Coupling structure is being used in building a<br />

hierarchical, modular system of models for<br />

the Coniferous Biome have been discussed .<br />

These models are hierarchical by virtue of th e<br />

identification of systems at one level o f<br />

organization as subsystems of the system s<br />

defined at the next higher level. They ar e<br />

modular by virtue of the uncoupled nature ;<br />

identification of the coupling variables at an y<br />

level allows complete flexibility of subsystem<br />

representation, provided that the integrity of<br />

the couplings is maintained .<br />

A general criterion for constructing sub -<br />

systems is provision of simple coupling s<br />

among subsystems with tight relations contained<br />

within . This principle is violated quite<br />

badly when trophic levels are used for sub -<br />

system definition . An alternate structure for<br />

consumers is being investigated in our program,<br />

as reported in the paper by Strand and<br />

Nagel .<br />

In applying the U-C structure to spatial<br />

units, another criterion is employed . Here it i s<br />

desirable to provide homogeneity of environment<br />

and process within strata, with variatio n<br />

among strata of no concern . This criterion i s<br />

sometimes in conflict with the first one o f<br />

minimal couplings, so that compromises are<br />

necessary .<br />

The paper has dealt primarily with aspect s<br />

of structure and criteria for application of th e<br />

structural form . Emphasis has been on th e<br />

first two of the three tasks which the U- C<br />

structure defines, to wit : (1) specification of<br />

the coupling variables, (2) elaboration of<br />

internal structure of subsystems . The third<br />

task, analysis and synthesis of behavior of the<br />

subsystems as objects, is given little attention ,<br />

but this probably is the most important, fro m<br />

the point of view of ecosystem theory . As<br />

mentioned earlier, the conceptualization of<br />

holistic system properties and behavior i s<br />

poorly advanced. The U-C structure provides<br />

an excellent basis for the attempt to develop<br />

this concept .<br />

One last point is made regarding the differences<br />

between the Universe-Couplin g<br />

approach and the State Variable approach .<br />

Since the State Variable model can als o<br />

assume hierarchical form (Goguen 1970), an d<br />

under common explicit models of subsyste m<br />

behavior the U-C model will reduce to a stat e<br />

variable model, one might question the practical<br />

validity of the distinction . Essentially, th e<br />

differences are that one (U-C) is structur e<br />

oriented and the other variable oriented an d<br />

that one (U-C) is oriented simultaneously to<br />

holistic and mechanistic representation an d<br />

the other solely to mechanism . In the application<br />

of ecosystem models so far produced, th e<br />

state variable approach seems adequate ,<br />

46


except for the retention of fine detail . Th e<br />

tendency in constructing models on this base<br />

has been to develop fine resolution mechanistic<br />

models in which the variables defined a t<br />

the finest resolution are in direct relation t o<br />

all the other variables in the system . Variabl e<br />

orientation seems to lead to variable sanctity .<br />

One might make a case for the position that<br />

the state variable approach was developed b y<br />

engineers who have no conceptualizatio n<br />

problem . They know what their variables are ,<br />

and the relations among them, and they ar e<br />

satisfied with mechanistic models .<br />

The U-C structure is seen as a means o f<br />

introducing holism, in addition to mechanism,<br />

into the model conceptualization process . By<br />

focusing on the subsystem structures, and b y<br />

attempting to describe the behavior of the<br />

system and subsystems at all levels in terms of<br />

their behavior as objects, each system an d<br />

subsystem is modeled at two levels, (1) holistically<br />

and (2) mechanistically in terms of th e<br />

holistic behavior of its subsystems . This<br />

approach is held to be much more promising ,<br />

from the point of view of development of<br />

ecosystem theory, than the currently popula r<br />

approach .<br />

Acknowledgments<br />

The work reported in this paper was supported<br />

in part by National Science Foundatio n<br />

Grant No. GB-20963 to the Coniferous <strong>Forest</strong><br />

Biome, U .S . Analysis of Ecosystems, International<br />

Biological Program. This is Contribution<br />

No . 23 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Clymer, A. B., and L . J. Bledsoe. 1970 . A<br />

guide to the mathematical modeling of a n<br />

ecosystem. In R. G . Wright and G . M. Van<br />

Dyne (eds.), Simulation and analysis of<br />

dynamics of a semi-desert grassland, p .<br />

75-99 . Colo . State Univ . Range Sci . Dep .<br />

Sci. Ser. No . 6 .<br />

Forrester, J . W . 1971 . Counterintuitive behavior<br />

of social systems . Technol. Rev . 73 :<br />

53-69 .<br />

Goguen, J. A . 1970 . Mathematical representation<br />

of hierarchically organized systems . In<br />

E . O . Attinger (ed .), Global systems<br />

dynamics, p . 112-129 . New York : Wiley .<br />

Klir, G. J. 1972 . The polyphonic general<br />

system theory . In G . J . Klir (ed .), Trends in<br />

general systems theory, p . 1-18 . New York :<br />

Wiley-Interscience .<br />

Klir, George J . 1969 . An approach to general<br />

systems theory. 323 p. New York: Van<br />

Nostrand Reinhold Co .<br />

Orchard, R . A .. 1972 . On an approach to general<br />

systems theory . In G. J . Klir (ed .) ,<br />

Trends in general systems theory, p .<br />

205-250. New York : Wiley-Interscience .<br />

Strand, M . A., and W. P. Nagel. 1972 . Preliminary<br />

considerations of the forest canop y<br />

consumer subsystem . In Jerry F. Franklin ,<br />

L. J. Dempster, and Richard H . Warin g<br />

(eds.), Proceedings-research on coniferou s<br />

forest ecosystems-a symposium, p . 71-77 ,<br />

illus. Pac. Northwest <strong>Forest</strong> & Range Exp .<br />

Stn., Portland, Oreg .<br />

47


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium.<br />

Bellingham, Washington-March 23-24, 197 2<br />

Hydrologic modeling in the<br />

Coniferous <strong>Forest</strong> Biome<br />

George W . Brow n<br />

Chairman, Hydrology Subcommittee<br />

Oregon State University<br />

Robert H . Burgy<br />

Hydrology Project Leade r<br />

University of California, Davi s<br />

R . Dennis Har r<br />

Hydrology Project Leade r<br />

Oregon State University<br />

J . Paul Riley<br />

Hydrology Project Leader<br />

Utah State University<br />

Abstract<br />

The objective of the hydrology program is to prepare a model which will provide predictions of the hydrologi c<br />

state of a coniferous watershed at any desired time and in any desired place, where state is defined by the inpu t<br />

needs of the other submodels or systems, particularly the producer and biogeochemical processes. Subsurface flo w<br />

is the dominant runoff mechanism in coniferous watersheds and one of the least understood processes in<br />

hydrology. Research projects in hydrology seek to understand this process using three different techniques . One<br />

project relates subsurface flow to soil properties using direct measurement techniques. Another project approaches<br />

the problem using simulation techniques. The third project utilizes systems analysis and statistical decompositio n<br />

of runoff events to make inferences about subsurface flow. These studies of hydrologic processes will b e<br />

incorporated into a hydrologic model and linked to studies of other systems. The first step in linking our mode l<br />

with those of other groups is watershed stratification, a problem now solved by our modeling efforts.<br />

Introduction<br />

Water is an essential component of any ecosystem.<br />

In the Pacific Northwest, water is a<br />

dominant element. The coniferous forests o f<br />

this region are noted as some of the best -<br />

watered terrestrial ecosystems in the United<br />

States; water is the major linkage which tie s<br />

the terrestrial portion of the coniferous ecosystem<br />

to the aquatic portion .<br />

Water performs several functions which<br />

foster this linkage . Water must be viewed as a<br />

carrier. It carries organic and inorganic nutrients<br />

between the several compartments of the<br />

terrestrial portion of the ecosystem and fro m<br />

the terrestrial to the aquatic portion . Wate r<br />

also carries sediment from the terrestrial t o<br />

the aquatic portion of the system .<br />

Water must also be viewed as a nutrient<br />

itself. It is an essential component of most<br />

biologic processes . The availability of water i n<br />

the soil governs both the initiation and termi -<br />

49


nation of any process as well as the rate a t<br />

which it proceeds .<br />

Objective of the Hydrology Progra m<br />

The objective of our program is to prepare<br />

a model which will provide predictions of th e<br />

hydrologic state of a coniferous watershed at<br />

any desired time and in any desired place ,<br />

where state is defined by the input needs of<br />

the other submodels or systems, particularl y<br />

the producer and biogeochemical processes .<br />

Structure of the Hydrology Modeling Effort<br />

Our modeling effort is organized to pa y<br />

particular attention to the many functions o f<br />

water in the forest ecosystem . A generalized<br />

model for water flow through a forest syste m<br />

was conceptualized long ago . This model is<br />

often called the hydrologic cycle . Rothacher<br />

et al. (1967) showed for the H . J. <strong>Andrews</strong><br />

<strong>Experimental</strong> <strong>Forest</strong> that water movement i n<br />

the forest soil and evapotranspiration are th e<br />

most significant processes governing water<br />

flow .<br />

Our studies focus upon subsurface wate r<br />

movement. One project measures subsurface<br />

flow directly in the study watershed . Another<br />

study is a computer simulation of a water -<br />

shed. This approach will provide yet anothe r<br />

avenue for assessing soil moisture and the sub -<br />

surface flow of water. The simulation model<br />

will be calibrated using 14 years of record o n<br />

watershed 2 . Then the model will be verifie d<br />

with the data available on watershed 10 . A<br />

third project seeks to develop techniques for<br />

predicting the subsurface flow componen t<br />

using a systems analysis technique for statistical<br />

decomposition of the hydrograph into its<br />

components . Other studies will contribut e<br />

submodels to the simulation of the hydrologi c<br />

system. The work of the Primary Producer<br />

group on a transpiration model as well as th e<br />

work of the Meteorology group on an evapotranspiration<br />

model will aid our modeling<br />

effort significantly .<br />

We are concurrently working toward a<br />

more spatially refined model, the character of<br />

which is determined as much by biologic a s<br />

hydrologic constraints . Our latest efforts hav e<br />

focused upon devising a system for stratification<br />

of watershed 10 which is amenable t o<br />

both hydrologic and biologic models .<br />

Other hydrology projects are included i n<br />

the biome effort. We hope to provide hydro -<br />

logic measurements as a part of the work a t<br />

Findley Lake and the evapotranspiratio n<br />

study at the Thompson site . We shall soo n<br />

begin to solicit and organize available dat a<br />

from coniferous watersheds throughout th e<br />

West in anticipation of extrapolating our<br />

model to other ecosystems .<br />

This has provided a broad overall view of<br />

the Coniferous Biome Hydrology Program-its<br />

structure to achieve a better understanding o f<br />

water flow through the ecosystem and its<br />

interaction with other system components . A<br />

more detailed description of the hydrolog y<br />

efforts follows . Each project focuses upon<br />

evaluating water flow in a forested watershed .<br />

Analytical techniques vary between projects ,<br />

but the ultimate goal of modeling subsurfac e<br />

flow mechanisms and watershed respons e<br />

links all projects together .<br />

Subsurface Movement<br />

of Water on<br />

Steep, <strong>Forest</strong>ed Slopes l<br />

With the exception of stream channel interception,<br />

the hydrographs of watersheds in the<br />

forested, steep topography of western Oregon<br />

reflect overall subsurface movement of water .<br />

Watershed response is rapid but without<br />

surface runoff (Barnett 1963, Rothacher et al .<br />

1967) . Although subsurface flow is by far th e<br />

major component of the hydrograph, virtuall y<br />

nothing is known about the process on stee p<br />

slopes. The objective in our study is to characterize<br />

the subsurface movement of water i n<br />

steeply forested topography .<br />

1 Authored by R. Dennis Harr, Assistant Professor,<br />

Oregon State University, Corvallis .<br />

50


The study areas are located at several of th e<br />

lower watersheds in the H . J. <strong>Andrews</strong> <strong>Experimental</strong><br />

<strong>Forest</strong> near Blue River, Oregon . Vegetation<br />

is typical of the low-elevation Douglasfir<br />

forest. Study slopes average about 75 per -<br />

cent. Soil depth is variable with maximu m<br />

depths in excess of 5 meters . Because of high<br />

porosities (70-80 percent) and large proportions<br />

of macropores, these soils drain rapidly .<br />

Permeabilities of 5,000 and 900 mm per hou r<br />

have been noted on nearby watersheds fo r<br />

surface soil and subsoil, respectively (Dyrnes s<br />

1969) .<br />

Methods<br />

Initial investigations are being directe d<br />

toward describing the physical properties of<br />

the porous medium through which wate r<br />

moves on its way to a stream . These field and<br />

laboratory investigations will indicate where<br />

water movement most likely occurs .<br />

Drilling with a portable power drill wil l<br />

follow a grid pattern over a small stream-toridge<br />

portion of slope . At each grid point soi l<br />

depth, depth and thickness of saprolite, an d<br />

depth to unweathered bedrock will be deter -<br />

mined. Additional drilling between initial gri d<br />

points will indicate in more detail the surfac e<br />

contour of the impermeable parent material .<br />

Aluminum tubing placed in each hole wil l<br />

provide access for measurement of ground -<br />

water level or soil moisture content .<br />

In the laboratory, undisturbed soil core s<br />

taken from various depths in soil pits locate d<br />

over the study area are being analyzed . Such<br />

properties as porosity, pore-size distribution ,<br />

stone content, permeability, and moisture retention<br />

characteristics are being evaluated .<br />

The type and amount of measurements to<br />

be made during and following winter stor m<br />

events in 1972-73 will depend on the in -<br />

formation gathered during initial field an d<br />

laboratory investigations now underway .<br />

Anticipated measurements include soil moisture<br />

content, vertical and lateral extent o f<br />

saturated flow, soil moisture tension, precipitation,<br />

and water outflow from the base of<br />

the slope . Drilling and tracer studies will<br />

attempt to define the source area for this<br />

water .<br />

Preliminary Results<br />

Although the study has just recently begun ,<br />

certain observations have provided qualitative<br />

information concerning the subsurface flow<br />

process on steep slopes. Precipitation moves<br />

downward under the influence of gravity until<br />

this movement is obstructed . In some parts of<br />

the study area this obstruction may be cause d<br />

by rock fragments which cause shallow ,<br />

localized saturation as evidenced in several<br />

soil pits during a period of heavy rain . Where<br />

rock fragments are not present, downward<br />

movement of water continues until the relatively<br />

impermeable parent material is reached .<br />

Here saturation occurs, flow acquires a horizontal<br />

component, and water begins movin g<br />

toward the stream.<br />

At some point on the slope this saturated<br />

flow is concentrated into pipelike subsurfac e<br />

channels. The cause of this concentration i s<br />

unknown but could conceivably result fro m<br />

the microrelief of the impermeable material ,<br />

from bedrock fractures, or from decayed root<br />

channels. At the toe of the slope the channels<br />

are spaced about 1-6 meters apart . They lie on<br />

the bedrock surface and appear associated<br />

with surface micro-relief. Shapes of their cross<br />

sections range from circular to flat rectangular.<br />

Width is also variable, ranging from 1<br />

centimeter to about a meter. Where these<br />

channels discharge into the stream channel ,<br />

they are separated by soil which may contai n<br />

a shallow saturated lower layer from whic h<br />

seepage occurs. Water velocity of the seepage<br />

appears to be several orders of magnitud e<br />

lower than that of the subsurface channels .<br />

The latter accounts for the greatest portion o f<br />

stormflow .<br />

The subsurface channels evident at the to e<br />

of the study slopes may be outlets of a sub -<br />

surface drainage system much like that de -<br />

scribed for other humid areas (Jones 1971) .<br />

Water can be observed discharging from suc h<br />

channels in roadcuts and recent soil slumps at<br />

various slope positions in the vicinity of th e<br />

study area. Such a subsurface drainage syste m<br />

could account for the rapid hydrologic<br />

response of these steep slopes .<br />

51


Computer Simulation<br />

of Fores t<br />

Watershed Hydrology 2<br />

A hydrologist is often faced with the need<br />

to predict system responses under variou s<br />

possible management alternatives . One approach<br />

to this problem is to apply the technique<br />

of computer simulation, whereby a<br />

quantitative mathematical model is developed<br />

for investigating and predicting the behavio r<br />

of the system. In this study, a computer<br />

model is being developed to simulate th e<br />

hydrologic responses of a forest watershed ,<br />

emphasizing the measurable variables related<br />

to the plant communities and soil types of th e<br />

watershed . The model represents the inter -<br />

related processes of the system by function s<br />

which describe the different components o f<br />

physical and biological phenomena in a<br />

watershed .<br />

Scope and Objectives of th e<br />

Simulation Stud y<br />

In the first phase of the study, the scope is<br />

being limited to the formulation of a fundamental<br />

model of watershed hydrology whic h<br />

takes precipitation as the basic input and<br />

evapotranspiration and streamflow as output s<br />

of the system . The various component processes<br />

within the system are linked by the conservation<br />

of mass principle. Depending upon<br />

energy levels, water can vary among its solid ,<br />

liquid, and vapor forms ; hence, the energy<br />

budget is used as an auxiliary tool for maintaining<br />

the water balance . That aspect of th e<br />

system involving water as a carrier of nutrients<br />

and sediments will be examined in a<br />

subsequent phase of the study . Under this<br />

next phase a water quality submodel will b e<br />

formulated and added to the quantity mode l<br />

now being developed .<br />

2 Authored by J . Paul Riley, Professor, Utah Wate r<br />

Research Laboratory, Utah State University, Logan ;<br />

George B . Shih, Research Engineer, Utah Water<br />

Research Laboratory, Utah State University, Logan ;<br />

and George E . Hart, Associate Professor . Fores t<br />

Science, Utah State University, Logan, Utah .<br />

The specific objectives of the current phase<br />

of the study are stated as follows :<br />

1. To develop and verify (calibrate an d<br />

test) a hydrologic simulation model for a<br />

small forested subwatershed on the H . J .<br />

<strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong> .<br />

2. To estimate through model sensitivit y<br />

studies the relative importance of variou s<br />

processes within the hydrologic syste m<br />

of the model, with particular emphasi s<br />

on evaluating the soil moisture and inter -<br />

flow components .<br />

Hydrologic System Models<br />

Several hydrologic simulation models ar e<br />

currently available . Examples which might be<br />

cited include Crawford and Linsley (1964) ,<br />

Sittner et al . (1969), and Riley et al.(1966) .<br />

However, in order to meet the needs of thi s<br />

study all existing models require some modifications<br />

and further development . Therefore ,<br />

on the basis of previous work at Utah State<br />

University a computer model is being developed<br />

to simulate the hydrologic behavior of<br />

forest watersheds . The model will be applicable<br />

to a wide variety of geographical area s<br />

and management problems . In this study, data<br />

from watershed 2 on the H . J. Andrew s<br />

<strong>Experimental</strong> <strong>Forest</strong> will be used to demonstrate<br />

the utility of the model . Figure 1 summarizes<br />

the geophysical features of the study<br />

area (Rothacher et al . 1967). Data requirements<br />

include air temperature, precipitation ,<br />

runoff hydrographs, and characteristics of th e<br />

watershed (average slope, degree and aspect ,<br />

vegetative cover, density of vegetative cover ,<br />

soil moisture holding characteristics, an d<br />

drainage density). Other observed records ,<br />

such as snow depth, soil moisture content will<br />

be used to check the performance of th e<br />

model in simulating various component processes<br />

of the system .<br />

Figure 2 illustrates the various componen t<br />

processes represented in this model, with th e<br />

boxes representing storage locations and th e<br />

lines transfer functions . Under this study, th e<br />

hydrology of the drainage area will be synthesized<br />

first as a lumped parameter model in<br />

which the entire watershed area is considere d<br />

as a single space unit. On this basis, a dis -<br />

52


K m<br />

Figure 1 . Watershed 2, H . J. <strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong>, Oregon . Area: 60.3 hectares ; aspect: NW ;<br />

average slope (percent) : 61 .1 ; elevation min .: 526 m; elevation max .: 1,078 m ; main channel length :<br />

1,108 m; drainage density: 4.3 km/km 2 ; precipitation (1952-62) : 2,400 mm/yr; runoff (1953-62) :<br />

1,560 mm/yr; average evapotranspiration (1959-62) : 540 mm/yr .<br />

tributed parameter model will be developed i n<br />

which the watershed will be divided into four<br />

space units, roughly corresponding to subwatersheds<br />

within the area . The model will<br />

compute continuous daily streamflow for<br />

each subarea and route the contribution of<br />

each down the streams to the gaging station ,<br />

where the computed and observed discharges<br />

will be compared. Other important outpu t<br />

functions from the model will include soi l<br />

moisture, actual evapotranspiration, and sno w<br />

depth. Several of the component processe s<br />

which are illustrated by figure 2 are discussed<br />

in the following sections .<br />

Interceptio n<br />

Interception is the part of precipitation<br />

that is caught temporarily by forest canopie s<br />

and then redistributed either to the atmosphere<br />

by evaporation or sublimation or to the<br />

forest floor. The amount of interceptio n<br />

depends upon storm size and intensity, an d<br />

canopy type and density. A report by<br />

Rothacher (1963) showed that throughfall in<br />

the study area was related to storm size b y<br />

the equation :<br />

Throughfall = 0 .8311 x (gross precipitation) - 0 .117 (1 )<br />

In equation 1 throughfall is, of course ,<br />

bounded by the condition that it must be<br />

greater than or equal to zero . Although larger<br />

amounts of snow may be temporarily intercepted<br />

than rain, there is strong evidence that<br />

most intercepted snow ultimately falls and be -<br />

comes part of the snowpack .<br />

53


CLOUD S<br />

I<br />

EVAPORATI<strong>ON</strong><br />

SUBLIMATI<strong>ON</strong><br />

INTERCEPTIO N<br />

0<br />

SNOW SNOW SNO W<br />

z<br />

STORAG E<br />

H<br />

z<br />

0<br />

H<br />

EVAPOTRANSPIRATI<strong>ON</strong><br />

WATER STORED <strong>ON</strong><br />

LAND SURFACE<br />

SNOWMEL T<br />

H<br />

EVAPOTRANSPIRATI<strong>ON</strong><br />

0 .<br />

H<br />

H<br />

w<br />

H<br />

SOIL MOISTURE<br />

OVERLAND FLO W<br />

INTERFLOW<br />

STREAM<br />

CHANNEL<br />

STORAG E<br />

SURFAC E<br />

OUTFLOW<br />

GROUNDWATER<br />

INFLOW<br />

GROUNDWATER<br />

STORAGE<br />

EFFLUENT FLOW -Aim.<br />

INFLUENT FLOW E-<br />

GROUNDWATER OUTFLO W<br />

Figure 2 . A flow diagram of the hydrologic system within a typical watershed area.<br />

54


An alternative way of considering interception<br />

quantities in a model is to express<br />

interception rate as a decaying function of<br />

time limited by an average interception storage<br />

capacity for the watershed canopy . This<br />

approach was incorporated into a watershed<br />

simulation model by Riley et al . (1966) .<br />

Forms of Precipitation<br />

Snow Storage and Melt<br />

Only two forms of precipitation, rain an d<br />

snow, are considered in this study, with a<br />

surface air temperature criterion being applie d<br />

to establish the occurrence of these two<br />

forms. Figure 3 (U .S. Army Corps of Engineers<br />

1956) shows that at a temperature of<br />

1.5°C there is a 50-percent chance that th e<br />

precipitation will be in the form of snow . A<br />

straight-line fit to figure 3 is used to deter -<br />

mine the portion of rain in a given day ac -<br />

cording to the following equation .<br />

Ta - Ts<br />

R=P<br />

(2 )<br />

T r- Ts<br />

in which<br />

R = estimated portion of total dail y<br />

precipitation occurring as rai n<br />

P = total daily precipitation<br />

Ta<br />

T s<br />

Tr<br />

= mean daily surface air temperatur e<br />

= mean daily air temperature below<br />

which all precipitation is assumed<br />

to occur as snow<br />

= mean daily air temperature above<br />

which all precipitation is assumed<br />

to occur as rain<br />

Precipitation falling as snow will be accumulated<br />

on the watershed until air temperatures<br />

rise sufficiently above the freezing poin t<br />

to initiate snowmelt .<br />

10 0<br />

1 0<br />

9 0<br />

2 0<br />

8 0<br />

3 0<br />

7 0<br />

4 0<br />

5 0<br />

6 0<br />

7 0<br />

8 0<br />

9 0<br />

100<br />

-1 .5 -1 .0 -0 .5Ts 0 0 .5 1 .0 2 .0 3 .0 4 .0 Tr<br />

SURFACE AIR TEM<strong>PE</strong>RATURE, ° C<br />

Figure 3 . Frequency distribution of precipitation in rain and snow forms .<br />

0<br />

55


LIQUID WATER-HOLDIN G<br />

CAPACITY OF SNOWPAC K<br />

Figure 4 . A flow chart of the snow accumulation and ablation processes.<br />

56


Snowmelt<br />

A flow chart of the snow accumulation an d<br />

ablation processes is shown by figure 4 . Rate<br />

of snowmelt depends primarily upon the rate<br />

of energy input to the snowpack . However,<br />

both the complex nature of snowmelt an d<br />

data limitations prevent a strictly analytica l<br />

approach to the simulation of this process ,<br />

and air temperatures are frequently applied a s<br />

an index of available energy . Examples of<br />

researchers who have used this approach ar e<br />

Pysklywec et al. (1968), Anderson and Craw -<br />

ford (1964), Amorocho and Espildor a<br />

(1966), and Eggleston et al. (1971). Becaus e<br />

temperature data are the only indicators o f<br />

energy levels available on watershed 2, a<br />

degree-day approach based upon the work o f<br />

Eggleston et al. (1971) will be used in th e<br />

model of this study to represent the snowmelt<br />

process at the surface of the snowpack . This<br />

component submodel includes mathematical<br />

relationships for various phenomena involve d<br />

in the snowmelt process . The submodel is<br />

applicable to any geographic location by determining<br />

appropriate constants for certain re -<br />

lationships through a verification procedure .<br />

The relationship for surface melt rate i s<br />

expressed as follows :<br />

RIs<br />

Prg<br />

Mrs =kmkvRIh Ta( 1-A ) +Ta 80 (3)<br />

in whic h<br />

km = a constant of proportionality<br />

kv<br />

RI h<br />

RI s<br />

= vegetation transmission coefficient<br />

for radiatio n<br />

radiation index for a horizontal surface<br />

at the same latitude as th e<br />

particular watershed or zone under<br />

study<br />

radiation index for a particula r<br />

watershed zone possessing a known<br />

degree and aspect of slop e<br />

T a = surface air temperature in °C<br />

A = albedo, or reflectivity, of the snowpack<br />

surfac e<br />

Prg<br />

precipitation reaching the snow surface<br />

in the form of rain, in<br />

centimeters<br />

Infiltration<br />

Rates of water supply on the ground surface,<br />

whether in the form of rainfall minus<br />

interception or snowmelt, must exceed infiltration<br />

rates before any surface runoff occurs .<br />

The infiltration rate depends on the physical<br />

and moisture characteristics of the soil, as<br />

well as the surface organic conditions, and it<br />

is often expressed in the form of Horton's<br />

exponential equation . However, the soils of<br />

watershed 2 are very porous and no overlan d<br />

flow has been observed . Thus, all precipitation<br />

reaching the ground surface is assumed to<br />

infiltrate into the soil and to move to th e<br />

stream channels as subsurface flow .<br />

Soil Moisture<br />

Soils on the study watershed are relatively<br />

deep and have a high porosity . Data on the<br />

physical properties of the watershed soils ar e<br />

available (Rothacher et al . 1967), and this<br />

information will be used to determine the soi l<br />

moisture holding characteristics .<br />

The computer model allows infiltrating<br />

water to satisfy first the available moistur e<br />

holding capacity of the soil within the roo t<br />

zone of the forest canopy . When the available<br />

soil moisture holding capacity is reached ,<br />

additional infiltration is assumed to percolate<br />

by gravitation either somewhat laterally with -<br />

in the root zone or downward to deeper soi l<br />

zones. Water which moves laterally usuall y<br />

reaches a surface channel within a relativel y<br />

short period of time, whereas deep percolation<br />

moves from the watershed more slowl y<br />

and sustains streamflow during dry seasons .<br />

From preliminary studies (Rothacher et al .<br />

1967), approximately 87 percent of the tota l<br />

annual precipitation reaches the ground surface,<br />

and about 75 percent of this quantity<br />

becomes surface runoff . From an analysis of<br />

streamflow hydrographs it is estimated tha t<br />

about 10 percent of the runoff comes fro m<br />

baseflow, which is contributed from dee p<br />

percolation . Thus, of the average annual<br />

precipitation of 2,400 mm which falls on th e<br />

watershed approximately 2,100 mm enter th e<br />

soil, 440 mm are abstracted by evapotranspiration,<br />

and the remaining 1,660 mm leav e<br />

57


the watershed as surface runoff, with 170 m m<br />

of this quantity occurring as baseflow . Th e<br />

soil moisture content computed by the mode l<br />

will be checked with observed data .<br />

Evapotranspiratio n<br />

Factors affecting evapotranspiration includ e<br />

temperature, solar radiation, wind, humidity ,<br />

and consumptive use by plants . However ,<br />

only temperature and humidity data are<br />

available for the watershed . Among the commonly<br />

used evapotranspiration equations<br />

(Veihmeyer 1964), the Penman equation i s<br />

perhaps the most rational, but the data requirements<br />

are extensive . For this reason, the<br />

modified Hargreaves (Veihmeyer 1964), will<br />

be used in this study . The equation is stated<br />

as follows .<br />

U =E Kd(0 .38-0.0038h)T a (4)<br />

in which<br />

U = the daily potential evapotranspiration<br />

in centimeters<br />

d = the daily daytime coefficient de -<br />

pendent upon latitud e<br />

h = the mean daily relative humidity at<br />

noon<br />

Ta = the mean daily surface air temperature<br />

in ° c<br />

K = a monthly consumptive use coefficient<br />

which is dependent upo n<br />

plant related characteristics, such a s<br />

species, growth stage, and densit y<br />

on the watershe d<br />

The influence of soil water on evapotranspiration<br />

has been the subject of much re -<br />

search and discussion . It is now generall y<br />

recognized that there is some reduction i n<br />

evapotranspiration rate as the quantity o f<br />

water within the root zone decreases . In this<br />

study it will be assumed that evapotranspiration<br />

occurs at the potential rate through a<br />

certain range of the available soil moisture . A<br />

critical moisture level is then reached at whic h<br />

actual transpiration begins to lag behind th e<br />

potential rate . Within this range of the avail -<br />

able soil moisture the relationship betwee n<br />

available water content and transpiration rat e<br />

will be assumed to be virtually linear . Thus,<br />

E = U, [Mes < Ms(t) < M cs] (5 )<br />

in which<br />

E = daily evapotranspiration adjuste d<br />

for the influence of soil moistur e<br />

levels<br />

M s quantity of water stored within th e<br />

root zone and available for plan t<br />

use at any time, t<br />

Mes limiting root zone available moisture<br />

content below which soil moisture<br />

tensions reduce evapotranspira -<br />

tion rates<br />

M cs root zone storage capacity of water<br />

available to plants<br />

M s (t)<br />

E = U [Mes > M s (t) >_ O] ( 6 )<br />

Mes<br />

Considering the pressure effect, the total rate<br />

of gravity water storage depletion through<br />

both interflow and deep percolation is<br />

assumed to be directly proportional to th e<br />

quantity of water in this form of storage remaining<br />

in the soil profile at any particula r<br />

time. The interflow portion of this depletion ,<br />

Nr, will be expressed as follows :<br />

Nr ( t) = Ki<br />

ddGs = Ki (Kg Gs (t) )<br />

in which<br />

K i = interflow depletion coefficien t<br />

Kg = gravity water depletion coefficien t<br />

-K t<br />

That is, Nr = KiKgGs(0)e g , in which gravity<br />

storage at time, t = 0 is represented b y<br />

Gs(0) and no input to Gs is assumed to occur<br />

between t = 0 and any other time, t . It is estimated<br />

that on watershed 2 about 90 percen t<br />

of the gravity water storage leaves the area as<br />

interflow, in which case KiKg = 0 .9 .<br />

Groundwater<br />

(7 )<br />

Water enters groundwater storage as dee p<br />

percolation from the overlying plant root<br />

zone. The rate of deep percolation, Gr, is<br />

numerically equal to the total rate of gravity<br />

water depletion within the root zone less th e<br />

interflow rate . Thus,<br />

58


d Gs<br />

Gr = (1 - Ki) d t<br />

By integrating equation 9 over a specific tim e<br />

period the accumulated inflow to the ground -<br />

water basin, Gam,, is estimated for this tim e<br />

period .<br />

Gw = f<br />

o<br />

G<br />

(1 - Ki) d dt (10)<br />

dt<br />

If the groundwater basin is considered as a<br />

linear reservoir, the outflow rate is given by<br />

the expression<br />

Qrg = K b Gw ( 11 )<br />

in whic h<br />

K b = a coefficient which is estimated<br />

from dry season streamflow hydrograph<br />

s<br />

Qrg = the outflow rate from the groundwater<br />

reservoir<br />

By combining equations 9 and 11, the net<br />

rate of storage change within the groundwater<br />

basin is derived a s<br />

Gr - Qrg d t<br />

By substituting equation 11 into equation 1 2<br />

and rearranging terms, the following relation -<br />

ship is obtained .<br />

d dtg = Kb Mr (t) - Qrg (t)l (13)<br />

The rate of discharge from the groundwate r<br />

basin as baseflow is obtained by solving equation<br />

13 for Qrg.<br />

Runoff<br />

d G am,<br />

(9)<br />

(12)<br />

The possible sources of streamflow at any<br />

reach within a channel are overland flow (surface<br />

runoff), interflow, groundwater, and up -<br />

stream input. Manning's equation is usuall y<br />

applied to compute overland and channel flow<br />

rates at any point . Under conditions on watershed<br />

2, however, surface runoff does not occur ,<br />

and channel routing on a daily time increment<br />

is not significant . Therefore, runoff rates at th e<br />

stream gage are given by summing the interflow<br />

and groundwater discharge rates .<br />

Model Verification<br />

Model verification includes calibration of<br />

the model parameters to a particular area ,<br />

testing the sufficiency of processes defined i n<br />

the model, and examining the prediction performance<br />

of the model . A self-calibration subroutine<br />

will be included in the model whereb y<br />

the program will search for optimal model<br />

parameter values . Under this procedure each<br />

water year is used as a unit for optimizatio n<br />

and the objective function is to minimize th e<br />

variance between observed and compute d<br />

streamflow (Shih 1971). The sufficiency of<br />

processes defined in the model is reflected in<br />

the dispersion of parameter values resultin g<br />

from each year of calibration . After th e<br />

model is calibrated, those years of data whic h<br />

were not used ' for calibration are used to<br />

examine the confidence level of prediction s<br />

by the model . A flow diagram of the mode l<br />

verification procedure is shown by figure 5 .<br />

Model Parameters<br />

Model parameters are the coefficients use d<br />

in defining the processes which have not bee n<br />

accurately measured or which cannot b e<br />

directly measured. By establishing the value s<br />

of these coefficients the general model i s<br />

fitted to the hydrologic system of a specifi c<br />

watershed . Depending upon the resolution of<br />

the model and the availability of data, the<br />

number of parameters to be calibrated may<br />

vary. In order to avoid using a large numbe r<br />

of degrees of freedom in the calibration<br />

process and to save computation time, the<br />

number of model parameters should be kept<br />

as few as possible. In this study, the preliminary<br />

model parameters to be calibrated are<br />

interception storage capacity (SI), snowmelt<br />

coefficient (Ks), soil moisture retentio n<br />

capacity (Mcs), gravity water depletion coefficient<br />

(Kg), and groundwater recession coefficient<br />

(Kb) .<br />

59


(AVAILABLE DAT A<br />

WATER YEA R<br />

HYDROLOGI C<br />

DATA<br />

CALCULATE WATERSHED COEFFICIENT S<br />

ESTIMATE THE INITIAL VALUE AN D<br />

LIMIT OF MODEL PARAMETERS SE T<br />

UP PARAMETER VALUE C<strong>ON</strong>STRAINT S<br />

ENTIRE WATERSHED AS SINGLE UNIT<br />

WATERSHED HYDROLOGI C<br />

SIMULATI<strong>ON</strong> MODE L<br />

1<br />

CALIBRATE THE<br />

MODEL PARAMETERS<br />

CHECK ACCURACY AN D<br />

PARAMETER VALUE DIS<strong>PE</strong>RSI<strong>ON</strong><br />

IMPROV E<br />

MODE L<br />

DAILY TIME INCREMENT<br />

C<strong>ON</strong>VERT MODEL COEFFICIENTS<br />

AND PARAMETERS FOR DAIL Y<br />

TIME INCREMEN T<br />

CALCULATE AND ESTIMAT E<br />

SUBWATERSHED COEFFICIENT S<br />

AND PARAMETERS<br />

CALIBRATI<strong>ON</strong>, RESULTS STUD Y<br />

AND COMPARIS<strong>ON</strong> WITH<br />

M<strong>ON</strong>THLY MODE L<br />

NO Ị<br />

IMPROV E<br />

DEFINITI<strong>ON</strong><br />

OF PROCESSE S<br />

SUBWATERSHE D<br />

MODEL<br />

NO<br />

SENSITIVIT Y<br />

ANALYSIS<br />

APPLICATI<strong>ON</strong><br />

Figure 5 . Flow diagram for verification procedure .<br />

I<br />

6 0


Calibration of Parameter s<br />

In general, it is anticipated that realisti c<br />

parameter values are established through th e<br />

calibration procedure . However, when streamflow<br />

is the only available component for<br />

checking the model, it is possible that several<br />

combinations of parameter values will yiel d<br />

satisfactory agreement between observed an d<br />

computed outflow hydrographs . The proble m<br />

of establishing unique parameter values i s<br />

approached on the basis of hydrologic judgment<br />

and by using "interior" observations ,<br />

such as snow depth and soil moisture, a s<br />

check points on model performance . Other<br />

ways of testing the model include the tim e<br />

distribution of output quantities, such a s<br />

stream flow, and known (or estimated )<br />

monthly or annual quantities . For example ,<br />

for watershed 2 it is estimated that interception<br />

storage is about 0 .5 cm, and total interception<br />

amounts to approximately 17 percen t<br />

of the annual precipitation .<br />

Under the self-calibration technique model<br />

parameter values are altered or purturbed in a<br />

random sequence and the resulting changes in<br />

the objective function are examined (Shi h<br />

1971). A computer flow chart for the calibration<br />

subroutine is shown by figure 6 . The<br />

entire program model, including the calibration<br />

subroutine, will be synthesized on a<br />

hybrid computer.<br />

Sensitivity and Management Studies<br />

Sensitivity<br />

A sensitivity analysis is performed b y<br />

changing one system variable while holdin g<br />

the remaining variables constant and notin g<br />

the changes in the model output functions . If<br />

small changes in a particular system parameter<br />

induce large changes in the output or respons e<br />

function, the system is said to be sensitive t o<br />

that parameter . Thus, through sensitivity<br />

analyses it is possible to establish the relative<br />

importance with respect to system respons e<br />

of various system processes and input functions.<br />

This kind of information is useful fro m<br />

the standpoint of system management, syste m<br />

modeling, and the assignment of priorities i n<br />

the collection of field data. Under this study,<br />

the verified model will be used to perfor m<br />

various sensitivity analyses for the hydrologi c<br />

system of watershed 2 .<br />

Management<br />

Opportunities for management of a fores t<br />

watershed are widely varied, and range fro m<br />

changes in logging practices to forms of soil<br />

treatment. Actual implementation of a<br />

management scheme depends upon benefit s<br />

gained as compared with possible disbenefits .<br />

The simulation model developed under thi s<br />

study will not make direct comparisons of<br />

benefits and disadvantages, but will predict<br />

changes in the system output associated wit h<br />

given management alternatives . Under this<br />

study the capability of the model will be<br />

demonstrated for rapidly testing many<br />

possible management alternatives .<br />

Much of the work discussed by this paper i s<br />

based upon past developments in watershed<br />

simulation at Utah State University . Hydro -<br />

logic modeling of the H . J. <strong>Andrews</strong> <strong>Experimental</strong><br />

<strong>Forest</strong> for Coniferous <strong>Forest</strong> Biom e<br />

has just begun, and the preceding discussio n<br />

has been influenced by a consideration o f<br />

particular conditions in the study area . How -<br />

ever, the model will be fundamental in<br />

concept, and therefore generally applicable in<br />

a geographic sense . Whenever feasible, th e<br />

model will use basic equations which are valid<br />

for short time intervals to define the variou s<br />

processes in the model . The output will the n<br />

be summed for application to daily or longe r<br />

time increments . For example, Horton 's infiltration<br />

equation is applicable in minute units ,<br />

but by summing these quantities, the model i s<br />

capable of calculating equivalent daily infiltration<br />

rates . For the area under this study, how -<br />

ever, it is assumed that water supply rates a t<br />

the ground surface do not exceed infiltratio n<br />

capacities, so that surface runoff does not<br />

occur, thus considerably simplifying th e<br />

model calibration process .<br />

System functions which are important t o<br />

forest management and other aspects of th e<br />

total project include evapotranspiration and<br />

soil moisture . These functions, along wit h<br />

streamflow, will be estimated by the mode l<br />

for use in other parts of the total system<br />

61


SET UP PARAMETER OPTIMIZIN G<br />

ORDER IN RANDOM SEQUENC E<br />

CHANGE THE <strong>PE</strong>RTURBATI<strong>ON</strong> DIRECTI<strong>ON</strong><br />

WITHIN GIVEN PARAMETER LIMI T<br />

ASSIGN OPPOSITE DIRECTIO N<br />

OF <strong>PE</strong>RTURBATI<strong>ON</strong> I N<br />

NEXT CYCLE<br />

PARAMETERS TESTE D<br />

WITHOUT SIGNIFICANT<br />

IMPROVEMENT<br />

L = L + 1<br />

RESULT S<br />

FROM THIS S E<br />

OF PARAMETER VALU E<br />

SATISFY TH E<br />

C<strong>ON</strong>STRAINTS<br />

i RETURN<br />

Figure 6 . Flow chart for the model calibration .<br />

I<br />

E2


model being developed under the Coniferou s<br />

Biome Program. The important underlying<br />

feature throughout the entire study will b e<br />

that all of the separately described hydrologi c<br />

processes and phenomena are interlinked into<br />

a total system . Thus, from the model, hope -<br />

fully, it will be possible to evaluate the relative<br />

importance of the various items, explore<br />

critical areas where data and perhaps theor y<br />

are lacking, and finally establish guidelines fo r<br />

the improved management of forest water -<br />

sheds .<br />

INPU T<br />

(lOVO<br />

(A)<br />

SYSTEM<br />

(KNOWN )<br />

SYSTEM SYNTHESI S<br />

OUTPUT<br />

^-TI €N NNowN 1<br />

Hydrologic Systems Analysis 3<br />

INPUT<br />

(KNOWN)<br />

SYSTE M<br />

(UNKNOWN)<br />

OUTPU<br />

T (KNOWN )<br />

The purpose of this research is to devise a<br />

technique for statistical decomposition of a<br />

hydrologic event such that system processe s<br />

such as precipitation, subsurface flow, an d<br />

evapotranspiration, which contribute to th e<br />

observed streamflow can be separated and<br />

described. This technique will therefore pro -<br />

vide one more avenue for determination of<br />

the subsurface flow process on forest soils .<br />

The technique chosen for this research is a<br />

form of systems analysis .<br />

Systems, Definitions and Basic Principles<br />

System may be defined as an aggregate of<br />

physical parts that do not change with time ,<br />

operating on an input to produce an output ,<br />

both being functions of time . The simplified<br />

representation of a watershed, given in figure<br />

2, can be considered as a system whose input<br />

is precipitation and runoff its output . Syste m<br />

"synthesis" is a technique employed when th e<br />

system is known in terms of a mathematical<br />

equation; the objective is to determine th e<br />

nature of the output for any class of inpu t<br />

(fig. 7) . In system "analysis" a syste m<br />

response function or kernel which best de -<br />

scribes a given input-output pair is derived<br />

(fig. 7) . The term "best" implies that the<br />

derived kernels are not unique . Combination s<br />

of both techniques can be used for the solution<br />

of hydrologic problems . A system can<br />

3 Authored by Z. G. Papazafiriou, Research<br />

Associate, and R . H . Burgy, Professor, University o f<br />

California, Davis .<br />

(C)<br />

(B)<br />

SYSTEM ANALYSI S<br />

MULTI-INPUT/OUTPUT SYSTEM<br />

Figure 7 . Illustration of systems .<br />

have one input and one output or many in -<br />

puts and outputs (fig. 7). A system i s<br />

"lumped parameter" if input and output ar e<br />

functions of a single variable . Otherwise, th e<br />

system is of the "distributed parameter" type .<br />

If the system response at any time, due to a<br />

given input, is uniquely determined, th e<br />

system is said to be "deterministic." If the<br />

system response is subject to uncertain influences,<br />

the system is "stochastic o r<br />

probabilistic . "<br />

A quantity z is a "functional" for the function<br />

x(t) in the interval (a,b), if it depend s<br />

upon all values taken by x(t), when t varies i n<br />

the interval (a,b) . An illustration of a functional<br />

is given in figure 8 . The output of a<br />

system is a functional of the input and, fo r<br />

the same reason, runoff is a functional o f<br />

precipitation. A system is "time invariant " if<br />

it does not change with time. Such system s<br />

can be represented by functionals . "Physically<br />

realizable" is a system whose output at time t<br />

depends only upon past values of the input .<br />

63


can be almost linear, but there is no linea r<br />

system which can be almost nonlinear . In<br />

general, linearity is a limiting case of non -<br />

linearity. Therefore, any theory or technique<br />

adequate for a general nonlinear system i s<br />

equally adequate for linear systems .<br />

Deterministic Linear Hydrologic Systems<br />

The theory behind most linear methods ca n<br />

be generalized in the following manner . Suppose<br />

that s and a are continuous variable s<br />

representing position in space, and t and r<br />

define position in time . Consider the linear<br />

P. D. E . of the general form<br />

L[g(s,t)] = f(s,t) (17)<br />

Figure 8 . Demonstration of a functional .<br />

Hydrologic systems are physically realizabl e<br />

where L is linear P . D. operator of arbitrary<br />

since their outputs (runoff) at time t depen d<br />

order, and g(s,t) some function which satisfie s<br />

only on the past values of their inputs (precipitation)<br />

. The "memory" of a system is th e<br />

equation 17 within a certain region R . Given<br />

the appropriate homogeneous boundary conditions<br />

along R, the solution of equation 1 7<br />

time period between some past time and th e<br />

present for which the output depends only<br />

can be written according to Hildebran d<br />

upon the input . If the output depends only<br />

(1958) as<br />

on the present value of the input, the syste m<br />

is said to be a "no-memory" system . If the<br />

g(s,t) =f f G(s,t ; a,r) f (a,r) da dr (18)<br />

output of a time invariant system is analyti c<br />

R<br />

about zero input at some time to, the syste m If<br />

is "analytic". Analyticity is very important ,<br />

f(s,t) = fs (s) ft (t) (19)<br />

since if a system is analytic, its output can b e<br />

equation 18 can be written in the for m<br />

expanded in Volterra series . Hydrologic<br />

systems are assumed to be analytic.<br />

g(s,t) =f f(r) [f G(s,t ; a,r) do] dr (20)<br />

A deterministic system H is said to be<br />

s<br />

"linear," if given the inputs X 1 (t) and X 2 (t)<br />

such that<br />

If fs (s) is spatially invariant, we may write<br />

y 1 (t) = H [AX 1(t)] (1 4)<br />

f<br />

ti ti<br />

g(s,t) = G (s,t ; r) f (r) dr (21)<br />

y 2 (t) = H [BX 2 (t)] (15)<br />

implies that<br />

Y1 (t) + Y2 (t) = Y[ X (t)] ( 16)<br />

= H [AX 1 (t) + BX 2 (t) ]<br />

= AH [X 1 (t)] + BH [X 2 (t)]<br />

that is, in a linear system, each member of a<br />

sequence of input values influences the out -<br />

put independently of every other . This is the<br />

well known principle of superposition. If a<br />

system does not satisfy the above condition i t<br />

is said to be "nonlinear . " A nonlinear system<br />

where<br />

ti ti<br />

f (r) = f(s,t), and G(s,t ; r) (22)<br />

= fG(s,t ; a,r) da<br />

s<br />

We can write equation 21 in differentia l<br />

equation form as<br />

d<br />

An(s,t)<br />

dtn't) + An 1(s,t) dn-lg(s,t) +<br />

dt n- 1<br />

+ A 0(s,t) g(s,t) = f(s,t) (23)<br />

64


Usually, we are interested in the output variable<br />

at some particular point in space ( a<br />

particular gaging station), in which case equation<br />

23 becomes<br />

An(t) dng(t)<br />

+ An-1(t)<br />

do-1g(t)<br />

+<br />

dtn dt n-1<br />

▪ + A0(t) g(t) = f(t)<br />

(24)<br />

which describes a spatially lumped parameter ,<br />

time-varying linear system. If we assume that<br />

the parameters in equation 24 are time in -<br />

variant, we obtai n<br />

An ddgnt) + An-1 do-lg(t) +<br />

▪<br />

Aog(t) = f(t)<br />

dtn- 1<br />

(25)<br />

which describes a time invariant, lumpe d<br />

parameter linear system . If we assume that<br />

the system is completely at rest at t=0, we can<br />

write equation 25 in the form of the con -<br />

volution equation<br />

t<br />

g(t) = f h(r) f(t-r) dr (26 )<br />

0<br />

which is the basis of the unit hydrograp h<br />

theory and many other hydrologic techniques.<br />

In equation 26, g(t) represents runoff,<br />

f(t) rainfall, and h(t) is the kernel, or in thi s<br />

case, the unit hydrograph .<br />

In summary, application of equation 26<br />

implies that the watershed behaves as a linear<br />

system, it is time invariant, the rainfall is<br />

uniformly distributed over the watershed<br />

area, and the watershed is completely at res t<br />

at the beginning of the rain . The "effective"<br />

precipitation is used as an input to th e<br />

system. This implies that we know some<br />

method for the separation of the runof f<br />

hydrograph into base flow and direct runoff<br />

(fig. 9) . This approach in fact is a combination,<br />

using system synthesis for the estimation<br />

of the effective precipitation, and syste m<br />

analysis for the estimation of runoff. Methods<br />

using this technique have been developed b y<br />

Snyder (1955), Eagleson et al. (1966), Nash<br />

(1957, 1960), O'Donnell (1960), Dooge<br />

(1965), and others .<br />

Deterministic Nonlinear Hydrologic System s<br />

As it was noted in a previous section, a<br />

time invariant analytic system can be expanded<br />

in Volterra series . Such an expansion<br />

can be written in the form<br />

y(t)=ho+ f h 1( T1) x ( t-r1) dr 1<br />

+ f f h 2 (T1 ,T2) X (t-T1) X(t-T 2 )dr 1 dr 2<br />

+f f hn(T 1 Tn) x(t-T 1 ) ...<br />

x(t-Tn) dr 1<br />

. . . drn<br />

+ (27)<br />

EFFECTIVE RAINFALL = DIRECT RUNOF F<br />

xUI<br />

Y(T)<br />

BASE FLOW<br />

m.'UtAut%A%atU<br />

Figure 9 . Estimation of effective rainfall through hydrograph separation .<br />

65


where hi are the kernels of the system an d<br />

ho=0 unless a source or sink is present. If th e<br />

system is physically realizable, and has finite<br />

memory (as it happens with hydrologi c<br />

systems), equation 27 can be written in th e<br />

form u<br />

r y ( t) = J h 1 (r 1 ) x(t-r1) dr 1<br />

U<br />

0<br />

u<br />

+ f ,/ 3 r h 2 ( T 1 ,T2) X4-T 1 ) X(t-T2 )dr1 dr2<br />

0 0<br />

+ . . . . . . . . . . . . . .<br />

+<br />

u<br />

f . . .<br />

u<br />

f hn (T 1 . . .Tn)x(t-T1) . . .<br />

0 0<br />

x(t-T n )dT i<br />

. . . dTn<br />

subject to the condition<br />

h i(t) = 0 for all r < 0<br />

(28)<br />

and where u is the length of the memory . If ,<br />

instead of continuous functions, discrete set s<br />

of data are used, equation 28 can be written<br />

in the form<br />

U<br />

y(T) = H 1(S1) X(T-S 1 )<br />

S 1 = 0<br />

U U<br />

+ E E H 2 (S1,S2) X(T-S1 )<br />

S 1 =0 S 2 =0 X(T-S 2 )<br />

U U<br />

+ E . . . E H n(S I . .. S n ) X(T-S 1 )<br />

Si =0 Sn = O X(T-S n ) (29)<br />

The first term in equation 28 or 29 represents<br />

a linear system, that is an ordinary con -<br />

volution integral. The other terms are a<br />

generalization of the convolution integral . In<br />

general, the i th term represents a pure subsystem<br />

of order i. Therefore our system y(t) i s<br />

composed of the summation of a linear and a<br />

series of nonlinear subsystems . If we represent<br />

the successive terms of the system by H i (t) ,<br />

H 2 (t), . . ., Hn(t), . . ., respectively, the system<br />

can be written in the for m<br />

y(t) = H1(t) + H 2 (t) + . . .. + Hn(t) + . . . . (30)<br />

An illustration of a nonlinear system is give n<br />

in figure 10 .<br />

A system is identified whenever its kernel s<br />

hn(t) are calculated. This evaluation is base d<br />

on the past behavior of the system. Theoretically,<br />

after knowing all the kernels, th e<br />

response of the system can be calculated fo r<br />

any given set of input values .<br />

Equation 28 (in the form given) is very<br />

generalized and its usage is limited and tim e<br />

consuming . It is desirable to reduce the operations<br />

involved . For example, given a sequence<br />

of inputs, sometimes sequential values ar e<br />

highly correlated, and the system itself may<br />

smooth out rapid fluctuations . Hydrologic<br />

systems demonstrate both of these characteristics<br />

. Thus, we may look for ways whic h<br />

can describe the sequence by a smalle r<br />

number of parameters . If the sequence x 1 . . . .<br />

Xi . . . . Xn represents observations at times<br />

t1 . . ..tj . . . .tn, respectively, they can be approximated<br />

by a polynomial of the for m<br />

k<br />

X k(t) = aa +alt+ . ..+aktk= a m tm (31)<br />

m= 0<br />

where k


LINEAR<br />

Y~(T)<br />

2nd ORDER<br />

Y2 (T)<br />

3rd ORDER<br />

Y 3 (T)<br />

INPUT<br />

X (T)<br />

OUTPUT<br />

Y (T )<br />

}<br />

nth ORDER<br />

Yi (T )<br />

Figure 10 . Illustration of nonlinear system representation by Volterra Series.<br />

the memory, an exact match could be made<br />

to each of the input values . However, n o<br />

economy in the description would have bee n<br />

affected . Instead, fewer polynomials can b e<br />

used. This introduces an error, but due to th e<br />

nature of the operation, it will be the leas t<br />

error possible .<br />

Methods using the above procedure hav e<br />

been introduced by Jacobi (1966), Harder<br />

and Zand (1969), and Brandstetter an d<br />

Amorocho (1970) . It is this method which is<br />

being considered for use in this study . There<br />

are certain distinct advantages associated with<br />

the process . The method is quite general and<br />

can handle a variety of problems such as runoff,<br />

chemical quality of runoff waters, and<br />

suspended sediment predictions, when inpu t<br />

values are properly weighted by functions<br />

describing the physical processes involved in<br />

each case. Once the response functions or<br />

kernels of the system are evaluated, they may<br />

be used for predictions given any sequence o f<br />

inputs. It requires a minimum amount of<br />

data, possibly 1 to 2 years of good records .<br />

Systems can be used in cascade, lik e<br />

predict predict<br />

precip.-runoff quality or sediment<br />

It can be used for quantitative evaluations o f<br />

the changes created by any type of watershe d<br />

management procedure by evaluating th e<br />

kernels of the original and the manage d<br />

hydrologic system. Emphasis will be given to<br />

establishing weighted functions for the bes t<br />

description of the physical process, and o n<br />

deriving tools for the greatest economizatio n<br />

of the procedure .<br />

67


Watershed Stratification :<br />

A Problem on Watershed 10 4<br />

W e have also begun to structure the<br />

hydrologic model for watershed 10 concurrently<br />

with the simulation study underwa y<br />

for watershed 2 and the watershed system s<br />

analysis. Our objective here is to prepare a<br />

fine-resolution hydrologic model which in -<br />

corporates the transpiration model from th e<br />

Primary Producers, the evapotranspiratio n<br />

model from Meteorology, and provides soi l<br />

moisture and subsurface water flow for th e<br />

Primary Producer and Bio-Geochemical Processes<br />

groups . The first step in structuring suc h<br />

a model is system stratification .<br />

One key to the analysis of complex system s<br />

is the compartmentalization of the syste m<br />

into homogeneous subsystems which can the n<br />

4 Authored by George W . Brown, Associate Professor,<br />

Oregon State University, Corvallis .<br />

be isolated for study . This should be done in<br />

such a way that the linkage between compartments<br />

is simple and direct .<br />

Another key is placement of compartmen t<br />

boundaries in such a way that the cells ar e<br />

easily uncoupled, or are coupled as a simpl e<br />

linear cascade . To do otherwise would complicate<br />

the modeling considerably. A hydrologic<br />

model that consists of a series of compartments<br />

arranged as a branching cascade is<br />

extremely difficult to manage . Water flow<br />

from an upper compartment must be some -<br />

how divided between lower compartments in<br />

the cascade. The basis for such division i s<br />

generally obscure and usually arbitrary .<br />

In our attempt to structure the hydrologi c<br />

model for watershed 10 at the next level o f<br />

resolution, we began by setting compartmen t<br />

boundaries along stream courses . This automatically<br />

decoupled the compartments, sinc e<br />

water does not cross the channel (fig . 11) .<br />

Next, it was necessary to consider the arrangement<br />

of the plant communities within<br />

Figure 11. Initial and secondary stratification of watershed 10, H . J. <strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong>, Oregon .<br />

68


the watershed. The hydrologic model and the<br />

primary producer model are obviously linked .<br />

The principal coupling variable betwee n<br />

models is the soil moisture profile . Soil moisture<br />

is that portion of the "hydrologic state "<br />

of the watershed which closely regulates plant<br />

growth. Plants, in turn, influence soil moisture<br />

by transpiration . Thus, superimpositio n<br />

of the primary producer 's vegetative structure<br />

upon the structure of the hydrologic system i s<br />

essential . This structure was combined wit h<br />

the initial hydrologic stratification and de -<br />

fined the subcompartment boundaries . Sub -<br />

compartment boundaries approximate th e<br />

vegetative type-map boundaries and are arranged<br />

into riparian, midslope and ridgeto p<br />

zones. These zones undoubtedly reflect th e<br />

changes in soil moisture regime within th e<br />

watershed . Also, this stratification allows u s<br />

to consider flow between subcompartments a s<br />

simple linear cascades .<br />

This final stratification for watershed 1 0<br />

will provide the basis for sampling schemes t o<br />

characterize soil moisture, water flow and<br />

other hydrologic and biologic processes necessary<br />

for the next round of model construction.<br />

It is essential to note that this stratification<br />

is compatible for modeling hydrologi c<br />

processes and is also compatible for linking<br />

the hydrologic model with that of the primary<br />

producers . It is the major achievement<br />

of our initial modeling effort and sets the<br />

stage for continued progress .<br />

Acknowledgments<br />

The work reported in this paper was supported<br />

in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferou s<br />

<strong>Forest</strong> Biome, U .S . Analysis of Ecosystems ,<br />

International Biological Program . This is Contribution<br />

No. 24 to the Coniferous <strong>Forest</strong><br />

Biome .<br />

Literature Cited<br />

Amorocho, J., and B . Espildora . 1966 . Mathematical<br />

simulation of the snow melting<br />

processes. Univ. Calif. Water Sci. & Eng .<br />

Pap. No. 3001, 156 p . Davis .<br />

Anderson, E. A., and N. H . Crawford . 1964 .<br />

The synthesis of continuous snowmelt run -<br />

off hydrographs on a digital computer .<br />

Stanford Univ ., Dep. Civil Eng. Tech. Rep .<br />

No. 36, 103 p . Palo Alto, Calif .<br />

Barnett, L. O . 1963. Storm runoff characteristics<br />

of three small watersheds in wester n<br />

Oregon. 84 p. M .S. thesis, on file at Colo .<br />

State Univ., Fort Collins .<br />

Brandstetter, A ., and J. Amorocho. 1970 .<br />

Generalized analysis of small watershed<br />

responses. Univ. Calif. Dep. Water Sci . &<br />

Eng . Pap. No. 1035, 204 p . Davis .<br />

Crawford, N. H., and R. K. Linsely . 1964 .<br />

Digital simulation in hydrology : Stanfor d<br />

Watershed Model IV . Stanford Univ . Dep .<br />

Civil Eng . Tech. Rep. No. 39, 210 p . Palo<br />

Alto, Calif .<br />

Dooge, J . C . I. 1965. Analysis of linear systems<br />

by means of Laguerre functions . J .<br />

Soc . Ind . Appl. Math. Cont., A. 2 :<br />

396-408 .<br />

Dyrness, C . T. 1969 . Hydrologic properties o f<br />

soils on three small watersheds in th e<br />

western Cascades of Oregon . USDA <strong>Forest</strong><br />

Serv. Res. Note PNW-111, 17 p . Pac .<br />

Northwest <strong>Forest</strong> & Range Exp . Stn., Portland,<br />

Oreg .<br />

Eagleson, P . S., R. Mejia, and F . March . 1966 .<br />

The computation of optimum realizabl e<br />

unit hydrographs. Water Resour . Res. 2(4) :<br />

755-765 .<br />

Eggleston, Keith O ., Eugene K . Israelsen, and<br />

J . Paul Riley . 1971 . Hybrid computer simulation<br />

of the accumulation and melt processes<br />

in a snowpack . Utah State Univ . Coll .<br />

Eng., Utah Water Res. Lab., Pap. No .<br />

PRWG 65-1, 77 p . Logan, Utah .<br />

Forsythe, G . E . 1957 . Generation and use of<br />

orthogonal polynomials for data-fittin g<br />

with a digital computer . J . Soc. Ind. Appl .<br />

Math. 5(9) : 74 .<br />

Harder, J . A., and S . Zand . 1969 . The identification<br />

of nonlinear hydrologic systems .<br />

Hydraul . Eng. Lab. Tech . Rep . HEL8-2, 8 3<br />

p. Univ. Calif., Berkeley .<br />

Hildebrand, F . B. 1958. Methods of applie d<br />

mathematics . 523 p. New York : Prentice -<br />

Hall .<br />

69


Jacobi, S . L. S. 1966. A mathematical mode l<br />

for nonlinear hydrologic systems . J. Geol .<br />

Res. 71 : 4811 .<br />

Jones, A. 1971 . Soil piping and stream channel<br />

initiation. Water Resour . Res. 7(3) :<br />

602-610 .<br />

Nash, J. E . 1957 . The form of the instantaneous<br />

unit hydrograph. Int. Assoc. Sci .<br />

Hydrol . Gen. Assem. Toronto, Proc ., 3 :<br />

114-121 .<br />

. 1960. A unit hydrograph study<br />

with particular reference to British catchments<br />

. Inst . Civil. Eng. (Proc.) 17 :<br />

249-282 .<br />

O ' Donnell, T . 1960. Instantaneous unit<br />

hydrograph derivation by harmonic analysis.<br />

Assoc . Sci. Hydrol. Publ. No. 51, p .<br />

546-557 .<br />

Pysklywec, D . W., K. S. Davar, D . I. Bray .<br />

1968 . Snowmelt at an index plot. Water<br />

Resour. Res . 45 : 937-946 .<br />

Riley, J. P ., D . G . Chadwick, and J . M .<br />

Bagley. 1966. Application of electroni c<br />

analog computer to solution of hydrologi c<br />

and river-basin-planning problem : Utah<br />

Simulation Model II . Utah State Univ . ,<br />

Utah Water Res . Lab. PRWG 32-1, 121 p .<br />

Logan, Utah .<br />

Rothacher, Jack. 1963 . Net precipitation<br />

under a Douglas-fir (Pseudotsuga menziesii)<br />

forest. <strong>Forest</strong> Sci . 9 : 4 .<br />

, C . T. Dyrness, and Richard L .<br />

Fredriksen . 1967. Hydrologic and related<br />

characteristics of three small watersheds in<br />

the Oregon Cascades . USDA <strong>Forest</strong> Serv .<br />

Pac. Northwest <strong>Forest</strong> & Range Exp . Stn . ,<br />

54 p. Portland, Oreg.<br />

Shih, C . C. 1971. A simulation model for predicting<br />

the effects of weather modificatio n<br />

on runoff characteristics . 116 p . Ph.D .<br />

thesis, on file at Utah State Univ ., Logan .<br />

Sittner, W. T., C. E. Schauss, and J . C . Monro .<br />

1969. Continuous hydrograph synthesi s<br />

with an API-type hydrologic model . Water<br />

Resour . Res. 5(10) : 1007-1022 .<br />

Snyder, W. M. 1955. Hydrograph analysis b y<br />

the method of least squares . Am . Soc . Civil<br />

Eng. Proc., J. Hydraul . Div. 81(793), 25 p .<br />

U .S. Army Corps of Engineers . 1956 . Summary<br />

report of snow investigations, sno w<br />

hydrology . 437 p . Portland, Oreg.<br />

U .S . Department of Agriculture . 1964 . Irrigation<br />

water requirements . Soil Conserv .<br />

Serv. Eng . Div . Tech. Release No . 21 . Washington,<br />

D .C .<br />

Veihmeyer, F . J. 1964 . Evapotranspiration . In<br />

Ven to Chow (ed.), Handbook of applie d<br />

hydrology, p . 11-1 to 11-23 . New York :<br />

McGraw-Hill .<br />

70


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 1972 .<br />

Preliminary consideration s<br />

of the forest canopy<br />

consumer subsystem<br />

M. A. Strand and W . P . Nagel<br />

Oregon State University<br />

Corvallis, Orego n<br />

Abstract<br />

A food chain beginning with herbaceous materials produced in the forest canopy is delineated as a<br />

subsystem. A conceptual model of energy flow through this subsystem includes assumptions and definitions<br />

relating to components of the canopy food chain, processes by which energy is transferred, and the energ y<br />

pathways. From this conceptual model, it is observed that the canopy food chain is directly coupled to othe r<br />

consumer subsystems through common predators and through the detritus component . Besides the direct<br />

relationship through the food base to primary production, this subsystem may also influence the plant syste m<br />

by its effects on the sites ofplant hormone production and on the fate of mobile nutrients in the ecosystem .<br />

Introduction<br />

In the past, modeling efforts concernin g<br />

consumers have been centered on populatio n<br />

dynamics, while more recently energetics hav e<br />

been receiving attention . Our consideration s<br />

of consumers will also concern energy transfer.<br />

We will focus on the couplings betwee n<br />

consumers and other parts of the ecosystem .<br />

These couplings may be direct (i .e., involve<br />

energy transfer to or from a consumer) o r<br />

indirect (i .e., involve the influence of on e<br />

component on energy transfer between othe r<br />

components of the ecosystem) . The majo r<br />

subdivisions of the ecosystem are discussed b y<br />

Overton (1972) .<br />

To facilitate our understanding of th e<br />

consumer system, it would be convenient t o<br />

subdivide it into a hierarchical set of units an d<br />

observe the couplings between them and their<br />

couplings with other ecosystem components .<br />

If our units are to be subsystems, then th e<br />

most logical divisions would occur aroun d<br />

groups of components with high degrees o f<br />

interconnections. The majority of linkage s<br />

with a subsystem's components should occu r<br />

within the subsystem ; couplings to components<br />

of other subsystems should be comparatively<br />

rare. The classic way to defin e<br />

consumer subsystems has been by trophi c<br />

levels; however, for our purpose this divisio n<br />

may not be the most meaningful one . By<br />

basing the division of the consumer system o n<br />

trophic level designations, the direct linkage s<br />

with regard to energy flow occur between th e<br />

subsystems and only indirect linkages, such a s<br />

competition, occur within them . We hypothesize<br />

that a better subdivision of the consumer<br />

system could be made by defining th e<br />

subsystems as food chains on the basis of a<br />

stratification of their primary food base .<br />

The first division in our hierarchy will b e<br />

into grazing and detritus food chains . Th e<br />

grazing food chain includes primary consumers<br />

feeding on plant material and a serie s<br />

of secondary consumers that prey on the<br />

herbivores and on each other . To further<br />

subdivide this food chain, the food base i s<br />

stratified into forest canopy herbaceous material,<br />

woody material, forest floor herbaceou s<br />

material, and roots . Food chains are define d<br />

by the food material eaten by the primar y<br />

consumers in the chain . Likewise, the detritu s<br />

food chain which recycles detritus to lowe r<br />

71


grades of detritus is divided according t o<br />

particle size. The two resulting food chain s<br />

use fine or coarse detritus as their primar y<br />

food sources. By uncoupling these six foo d<br />

chains (four grazing and two detritus) with<br />

regard to their food bases, the first orde r<br />

consumers are reasonably unique to a particular<br />

food chain subsystem, while commo n<br />

predators couple the subsystems together .<br />

The purposes of this paper will be to defin e<br />

the forest canopy food chain and to develop a<br />

conceptual model of it as an example of th e<br />

type of preliminary consideration which is a<br />

prerequisite to mathematical modeling.<br />

The forest canopy food chain forms a<br />

subsystem that begins with herbaceous material<br />

produced by trees and follows it throug h<br />

its many transfers to detritus . The food bas e<br />

is primarily photosynthetic tissue, needles an d<br />

leaves of trees ; however, buds, young twigs ,<br />

immature cones, and seeds will also be included,<br />

since consumption of these foods i s<br />

an interrelated process . For present considerations,<br />

we will confine our concern to an<br />

old-growth, coniferous forest canopy . Successional<br />

changes, epidemic outbreaks, an d<br />

environmental gradients will not be include d<br />

in our discussions . It is assumed that variations<br />

in energy flow from year to year are no t<br />

appreciable, so there are periods of short-ter m<br />

stability . The conceptual model we will pro -<br />

pose is a description of this type of stabl e<br />

condition .<br />

Our discussion will consist of four parts :<br />

the components of the subsystem, the processes<br />

involved in energy transfer, energy path -<br />

ways, and the interconnections between thi s<br />

and other food chain subsystems as well a s<br />

the subsystems defined by Overton (1972) i n<br />

this symposium . In each section we wil l<br />

present a set of assumptions and definition s<br />

that represent our notion of this subsystem .<br />

The set will form our conceptual model .<br />

Components<br />

The components of a subsystem are the<br />

locations of energy storage and vehicles o f<br />

energy transfer along the food chain . Components<br />

are defined by ecological function<br />

rather than taxonomic criteria . However great<br />

species diversity might be in a forest canopy ,<br />

functional roles are common to many species ,<br />

enabling them to be regarded as a few larg e<br />

populations. To avoid the pitfall of over -<br />

generalization about the superpopulations, a<br />

brief discussion of the variations of behavio r<br />

relevent to energy transfer of our components<br />

is included .<br />

Primary consumers are the key links in th e<br />

food chains, diverting energy produced by th e<br />

plants into the animal system . In the fores t<br />

canopy most consumers are insect grazers<br />

which feed mainly on new tissue from expanding<br />

needles or leaves (Keen 1952) . The<br />

feeding stage either hatches or emerges near a<br />

feeding site, so that food finding initially i s<br />

not a problem . Many of them feed gregariously<br />

when larvae are small, but disperse t o<br />

"feed singly later. The main disseminatio n<br />

phase, however, is the adult where female s<br />

seek out new feeding sites for oviposition .<br />

The feeding period occurs primarily in th e<br />

spring and early summer with resting phase s<br />

(pupae, eggs, or overwintering larvae) beginning<br />

in late August or early September . A<br />

second group of primary consumers include s<br />

insects with sucking mouth parts . They act as<br />

physiological sinks and remove dissolve d<br />

nutrients from the xylem sap (Way an d<br />

Cammell 1970). Only one significant vertebrate<br />

foliage feeder has been reported in th e<br />

forest canopy-the red tree mouse (Phenacomys<br />

longicaudus) (Maser 1966). It eats<br />

needles of Douglas-fir (Pseudotsuga menziesii )<br />

leaving the central xylem strand . Smal l<br />

branches are cut and carried to the nest where<br />

they are consumed .<br />

Plant reproductive tissues are also consumed.<br />

Immature cones are attached mainl y<br />

by cone and seed feeding insects ; some fee d<br />

exclusively on the seeds, while others feed o n<br />

the cone preventing seed development . Mature<br />

seed may be consumed prior to dissemination<br />

by certain birds and squirrels . Thes e<br />

omnivorous birds use seeds as a primary foo d<br />

source during the winter ; they remove undisseminated<br />

seeds while the cones are stil l<br />

attached to the tree (Isaac 1943) . Squirrel s<br />

cut cones and extract the seeds on th e<br />

ground ; the subsequent fate of the seeds i s<br />

72


considered as part of another food chai n<br />

subsystem which utilizes forest floor herbaceous<br />

material .<br />

Predators and parasites constitute the nex t<br />

links in the food chain . Small invertebrate s<br />

are consumed generally by other invertebrates<br />

. Parasitic forms may feed on larvae ,<br />

pupae, or eggs of the host . Often several<br />

generations of parasites are produced for<br />

every single host generation ; hyperparasitis m<br />

of several degrees is possible. Predaceous<br />

invertebrates include both spiders and insects .<br />

Some, as the orb-spinning spiders, wait fo r<br />

mobile forms to be trapped, while other s<br />

move about actively in search of food an d<br />

feed on nearly any invertebrate they ca n<br />

capture (Graham 1952) .<br />

During the mating and nesting season ,<br />

omnivorous birds feed on insects with eg g<br />

hatch often being correlated with peak insec t<br />

abundance (Lack 1954) . Very small larvae are<br />

usually not eaten since they are inconspicuou s<br />

and many would be required to satisfy a<br />

bird 's energy needs . Flycatchers, which ar e<br />

not abundant in the winter, breed in the<br />

forest canopy and feed mainly on adult<br />

insects . Other insectivorous birds search th e<br />

foliage for the larger larvae . Predation of adult<br />

birds is rare, but nest predators may be foun d<br />

consuming both eggs and young nestlings .<br />

Other predatory birds are also rare components<br />

of the forest canopy ; however, th e<br />

spotted owl (Strix occidentalis) is an important<br />

predator of vertebrate foliage feeder s<br />

(Nussbaum 1972 1 ) .<br />

The components of the forest canopy food<br />

chain are restricted to those animals that ca n<br />

complete at least the feeding stages of their<br />

life cycles above the shrub stratum. In our<br />

considerations of the forest canopy, we hav e<br />

restricted ourselves to endemic population<br />

conditions and have not considered the larg e<br />

numbers of defoliators found in some areas .<br />

The actual population densities found under<br />

"normal" circumstances is not known for th e<br />

old-growth stands, and estimates of thes e<br />

densities will be a subject of future research .<br />

We will consider herbivores to be predator<br />

limited and predators to be food limited . All<br />

' Personal communication .<br />

components will be assumed to be balanced<br />

with respect to immigration and emigration<br />

except for seasonally migrating birds .<br />

In summary, we have designated nine<br />

superpopulations or components in the forest<br />

canopy food chain subsystem : grazing insects ,<br />

sucking insects, vertebrate foliage feeders ,<br />

seed and cone insects, omnivorous birds ,<br />

parasitic invertebrates, predaceous invertebrates,<br />

nest predators, and other predator y<br />

birds. Energy flow will be followed between<br />

these components ; possible transfers within<br />

them will not be considered explicitly .<br />

Processes<br />

Certain processes are essential for the transfer<br />

of energy, and we have defined five main<br />

processes for the food chain subsystems :<br />

consumption, elimination, respiration, assimilation,<br />

and death . The definitions we use may<br />

not be the same as those found in ecologica l<br />

literature, but our definitions are consistent<br />

with the purpose of defining direct couplings<br />

within and between the food chains and other<br />

subsystems of the ecosystem .<br />

The concept of consumption is often<br />

equated with ingestion ; however, the importance<br />

of consumers may not be measured by<br />

what they ingest alone . Shelter building an d<br />

wasteful feeding behavior may result in fa r<br />

greater death of the food resource than th e<br />

amount ingested. Hence consumption wil l<br />

include ingestion as well as other activities o f<br />

the animals that result in the loss of life o f<br />

their food. All forms of waste production b y<br />

animals after the food is ingested is include d<br />

in the process of elimination . Thus, energ y<br />

lost as a result of indigestion, metaboli c<br />

wastes, or losses of integument will be transferred<br />

to detritus via elimination. Assimilatio n<br />

is defined as individual secondary production .<br />

It is the process by which energy from on e<br />

trophic level is incorporated into the tissue of<br />

the next. The process of respiration include s<br />

the energy loss to the atmospheric sink as th e<br />

result of maintenance metabolism and o f<br />

work. Death includes mortality losses as a<br />

result of factors other than consumption b y<br />

another component of a food chain. Both the<br />

73


processes of death and elimination result i n<br />

the production of detritus, but they must b e<br />

considered separately since they differ in their<br />

consequences to the food chain components .<br />

Elimination does not affect the potentia l<br />

activity of a component, whereas, death may .<br />

The purpose of defining these five processes<br />

is to identify the means by whic h<br />

components are directly coupled with eac h<br />

other or with other subsystems of the ecosystem.<br />

Each one is influenced by indirec t<br />

couplings . For example, feeding behavior ,<br />

population age structure, and energy demand s<br />

are indirect couplings between the feeding<br />

population and the process of consumption .<br />

Nutritional quality of the food may influenc e<br />

the processes of assimilation and elimination .<br />

These indirect couplings may be thought of as<br />

informational transfers and may be as impor -<br />

tant to the system as the direct couplings .<br />

Energy Pathways<br />

The general energy pathway is diagramme d<br />

in figure 1 . The boxes are used to show sites<br />

of energy storage . They are associated with a<br />

set of variables which are used to describe the<br />

site. The consumed tissue compartment is a<br />

hypothetical site but is included to separate<br />

the processes of consumption and assimilation<br />

. Circles are used to indicate the processe s<br />

which are employed to transfer energy between<br />

sites . Solid arrows mean true flows of<br />

materials or energy (direct couplings), whil e<br />

dotted arrows and diamonds indicate lines o f<br />

influence (indirect couplings) from an energ y<br />

storage site to a process .<br />

= Consumption<br />

WO<br />

0<br />

0<br />

= Waste (Subproces s<br />

of consumption<br />

= Assimilatio n<br />

Respiration<br />

= Death<br />

= Elimination<br />

4<br />

Food quality an d<br />

availabilit y<br />

Population ag e<br />

structure<br />

Food quality<br />

Food qualit y<br />

Population feeding an d<br />

sheltering behavio r<br />

Abiotic environmen t<br />

Figure 1 . Basic energy pathway through food chain subsystem . See text for explanation of symbols .<br />

74


PRIMARY PRRODUCTIO N<br />

■<br />

FOLIAGE<br />

C<strong>ON</strong>ES AN D<br />

SEEDS<br />

r<br />

GRAZIN G<br />

VERTEBRATE S<br />

GRAZIN G<br />

INSECTS<br />

SUCKIN G<br />

INSECTS<br />

SEED AN D<br />

C<strong>ON</strong>E INSECTS<br />

f<br />

V<br />

PREDACEOU S<br />

BIRDS<br />

PARASITI C<br />

INVERTEBRATES<br />

PREDACEOUS<br />

INVERTEBRATE S<br />

Figure 2 . The major energy pathways between components of the forest canopy food chain .<br />

The paths of energy (fig. 2) may be<br />

followed beginning in the spring when th e<br />

buds of the conifers swell and insect herbivores<br />

emerge and feed . While they are small ,<br />

the grazing and sucking insects are preyed<br />

upon by other invertebrates . Later they are<br />

preyed upon by birds that switch from<br />

conifer seeds to insect larvae during thei r<br />

breeding season . Energy leaves the system in<br />

the form of food for predators in other foo d<br />

chains, migrating birds, detritus, and heat .<br />

The amounts of energy following these routes<br />

is not known yet . Consumption rates, relative<br />

densities of consumers, food quality variability,<br />

assimilation efficiencies, an d<br />

migration habits are also not known for mos t<br />

of the consumers . These problems will require<br />

further research before the model may b e<br />

quantified. The two diagrams and the definitions<br />

and assumptions that were outlined i n<br />

the preceding sections represent our curren t<br />

conceptual model of the canopy food chain .<br />

Interconnections with<br />

Other Subsystem s<br />

The conceptual model presented here of<br />

the forest canopy consumer subsystem ma y<br />

be used to describe possible interrelation s<br />

between it and other subsystems. By definition,<br />

this subsystem contains consumers tha t<br />

feed primarily in the forest canopy ; therefore ,<br />

we would not expect it to be tightly coupled<br />

to other food chains . However, since the food<br />

base defines the subsystem and, in the case of<br />

the canopy food chain, herbivores are mor e<br />

specific in food habits than the carnivores, it<br />

is evident that direct couplings will probabl y<br />

result from the feeding behavior of th e<br />

secondary consumers . For example, adul t<br />

insects from other subsystems may serve as<br />

food for flycatcher birds which live primaril y<br />

in the canopy . Predaceous invertebrates from<br />

the forest floor feed upon canopy specie s<br />

75


when they move to the floor for overwintering<br />

or pupation. The food chain with primary<br />

consumers that feed on fine detritus i s<br />

coupled to this subsystem through the detritus<br />

output, and the canopy food chain ma y<br />

influence the energy flow rates within th e<br />

detritus food chain by the quality of its<br />

detrital outputs .<br />

Besides these connections with other foo d<br />

chain subsystems, the canopy consumers ar e<br />

closely linked to the primary productio n<br />

system . The consumers are influenced by th e<br />

quantity and quality of herbaceous materia l<br />

produced by the trees . However, the influence<br />

of the canopy food chain on primary production<br />

is more subtle. Although consumption o f<br />

foliage or cones is not considered to be at a<br />

rate to affect production in the old-growt h<br />

canopy of the model, the preferential consumption<br />

of new needles and expanding cone s<br />

may affect the nutrient capital of the trees .<br />

These consumption sites are physiologica l<br />

sinks, and mobile nutrients are actively moved<br />

to them from senescent foliage and storag e<br />

sites (Sweet and Wareing 1966) . The loss o f<br />

these sinks may mean the return of certain<br />

nutrients to the soil solution rather than thei r<br />

retention by the trees (Rafes 1970) . Also ,<br />

consumption of bud tips reduces the sites o f<br />

growth hormone production ; this hormone i s<br />

necessary for cambial division, and loca l<br />

deficiencies may result from the gregariou s<br />

feeding habits of some herbivores (Kozlowsk i<br />

1969). These indirect couplings may prove t o<br />

be the most significant role of the food chai n<br />

subsystems in the old-growth forest .<br />

Summary<br />

The consumer system is composed of a<br />

series of grazing and detritus food chains . Th e<br />

primary consumers are thought to be reason -<br />

ably unique to each food chain, and direc t<br />

energy transfers between the food chain s<br />

occur via common predators . The processes of<br />

energy transfer that link the components of<br />

the food chains with each other and the rest<br />

of the ecosystem are : consumption, assimilation,<br />

elimination, respiration, and death. Indirect<br />

links are formed by informational transfers<br />

which influence energy flows .<br />

The forest canopy food chain includes the<br />

primary consumers that feed on herbaceous<br />

material produced in the canopy and the<br />

series of related predators . This food chain is<br />

directly linked to primary production by<br />

consumption of herbaceous materials, to th e<br />

other food chains by common predators, and<br />

to the detritus component through production<br />

of detrital material . Primary productio n<br />

influences the food chain processes by variations<br />

in nutritional quality, spatial arrangement,<br />

and quantity of material produced . Th e<br />

influences of the food chain on primar y<br />

production are more subtle . Preferential feeding<br />

habits of many canopy grazers affect th e<br />

number of sites of plant hormone productio n<br />

and the fate of mobile nutrients in th e<br />

ecosystem .<br />

Acknowledgments<br />

The work reported in this paper wa s<br />

supported by National Science Foundatio n<br />

Grant No. GB-20963 to the Coniferous Fores t<br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contributio n<br />

No . 25 from the Coniferous <strong>Forest</strong> Biome .<br />

76


Graham, S. A. 1952. <strong>Forest</strong> entomology . 35 1<br />

p., illus. New York : McGraw-Hill Book Co . ,<br />

Inc .<br />

Isaac, L. A. 1943. Reproductive habits of<br />

Douglas-fir . 107 p ., illus. Washington, D .C . :<br />

Charles Lathrop Pack For . Found .<br />

Keen, F . P. 1952. Insect enemies of western<br />

forests. U .S. Dep. Agric. Misc. Publ . 273 :<br />

75-126 .<br />

Kozlowski, T . T. 1969. Tree physiology ' an d<br />

forest pests . J. For. 67 : 118-122 .<br />

Lack, David . 1954. The natural regulation of<br />

animal numbers. 343 p., illus. Oxford ,<br />

England: Clarendon Press .<br />

Maser, C . O. 1966 . Life histories and ecology<br />

of Phenacomys albipes, Phenacomys longicaudus,<br />

Phenacomys siluicola . M .S. thesis<br />

on file, Oreg . State Univ., Corvallis .<br />

Overton, W . S. 1972 . Toward a general model<br />

structure for a forest ecosystem . In Jerry F .<br />

Literature Cited<br />

Franklin, L. J. Dempster, and Richard H .<br />

Waring (eds .), Proceedings-research on<br />

coniferous forest ecosystems-a symposium,<br />

p . 37-47, illus . Pac . Northwest <strong>Forest</strong><br />

& Range Exp . Stn., Portland, Oreg .<br />

Rafes, P. M. 1970. Estimation of the effect s<br />

of phytophagous insects on forest production<br />

. In D. E. Reichle (ed .), Analysis of<br />

temperate forest ecosystems, p . 100-106 .<br />

Berlin : Springer-Verlag .<br />

Sweet, G . B., and P. F. Wareing. 1966. Role<br />

of plant growth in regulating photosynthesis.<br />

Nature 210 : 77-79 .<br />

Way, M . J., and M. Cammell . 1970. Aggregation<br />

behavior in relation to food utilizatio n<br />

by aphids . In A . Watson (ed.), Animal<br />

populations in relation to their food re -<br />

sources, p. 229-247. Oxford, England :<br />

Blackwell Sci . Pub .<br />

77


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

An environmental grid fo r<br />

classifying coniferous<br />

forest ecosystems<br />

R . H . Waring,<br />

K . L . Reed,<br />

W. H . Emmingham<br />

School of <strong>Forest</strong>ry<br />

Oregon State Universit y<br />

Corvallis, Orego n<br />

Abstract<br />

To develop models which will predict primary production and forest composition across the Biome, we<br />

suggest that the environment be defined operationally-as it affects certain plant responses . Models can be<br />

developed which indicate how water, nutrients, light, temperature, and mechanical factors interact through th e<br />

plant to control primary production and plant composition. Plant responses including water stress, stomatal<br />

behavior, foliar nutrition, and phenology, among others, were coupled to environmental variables to create plan t<br />

response indices for soil moisture, temperature, transpiration, and soil fertility . By correlating indicator species<br />

with plant response indices, an ecosystem can be defined environmentally without direct measurements . Other<br />

processes including biogeochemical cycling, consumer population dynamics, and hydrologic functions may als o<br />

be related to the environmental grid.<br />

Introduction<br />

Diversity in environment and vegetation i s<br />

characteristic of the Coniferous <strong>Forest</strong> Biome .<br />

Such diversity is esthetically pleasing but presents<br />

formidable problems of classification .<br />

Classification is necessary, however, if we ar e<br />

to understand the processes which affect the<br />

composition and allow for continued productivity<br />

of the forests and related water, wild -<br />

life, and recreational resources .<br />

If we could understand how the environment<br />

influences forest growth and composition,<br />

we would be well on the way toward<br />

reassessing our past management successe s<br />

and failures . We could also make our detailed<br />

ecosystem studies more widely applicable in<br />

both time and space, thus providing a frame -<br />

work for intelligent land management. Th e<br />

problem lies (a) in identifying which proper -<br />

ties of the environment should be measured ,<br />

(b) coupling these to important biological<br />

processes, and (c) finally in predicting certai n<br />

key properties of ecosystems .<br />

In this paper we attempt to establish a conceptual<br />

basis for evaluating environment, then<br />

test the general approach, and finally suggest<br />

necessary modifications to permit extensio n<br />

across the Coniferous <strong>Forest</strong> Biome .<br />

General Theory<br />

The Operational Environment<br />

Excellent descriptions of vegetation and<br />

regional physiography exist for many region s<br />

of the world . In the Pacific Northwest the<br />

most thorough treatments have been provide d<br />

by Daubenmire (1952), Krajina (1965), and<br />

Franklin and Dyrness (1969) . Detailed<br />

physiographic classification such as that developed<br />

by Hills (1959) has proven most use-<br />

79


ful in regions with relatively uniform climate .<br />

Probably the most helpful classifications<br />

from a standpoint of predicting the nature o f<br />

forest ecosystems, however, have been thos e<br />

related to environmental gradients (Warming<br />

1909, Sukachev 1928, Pogrebnjak 1929 ,<br />

Bakuzis 1961, Ellenberg 1950, 1956, Row e<br />

1956, Whittaker 1956, 1960, Loucks 1962 ,<br />

Waring and Major 1964) . Unfortunately, al<br />

previously defined environmental gradient s<br />

cannot be directly applied outside the particular<br />

region where they were developed . We feel<br />

environmental gradients can be more widel y<br />

utilized only when the environment i s<br />

coupled to basic processes controlling plan t<br />

growth and composition . We believe a key t o<br />

expanding the gradient analysis approach is t o<br />

focus more closely upon the basic physiological<br />

behavior of plants .<br />

Fortunately a foundation for a process -<br />

oriented approach has already been laid .<br />

Mason and Langenheim (1957) defined th e<br />

idea of an "operational environment" as on e<br />

which directly affects an organism. Implicit i n<br />

their concept of environment is the recognition<br />

of specific plant responses to environmental<br />

stimuli. In fact, these authors felt tha t<br />

"environment" could not exist independently<br />

in an ecological sense . It must have an effec t<br />

upon an organism . Although they did not de -<br />

fine measurable environmental stimuli, thei r<br />

very definition of an "operational environment"<br />

focuses attention upon the influenc e<br />

rather than on the origin of the stimulus .<br />

Thus, in an interpretive ecological sense, i t<br />

is more important to assess the availability o f<br />

water to plant roots than the origin of tha t<br />

water. This distinction greatly reduces the<br />

number of factors requiring consideration, for<br />

although altitude, slope, and other physiographic<br />

features are correlated with vegetation,<br />

such indirectly operating factors may be<br />

ignored if the mode of action can be identified<br />

and measured . Ellenberg (1956), Bakuzis<br />

(1961), Waring and Major (1964), and other s<br />

have suggested that vegetation responds t o<br />

changes in water, temperature, light, and<br />

chemical and mechanical factors. Although<br />

these factors do not operate independently ,<br />

they cannot completely substitute for one<br />

another .<br />

General Approac h<br />

We still have to translate the concept of an<br />

"operational environment" into a design adequate<br />

for research . As a first step we can diagram<br />

the flow of material through a plant into<br />

various compartments and identify the controls<br />

on the rate of flow imposed by the<br />

environment, the organism, or the amount o f<br />

material in a given compartment. Figure 1<br />

Figure 1 . General model of primary production .<br />

Material inputs of H 2 0, CO 2 . and energy flow int o<br />

the system to form carbohydrates . The rate of incorporation<br />

(photosynthesis) is a function of<br />

various plant responses. These responses are<br />

triggered by a host of environmental stimuli :<br />

temperature, light, humidity, soil water potential ,<br />

soil fertility, and mechanical stress. The plant<br />

responses interact to provide two functions: (1) a<br />

control on carbohydrate storage and (2) a contro l<br />

on growth. Losses from the system are throug h<br />

respiration (R), death of roots and other organ s<br />

and in litter (L), and consumption by animals (C) .<br />

presents such a diagramatic model . Th e<br />

rectangles are the compartments, and the flo w<br />

of materials is indicated by solid lines, whil e<br />

dotted lines indicate the transfer of information<br />

through valves controlling the rate of<br />

material flow from one compartment to an -<br />

80


other. The plant integrates environmenta l<br />

stimuli. Arrows indicate the direction of flow ,<br />

and circles are special functions regulated b y<br />

information from the plant integrator .<br />

We interpret environmental variables as<br />

being external to the system and impingin g<br />

upon it . The system, in this case the organism ,<br />

responds only in accordance to the externa l<br />

stimuli. Modifications in the environmen t<br />

brought about by the organism itself are no t<br />

distinguished from those originating throug h<br />

other means .<br />

The amount of material transfer from environment<br />

to plant is determined by the kin d<br />

and magnitude of the response ; i .e., the<br />

amount of CO2 converted into carbohydrate<br />

is determined by the photosynthetic rate . The<br />

amount of carbohydrate utilized in th e<br />

growth of foliage, stem, and roots is also<br />

affected by the stage of plant development<br />

and other plant responses . As long as the carbohydrate<br />

pool does not become exhausted ,<br />

or one of the plant responses is not death,<br />

biomass can accumulate .<br />

When the environment initiates a letha l<br />

plant response, the transfer of carbohydrate s<br />

is stopped and all of the biomass compartments<br />

empty . Biomass loss also occur s<br />

through respiration, consumption by animals,<br />

the normal fall of litter, and death of roots<br />

and twigs .<br />

Unfortunately, we do not, at present, have<br />

adequate knowledge for complete evaluatio n<br />

of such an energy flow model, even for a<br />

single species . We hope to acquire muc h<br />

needed data through the Biome program. Still<br />

it is safe to say that we will lack complete<br />

data on most species to predict accurately th e<br />

system 's response to external variables . Our<br />

approach, therefore, is to study physiologica l<br />

responses in relation to the environment ,<br />

develop functional models, and from thes e<br />

create plant response indices as a means of<br />

interpreting environmental gradients . Further ,<br />

we propose to correlate ecosystem propertie s<br />

with these plant response indices . Figure 2<br />

ENVIR<strong>ON</strong>MENTA L<br />

STIMUL I<br />

Absolute Humidit y<br />

Radiation<br />

Air & Soil Temperature<br />

Soil Wate r<br />

Soil Fertility<br />

Wind and Snow<br />

PLANT RESP<strong>ON</strong>SE<br />

INDICES<br />

ECOSYSTE M<br />

CHARACTERISTIC S<br />

Plant Moisture Stress<br />

Productivity<br />

Transpiratio n<br />

Compositio n<br />

PHYSIOLOGICAL<br />

CALIBRATI<strong>ON</strong><br />

j---* Photosynthesi s _<br />

Hydrologic Stat e<br />

PROCESS MODELS<br />

VALIDATI<strong>ON</strong><br />

Nutritional<br />

Decomposition<br />

4<br />

PREDICTI<strong>ON</strong><br />

Temperature Growth Consumer Population s<br />

Mechanical Stress<br />

/<br />

/ PHYSIOLOGICA L<br />

/ BEHAVIO R<br />

Stomatal Respons e<br />

Cambial Activity<br />

Photosynthesi s<br />

Transpiration<br />

Plant Moisture Stres s<br />

Phenology<br />

Foliar Nutritio n<br />

Figure 2 . Physiological process models couple environmental stimuli to the physiological behavior of selecte d<br />

conifers . From an understanding of the process models, plant response indices are derived which reflec t<br />

major environmental gradients . These indices locate ecosystems within an environmental grid and throug h<br />

correlation permit certain ecosystem characteristics to be predicted .<br />

8 1


epresents the basic linkages between environmental<br />

stimuli, physiological behavior, an d<br />

the generation of plant response indices whic h<br />

are to be correlated with ecosyste m<br />

characteristics .<br />

The Organis m<br />

If we strive for both a predictive model and<br />

an understanding of the general processes con -<br />

trolling plant composition, we must compromise<br />

by linking observations of plant distribution<br />

and growth for many species to th e<br />

physiological behavior of certain trees . For i n<br />

this Biome, trees represent the dominant for m<br />

in which energy accumulates in forest ecosystems<br />

.<br />

The size of reference plants must also b e<br />

standardized . Although establishment is critical<br />

for determining the composition of an y<br />

plant community, young seedlings represen t<br />

both adapted and ill-adapted plants . Mature<br />

plants, on the other hand, are usually so well<br />

established that their responses only hint at<br />

the critical selective factors operating at th e<br />

time of establishment and early growth . As a<br />

compromise, in this Biome we study recently<br />

established conifers between 1-2 m tall . Few<br />

such trees will die during the course of measurements<br />

; yet they are small enough to<br />

reflect the major environmental forces actin g<br />

at the critical time of establishment .<br />

Having decided upon the important environmental<br />

factors and the size and kind of<br />

reference plant, we are still faced wit h<br />

coupling the environment to plant responses .<br />

The most commonly selected response has<br />

been growth in some form or another, bu t<br />

growth reflects the total integration of al l<br />

environmental influences and a host of different<br />

environmental combinations can resul t<br />

in similar productivity . We choose response s<br />

related to the entire plant, but those whic h<br />

may be more closely identified with specifi c<br />

environmental stimuli .<br />

Physiological Behavior<br />

Because plants are a complex system, no<br />

response is completely unrelated to another .<br />

Therefore most responses cannot be inter -<br />

preted without knowledge of others, and<br />

cause and effect relationships can rarely b e<br />

implied. The most important behavior<br />

includes :<br />

1. Phenological development . This reflects<br />

past environment and serves to define<br />

periods during which the sensitivity of<br />

the plant is similar. Lack of attention to<br />

phenological development is a majo r<br />

reason why otherwise good environmental<br />

data are often difficult to interpret<br />

ecologically (Azzi 1955) .<br />

2. Carbon dioxide exchange provides a<br />

ready means of quantifying the effect s<br />

of light and temperature upon photo -<br />

synthesis .<br />

3. Plant moisture stress (plant water potential)<br />

before dawn is an excellent measur e<br />

of how a reference plant responds to soil<br />

water status.<br />

4. Stomatal resistance reflects the influenc e<br />

of evaporative stress and soil drought, a s<br />

well as certain other environmental variables<br />

.<br />

5. Foliar nutrition indicates the interactio n<br />

between supply of nutrients in the soil<br />

and the demand for nutrients .<br />

Results of Past Studie s<br />

To help clarify the above discussion, we<br />

draw upon our experience during the last 8<br />

years in an environmentally and floristicall y<br />

diverse region of southwestern Oregon (Whit -<br />

taker 1960, 1961 ; Waring 1969) .<br />

At all or some of the 25 forest stand s<br />

described briefly in table 1, environmental<br />

data on air and soil temperature, radiation,<br />

humidity, soil moisture, and soil fertility were<br />

recorded . Physiological response studies o n<br />

two widely distributed conifers, Douglas-fi r<br />

(Pseudotsuga menziesii) and Shasta red fir<br />

(Abies magnifica var . shastensis), included observations<br />

of phenology, plant water potential,<br />

stomatal resistance, and foliar nutrition .<br />

Productivity was estimated by measuring terminal<br />

elongation or total height of dominant<br />

old-growth trees. Details of this work hav e<br />

been published elsewhere (Waring and Clear y<br />

1967, Waring 1969, Atzet and Waring 1970 ,<br />

82


Table 1.-Description of stands<br />

Stand Elevation 1 Slope Aspect Parent material Dominant vegetation<br />

Meters<br />

Percent<br />

3 780 45 N Granite Douglas-fir, black oak,<br />

ponderosa pine<br />

21 550 75 N Metavolcanic Douglas-fir, black oak,<br />

Oregon white oa k<br />

8 1,280 40 SW Granite Ponderosa pine, Douglas-fi r<br />

12 1,585 70 WSW Green schist Ponderosa pine, Douglas-fir<br />

1 1,490 25 W Granite White fir, ponderosa pine ,<br />

Douglas-fir<br />

11 1,370 35 SW Granite Ponderosa pine, sugar pine ,<br />

white fir, Douglas-fi r<br />

13 1,340 20 W Green schist Douglas-fir, ponderosa pine ,<br />

sugar pine, white fir<br />

19 1,250 70 W Mica schist Douglas-fir, sugar pine,<br />

white fir<br />

2 1,675 60 WNW Granite White fir, Douglas-fir<br />

9 1,550 55 NNW Metavolcanic White fir, sugar pine ,<br />

Shasta red fir<br />

14 1,585 45 E Green schist Douglas-fir, white fir<br />

22 1,460 50 N Metasedimentary Douglas-fir, white fir<br />

7 1,920 20 N Granite Shasta red fi r<br />

16 1,890 40 SW Mica schist Shasta red fi r<br />

17 1,830 10 E Metavolcanic Shasta red fi r<br />

6 2,040 35 NNE Granite Mountain hemlock, Shasta red fi r<br />

15 2,010 10 NE Mica schist Mountain hemloc k<br />

18 2,135 30 NE Granite Mountain hemlock<br />

24 2,040 45 NNE Metavolcanic Mountain hemlock<br />

4 1,920 65 SE Ultrabasic Jeffrey pine, incense-cedar ,<br />

western white pin e<br />

5 1,710 65 SE Ultrabasic Jeffrey pine, incense-ceda r<br />

25 1,740 5 SE Ultrabasic Jeffrey pine, white fir,<br />

incense-cedar, Douglas-fi r<br />

20 760 70 NNW Mica schist Douglas-fir, Pacific ye w<br />

23 1,400 10 N Granite Engelmann spruce, Douglas-fir ,<br />

white fi r<br />

10 1,740 55 N Metavolcanic Brewer spruce, Shasta red fir,<br />

mountain hemlock<br />

8 3


Cleary and Waring 1969, Reed 1971, Emmingham<br />

1971, Waring and Youngberg 1972) .<br />

In this article we will restrict discussion to<br />

interpretation of results, concentrating upo n<br />

the operational effects of water, temperature ,<br />

and soil fertility . Mechanical stress from sno w<br />

creep or ice storms was important as was win d<br />

(Tranquillini 1970), but such adverse mechanical<br />

forces are closely associated wit h<br />

temperature gradient. Light, too, was important<br />

in understanding succession an d<br />

accounted for more than 75 percent of th e<br />

variation in terminal elongation of Abies<br />

when other environmental factors were<br />

restricted to narrow ranges (Emmingham<br />

1971) . In our stands, which were mature<br />

forests, maximum height of mature trees wa s<br />

used as an index to productivity . In such a<br />

situation, earlier reductions in growth due to<br />

shading or mechanical damage could be ignored<br />

.<br />

Measurement and Interpretation of<br />

Plant Water Stress<br />

During the growing season, plant moistur e<br />

stress measured by the Scholander pressur e<br />

chamber (Scholander et al . 1965, Waring an d<br />

Cleary 1967, Boyer 1967) provided a goo d<br />

estimate of plant water potential (Waggone r<br />

and Turner 1971) . We took measurements a t<br />

monthly intervals, with more frequenc y<br />

during September when the vegetation wa s<br />

under the greatest stress .<br />

Plant moisure stress usually increases fro m<br />

a predawn minimum to some maximum leve l<br />

in the afternoon . Predawn stress represent s<br />

the nearest equilibrium between soil and plant<br />

water potential . Not only the diurnal patter n<br />

but also the seasonal changes in plant wate r<br />

stress are important . The first reflects an imbalance<br />

between transpiration and water<br />

uptake ; the latter can be related to the avail -<br />

ability of soil water over a season . In figure 3 ,<br />

three contrasting forest types are shown t o<br />

have different seasonal water stress curves .<br />

The type dominated by Engelmann spruce<br />

(Picea engelmannii) was restricted to mois t<br />

sites, while oak (Quercus kelloggii) grew<br />

where stress was sufficient to bring about<br />

cessation of cambial activity in the reference<br />

Figure 3. Seasonal changes in minimum plant moisture<br />

stress associated with different forest ecosystems<br />

(Waring 19701 . All data are from 1- t o<br />

2-m-tall Douglas-fir. For this particular set of data ,<br />

the end of season Plant Moisture Stress Inde x<br />

would be 30 for the Oak Type, 18 for the Pin e<br />

Type, and 7 for the Spruce Type .<br />

plants. Where no rainfall occurs during the<br />

summer, the minimum night moisture stres s<br />

at the end of the growing season is an index<br />

to the plant moisture conditions throughout<br />

the entire season . From figure 3, we see the<br />

oak type had a plant moisture stress index o f<br />

30 ; the pine, an index of around 18 ; and the<br />

spruce, 7 atmospheres . Where summer precipitation<br />

is important, a mathematical description<br />

of the seasonal trend is desirable (Ree d<br />

1971) . Winter drought due to cold or frozen<br />

soils should also be similarly evaluated .<br />

Measurement and Interpretatio n<br />

of Temperatur e<br />

At each of the 25 forest stands, air temperature<br />

and soil temperature were recorded<br />

on 30-day thermographs because both root<br />

and shoot temperatures are important . At the<br />

time, we did not have photosyntheti c<br />

response information to help interpret th e<br />

effect of various combinations of temperatur e<br />

and light. Because winter temperatures are<br />

often below freezing, winter photosyntheti c<br />

activity is probably less in the study regio n<br />

than in the milder climate along the Pacifi c<br />

84


coast. Without the benefit of good photo -<br />

synthetic data, we used laboratory studie s<br />

conducted by our colleague, D . P. Lavender .<br />

Lavender found that the dry weight increas e<br />

of Douglas-fir seedlings was closely related to<br />

both air and soil temperatures . With these<br />

data a temperature-plant index was develope d<br />

by summing the potential growth possible for<br />

each day during the growing season (Cleary<br />

and Waring 1969) . Other variables were<br />

assumed nonlimiting. In the field, the Temperature<br />

Growth Index ranged from 30 near<br />

timberline to nearly 100 on oak and pin e<br />

forests. Douglas-fir was not found where th e<br />

index was below 40 .<br />

To visualize this information more effectively,<br />

the distributions of selected conifer s<br />

are presented in relation to these two rathe r<br />

simple plant response indices (fig. 4). The distributional<br />

patterns closely reflect the adapta -<br />

I . I . I I<br />

10 20 30<br />

PLANT MOISTURE STRESS, ATM .<br />

Figure 4 . Distribution of natural regeneration in relation<br />

to gradients of moisture and temperature de -<br />

fined by plant response indices (Waring 1970) .<br />

Validation stands, symbolized by 0, had vegetation<br />

predicted by the intercept of their plant<br />

response indices .<br />

tion of the various conifers . In a local region ,<br />

the variation in distribution of different<br />

species provides a means of predicting the environment<br />

through association with measurements<br />

taken on reference plants . The plan t<br />

response indices were correctly predicte d<br />

from knowledge of plant distributions fo r<br />

three validation stands, indicated as divide d<br />

circles in figure 4 . It is significant that suc h<br />

predictions are possible without physiological<br />

observations on species other than the reference<br />

plants and without special attention t o<br />

events controlling establishment . Once such<br />

relationships are established, the vegetatio n<br />

can provide understanding of the operational<br />

environment without requiring additional<br />

measurements of any kind . We shall expand<br />

this idea later .<br />

Measurement and Interpretation of<br />

Stomatal Respons e<br />

Conifer stomata are most difficult to observe,<br />

usually being sunken and occluded b y<br />

wax . The resistance which they offer to th e<br />

transfer of water vapor from the interior of<br />

the needles can be assessed by determinin g<br />

the rate of water vapor movement with a<br />

diffusion porometer (Waggoner and Turne r<br />

1971). The aperture of stomata may be estimated<br />

by observing the pressure necessary t o<br />

force a 50-percent ethanol solution through<br />

the pores (Fry and Walker 1967). Th e<br />

diffusion resistance is then determined by<br />

calibrating these pressures with reduction i n<br />

transpiration under known vapor pressure<br />

gradients (Reed 1971) . The latter procedure<br />

was followed in our past fieldwork . We found<br />

that stomatal resistance increased as the soi l<br />

moisture became less available during th e<br />

growing season . Further increase in stomatal<br />

resistance was possible during the day i f<br />

evaporative stress was high . These relationships<br />

were quantified and developed into a<br />

simulation model by Reed (1971) . Wit h<br />

knowledge of temperature, humidity, and<br />

nocturnal plant water stress, the model predicted<br />

on a daily basis, both potential transpiration<br />

(PT) and transpiration (T) wit h<br />

stomatal control . These values were summed<br />

for the entire season and confirmed the obser -<br />

85


vation that species such as Brewer spruc e<br />

(Picea breweriana), Port-Orford-cedar<br />

(Chamaecyparis lawsoniana), vine maple<br />

(Ater circinatum), and rhododendron<br />

(Rhododendron macrophyllum) are restricte d<br />

to areas with low potential transpiration . Th e<br />

most valuable index, however, was the rati o<br />

of actual to potential transpiration (T/PT)<br />

Which integrates for water, the demand, th e<br />

supply, and the control b the plant . Where<br />

sot moisture was never inadequate and th e<br />

evaporative stress remained low, the ratio wa s<br />

1 .0. Where water became limiting and hig h<br />

evaporative stress was common, ratios of 0 . 3<br />

were calculated . Under these conditions, both<br />

forest composition and growth were dramatically<br />

affected .<br />

Measurement and Interpretation of Soi l<br />

Fertility and Plant Nutritio n<br />

Soil profiles from each stand were reconstructed<br />

to a depth of 60 cm and taken into<br />

controlled environment chambers where seedlings<br />

of Douglas-fir and Shasta red fir were<br />

grown for a period of 5 months (Waring an d<br />

Youngberg 1972) . The dry weight yields at<br />

the end of the experiment were used as a bio -<br />

assay of soil fertility, representing the supply<br />

of nutrients available to conifers . Foliar analyses<br />

were made on reference trees, first durin g<br />

the time of maximum demand when ne w<br />

foliage was being produced and again after al l<br />

shoot and diameter growth had ceased in th e<br />

fall. In the first period, 1-year-old foliage wa s<br />

assessed because it represents a major sourc e<br />

of mobile nitrogen, phosphorus, and potassium.<br />

Three categories of nutrient availability<br />

can be identified : (1) where no nutritional<br />

stress occurs during the year ; (2) where nutrition<br />

is adequate only when shoot and<br />

diameter growth has ceased ; and (3) where<br />

nutrition is inadequate year around (Warin g<br />

and Youngberg 1972) . These three categorie s<br />

may be further refined and are of consider -<br />

able value in reaching decisions concernin g<br />

fertilization. The composition of vegetatio n<br />

is, however, insensitive to this classification of<br />

nutrient availability .<br />

Only on soils where an imbalance of nutrients<br />

or toxic amounts of certain heavy metals<br />

were present did special plant communitie s<br />

develop. In this paper, therefore, we hav e<br />

assigned plants to one of three categories of<br />

tolerance to infertile soils developed from<br />

ultrabasic parent materials : In class 1 ar e<br />

those which are tolerant ; class 2 is made up of<br />

those species that are intolerant ; and in class 3<br />

are those plants both tolerant and competitively<br />

restricted to ultrabasic soils .<br />

Vegetation as an Index to Environmen t<br />

Can we use plants to define their environment,<br />

or more precisely, the environment expressed<br />

through reference plants? To test thi s<br />

idea, we selected 47 species from more than<br />

600 in the local flora and recorded thei r<br />

distributional limits in relation to various<br />

environmental plant indices (table 2) . For<br />

some plants we knew only the approximat e<br />

ranges and occasionally we had insufficien t<br />

data to set particular index limits .<br />

In forest ecosystems where plant composition<br />

is known, the range of environmenta l<br />

indices may be predicted from information i n<br />

table 2 . In table 3, the plant response indice s<br />

were thus predicted for the 25 stand s<br />

described in table 1 . Therefore for the blac k<br />

oak forest, stand 3, where Rhus diversiloba ,<br />

Arbutus menziesii, Lonicera hispidula, and<br />

Quercus chrysolepis grew, one can assess fro m<br />

table 2 that the Temperature Growth Inde x<br />

lies between 98 and 96, the Plant Moistur e<br />

Stress Index at 25 .4, and the ratio of transpiration<br />

to potential transpiration as 0 .29 .<br />

Plants that were exclusively adapted t o<br />

ultrabasic soils, such as Jeffrey pine, isolated<br />

those ecosystems (stands 4, 5, and 25) . Those<br />

forests with oak (stands 3, 8, and 21 )<br />

exhibited the greatest water stress, th e<br />

warmest environments, and the greatest control<br />

of transpiration . At the other extreme ,<br />

mountain hemlock forests (stands 6, 18, 15 ,<br />

and 24) had the coolest environments and<br />

lowest potential transpiration .<br />

In figure 5, the midpoint of the predicted<br />

Temperature Growth Index from table 3 is<br />

plotted against the Temperature Growt h<br />

Index calculated from temperature records .<br />

The regression has an r2 of 0 .93, and is highl y<br />

86


Generic name<br />

Table 2.-Ecological distribution of selected species in relation t o<br />

plant response indices'<br />

PT , PT , T/PT , T/PT , PMS , PMS , TGI, TGI, Soil<br />

maxi- mini - maxi- mini- maxi- mini- maxi- mini- tolermum<br />

mum mum mum mum mum mum mum anc e<br />

ANACARDIACEAE<br />

Rhus diversiloba 30.0 17 .7 0 .42 0 .29 25 .4 20 .3 100 80 1<br />

BERBERIDACEA E<br />

Achlys triphylla 16.8 1 .00 .57 16 .2 80 60 2<br />

CAPRIFOLIACEAE<br />

Lonicera hispidula<br />

COMPOSITA E<br />

30.0 17 .7 .29 30.0 25 .4 100 95 1<br />

Arnica latifolia 13.5 7 .5 1 .00 .46 19 .1 5 .2 75 35 2<br />

CUPRESSACEA E<br />

Libocedrus decurrens<br />

ERICACEAE<br />

19.5 12.2 1 .00 .40 30 .0 5 .2 85 52 1<br />

Arbutus menziesii 30.0 - 1 .00 .29 30 .0 5 .2 98 68 1<br />

Arctostaphylos uiscida 30.0 17 .7 .42 .29 25 .4 20 .3 100 90 1<br />

Rhododendron macrophyllum 12.5 - 1 .00 .57 16 .2 - 100 2<br />

Vaccinium membranaceum 12.5 .63 .46 19 .1 12 .8 60 45 2<br />

Vaccinium scoparium 10.3 1 .00 .46 19 .1 - 50 2<br />

FAGACEA E<br />

Castanopsis chrysophylla 19.5 - 1.00 .51 15 .3 - 85 40 1<br />

Quercus chrysolepis 30.0 12 .0 1.00 .63 30 .0 8 .0 100 78 1<br />

Quercus kelloggii 30.0 17 .6 .42 .29 30 .0 15 .0 98 80 1<br />

Quercus sdeeriana 16.8 - 1 .00 .46 19 .1 - 70 50 1<br />

Quercus uaccinifolia 30.0 .40 - 30.0 100 - 1<br />

GARRYACEA E<br />

Garrya fremontii 30.0 12 .7 .62 .40 30 .0 12 .6 70 50 1<br />

LABIATA E<br />

Monardella odoratissima 21 .4 .62 .40 20 .3 12 .6 100 35 1<br />

LEG UMIN OS A E<br />

Lathyrus polyphyllus 13.1 - 1 .00 .63 12 .8 5 .2 85 45 2<br />

Lupinus leucophyllus 30.0 12 .7 .86 - 30.0 8 .4 70 50 3<br />

LILIACEAE<br />

Clintonia uniflora 13 .1 7 .5 1.00 .51 16 .2 5 .2 85 35 2<br />

Disporum hooheri 16 .8 12 .2 1.00 .51 16 .2 5 .2 85 45 2<br />

Xerophylium tenax 19 .5 - 1 .00 .40 19 .1 8 .4 80 40 1<br />

PINACEA E<br />

Abies concolor 19.5 - 1 .00 .46 19 .1 5 .2 84 47 1<br />

Abies magnifica 12.5 7 .5 1 .00 .46 19 .1 5 .2 59 34 2<br />

var . shastensis<br />

Picea breweriana 10.3 1 .00 .46 19 .1 52 1<br />

Picea engelmannii 11 .0 - 1.00 .63 12 .8 - 47 - 2<br />

Pinus jeffreyi 25.0 12 .7 .61 - 25 .0 8 .4 100 52 3<br />

Pinus lambertiana 21 .4 11 .0 1 .00 .29 25 .4 5 .2 96 45 1<br />

Pin us ponderosa 21 .4 13 .5 1 .00 .29 25 .4 5 .2 98 68 2<br />

Pinus monticola 19.5 7 .8 1 .00 .46 19 .1 5 .2 70 44 1<br />

Pseudotsuga menziesii 21 .4 - 1 .00 .29 25 .4 5 .2 98 47 1<br />

Tsuga mertensiana 11 .0 7 .5 1 .00 .46 19 .1 5 .2 59 34 2<br />

POLEM<strong>ON</strong>IACEA E<br />

Phlox adsurgens 16.8 10 .3 1.00 .46 19 .1 5 .2 85 52 2<br />

Polemonium californicum 7 .8 - 1 .00 .98 5 .2 - 50 30 2<br />

POLYG<strong>ON</strong>ACEAE<br />

Polygonum dauisiae 7 .8 1 .00 .98 5 .2 40 30 2<br />

POLYPODIACEAE<br />

Onychium densum<br />

RANUNCULACEAE<br />

30.0 12 .7 .62 18 .1 5 .2 80 50 3<br />

Anemone deltoidea 19 .5 7 .8 1 .00 .51 19 .1 5 .2 85 45 1<br />

Anemone lyallii 19 .5 7 .8 1 .00 .51 15 .3 5 .2 75 45 2<br />

ROSACEAE<br />

Amelanchier pallida 16.8 .62 .46 19 .1 75 50 1<br />

Rubus lasiococcus 12.5 1 .00 .46 19 .1 - 60 - 2<br />

Rubus paruiflorus<br />

SAXIFRAGACEA E<br />

16 .8 1 .00 .51 16 .2 5 .2 85 50 2<br />

Ribes viscosissimum 13 .1 7 .8 1 .00 .51 16 .2 5 .2 85 40 2<br />

Tiarella unifoliata 13 .1 - 1 .00 .63 12 .0 - 85 47 2<br />

SCROPHULARIACEAE<br />

Pedicularis racemosa<br />

VALERIANACEAE<br />

12.5 7 .5 1 .00 .51 16 .2 5 .2 65 35 2<br />

Valeriana sitchensis<br />

VIOLACEA E<br />

7 .8 1 .00 .98 5 .2 45 30 2<br />

Viola glabella<br />

Viola sempereirens<br />

13 .1<br />

13 .1<br />

7 .5<br />

-<br />

1 .00<br />

1 .00<br />

.63<br />

.57<br />

12 .8<br />

16 .2<br />

5 .2<br />

5 .2<br />

95<br />

85<br />

35<br />

45<br />

2<br />

2<br />

'<br />

P otential transpiration (PT) is calculated for April through September based on a minimum stomatal resistance of 4 se c<br />

cm Transpiration is expressed in g cm - 2<br />

The ratio of simulated transpiration (T) to potential (PT) reflects the degree of stomatal control exhibited by a referenc e<br />

conifer.<br />

Plant Moisture Stress (PMS) is expressed in atm and represents predawn measurements on reference conifers near the end<br />

of the summer dry season (Sept .) .<br />

A Temperature Growth Index (TGI) reflects the potential for Douglas-fir seedling growth as a function of air and soil<br />

t emp erature for the entire growing season.<br />

The soil tolerance index indicates whether the species is exclusively restricted (class 3), tolerant (class 1), or excluded (class<br />

2) from infertile ultrabasic soils .<br />

Values of 30 in PT and PMS columns indicate species are known to occupy environments more extreme than possible fo r<br />

coniferous forest .<br />

8 7


Table 3.-Predicted plant response indices for 25 forest ecosystems based on th e<br />

inclusive limits of species present with known ecological distributions '<br />

<strong>Forest</strong> type 2<br />

Stand<br />

PT , PT, T/PT , T/PT , PMS , PMS, TGI , TGI , Soi l<br />

maxi - mini- maxi - mini- maxi - mini- maxi- mini - tolermum<br />

mum mum mum mum mum mum mum anc e<br />

Black oak 3 21 .4 17 .7 0 .29 0 .29 25 .4 25 .4 98 95 1<br />

Black oak 21 3 21 .4 17 .7 .29 .29 25 .4 25 .4 96 95 1<br />

Ponderosa pine 8 21 .4 17 .7 .42 .40 20 .3 20 .3 96 90 1<br />

Ponderosa pine 12 3 16 .8 13 .5 .62 .46 19 .1 12 .6 70 68 1<br />

Mixed conifer 1 13 .5 13 .5 .62 .51 15 .3 5 .2 75 68 2<br />

Mixed conifer 11 16 .8 13 .5 1 .00 .57 15 .3 5 .2 70 68 1<br />

Mixed conifer 13 13 .1 12 .2 1 .00 .63 12 .8 8 .4 80 68 1<br />

Mixed conifer 19 13 .1 12 .2 1 .00 .63 12 .8 5 .2 84 68 2<br />

White fir 2 12 .5 12 .2 .62 .51 15 .3 12 .6 65 50 1<br />

White fir 9 12 .5 12 .2 .63 .57 16 .2 12 .8 59 60 2<br />

White fir 14 3 13 .1 11 .0 1 .00 .63 12 .8 8 .4 70 52 2<br />

White fir 22 3 13 .1 10 .3 .62 .63 12 .8 5 .2 70 52 2<br />

Shasta red fir 7 3 7 .8 7 .8 1 .00 .98 5 .2 5 .2 45 45 2<br />

Shasta red fir 16 12.5 7 .5 1 .00 .51 15 .3 5 .2 59 40 1<br />

Shasta red fir 17 7 .8 7 .8 1 .00 .98 5 .2 5 .2 45 45 2<br />

Mountain hemlock 6 7 .8 7 .5 1 .00 .98 5 .2 5 .2 45 35 2<br />

Mountain hemlock 15 3 11 .0 7 .5 1 .00 .46 19 .1 5 .2 59 34 2<br />

Mountain hemlock 18 3 7 .8 7 .5 .62 .46 19 .1 12 .6 40 35 2<br />

Mountain hemlock 24 3 7 .8 7 .8 1 .00 .98 5 .2 5 .2 45 44 2<br />

Jeffrey pine 4 16 .8 12 .7 .61 .51 15 .3 12 .6 70 52 3<br />

Jeffrey pine 5 19 .5 12 .7 .40 .40 18 .1 12 .6 70 52 3<br />

Jeffrey pine 25 3 19.5 12 .7 .61 .51 15 .3 8 .4 70 52 3<br />

Yew 20 3 17 .7 16 .8 .51 .42 16 .2 15 .0 85 80 2<br />

Engelmann spruce 23 11 .0 11 .0 .63 .63 12 .8 12 .8 47 47 2<br />

Brewer spruce 10 10.3 7 .8 .62 .46 19 .1 12 .8 50 50 2<br />

' Potential transpiration (PT) is calculated for April through September based on a minimum stomatal resistance of 4 sec<br />

cm^1 . Transpiration expressed in cm- 2<br />

The ratio or simulated transpiration (T) to potential (PT) reflects the degree of stomata] control exhibited by a referenc e<br />

conifer.<br />

Plant Moisture Stress (PMS) is expressed in atm and represents predawn measurements on reference conifers near the en d<br />

of the summer dry season (Sept .) .<br />

A Temperature Growth Index (TGI) reflects the potential for Douglas-fir seedling growth as a function of air and soil<br />

temperature for the entire growing season .<br />

The soil tolerance index indicates whether the species is exclusively restricted (class 3), tolerant (class 1), or excluded (clas s<br />

2) from infertile ultrabasic soils .<br />

2 From R . H . Waring, <strong>Forest</strong> plants of the eastern Siskiyous : their environmental and vegetational distribution. Northwes t<br />

Sci . 43 : 1-17, 1969 .<br />

'Simulation of PT and T/PT was not possible because of inadequate data . The predicted values appear reasonable whe n<br />

compared with other stands with similar vegetation, temperature, and plant moisture stress indices .<br />

88


x<br />

-0 o 90<br />

LL<br />

L<br />

70<br />

L<br />

a)<br />

a<br />

E<br />

50<br />

a )<br />

t<br />

-o a )<br />

L<br />

a- 30<br />

30 50 70 90<br />

Observed Temperature Growth Inde x<br />

Figure 5 . Relationship between the temperature index<br />

derived from the growth response of Douglas -<br />

fir to temperature and the midpoint values predicted<br />

from the presence of indicator species in a<br />

given ecosystem (table 3) . Validation stands, 0 ,<br />

with one exception fell within the 95-percent<br />

confidence limits . r2 = 0.93 .<br />

30<br />

significant . Similarly, figure 6 presents the regression<br />

for plant moisture stress where the r 2<br />

was 0.77, also highly significant. The symbol<br />

8 represents validation stands which generall y<br />

fell within the 95-percent confidence intervals<br />

.<br />

From the comparisons made earlier, we<br />

showed that environmental plant indices ca n<br />

predict the forest composition, and conversely<br />

from the above relationships, we not e<br />

that environmental indices may be predicted<br />

from knowledge of the distribution of specifi c<br />

plants .<br />

Environmental Plant Indices as Predictor s<br />

of <strong>Forest</strong> Productivity<br />

In this study the only index to productivit y<br />

was maximum tree height . Particularly on th e<br />

infertile ultrabasic soils, this index overestimates<br />

productivity, for although individua l<br />

trees reach considerable height, the density o f<br />

trees remains low. Reed and Waring l reporte d<br />

that the ratio of actual to potential transpiration<br />

(T/PT) combined with the Temperatur e<br />

Growth Index (TGI) accounted for 96 per -<br />

cent of the variation observed in the maxi -<br />

mum height of Douglas-fir . For all dominant<br />

trees, these and related indices accounted fo r<br />

86 percent of the variation in height . The<br />

model for analytic solution is :<br />

Maximum height (in meters) =<br />

- 102.3 + 0.334 PT + 316.4 T/PT + 0.405 TGI<br />

- 186.8 (T/PT) 2 - 3.25 (T/PT)PT + 0.631 (T/PT)TGI<br />

This is encouraging, considering the limited<br />

sampling of tree heights and the expecte d<br />

differences among species . The model was not<br />

improved by including the index to soil fertility,<br />

again suggesting that total height may<br />

be an inadequate measure of productivity .<br />

Figure 6. Relationship between the moisture stress<br />

index during the period when soil water may b e<br />

limiting to reference conifers and the midpoin t<br />

values predicted from the presence of indicato r<br />

species in a given ecosystem (table 3) . Validation<br />

stands, 0, fell within the 95-percent confidence<br />

limits . r2 = 0.77 .<br />

30<br />

Data Requirement s<br />

If the environmental grid is to be applie d<br />

across the Coniferous <strong>Forest</strong> Biome, then a<br />

coordinated effort must be made to test an d<br />

improve the approach . Present efforts by<br />

1 Unpublished data .<br />

89


Walker, Reed, Scott, and Webb are channele d<br />

toward providing additional physiological<br />

information on Douglas-fir, one of four pro -<br />

posed reference plants . These investigators are<br />

concerned with photosynthesis, water uptake ,<br />

transpiration, and translocation . Lavender an d<br />

Hermann are responsible for providing functional<br />

relationships between foliage, stem, an d<br />

roots. Through stem and twig analysis and b y<br />

aging foliage, we hope to gain measures of net<br />

primary production .<br />

Other reference plants include Pinus<br />

ponderosa, Tsuga heterophylla, and Abies<br />

lasiocarpa ; additional studies on these specie s<br />

are necessary .<br />

To extend and validate the genera l<br />

approach described in this paper, we will re -<br />

quire certain kinds of information at selecte d<br />

forest environments across the Biome . <strong>Forest</strong><br />

stands will be selected which represent equilibrium<br />

stages ; i.e., approach the maximu m<br />

accumulation of biomass, on homogeneou s<br />

areas at least 30 m 2 .<br />

Within these stands, bud break will b e<br />

recorded on 1- to 2-m reference conifers b y<br />

observing five lateral branches on each of a t<br />

least five trees. Cambial activity will b e<br />

assessed by using the pin-into-cambium technique<br />

which identifies the termination of cel l<br />

divisions (Wolter 1968) . Stand compositio n<br />

and structure will be estimated from basa l<br />

area and height estimates of trees and under -<br />

story sampling of shrub and herbaceous cover .<br />

Nocturnal moisture stress will be measure d<br />

on three to five reference trees at 2-week intervals<br />

throughout the growing season and at<br />

less frequent intervals during the dorman t<br />

season, unless winter dessication is likely .<br />

Temperature under the stand near referenc e<br />

trees will be recorded at 1 m above and 20 c m<br />

below ground level. Latitude location o f<br />

stands will allow for calculation of day length<br />

for different seasons . Nutrition will b e<br />

assessed from a composite sample of 1-yearold<br />

foliage collected from the lateral branches<br />

of at least five trees at the time of new shoot<br />

elongation and during the late fall .<br />

Short wave radiation will be measured in a<br />

nearby open area or above the stand. Data<br />

will be expressed in cal cm -2 day -1 . Absolut e<br />

humidity will be measured daily at solar noon<br />

during the growing season and if possible continuously<br />

year around. Precipitation will include<br />

both monthly values and average maxi -<br />

mum snow depth . Estimates of wind velocity<br />

are desired where conditions appear to warrant<br />

measurements .<br />

These data will not only be valuable to hel p<br />

construct the environmental grid but wil l<br />

serve to interpret process studies in decomposition,<br />

consumer population dynamics, and<br />

other ecosystem properties .<br />

Acknowledgments<br />

All field investigations reported in thi s<br />

paper were generously funded under an extended<br />

McIntire-Stennis Federal Grant . Th e<br />

simulation of transpiration, regression equations<br />

for productivity, and a computer program<br />

for defining plant response indices o n<br />

the basis of overlapping plant distribution s<br />

were developed under support by Nationa l<br />

Science Foundation Grant No . GB-20963 t o<br />

the Coniferous <strong>Forest</strong> Biome, U .S. Analysis o f<br />

Ecosystems, International Biological Program .<br />

This is Contribution No . 26 from the Coniferous<br />

<strong>Forest</strong> Biome .<br />

Literature Cited<br />

Atzet, T., and R. H . Waring. 1970. Selectiv e<br />

filtering of light by coniferous forests an d<br />

minimum light energy requirements fo r<br />

regeneration . Can. J. Bot. 48 : 2163-2167 .<br />

Azzi, G . 1955. Agricultural ecology . 424 p.<br />

London : Constable and Co .<br />

Bakuzis, E. V. 1961 . Synecological coordinates<br />

and investigations of forest ecosystems<br />

. 13th Congr. Int. Union For. Res .<br />

Org. (Vienna) Proc . 2, Sec . 21½ .<br />

Boyer, J . S . 1967 . Leaf water potentials measured<br />

with a pressure chamber. Plant<br />

Physiol. 42: 133-137 .<br />

Cleary, B . D., and R. H. Waring. 1969. Temperature<br />

: collection of data and its analysis<br />

for the interpretation of plant growth an d<br />

distribution . Can . J. Bot . 47 : 167-173 .<br />

Daubenmire, R . 1952. <strong>Forest</strong> vegetation of<br />

northern Idaho and adjacent Washington ,<br />

90


and its bearing on concepts of vegetatio n<br />

classification . Ecol . Monogr . 22 : 301-330 .<br />

E l lenberg, H. 1950. Landwirtschaftliche<br />

Pflanzensoziologie . Bd . I . Unkrautgemeinschaften<br />

als Zeiger fur Klima and<br />

Boden . 141 p . Ulmer, Stuttgart .<br />

1956 . Aufgaben and Methoden<br />

der Vegetationskunde . In H . Walter (ed .) ,<br />

Einfuhrung in die Phytologie. Bd. IV .<br />

Grundlagen der Vegetationsgliederung . Tel<br />

1 . 136 p . Ulmer, Stuttgart .<br />

Emmingham, W . H . 1971 . Conifer growth and<br />

plant distribution under different light environments<br />

in the Siskiyou Mountains of<br />

southwestern Oregon . 50 p. M .S. thesis, on<br />

file at Oreg . State Univ ., Corvallis .<br />

Franklin, J. F., and C. T. Dymess. 1969 .<br />

Vegetation of Oregon and Washington .<br />

USDA <strong>Forest</strong> Serv . Res . Pap . PNW-80, 21 6<br />

p., illus. Pac. Northwest <strong>Forest</strong> & Range<br />

Exp . Stn., Portland, Oreg .<br />

Fry, K. E., and R . B . Walker. 1967 . A pressure<br />

infiltration method for estimatin g<br />

stomatal opening in conifers . Ecology 48 :<br />

155-157 .<br />

Hills, G. A. 1959. A ready reference to th e<br />

description of the land of Ontario and it s<br />

productivity. 142 p. Ont . Dep. Lands &<br />

<strong>Forest</strong>s .<br />

Krajina, V . J. 1965. Biogeoclimatic zones and<br />

classification of British Columbia . In V . J .<br />

Krajina (ed .), Ecology of western North<br />

America, vol . 1, p. 1-17 . Univ . Brit . Columbia,<br />

Vancouver .<br />

Loucks, O. L. 1962. Ordinating forest communities<br />

by means of environmental scalars<br />

and p h y t o sociological indices. Ecol .<br />

Monogr . 32 : 137-166 .<br />

Mason, H. L., and J . H. Langenheim . 1957 .<br />

Language analysis and the concept of environment.<br />

Ecology 38 : 325-339 .<br />

Progrebnjak, P . S. 1929 . Uber die Methodik<br />

der Standortsuntersuchungen in Verbindung<br />

mit den Waldtypen. Int. Congr .<br />

<strong>Forest</strong> Exp. Stn. (Stockholm) Proc., p .<br />

455-471 .<br />

Reed, K. L. 1971. A computer simulation<br />

model of seasonal transpiration in Douglasfir<br />

based on a model of stomatal resistance .<br />

132 p . Ph .D . thesis, on file at Oreg . State<br />

Univ., Corvallis .<br />

Rowe, J . S . 1956. Use of undergrowth plan t<br />

species in forestry . Ecology 37 : 461-472 .<br />

Scholander, P . F., H . T. Hammel, D . Broad -<br />

street, and E . A. Hemmingsen. 1965. Sap<br />

pressure in vascular plants . Science 148 :<br />

339-346 .<br />

Sukachev, V. N . 1928. Principles of classification<br />

of the spruce communities of European<br />

Russia. J. Ecol . 16 : 1-18 .<br />

Tranquillini, W ., 1970 . Einfluss des Winde s<br />

auf den Gaswechsel der Pflanzen Umsha u<br />

in Wissenschaft and Technik 26: 860-861 .<br />

Waggoner, P . E ., and N. C . Turner . 1971 .<br />

Transpiration and its control by stomata in<br />

a pine forest . Conn. Agric. Exp. Stn . Bull .<br />

726, 87 p. New Haven .<br />

Waring, R. H. 1969 . <strong>Forest</strong> plants of the east -<br />

ern Siskiyous : their environmental and<br />

vegetational distribution. Northwest Sci .<br />

43 : 1-17 .<br />

1970. Matching species to site .<br />

In R. K. Hermann (ed .), Regeneration o f<br />

ponderosa pine, p . 54-61 . Oreg. State<br />

Univ., Corvallis .<br />

and B. D. Cleary . 1967. Plan t<br />

moisture stress : evaluation by pressure<br />

bomb. Science 155 : 1248-1254 .<br />

and J. Major . 1964. Some vegetation<br />

of the California coastal redwoo d<br />

region in relation to gradients of moisture ,<br />

nutrients, light, and temperature . Ecol .<br />

Monogr . 34: 167-215 .<br />

and C. T. Youngberg . 1972 .<br />

Evaluating forest sites for potential growt h<br />

response of trees to fertilizer. Northwest<br />

Sci . 46 : 67-75 .<br />

Warming, E . 1909. Oecology of plants . 422 p.<br />

London : Oxford Univ . Press .<br />

Whittaker, R . H. 1956. Vegetation of th e<br />

Great Smoky Mountains . Ecol. Monogr .<br />

26 : 1-80 .<br />

1960. Vegetation of the Siskiyou<br />

Mountains, Oregon and California .<br />

Ecol . Monogr . 30 : 279-338 .<br />

1961 . Vegetation history of th e<br />

Pacific Coast States and the "central" significance<br />

of the Klamath region . Madron o<br />

16 : 5-23 .<br />

Wolter, K . E . 1968. A new method for marking<br />

xylem growth . <strong>Forest</strong> Sci . 14 : 102-104 .<br />

91


Water and Nutrien t<br />

Movement Through Ecosystems<br />

93


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Modeling water movement within<br />

the upper rooting zone of a<br />

Cedar River soil<br />

W. H . Hathewa y<br />

P. Machn o<br />

an d<br />

E . Hamerl y<br />

University of Washingto n<br />

Seattle, Washington 9819 5<br />

Abstract<br />

The Richards equation for unsaturated soil water flow is used to represent flow in a natural forest soil . Th e<br />

differential equation and the corresponding finite difference technique used to obtain an approximate solutio n<br />

are discussed. Independent estimates of soil moisture conductivity and of initial soil water content at severa l<br />

depths are used as inputs to the finite difference equation. Conductivity was estimated by laboratory techniques,<br />

and initial conditions in the field were measured by tensiometer. A tension lysimeter system provided<br />

estimates of soil water flow. Predicted values were compared with values measured in the field . Results sugges t<br />

that the model gives a satisfactory representation of actual soil water flow despite considerable variability in<br />

forest soil properties.<br />

Introduction<br />

In the Coniferous <strong>Forest</strong> Biome, we hav e<br />

been interested in models which represent th e<br />

movement of water through unsaturated soil s<br />

for use as possible components of a larger<br />

computer model which describes water relations<br />

within a local soil-plant-atmospher e<br />

system. Other components of the larger<br />

model include uptake of water by the roots of<br />

trees, conduction of water through the vascular<br />

system to the leaves, and transpiratio n<br />

from the leaves to the surrounding atmosphere<br />

. When we find that existing submodel s<br />

can be used-possibly with slight modifications-in<br />

our larger soil-plant-atmospher e<br />

model, we prefer to use them rather than t o<br />

invent them anew . Where existing models ar e<br />

not well suited to our objectives, we prefer t o<br />

develop new ones . For example, we are currently<br />

constructing a model which represent s<br />

the flow of water through the vascular syste m<br />

of a tree. We evaluate existing or new sub -<br />

models on the basis of (1) their behavior in<br />

isolation and (2) their compatibility with<br />

other components of the system . In the<br />

present paper we report our experiences with<br />

the Richards equation, a model which de -<br />

scribes the flow of water in an unsaturate d<br />

soil .<br />

The Model<br />

The Richards equation (Richards 1931) is<br />

essentially a modification of the well-known<br />

Darcy relationship for saturated flow i n<br />

porous media. Because the Richards model<br />

allows for variation in hydraulic conductivity<br />

with changes in soil water content, it is wel l<br />

suited for the description of water flow in<br />

unsaturated media . Darcy 's law with variabl e<br />

conductivity and negative hydraulic head can<br />

be written (Rose 1966 )<br />

v=K aia 6 ae aZ<br />

K (1)<br />

95


Here v is the volume of water crossing a uni t<br />

area per unit time (cm/min) ,<br />

K is the hydraulic conductivity (cm/min) ,<br />

is the soil water pressure or suctio n<br />

(hydraulic head) (cm) ,<br />

0 is the soil water content (volume of<br />

water per unit volume of soil), an d<br />

z is the height above datum (referenc e<br />

level) of the point under consideratio n<br />

(cm) .<br />

By introducing the concept of soil-wate r<br />

diffusivity, defined b y<br />

purposes of calculations . Soil moisture specified<br />

as a function of time is required at th e<br />

upper and lower boundaries of the soil column<br />

. Their solution permits the simultaneous<br />

calculation o f<br />

(Dj +<br />

Dj+) )<br />

2 (oJ+1 zj+1 - 01 zj<br />

This expression approximates D a8 and hence<br />

v, the volume flux of water, apart from th e<br />

constant K (cf. equation 2) .<br />

D = -K a alP d ,<br />

where D is measured in c m 2 /min, Childs and<br />

Collis-George (1950) cast the Darcy equatio n<br />

in a form similar to Fick 's law of diffusion :<br />

v = -Da 8 -K (2 )<br />

Applying the equation of continuity ,<br />

ad + av<br />

at az<br />

= 0 (which implies that there are n o<br />

sources or sinks of water in the system), one<br />

obtains the full Richards soil flow equation :<br />

30 __ a ad aK<br />

at az (D az ) + a z<br />

(3 )<br />

Soil moisture is obtained as a function o f<br />

depth (z) and time (t) by integrating (3) .<br />

Accordingly the Richards model is compatibl e<br />

with our larger soil-plant-atmosphere model ,<br />

which requires soil moisture at various depth s<br />

and times as an input to a submodel whic h<br />

represents the uptake of water by the roots o f<br />

a tree. Our current uptake model is that of<br />

Gardner (1960), a diffusion-type flow equation<br />

in which time rate of change of soil wate r<br />

concentration (0) at any point (z) in the soi l<br />

is related to the moisture gradient betwee n<br />

that point and an absorbing root .<br />

Since D and K depend on 0, (3) is non -<br />

linear in 0, and a solution is possible only i f<br />

numerical approximations are used . Remso n<br />

et al . (1965) developed a computer program<br />

which gives approximate results when initial<br />

moisture content is specified at each of n+ 1<br />

equally spaced points or nodes into which a<br />

soil column zn centimeters thick is divided for<br />

Available<br />

Observational Data<br />

At the Allan Thompson Research Center ,<br />

Cedar River Watershed, the downward flow o f<br />

water through soil profiles has been measure d<br />

for several years by a tension lysimeter syste m<br />

(Gessel and Cole 1965) . A prescribed suctio n<br />

on the porous lysimeter plate causes drainage<br />

through the plate to approximate the drainag e<br />

in the adjacent soil . Results are recorded i n<br />

permanent form on paper tape by an automatic<br />

data recording system . Our data com e<br />

from a series of controlled experiments, carried<br />

out in 1970, which were designed t o<br />

study the effects of rainfall on wetting front s<br />

in the rooting zone of a Douglas-fir stand .<br />

Lysimeter plates were installed at depths o f<br />

11 and 25 cm and rainfall was simulated by a<br />

below-canopy sprinkler system . Soil moisture<br />

was measured by tensiometer at a point mid -<br />

way between the soil surface and the uppe r<br />

lysimeter plate, and water flux through th e<br />

lysimeter was recorded by a flow meter connected<br />

to the data recorder .<br />

For the present study only the data for th e<br />

11-cm plate were used. Since an abrup t<br />

change in the physical characteristics of th e<br />

soil occurs between the 11- and 25-cm levels ,<br />

a different set of estimates of the conductivity<br />

and diffusivity parameters is require d<br />

for the lower layer . A model coupling th e<br />

flow of water through the two layers has bee n<br />

developed but has not yet been tested .<br />

96


Parameter Estimation<br />

In order to estimate conductivity (K) and<br />

diffusivity (D), both of which are functions of<br />

soil water content (0), we used standar d<br />

laboratory and computational procedures<br />

(Richards 1948, Green and Corey 1971) . The<br />

so-called moisture release curve, a plot of ,1i<br />

against 0, was also obtained in the laboratory .<br />

Curves for K, D, and as functions of 0 are<br />

shown in figure 1 .<br />

Our Cedar River core samples of the<br />

Everett gravelly sandy loam were biased to an<br />

unknown degree by the removal of the large<br />

stone fraction before the laboratory analysis .<br />

These stones constitute perhaps up to 20 per -<br />

cent of the volume of the Everett soil series .<br />

Because these stones tend to impede the flo w<br />

of water in unsaturated soils-the path aroun d<br />

them is longer than it would be for smaller<br />

obstructions-we believe our estimates of K<br />

may be somewhat too large . This decrease in<br />

conductivity associated with the presence o f<br />

large stones is especially noteworthy if th e<br />

soil is highly unsaturated because water<br />

strongly held by matric forces flows only<br />

along capillary paths around the stones an d<br />

not directly through the interstices of aggregations<br />

of large stones and cobbles .<br />

Since conductivity of water in unsaturate d<br />

soils is extremely difficult to measure directly,<br />

it has become common practice to estimate<br />

this parameter from pore-size distribution<br />

data, or equivalently, from the soil<br />

moisture release curve. Green and Corey<br />

(1971) have reviewed three methods fo r<br />

calculating K based on pore-size distribution<br />

and found that all give good results whe n<br />

compared with measured data . Our curve for<br />

K as a function of 0 is based on laboratory<br />

measurements of K for the saturated Everett<br />

soil and adjusted for unsaturated soils by us e<br />

of the Marshall pore-interaction relationship<br />

with a matching factor (Marshall 1958) .<br />

The shape of the moisture release curve i s<br />

also affected by sampling procedures. Removal<br />

of large stones clearly biases the result s<br />

in the direction of a soil with smaller pores .<br />

Accordingly, the correct relationship is some -<br />

what to the left of the one determined in the<br />

laboratory . This implies that the slope aO i s<br />

different from that indicated by our laboratory<br />

data and that our estimate of the dif -<br />

fusivity D = -K is biased .<br />

ao<br />

Computational Procedures<br />

We used the Remson computer program ,<br />

which provides an approximation to the solution<br />

of (3). As we have already noted, it is<br />

possible to calculate at the same time a n<br />

approximation to the volume flux of water ,<br />

v = -D ae, at any depth in the soil. For thi s<br />

calculation initial moisture content must b e<br />

specified at a number of equally space d<br />

points, called nodes, along the vertical soil<br />

profile. Our nodes were set at depths of 1, 3 ,<br />

5, 7, 9, and 11 cm. Initial moisture content<br />

was measured at approximately 5 .5 cm by<br />

tensiometer. The lysimeter plate at 11 cm<br />

depth provided another point of known<br />

moisture content . Because suction at the<br />

lysimeter plate was maintained at a constant<br />

level, the moisture content at that point coul d<br />

be obtained directly from the empirically<br />

determined moisture release curve. Estimate s<br />

of initial values at other nodes were obtaine d<br />

by linear extrapolation . Moisture content at<br />

the upper boundary (the soil surface) varie d<br />

with incoming precipitation .<br />

Results and Discussion<br />

Observed flow of water through the lysimeter<br />

plate is compared to that predicted by<br />

the flow equation in figures 2, 3, and 4 which<br />

summarize the outcomes of three field experiments,<br />

and the corresponding computer runs .<br />

The results follow a consistent pattern . Predicted<br />

and observed flows begin, peak, an d<br />

subside at about the same time . Maximu m<br />

predicted flow is about 25 percent greate r<br />

than observed. At lower levels of flow th e<br />

correspondence is closer .<br />

We conclude that the Richards flow equa -<br />

97


2<br />

10 7100 .160 .220 .280 .340 .400 .460 .520 .580 •640 -70 0<br />

2<br />

10 100 .160 .220 .280 .340 .400 .460 .520 .580 .640 .700<br />

100 0<br />

90 0<br />

t.u<br />

I- 60 0<br />

70 0<br />

0 60 0<br />

50 0<br />

U<br />

40 0<br />

30 0<br />

0 20 0<br />

100<br />

0<br />

-10 IQQ .200 .300 400 .500 .600 .70 0<br />

WATER FILLED POROSITY - THETA . (CU . CM . WATER)/(CU . CM . SOIL )<br />

Figure 1 . Soil water diffusivity (top), conductivity (middle), and suction (bottom) a s<br />

functions of soil moisture content (0) . Everett gravelly sandy loam . M indicates mea n<br />

values used in model . Three samples were used to obtain the moisture-release curve .<br />

98


F ,<br />

20 .0 40 .0 60 .0 80 .0 100 .0 120 .0 140 .0 160 .0 180 .0 200 . 0<br />

ELAPSED TIME . (MIN .)<br />

Figure 2 . Precipitation (P), observed soil water flow (F), and modeled<br />

soil water flow (M) at 11 cm. Lysimeter suction was 50 cm of water .<br />

.0600<br />

120 .0 140 . 0<br />

Figure 3 . Precipitation (P), observed soil water flow (F), and modeled<br />

soil water flow (M) at 11 cm. Lysimeter suction was 88 cm of water .<br />

99


.0500<br />

.040 0<br />

z .030 0<br />

E<br />

L.)<br />

0 .020 0<br />

.0140<br />

I<br />

4 0 20 .0 40 .0 60 .0 80 . 0 100 . 0 120 . 0 146 . 0<br />

ELAPSED TIME . (MIN . ]<br />

Figure 4 . Precipitation (PI, observed soil water flow (F), and modele d<br />

soil water flow (M) at 11 cm . Lysimeter suction was 146 cm of water .<br />

tion provides a satisfactory description of th e<br />

flow of water in the upper rooting zone o f<br />

our Douglas-fir stand . Since the Everett<br />

gravelly sandy loam is extremely variable, our<br />

results probably should not be extrapolate d<br />

much beyond the local area in which the y<br />

were obtained .<br />

We have not been able to account for th e<br />

25 percent discrepancy between predicted<br />

and observed peak flows . A number of possible<br />

explanations have occurred to us . It can<br />

be argued that the lysimeter system is not a<br />

perfect device for measuring water flow in a<br />

vertical column of soil . For example, in the<br />

early stages of a flow experiment, acceleratio n<br />

of flow due to lysimeter suction may be<br />

greater than that due to gravity. Perhaps more<br />

significant may be variations in flow associated<br />

with the lack of homogeneity of an i n<br />

situ forest soil. We suspect that stones or<br />

roots in soil above the lysimeter plate diverte d<br />

water toward or away from the plate . Perhaps<br />

more important, our estimates of initial value s<br />

of 0 at several nodes are not accurate.<br />

It is probably more important to emphasiz e<br />

that the field experiments described here wer e<br />

conducted for purposes rather far removed<br />

from our present one of evaluating th e<br />

Richards equation as a possible component of<br />

a soil-plant-atmosphere model . We also wish<br />

to point out that the Richards equation has<br />

not received adequate evaluation in field situations,<br />

perhaps because the stringent under -<br />

lying assumptions are rarely satisfied . Th e<br />

rather close correspondence between model<br />

and observation that we have obtained suggests<br />

that the model may be more robust tha n<br />

has been generally believed .<br />

Acknowledgments<br />

The work reported in this paper was supported<br />

by National Science Foundation Grant<br />

No. GB-20963 to the Coniferous <strong>Forest</strong> Biome ,<br />

U .S. Analysis of Ecosystems, Internationa l<br />

Biological Program . This is Contribution No .<br />

27 to the Coniferous <strong>Forest</strong> Biome .<br />

100


Childs, E. C., and N . Collis-George . 1950 . The<br />

permeability of porous materials . Roy . Soc.<br />

London Proc. A 201 : 392-405 .<br />

Gardner, W. R. 1960. Dynamic aspects of<br />

water availability to plants . Soil Sci . 85 :<br />

63-72 .<br />

Gessel, S . P., and D . W. Cole. 1965 . Influence<br />

of removal of forest cover on movement o f<br />

water and associated elements through soil .<br />

J . Ann . Water Works Assoc . 57 :<br />

1301-1310 .<br />

Green, R . E., and J . C . Corey . 1971 . Calculation<br />

of hydraulic conductivity : a further<br />

evaluation of some predictive methods . Soil<br />

Sci. Am. Proc . 35 : 3-8 .<br />

Literature Cited<br />

Marshall, T. J. 1958 . A relation between<br />

permeability and size distribution of pores .<br />

J. Soil Sci . 9 : 1-8 .<br />

Remson, I ., R. L. Drake, S . S. McNeary, and<br />

E. M. Wallo . 1965. Vertical drainage of an<br />

unsaturated soil . J. Hydraul . Div., Am . Soc .<br />

Civil Eng., Proc . 91 : 55-74 .<br />

Richards, L . A. 1931. Capillary conductio n<br />

through porous mediums . Physics 1 :<br />

318-333 .<br />

1948 . Porous plate apparatus fo r<br />

measuring moisture retention and trans -<br />

mission by soil . Soil Sci . 66 : 105-110.<br />

Rose, C . W. 1966 . Agricultural physics . 226 p .<br />

London: Pergamon Press .<br />

101


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Elemental transport changes occurring<br />

during development of a second-growth<br />

Douglas-fir ecosystem<br />

Charles C . Grie r<br />

Research Associat e<br />

an d<br />

Dale W . Cole<br />

Associate Professo r<br />

College of <strong>Forest</strong> Resources<br />

University of Washingto n<br />

Seattle, Washington 98195<br />

Abstract<br />

Mineral cycling processes in a second-growth Douglas-fir ecosystem have been monitored for nearly 10 year s<br />

at the A. E. Thompson Research Center in western Washington. In this interval, substantial year-to-yea r<br />

differences in quantities of elements transferred have been observed. For example, a comparison of transfers<br />

during the 1964-65 and 1970-71 measurement years showed that input of calcium by precipitation increase d<br />

roughly 300 percent while calcium transfer by throughfall, stem flow, and leaching from the forest floo r<br />

increased roughly 600 percent, 350 percent and 220 percent, respectively . In the same interval, the mass an d<br />

elemental content of the standing crop increased roughly 15 percent while forest floor mass and elementa l<br />

content remained constant. A portion of this difference is related to the increased mass accumulated in th e<br />

standing crop. However, the major factor causing this large difference is climatic variation . These differences in<br />

elemental transfer indicate the need for continuous observation of transfer processes to resolve the effects of<br />

ecosystem development on mineral cycling against this background of climatic variation .<br />

Introduction<br />

The growth and development of a coniferous<br />

forest ecosystem does not normally occu r<br />

in a steady linear fashion . Rather, intervals of<br />

rapid stand growth may be interrupted b y<br />

periods of slower growth reflecting changes in<br />

growing conditions . In turn, the cycling of<br />

elements within a forest ecosystem can be expected<br />

to reflect these changes .<br />

Intensive study of the mineral cycling processes<br />

in second-growth Douglas-fir stands ha s<br />

been in progress at the A. E. Thompson Re -<br />

search Center since 1961 . During this time, a<br />

substantial body of information has been collected<br />

concerning the amounts, pathways ,<br />

rates of transfer, and mechanisms controlling<br />

transfer of elements by the mineral cyclin g<br />

processes of this ecosystem. Between 196 1<br />

and 1966, research was concentrated largely<br />

on detailed descriptions of ecosystem components<br />

(Cole, Gessel, and Dice 1967) .<br />

Research at the Thompson site since 196 6<br />

has had two major objectives : (1) to provide<br />

information describing the changes in mineral<br />

cycling processes as the stand grows an d<br />

(2) to examine the year-to-year variation i n<br />

mineral cycling processes . More recently, a<br />

third objective has been to examine the relationships<br />

between climatic factors and th e<br />

various mechanisms controlling rates of transfer<br />

along the different transfer pathways .<br />

During the course of research at th e<br />

Thompson site, a number of changes in th e<br />

103


patterns of elemental cycling have been observed<br />

. Some of these changes are related t o<br />

changes in the structure or mass of the ecosystem;<br />

other changes are related to short -<br />

term variations in climatic factors . The<br />

purpose of this paper is to examine several of<br />

the changes in mineral cycling processes a s<br />

they have been observed during the 10 years<br />

of research at the Thompson Research Center .<br />

<strong>Experimental</strong> Area<br />

These studies were conducted at the A . E .<br />

Thompson Research Center, an area specifically<br />

developed for the study of elemental<br />

cycling in second-growth Douglas-fir stands . It<br />

is located about 64 km southeast of Seattle ,<br />

Washington, at an elevation of 215 m in the<br />

foothills of the Washington Cascades . A full<br />

description of the geology, soils, vegetation ,<br />

and climate is given by Cole and Gessel<br />

(1968) .<br />

The study site is located on a glacial out -<br />

wash terrace along the Cedar River. This outwash<br />

terrace was formed during the recessio n<br />

of the Puget lobe of the Fraser ice sheet abou t<br />

12,000 years ago .<br />

The soil underlying the research plots is<br />

classified as a typic haplorthod (U .S. Department<br />

of Agriculture 1960) and is mapped a s<br />

Everett gravelly, sandy loam. This soil contains<br />

less than 5 percent silt plus clay an d<br />

normally contains gravel amounting to 50-8 0<br />

percent of the soil volume. The forest floor is<br />

classified as a duff-mull (Hoover and Lun t<br />

1952) and ranges from 1 cm to 3 cm thick .<br />

This forest floor represents the accumulatio n<br />

since 1931 when the present stand was established<br />

following logging (around 1915) an d<br />

repeated fires .<br />

The present overstory vegetation is a<br />

planted stand of Douglas-fir (Pseudotsuga<br />

menziesii (Mirb.) Franco) which was established<br />

about 1931 . Currently, the trees average<br />

about 19 m high with a crown density o f<br />

about 85 percent .<br />

The principal understory species are salal<br />

(Gaultheria shallon Pursh), Oregon grap e<br />

(Berberis nervosa (Pursh) Nutt .), red huckleberry<br />

(Vaccinium parvifolium Smith), and<br />

twinflower (Linnaea borealis L . ssp . americana<br />

(Forbes) Rehder) . Various mosses are<br />

the principal understory vegetation beneath<br />

the denser portions of the canopy .<br />

The climate is typical of foothill condition s<br />

in the Puget Sound Basin . Temperatures have<br />

ranged from -18° C to 38° C, but these extremes<br />

are seldom reached . The average temperature<br />

for July is 16 .7° C and for January is<br />

2.8° C. The average annual precipitation i s<br />

136 cm, almost all falling as rain . Precipitation<br />

rates are generally less than 0 .25 cm/hour<br />

and over 70 percent falls between Octobe r<br />

and March .<br />

Field Plots<br />

The facilities at the research site are designed<br />

to provide precision measurement of<br />

the flux of elements within this forest ecosystem.<br />

Some transfers are monitored continuously;<br />

others, with less potential for rapid<br />

change, are measured at regular intervals . A<br />

detailed description of the collection facilitie s<br />

at this site is given by Cole and Gessel (1968) .<br />

The following section will briefly describe<br />

these facilities with emphasis on recen t<br />

changes in the system .<br />

Sampling<br />

Precipitation collections are made at two<br />

locations in the stand . One collector is at the<br />

top of a 30-m tower and is fitted with a flowmeter;<br />

the other is an automatic collecto r<br />

(Wong Laboratories) which opens at the onse t<br />

of precipitation and is mounted on a 10- m<br />

tower in a clearcut adjacent to the plots .<br />

Canopy drip (foliar leaching) is collected i n<br />

six randomly located screened funnel s<br />

mounted in the necks of collection flasks .<br />

Stemflow is diverted from the stem of the six<br />

sampled trees at breast height (1 .35 m) by<br />

soft rubber collars and is collected in covered<br />

160-liter polyethylene trash cans . Litter is collected<br />

on eight randomly located 0 .21 m 2<br />

plastic litter screens. Leaching through the<br />

forest floor and soil is measured using the<br />

tension lysimeter system developed by Col e<br />

(1968). This system permits direct measure -<br />

104


ment of the flux of ions through the soil . I n<br />

this plot, 3 lysimeters per soil horizon are located<br />

in the soil at the base of the forest<br />

floor, at the base of the A l horizon (4 cm)<br />

and B 2 horizon (30 cm) and at 100 cm in th e<br />

C horizon .<br />

Instrumentatio n<br />

Meteorological measurements including ai r<br />

and soil temperature profiles, are made on a<br />

continuous basis utilizing a data logging system.<br />

Details of this instrumentation are give n<br />

by Cole and Gessel (1968) .<br />

The data logging system also permits th e<br />

continuous monitoring of the pH and conductivity<br />

of water moving through this ecosystem.<br />

These facilities are described by Col e<br />

(1968). Volumes of water flow between different<br />

components of the system are measured<br />

using the resistance flowmeter describe d<br />

by Cole (1968) .<br />

Chemical Analysis<br />

Analytical methods used on samples fro m<br />

this site have changed over the years. For<br />

methods used in earlier studies, the reade r<br />

may consult the following reports : Rahma n<br />

(1964); Cole and Gessel (1965) ; and Cole ,<br />

Gessel, and Dice (1967) . The methods currently<br />

in use will be briefly described in the<br />

following paragraphs .<br />

Water Analysis<br />

Determinations of calcium, magnesium ,<br />

potassium, and sodium are normally made<br />

directly on water samples using an Instrumentation<br />

Laboratories-353 atomic-absorptio n<br />

spectrophotometer . Calcium is determined in<br />

a nitrous oxide-acetylene flame ; and magnesium,<br />

potassium, and sodium are deter -<br />

mined in an air-acetylene flame . Nitrogen an d<br />

phosphorus are determined from a hydroge n<br />

peroxide-sulfuric acid digest (Linder and Harley<br />

1942) of a 10-fold concentration of th e<br />

solutions. Nitrogen is determined using a<br />

micro-Kjeldahl distillation (Jackson 1958) .<br />

Phosphorus is determined using the chlorostannous-reduced<br />

molybdophosphoric blue<br />

color method of Jackson (1958). Total io n<br />

concentration is estimated from specific conductance<br />

using an empirical method (Logan<br />

1961) and pH is determined instrumentall y<br />

using established methods . Further details of<br />

water analysis are given by Grier (1972) .<br />

Soil Analysis<br />

Total soil nitrogen is determined using th e<br />

standard Kjeldahl digest and distillation o f<br />

Jackson (1958) . Cation exchange capacity i s<br />

determined using the neutral ammonium acetate<br />

leaching method (Jackson 1958) . Ex -<br />

changeable cations are determined spectrophotometrically<br />

from the ammonium acetat e<br />

leachate . Soil carbon is determined instrumentally<br />

using a Leco carbon analyzer . Phosphorus<br />

is extracted from the soil by digestio n<br />

in 36 N sulfuric acid and 30 percent hydroge n<br />

peroxide ; concentrations are then determine d<br />

following the same procedures as for water<br />

samples .<br />

Tissue and <strong>Forest</strong> Floor Analysi s<br />

Elemental assays of forest floor and plant<br />

tissue are done using methods described by<br />

Grier and McColl (1971) .<br />

Results and Discussion<br />

Table 1 shows how organic matter, calcium,<br />

potassium, nitrogen, and phosphoru s<br />

were distributed among different component s<br />

of the ecosystem at the Thompson site at th e<br />

end of 1965 (Cole, Gessel and Dice 1967) .<br />

Since the time of this determination, transfer s<br />

between these various components have bee n<br />

monitored to assess the changes and rates of<br />

change of elemental distribution in the stand .<br />

In the interval between 1966 and the present,<br />

the volume of woody tissue in the stan d<br />

has increased by an average of 11 m 3 /ha/yf ' .<br />

This represents approximately a 15-percent<br />

increase in volume since 1966 . Increases in<br />

mass and elemental content of most components<br />

of the standing crop are estimatedusing<br />

estimation procedures developed fo r<br />

105


these stands by Dice (1970)-to be proportional<br />

to the increase in woody tissue mass .<br />

However, the foliar mass has probably remained<br />

about constant in this interval . Thus<br />

at the present time, the quantities of elements<br />

and organic matter in the standing crop at this<br />

site are about 15 percent greater than show n<br />

in table 1 .<br />

In the same interval, other components of<br />

the ecosystem have shown relatively little<br />

change. For example, the forest floor was intensively<br />

sampled in 1961, 1 1965 (Cole, Gessel,<br />

and Dice 1967), and 1969 (Grier and Mc -<br />

Coll 1971) . Results of these studies are summarized<br />

in table 2 . With the exception o f<br />

magnesium content, the forest floor of thi s<br />

D . W . Cole . Unpublished data on file at the Colleg e<br />

of <strong>Forest</strong> Resources, University of Washington, Seattle .<br />

Table 1 .-Distribution of N, P, K, Ca, and organic matter (g/m 2 ) in a 35-year-old second-growt h<br />

Douglas-fir ecosystem at the Thompson Research Center (Cole, Gessel, and Dice 1967 )<br />

Component N P K Ca<br />

Organic<br />

matte r<br />

TREE<br />

Foliage current 2.4 0 .5 1 .6 0.7 199 . 0<br />

older 7.8 2 .4 4.6 6.6 710 . 7<br />

Branches current .4 .1 .3 .2 51 . 3<br />

older 4.0 .9 3.2 6.5 1,337 . 3<br />

dead 1.7 .2 .3 3.9 814 . 5<br />

Wood current 1.0 .2 1 .0 .4 748 . 5<br />

older 6.7 .7 4.2 4.3 11,420 . 2<br />

Bark 4.8 1.0 4.4 7.0 1,872 . 8<br />

Roots 3.2 .6 2.4 3.7 3,298 . 6<br />

Total tree 32 .0 6.6 22 .0 33 .3 20,452 . 9<br />

SUBORDINAT E<br />

VEGETATI<strong>ON</strong> .6 .1 .7 .9 101 . 0<br />

FOREST FLOO R<br />

Branches .5 .1 .4 .8 142 . 3<br />

Needles 3.5 .4 .5 2.7 300 . 5<br />

Wood 1.4 .2 .8 1 .7 634 . 5<br />

Humus 12 .1 1.9 1 .5 8.5 1,199 . 9<br />

Total forest floor 17 .5 2.6 3.2 13 .7 2,277 . 2<br />

SOIL<br />

0-15 cm<br />

80 .9 116 .7 7 .9 31 .3 3,837 . 2<br />

15-30 cm 86 .8 119 .5 6 .6 19 .6 3,693 . 5<br />

30-45 cm 76 .1 98 .0 5 .2 15 .2 2,829 . 0<br />

45-60 cm 37 .1 53 .6 3 .7 8.0 795 . 5<br />

Total soil 280 .9 387 .8 23 .4 74 .1 11,155 . 2<br />

TOTAL ECOSYSTEM 331 .0 397 .1 49 .3 122 .0 33,986 .3<br />

106


Table 2 .-<strong>Forest</strong> floor composition changes during an 8-year period in<br />

a Douglas-fir ecosystem at the Thompson Research Center<br />

Year<br />

sampled<br />

Dry<br />

weight<br />

N<br />

g/m 2<br />

P Ca Mg K<br />

1961 1,542 15 .4 2.2 11 .2 3 .8 2 . 4<br />

1964 1,500 16 .1 2.4 12 .0 3.7 2 . 4<br />

1969 1,430 14 .0 12 .3 1.9 2 .7<br />

ecosystem has changed only slightly since 196 1<br />

indicating that litterfall is balanced by decomposition.<br />

Similarly, there has been no significant<br />

change in elemental capital in the soi l<br />

(table 3) due both to the large reserves of mos t<br />

nutrients in this soil and to efficient cyclin g<br />

processes within the ecosystem . The decreas e<br />

in magnesium capital in the forest floor between<br />

1965 and 1969 may indicate increased<br />

rates of internal redistribution of magnesiu m<br />

within the standing crop . However, this i s<br />

doubtful in view of the small change in magnesium<br />

capital in the soil (table 3) .<br />

The quantities of cations transferred between<br />

components of the ecosystem generall y<br />

increased between 1964-65 and 1970-71 (table<br />

4). Comparison of some transfers during th e<br />

1970-71 measurement year with those of th e<br />

1964-65 measurement year (table 4) reveal s<br />

some rather substantial differences in both th e<br />

overall amounts transferred and in the relativ e<br />

importance of the various pathways .<br />

For example, input of calcium and potassium<br />

by precipitation is roughly four-fol d<br />

greater now than in 1964-65. This increas e<br />

may reflect a general increase in atmospheri c<br />

pollution in the Puget Sound basin . The quantities<br />

of calcium and potassium returned t o<br />

the soil from the aboveground vegetation hav e<br />

also increased since the 1964-65 measurement.<br />

As an example, return of calcium an d<br />

potassium by stemflow has increased 25- and<br />

35-fold, respectively, since measurement during<br />

1964-65 . Increases in return by foliar<br />

leaching also were observed (table 4) bu t<br />

these were not of the same magnitude as th e<br />

Increased return by stemflow. No explanation<br />

has been found for the differential increases<br />

in return by stemflow and foliar leaching .<br />

Return of elements by litterfall also in -<br />

creased between 1964-65 and 1970-71 (table<br />

4) . These differences are probably due both<br />

to normal year-to-year fluctuation of litterfal l<br />

and also to the increasing proportion o f<br />

branch material observed as a component o f<br />

the litter .<br />

Increases were also observed in leaching of<br />

elements through the soil profile . Leaching of<br />

calcium and potassium from the forest floo r<br />

roughly doubled between 1964-65 an d<br />

1970-71 (table 4) . A portion of this increase<br />

may result from increased rates of litter de -<br />

composition as well as the increased input in<br />

precipitation .<br />

As previously noted, the forest floor mas s<br />

appears to have approached a steady state<br />

condition (table 2) ; the present mass representing<br />

a quasi-equilibrium state, achieve d<br />

since the stand was established in 1931, afte r<br />

repeated fires through the area . On the other<br />

hand, litterfall quantity has increased durin g<br />

the period of observation of this stand. In<br />

table 5, monthly litterfall measured at thi s<br />

site in 1962-63 (Rahman 1964) is compared<br />

with measurements made during 1970-71 ; the<br />

annual mass of litterfall has apparently increased<br />

roughly 50 percent over this 8-yea r<br />

period. Thus the increased quantities of ele-<br />

10 7


Table 3.-Elemental capital in the soil of a second-growth Douglas-fir ecosystem<br />

at the Thompson Research Center as determined in 1965 and 197 1<br />

Depth<br />

g/m 2<br />

N P K Ca Mg<br />

A) 1965 :<br />

0-15 81 117 7.9 31 3 . 7<br />

15-30 87 120 6.6 20 3 . 5<br />

30-45 76 98 5.2 15 2 . 5<br />

45-60 37 54 3.7 8 1 . 2<br />

Total soil 281 389 23 .4 74 10 . 9<br />

B) 1971 :<br />

0-15 86 123 8.3 40 3 . 2<br />

15-30 82 111 6.6 21 3 . 3<br />

30-45 71 104 5.3 10 2 . 7<br />

45-60 40 52 4.2 11 1 . 5<br />

Total soil 279 390 24 .4 82 10 .7<br />

Table 4. Transfers of elements between components of a Douglas-fir ecosyste m<br />

at the Thompson Research Center during two stages of development '<br />

Item<br />

Calcium (g/m 2 ) Magnesium (g/m 2 ) Potassium (g/m 2 ) Sodium (g/m 2 )<br />

1 2 1 2 1 2 1 2<br />

Precipitation 0.28 0.93 ND 0.22 0.08 0.47 ND 1 .6 8<br />

Throughfall .35 2.05 ND .49 1 .07 2 .83 ND 2 .5 9<br />

Stemflow .11 3.85 ND .60 .16 4.00 ND 2 .3 0<br />

Litterfall 1 .11 1.78 ND .20 .27 .47 ND .0 7<br />

Leaching fro m<br />

forest floor 1 .74 3.87 ND 1 .10 1 .05 2.60 ND 2 .1 8<br />

1 1=transfers measured during 1964-65 measurement year . Reported by Cole, Gessel and Dice (1967) .<br />

2=transfers measured during 1970-71 measurement year .<br />

108


Table 5 .-Monthly amounts of litterfall during two intervals in th e<br />

development of a second-growth Douglas-fir ecosyste m<br />

Month<br />

Quantity (g/m 2 )<br />

1962-63 1970-7 1<br />

October 73 .0 89 .0<br />

November 17 .9 44 . 5<br />

December 9.2 14 . 2<br />

January 6.5 10 . 0<br />

February 15 .9 3 . 8<br />

March 6 .2 13 . 8<br />

April 7 .4 5 . 8<br />

May 12 .8 ( I )<br />

June 11 .4 ( 1 )<br />

July 5 .2 39 .1 1<br />

August 5 .5 21 . 0<br />

September 12 .3 41 .6<br />

' Litterfall amounts for May, June, and July 1971 are combined in July collection .<br />

ments leached from the forest floor are a t<br />

least partially due to increased litter decomposition<br />

rates .<br />

Many of the changes in amount and relative<br />

importance of the transfer pathways sinc e<br />

1965, are probably due to changes in the mass<br />

and structure of this forest ecosystem . How -<br />

ever, many of the changes are completely ou t<br />

of proportion to estimates of increased mas s<br />

of the standing crop .<br />

Observations made since 1966 at th e<br />

Thompson site indicate that the return of elements<br />

to the soil and their subsequent distribution<br />

through the soil profile is sensitive t o<br />

certain climatic factors ; primarily precipitation<br />

and temperature . Variations in these<br />

climatic factors have been observed to cause<br />

substantial variation in quantities transferred<br />

along most pathways . As an example, the re -<br />

turn of elements to the soil by stemflow an d<br />

foliar leaching were found to be partially<br />

regulated by the distribution of precipitation ;<br />

especially precipitation occurring during th e<br />

summer. Figure 1 shows the monthly return<br />

of potassium by stemflow from one tree fo r<br />

1970 and 1971 ; the total amounts transferred<br />

being 0 .50 gm during 1970, and 0 .75 gm during<br />

1971 . Much of the greater amount re -<br />

turned during 1971 was returned during a fe w<br />

summer rainstorms . Foliar leaching, although<br />

not shown, exhibits a similar pattern of behavior.<br />

Apparently the extent of removal of<br />

potassium and other elements by leachin g<br />

from Douglas-fir foliage is related to its<br />

phenological stage . This suggests that precipitation<br />

occurring in certain critical parts of the<br />

year may remove elements that would other -<br />

wise remain in the plant .<br />

Temperature, the other major factor causing<br />

variation in year-to-year transfer rates ,<br />

exerts its control on transfers primaril y<br />

through its effect on organisms active in de -<br />

composition . Decomposer activity regulate s<br />

the availability of nutrient elements in tw o<br />

ways ; first, by regulating the rate of mineralization<br />

of elements and second, by its influence<br />

on levels of the mobile bicarbonate<br />

anion in the soil solution .<br />

10 9


(A . )<br />

f} Q<br />

2 0<br />

Figure 1 .<br />

A. Volume of stemflow from a 15-cm diameter Douglas-fir during 2 successive years .<br />

B. Quantity of potassium transferred to soil from a single 15-cm diameter Douglas-fir during 2 successive years .<br />

.2<br />

i<br />

N<br />

U<br />

N<br />

O<br />

U<br />

0<br />

20<br />

15<br />

4 JUNE<br />

6 - 7 JUN E<br />

1 k i<br />

1 1 k l 1<br />

2400<br />

O6DO 1200 1800<br />

TB[ OF DAY<br />

Figure 2 . Temperature effects on CO2 evolution from forest floor at Thompson Research Center (Ballard 1968) .<br />

110


Figure 2 illustrates how CO 2 production<br />

from the forest floor is related to temperatur e<br />

in this ecosystem (Ballard 1968) . Temperature<br />

effects on mineralization rates are probably<br />

directly proportional to the temperatur e<br />

effects on CO 2 production rates .<br />

The bicarbonate ion is the major anion i n<br />

soil solutions at the Thompson site. Moreover ,<br />

the concentration of this ion in the soil solution<br />

has been shown to be the major facto r<br />

controlling rates of cation leaching in these<br />

soils (McColl 1969) . Temperature effects o n<br />

the levels of the bicarbonate ion are related t o<br />

decomposer activity, since the equilibrium between<br />

CO2 in the soil atmosphere and HCO3<br />

in the soil solution is related to CO 2 concentration<br />

. Consequently, leaching rates in th e<br />

soil of the Thompson site are strongly relate d<br />

to temperature (McColl 1969) .<br />

Figure 3A illustrates the relationship between<br />

estimated ion concentration in fores t<br />

floor leachates and air temperature . As can be<br />

seen from these data, the ion concentratio n<br />

closely follows temperature during the year .<br />

Monthly variations in mean ion concentration<br />

of forest floor leachates for 3 successive years<br />

(figure 3B) are primarily the result of differences<br />

in mean monthly temperature .<br />

Total transfer by leaching is regulated both<br />

by temperature and the amount and date of<br />

-1C<br />

~'' 0''404%4P` EC I P . C<strong>ON</strong>DUCTANC E<br />

<strong>ON</strong>THLY PRECIP .<br />

~. _ _ -,'•<br />

■<br />

JAti FEB MAR APR NAY JUNE JULY AUG SEPT OCT NOV DE C<br />

0<br />

0<br />

JAN FEB MAR APR MAY JUNE JULY AUG SEPT OCT NOV DE C<br />

Figure 3 .<br />

A. Mean monthly precipitation, air temperature, and specific conductance of precipitation and forest floo r<br />

leachates for the Thompson site during 1970 .<br />

B. Specific conductance of forest floor leachates for 3 consecutive years at the Thompson site .<br />

11L


Figure 4 . Rates of calcium leaching through soil profile and rate of precipitation during 1970-71 growing season .<br />

occurrence of precipitation . The largest quantities<br />

of elements are transferred during<br />

periods of relatively high temperatures and<br />

ample precipitation . Figure 4 illustrates the<br />

relationship between leaching of calcium from<br />

the forest floor and A horizon of the minera l<br />

soil in 1970-71 and precipitation in tha t<br />

interval . As can be seen from these data, the<br />

greatest quantities of calcium were transferre d<br />

during the spring and autumn when decomposer<br />

activity was relatively high and sufficient<br />

precipitation fell to flush the decomposition<br />

products into the soil .<br />

Conclusions<br />

Mineral cycling processes within a forest<br />

ecosystem are responsive to changes in th e<br />

mass and structure of the standing crop an d<br />

also to the shorter term effects of climatic<br />

variation . Of the two, the effects of climatic<br />

variation are the most readily seen . For this<br />

reason, any efforts to describe mineral cycling<br />

processes within a forest ecosystem must include<br />

sufficient data that effects of changes in<br />

mass and structure of the ecosystem can be<br />

resolved and separated from the large year-toyear<br />

effects of climatic variability .<br />

Present and future mineral cycling research<br />

at the Thompson site is directed towards<br />

separating and describing the possibly inter -<br />

dependent effects of growth and climate on<br />

mineral cycling processes in second-growt h<br />

Douglas-fir .<br />

Acknowledgments<br />

The work reported in this paper was sup -<br />

ported in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferous<br />

<strong>Forest</strong> Biome, U .S. Analysis of Ecosystems ,<br />

International Biological Program. This i s<br />

Contribution No. 28 to the Coniferous <strong>Forest</strong><br />

Biome, IBP .<br />

112


Ballard, T. M. 1968 . Carbon dioxide production<br />

and diffusion in forest floor materiala<br />

study of gas exchange in biologically<br />

active porous media. 120 p . Ph .D . thesis on<br />

file, Univ . Wash., Seattle .<br />

Cole, D . W . 1968 . A system for measuring<br />

conductivity, acidity, and rate of flow in a<br />

forest soil . Water Resour. Res. 4 :<br />

1127-1136 .<br />

and S . P . Gessel . 1965 . Movement<br />

of elements through a forest soil a s<br />

influenced by tree removal and fertilizer<br />

additions . In C. T . Youngberg (Ed .) <strong>Forest</strong>soil<br />

relationships in North America, p .<br />

95-104. Corvallis, Oreg. : Oreg. State Univ .<br />

Press .<br />

and S. P . Gessel . 1968 . Cedar<br />

River research-a program for studyin g<br />

pathways, rates and processes of elemental<br />

cycling in a forest ecosystem. For. Resour .<br />

Monogr . Contrib . No . 4, 53 p . Univ. Wash . ,<br />

Seattle .<br />

S. P. Gessel, and S . F. Dice .<br />

1967 . Distribution and cycling of nitrogen ,<br />

phosphorus, potassium, and calcium in a<br />

second-growth Douglas-fir ecosystem . In<br />

Primary productivity and mineral cycling in<br />

natural ecosystems-symposium, p .<br />

193-197. Assoc . Advan. Sci. 13th Annu .<br />

Meet. Orono : Univ . Maine Press .<br />

Dice, S . F. 1970 . The biomass and nutrien t<br />

flux in a second-growth Douglas-fir ecosystem<br />

(a study in quantitative ecology) .<br />

165 p . Ph .D . thesis on file, Univ . Wash . ,<br />

Seattle .<br />

Literature Cited<br />

Grier, C . C. 1972 . Effects of fire on the movement<br />

and distribution of elements within a<br />

forest ecosystem. 167 p. Ph.D. thesis on<br />

file, Univ . Wash ., Seattle .<br />

and J. G . McColl . 1971 . <strong>Forest</strong><br />

floor characteristics within a small plot i n<br />

Douglas-fir in western Washington . Soil Sci .<br />

Soc. Am. Proc . 35 : 988-991 .<br />

Hoover, M . D., and A . H . Lunt . 1952 . A key<br />

for the classification of forest humus types .<br />

Soil Sci . Soc. Am. Proc . 16:368-370 .<br />

Jackson, M. L. 1958 . Soil chemical analysis .<br />

498 p. Englewood Cliffs, N .J . : Prentice-<br />

Hall, Inc .<br />

Linder, R . C., and C . P . Harley . 1942 . A rapid<br />

method for the determination of nitroge n<br />

in plant tissue . Science 96 : 565-566 .<br />

Logan, J. 1961 . Estimation of electrical conductivity<br />

from chemical analysis of natural<br />

waters. J . Geophys . Res . 66: 2479-2483 .<br />

McColl, J . G . 1969. Ion transport in a forest<br />

soil : models and mechanisms . 214 p . Ph.D .<br />

thesis on file, Univ . Wash., Seattle .<br />

and D. W. Cole. 1968 . A mechanism<br />

of cation transport in a forest soil .<br />

Northwest Sci . 42 : 134-140 .<br />

Rahman, A . H. M . M. 1964. A study of the<br />

movement of elements from tree crowns b y<br />

natural litterfall, stemflow and leaf wash .<br />

117 p. M.F. thesis on file, Univ . Wash . ,<br />

Seattle .<br />

U .S. Department of Agriculture. 1960. Soil<br />

classification-a comprehensive system-7th<br />

approximation. 265 p. Washington, D .C . :<br />

U .S. Govt. Print. Office .<br />

113


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Nutrient budget of a Douglasfir<br />

forest on an experimenta l<br />

watershed in western Oregon<br />

R . L . Fredrikse n<br />

Soil Scientist, <strong>Forest</strong>ry Sciences Laborator y<br />

Pacific Northwest <strong>Forest</strong> and Range Experiment Statio n<br />

<strong>Forest</strong> Service, U .S . Department of Agricultur e<br />

Corvallis, Orego n<br />

Abstract<br />

Annual loss of nitrogen, phosphorus, silica, and the cations sodium, potassium, calcium, and magnesiu m<br />

followed the same pattern as annual runoff which is heavily dominated by winter rainstorms arising from th e<br />

Pacific Ocean. Even though 170 and 135 cm of water passed through this Douglas-fir ecosystem for the 2 year s<br />

reported here, this ecosystem conserved nitrogen effectively as indicated by an average annual dissolved nitroge n<br />

outflow of 0.5 kg/ha from an annual average input of 1 .0 kg/ha in precipitation . There was a small annual ne t<br />

loss of phosphorus (0.25 kg/ha) . Cation input in precipitation was less than 10 percent of sources from minera l<br />

weathering-indicating that mineral weathering was the principal source of cations to the system. Average<br />

annual net losses of calcium, sodium, magnesium, and potassium were : 47, 28, 11, and 1 .5 kg/ha, respectively.<br />

Silica loss of 99 kg/ha-yr was the largest of all constituents and came entirely from within the forest system .<br />

Although loss of sediment was low during the period of study, loss of nutrients by soil erosion may become of<br />

major importance over a longer time scale due to widely spaced unsampled catastrophic erosion .<br />

Introduction<br />

The purpose of this study was to measur e<br />

the inputs, losses, and retentions of specifi c<br />

plant nutrients .<br />

<strong>Forest</strong>ed watersheds have been used fo r<br />

decades to study the hydrologic cycle .<br />

Bormann and Likens (1967) indicated tha t<br />

small watersheds are suitable for studies of<br />

chemical cycles if the ecosystem is located o n<br />

a watershed with relatively impermeable bed -<br />

rock . Then, all annual chemical fluxes int o<br />

and out of the forest system can be attribute d<br />

to a definable forest ecosystem, and error s<br />

due to deep seepage can be minimized . Small<br />

watersheds containing soil-plant systems hav e<br />

been generally accepted as useful units fo r<br />

ecological research .<br />

Nutrient cycles begin with the establishment<br />

of vegetation, and the quantity o f<br />

nutrients cycled undoubtedly increases as the<br />

forest passes through successional phases . At<br />

all points, along this developmental sequence ,<br />

the supply of nutrients in the soil-plan t<br />

system is regulated by the balance betwee n<br />

(1) the inputs from the atmosphere an d<br />

mineral weathering and (2) the outflow b y<br />

soil erosion or chemicals dissolved in stream<br />

water. A continued level of fertility of th e<br />

system therefore depends upon the retentio n<br />

of nutrients in the cycle from loss by leachin g<br />

and soil erosion .<br />

Nutrient budget studies were begun on two<br />

small watersheds at the H . J. <strong>Andrews</strong> <strong>Experimental</strong><br />

<strong>Forest</strong> in 1967 . Only the initial results<br />

from one of these and the chemicals in solution<br />

will be reported here . The followin g<br />

questions were asked :<br />

1 . What quantities of nitrogen, phosphorus ,<br />

the cations (sodium, potassium, calcium ,<br />

magnesium), and silica reach the site i n<br />

precipitation?<br />

115


2. How effectively are these chemicals retained<br />

by the forest ?<br />

3. What hypotheses can be made about th e<br />

processes that regulate the gain or loss o f<br />

these chemicals from the forest ?<br />

The 10.1-hectare study watershed, designated<br />

number 10 in <strong>Forest</strong> Service studies ,<br />

occupies strongly dissected topograph y<br />

characteristic of the west side of the Cascade<br />

Range. It rises from 430 m elevation at the<br />

stream gaging station to 670 m at the highes t<br />

point on the back ridge. The overall slope is<br />

44 percent, but side slopes and the headwal l<br />

range up to 90 percent due to deep incision of<br />

the basin into the main ridge . Soils, fro m<br />

weathered tuff and breccia materials, hav e<br />

weakly developed profiles that often overlie<br />

saprolite up to 6 m deep . Tree roots are mos t<br />

abundant in the surface 0 .6 m and probably<br />

do not penetrate much deeper than 2 .4 m .<br />

The saprolite is porous and therefore can<br />

transmit water . The average annual precipitation<br />

is 230 cm .<br />

<strong>Forest</strong> vegetation on the watershed consist s<br />

of remnants of a 450-year-old Douglas-fi r<br />

(Pseudotsuga menziesii) stand and islands of<br />

various younger age classes which originate d<br />

when the old-growth stand was broken up by<br />

windthrow, disease, or fire . Because the basin<br />

is oriented toward the southwest, the sid e<br />

slopes within it face south, west, and north . A<br />

considerable diversity in vegetation on thes e<br />

slopes has recently been classified for th e<br />

<strong>Experimental</strong> <strong>Forest</strong> by Dyrness, Franklin ,<br />

and Moir (personal communication) . Douglasfir<br />

climax forests occupy ridgetop and upper<br />

slope south aspect positions on the watershed .<br />

A dense evergreen shrub layer, compose d<br />

mainly of golden chinkapin (Castanopsis<br />

chrysophylla), Pacific rhododendron (Rhododendron<br />

macrophyllum), and salal (Gaultheria<br />

shallon) codominates the site beneath a spars e<br />

stand of old-growth trees. Western hemloc k<br />

climax forests make up the remainder of th e<br />

vegetation on the watershed . On these habitats,<br />

the dominant overstory is mainl y<br />

Douglas-fir and western hemlock (Tsuga<br />

heterophylla) . Proceeding from west to north<br />

aspects along a gradient of increasing moisture,<br />

the shrub cover shifts from vine maple<br />

(Acer circinatum)-salal, to rhododendron -<br />

Oregon grape (Mahonia nervosa) and finally<br />

to vine maple-western sword fer n<br />

(Polystichum munitum) on moist, northfacing<br />

middle and lower slopes .<br />

Methods<br />

Precipitation and streamflow are sample d<br />

by the following techniques . Precipitation at<br />

the top and bottom of the watershed i s<br />

measured by standard rain gages ; samples for<br />

chemical analysis are obtained from a specia l<br />

collector mounted on a tower 18 m tall in th e<br />

center of a nearby clearcut . At the outlet o f<br />

the watershed, an H-flume gages streamflo w<br />

and a proportional water sampler, capable of<br />

sampling both dissolved and suspended materials<br />

from the stream, composites discret e<br />

water samples in a polyethylene carbo y<br />

(Fredriksen 1969) . The samples are remove d<br />

at approximately 3-week intervals throughou t<br />

the year. In 1970 and 1971, grab sample s<br />

were taken from low summer flows ; and during<br />

the winter of 1971-72, a series of gra b<br />

samples were collected from streamflow during<br />

storm runoff events . In the laboratory ,<br />

solids are removed by filtration through a fine<br />

glass fiber filter (Whatman GF/C) from bot h<br />

precipitation and streamflow samples and<br />

chemical analyses were performed on bot h<br />

fractions. The fraction that passes the filter i s<br />

considered to be dissolved although we recognize<br />

that the filtered water may contain som e<br />

particles smaller than 2 microns in diameter .<br />

Since the concentration of each chemica l<br />

constituent is an estimate of the mean for th e<br />

time during which the sample was collected ,<br />

the input or loss for each sampling period is<br />

the product of precipitation or streamflow,<br />

respectively, and the concentration of eac h<br />

constituent .<br />

Chemical analyses were done in duplicat e<br />

as follows : Ammonium and dissolved organi c<br />

nitrogen by distillation and digestion, respectively,<br />

on 1/2-liter samples and detection b y<br />

Nesslerization; nitrite by the sulphanilamid e<br />

method ; nitrate by reduction and detection a s<br />

nitrite ; orthophosphorus by the molybdat e<br />

blue method ; total phosphorus by the<br />

molybdate blue method following a per -<br />

116


sulfate-sulfuric acid digestion in the autoclave ;<br />

reactive silica by the molybdate yello w<br />

method ; cations sodium and potassium by<br />

flame emission ; and calcium and magnesium<br />

by atomic absorption following addition of<br />

lanthanum as a masking agent .<br />

Detection limits (mg/1) for the following<br />

chemicals were : organic nitrogen, 0 .005 ;<br />

nitrate and nitrite, 0 .0005 ; phosphorus ,<br />

0.005; sodium, 0.05 ; potassium, 0 .1 ; calcium ,<br />

0.4; magnesium, 0.05 ; and silica, 0 .1 .<br />

Results<br />

Nutrient Input in Precipitatio n<br />

Total amounts of nutrients dissolved i n<br />

precipitation and the annual rainfall are give n<br />

in table 1 . Nitrogen enters the forest syste m<br />

mainly in organic substances ; nitrate average d<br />

only 13 percent of all nitrogen forms for th e<br />

2-year period. In that period, only trac e<br />

amounts of ammonia nitrogen were detected ,<br />

and a small quantity of nitrite in 1970. The<br />

phosphorus came mainly in organic form in<br />

two collections in the fall of 1969 . Of the<br />

cations, input of calcium was the greates t<br />

followed by sodium, magnesium, and potassium<br />

in descending order . Traces of silica were<br />

found in only five of the 14 collections in<br />

1970 .<br />

There was also an input of particulate<br />

matter from the atmosphere that measure d<br />

upwards of 3 to 5 kg/ha-yr . Although the<br />

average dustfall for 3 years was nearly equally<br />

divided between the dry season (April -<br />

October) and the wet season (November-<br />

March), the proportion in the wet season<br />

varied from 33 to 65 percent of the total fo r<br />

individual years .<br />

Nutrient Loss in Streamflo w<br />

Nitrogen is transported in the stream al -<br />

most entirely in the form of organic matter<br />

(table 2). Though water samples from thi s<br />

watershed were contaminated with nitrat e<br />

from the exhausts of propane heaters, the<br />

nitrate nitrogen outflow from another un-<br />

Table 1.-Nutrient input in precipitation for th e<br />

period October 1 through September 3 0<br />

Constituent in<br />

rainwater<br />

1969 1970<br />

-------------------- hg/ha-yr<br />

NO 3 -N 0 .06 0 .2 0<br />

N0 2 -N 0 .0 1<br />

Organic N 1 .02 .69<br />

All N 1 .08 .9 0<br />

Ortho P ( 1 ) .0 1<br />

Total P ( 1 ) .2 7<br />

Na 1 .17 2 .3 4<br />

K .27 .1 1<br />

Ca 7 .65 2 .3 3<br />

Mg .72 1 .3 2<br />

Si (' ) ( 2 )<br />

Precipitation (cm) 251 215<br />

1 Missing data.<br />

2 Trace .<br />

11 7


Table 2 .-Nutrient budget for dissolved constituents of precipitation an d<br />

runoff for the period October 1 through September 3 0<br />

Element<br />

1969 1970<br />

Input Outflow Net Input Outflow Net<br />

N L08 +0 .50 0 .90 0 .38 +0 .5 2<br />

Total P (') (') 1 (') .27 .52 -.2 5<br />

Na 1 .17 33 .66 -32.49 2 .34 25 .72 -23 .3 8<br />

K .27 1 .24 -.97 .11 2 .25 -2 .1 4<br />

Ca 7 .65 53 .65 -46 .00 2 .33 50 .32 -47 .9 9<br />

Mg .72 12 .70 -11 .98 1 .32 12 .44 -11 .1 2<br />

Si (') (') (') (2 ) 99 .3 -99 . 3<br />

H2 0-cm 251 170 81 215 135 80<br />

' Missing data .<br />

2 Trace .<br />

disturbed watershed nearby was 0 .00 4<br />

kg/ha-yr . A similar nitrogen outflow can be<br />

expected for the study watershed . Ammonium<br />

and nitrite do not occur in measurable<br />

concentrations .<br />

Total phosphorus outflow, both organi c<br />

and inorganic, was about in the same order o f<br />

magnitude as outflow for nitrogen. The<br />

organic form predominates in spring, summer,<br />

and fall seasons but reaches minimum value s<br />

in midwinter when the forest and soils are<br />

repeatedly flushed by rainwater . Orthophosphorus<br />

concentration remained nearl y<br />

constant throughout the year, varying fro m<br />

0.01 to 0 .02 mg/l.<br />

Cation and silica outflows were much large r<br />

than those of nitrogen and phosphorus . Potassium<br />

export was much smaller than that of<br />

other cations . Silica outflow was the largest of<br />

all the dissolved constituents . Transport of<br />

suspended sediment was 67 and 37 kg/ha-y r<br />

for 1969 and 1970, respectively .<br />

Net Loss from the Watershed<br />

The nutrient budget in table 2 results from<br />

the subtraction of the nutrient outflow in the<br />

stream and the input in precipitation . Net<br />

loss or gain is indicated by positive and negative<br />

values, respectively .<br />

The small annual nitrogen gain of about 0 . 5<br />

kg/ha was nearly the same for the 2 years . For<br />

all other mineral constituents of runoff, more<br />

was lost than was gained in precipitation .<br />

Phosphorus losses are relatively small and in<br />

the same order of magnitude as nitroge n<br />

losses. The cation and silica losses were fro m<br />

one to two orders of magnitude larger than<br />

those of nitrogen and phosphorus . Of th e<br />

cations, calcium losses were greatest followe d<br />

in descending order by sodium, magnesium ,<br />

and potassium. Of all constituents, silic a<br />

losses were greatest . Though precipitation an d<br />

runoff were greater in 1969, the use of water<br />

by the forest was nearly equal for the 2 years .<br />

118


Time Trends of Outflow and Concentratio n<br />

of Selected Chemicals<br />

Movement of water in the forest system i s<br />

obviously the dominant factor regulating th e<br />

flow of chemicals . However, moderation o f<br />

the annual chemical outflow pattern can b e<br />

expected from processes such as organi c<br />

decomposition, uptake into plants, an d<br />

mineral weathering . These processes may b e<br />

regulated by the hydrologic and thermal state<br />

of the forest system.<br />

The time trends of soil temperature an d<br />

water content are indicated on figure 1 . Be -<br />

cause precipitation occurs mainly during th e<br />

fall and winter months with less than 2 percent<br />

of the annual total in July and August,<br />

substantial withdrawals of water from the soi l<br />

are made by the forest during the summer .<br />

Water content of the top 30 cm of soi l<br />

reaches minima, near the wilting point ,<br />

usually in late August or early September .<br />

Fall rainstorms normally recharge the soi l<br />

mantle to field capacity by the end o f<br />

December, although recharge of the surfac e<br />

soil may occur from 1 to 2 months earlier .<br />

The soil temperature cycle is the inverse o f<br />

the soil water cycle . Maxima near 14° C occur<br />

in August and minima from 0° to 2° C are<br />

common from December through March .<br />

Precipitation and runoff for 1970 are illustrated<br />

in figure 2 for the periods that composite<br />

streamflow and precipitation samples were<br />

taken. Precipitation exceeded runoff by a<br />

wide margin in the fall of 1969 until soils<br />

reached field capacity sometime in late<br />

December or early January . The soil remained<br />

near field capacity until precipitation fell to<br />

low levels in early May . Wetting fronts from<br />

precipitation and snowmelt can pass through<br />

the soil mantle of the watershed during this<br />

period. Runoff declined rapidly with the in -<br />

crease of evapotranspiration and reduction of<br />

precipitation in the spring and summe r<br />

months .<br />

Organic nitrogen, calcium, and silica wer e<br />

chosen to illustrate the differences that exis t<br />

in annual concentration and outflow patterns<br />

(figs . 3, 4, and 5) . The influence of water ca n<br />

be seen immediately from the general correspondence<br />

of the stream runoff and chemical<br />

outflow cycles . Minimum chemical outflo w<br />

occurs in summer and fall while streamflow i s<br />

low; maximum chemical outflow comes with<br />

high runoff in the winter months. However ,<br />

the concentration patterns vary considerabl y<br />

between the three, and there are notabl e<br />

differences in outflow between organic nitrogen<br />

and the other two (calcium and silica)<br />

during the winter season .<br />

Organic Nitroge n<br />

The high concentration of organic nitrogen<br />

in October 1969 (fig. 3) came at a time when<br />

soils were dry (fig. 1) and could store all<br />

precipitation except that which falls directly<br />

into the stream or runs off rock and shallo w<br />

soil areas adjacent to the stream . After a<br />

decline in early November (fig . 3), the concentration<br />

and outflow rose to a peak in<br />

December when soils probably had reached<br />

field capacity (figs . 1 and 2) and wetting<br />

fronts from precipitation could pass through<br />

the soil mantle . Organic nitrogen outflo w<br />

remained high and reached a secondary peak<br />

in late January at peak runoff (fig . 3). As<br />

precipitation and runoff declined in February ,<br />

concentration and outflow reached low value s<br />

(figs. 2 and 3). The concentration and outflow<br />

rose again in late winter and early sprin g<br />

as precipitation increased (figs . 2 and 3) . Outflow<br />

and concentration declined in April an d<br />

May as precipitation declined and soil wate r<br />

storage dropped below field capacity (figs . 1 ,<br />

2, and 3) . Increased concentration in summe r<br />

had little effect on outflow due to very low<br />

runoff (fig . 3) .<br />

Though the forest system retains mor e<br />

organic nitrogen than it loses (table 2), th e<br />

gain is not uniform throughout the year .<br />

From mid-December to mid-February, th e<br />

forest system lost more in outflow than i t<br />

gained in precipitation; for the remaining 1 0<br />

months, there was a net retention of organi c<br />

nitrogen (fig . 6). For the 2-month midwinte r<br />

period, the outflow in runoff (0 .28 kg/ha)<br />

exceeded the input in precipitation<br />

(0.16 kg/ha) by 0 .12 kg/ha. There was an in -<br />

creasing margin during the 2 months .<br />

Contributions to the stream from the fores t<br />

119


Figure 1 . Water content of the surface 30 cm and temperature at 15 cm in the soil of an old-growth Douglas-fi r<br />

forest.<br />

120


30_<br />

20_<br />

E<br />

U<br />

10-<br />

0<br />

Figure 2 . Precipitation and runoff for an undisturbed Douglas-fir forest for each sampling period, 1969-70 .<br />

121


0.12_ _30<br />

rn<br />

z<br />

0<br />

Q<br />

z<br />

w<br />

0<br />

z<br />

0<br />

0<br />

z<br />

Q<br />

c<br />

o.10_<br />

0.08_<br />

0.06_<br />

rn 0.04_,<br />

-<br />

.-.Outflow<br />

Concentratio n<br />

___ Runoff<br />

_20<br />

Jo<br />

E<br />

U<br />

0<br />

u_ 0.02_<br />

I<br />

D<br />

0<br />

0.00<br />

2 r 0<br />

Figure 3 . Concentration and outflow of organic nitrogen dissolved in stream water and runoff per samplin g<br />

period from an undisturbed Douglas-fir forest, 1969-70.<br />

122


_20<br />

E<br />

U<br />

_10<br />

0<br />

10' '12 '<br />

4- 1969-*<br />

' 2 1 1 4 1<br />

1970<br />

1 6 ' 1 8<br />

0<br />

Figure 4 . Concentration and outflow of calcium dissolved in stream water and runoff per sampling period fro m<br />

an undisturbed Douglas-fir forest, 1969-70 .<br />

123


_3 0<br />

. Outflow<br />

Concentration<br />

_ __ Runoff<br />

E<br />

U<br />

0<br />

Figure 5 . Concentration and loss of silica dissolved in stream water and runoff per sampling period from a n<br />

undisturbed Douglas-fir forest, 1969-70 .<br />

124


Figure 6. Input in precipitation and loss in runoff of dissolved organic nitrogen for each sampling period ,<br />

1969-70 .<br />

125


0.10_<br />

0.08 _<br />

Pea k<br />

0.04_<br />

Concentration in precipitation<br />

0.02<br />

I I I I I I<br />

0.2 0.4 0.6 0.8 1 .0 1 .2<br />

STR EAM FLOW (m 3/ km 2 -sec )<br />

Figure 7 . Concentration of organic nitrogen dissolved in streamfiow during a storm runoff event Novembe r<br />

25-29, 1971 .<br />

126


and precipitation can be illustrated from a se t<br />

of samples taken during a storm runoff even t<br />

in late November 1971 (fig. 7). The rapi d<br />

initial rise of organic nitrogen concentratio n<br />

can be attributed to flushing of availabl e<br />

sources within the vegetation, soil, and atmosphere<br />

. The concentration declined as source s<br />

within the system were depleted and stream -<br />

flow rose by about one order of magnitude t o<br />

peak flow. The concentration rapidly declined<br />

on the recession curve to values slightly larger<br />

than the mean concentration in precipitation .<br />

Calcium<br />

Calcium outflow closely followed the run -<br />

off cycle (fig. 4) . The twofold to threefold<br />

increase in concentration from winter to the<br />

warm seasons had only minimal control o n<br />

calcium outflow compared with the hundredfold<br />

variation in annual runoff. The abrupt<br />

concentration drop in late November and th e<br />

rise in May corresponds approximately to th e<br />

time that the soil reaches field capacity in th e<br />

winter and falls below field capacity in the<br />

spring (fig . 1) .<br />

The calcium and bicarbonate carbon con -<br />

tent of the streamflow follow a similar pat -<br />

tern (fig. 8). Winter season values were nearl y<br />

constant whereas warm season stream con -<br />

tents rose sharply for streamflow values les s<br />

than 0.1 m 3 /km 2 -sec. The change in slop e<br />

(fig. 8) corresponds approximately to th e<br />

change in streamflow regimes (fig . 4) and th e<br />

time that soil water content rises to fiel d<br />

capacity in the winter and departs from it i n<br />

summer (fig . 1) .<br />

Silica<br />

The occurrence of minimum silica concentration<br />

in the dry and warm seasons of th e<br />

year and maxima in the winter and earl y<br />

spring (fig. 5) strongly contrasts calcium concentration-time<br />

trends (fig . 4). The silic a<br />

concentration increase in December an d<br />

January corresponds to the time of expecte d<br />

recharge of the soil to field capacity . Th e<br />

continued elevation of silica concentratio n<br />

through the winter season suggests tha t<br />

soluble silica is either released by weathering<br />

processes at that time or that it accumulated<br />

during the warm season of the year and i s<br />

removed when wetting fronts flush the soil<br />

mantle in midwinter . The reduction of concentration<br />

in late April corresponds to th e<br />

time that subsoils drain as the forest responds<br />

to evaporative demand and precipitation<br />

lessens .<br />

Discussion<br />

Organic nitrogen has seldom been identified<br />

in precipitation . Tarrant et al. (1968)<br />

found values of nitrogen input (1 .49 kg/ha-yr )<br />

similar to those reported here (table 1) and in<br />

the same proportions of organic to nitrat e<br />

nitrogen (87 and 13 percent, respectively) .<br />

Organic nitrogen from 0 .02 to 0.20 mg/1 was<br />

also found in New Zealand snows by Wilson<br />

(1959). Only trace amounts of ammoniu m<br />

were found in the present study and in tha t<br />

by Tarrant et al . (1968). However, Moodie<br />

(1964) reported from 1 to 6 kg/ha-yr of<br />

ammonium nitrogen at several agricultural<br />

sites in western Washington. Values of input<br />

for cations, sodium, potassium, calcium, an d<br />

magnesium (table 1) were generally similar to<br />

those reported by Moodie .<br />

Atmospheric particulate matter may con -<br />

tribute substantially to nitrogen and phosphorus<br />

input . A large proportion of this i s<br />

undoubtedly of local terrestrial origin durin g<br />

the dry season and arises from road dust ,<br />

pollen, and smoke particles . From November<br />

through March, when the local environment i s<br />

regularly dampened by precipitation, particulate<br />

matter is thought to be carried from th e<br />

Pacific Ocean in prevailing westerly winds .<br />

Presumably, the finest particles pass the filter s<br />

we use to separate solids from water and con -<br />

tribute to the "dissolved nutrients ." Nearly al l<br />

the phosphorus and 72 percent of the nitrogen<br />

resulted from a digestion of substances in<br />

rainwater collected during the dry seasons<br />

(table 1) .<br />

This forest system retains nitrogen ver y<br />

effectively . The net gain of 0 .5 kg/ha (table 2 )<br />

is nearly equal to that reported by Cole et al .<br />

(1967). The fact that this result was duplicated<br />

on two Douglas-fir stands, very divers e<br />

127


8J<br />

Calciu m<br />

0<br />

I I I I<br />

0.25 0.50 0.75 1.00<br />

STREAM FLOW (m 3/ K km 2 -sec )<br />

Figure 8 . Concentration of calcium and alkalinity as bicarbonate carbon in streamflow, 1970-72 .<br />

1 28


in terms of soil and stand structure, suggests<br />

that this retention property may be genera l<br />

for Douglas-fir forests west of the Cascade<br />

Range in the Northwestern United States .<br />

Although an explanation for this retentio n<br />

must await the completion of research now in<br />

progress, it appears from the small organic<br />

nitrogen outflow that most of the solubl e<br />

organic substances released from decomposition<br />

of detritus and organism excretions are<br />

rapidly incorporated by the organisms of th e<br />

forest. Nitrate nitrogen outflow is also regulated<br />

to levels much less than the input i n<br />

precipitation. Either nitrification rates are<br />

very low in the soils or the nitrate is retaine d<br />

and rapidly incorporated by the fores t<br />

system. Nitrification is active in the acid soil s<br />

of conifer forests on the Oregon coast (Bolle n<br />

and Lu 1968), but this information is lacking<br />

for soils of the H . J . <strong>Andrews</strong> <strong>Experimental</strong><br />

<strong>Forest</strong> .<br />

The status of nitrogen capital of a Douglas -<br />

fir forest is the result of a number of processes<br />

including input in precipitation, fixation<br />

by free living organisms such as blue-green<br />

algae and soil bacteria, and fixation by micro -<br />

organisms living in symbiotic association wit h<br />

a host plant. If the forest gains 0 .5 kg/ha<br />

annually, as is indicated by this short perio d<br />

of record, 8,000 years would be required t o<br />

accumulate the 4,000 kg/ha of organic nitrogen<br />

found in the more fertile soils of the<br />

study watershed. Additions of nitroge n<br />

symbiotically fixed by Ceanothus velutinu s<br />

may be somewhat larger . If Ceanothu s<br />

velutinus stands persist for 20 years and fi x<br />

20 kg/ha annually as Zavitkovski and Newto n<br />

(1968) suggest, and if generations of thes e<br />

brush stands are induced by fire at intervals o f<br />

200 years, then 16,000 kg/ha of nitrogen<br />

would accumulate in 8,000 years. From this,<br />

deductions must be made for losses by soi l<br />

erosion and microbial and fire volatilization .<br />

In any case, inputs of nitrogen in precipitation<br />

must be considered of importance compared<br />

with other mechanisms of gain and los s<br />

by the forest .<br />

Phosphorus outflow from the forest wa s<br />

similar to that of nitrogen (table 2) . Loss<br />

occurs as both organic phosphorus and orthophosphorus<br />

. The organic form predominates<br />

in the warm seasons of the year when there i s<br />

a rapid biological turnover of phosphorus b y<br />

the forest. Concentration of the ortho form re -<br />

mains fairly constant throughout the year and<br />

is thought to arise from mineral weathering .<br />

The losses of the cations sodium, potassium,<br />

calcium, and magnesium were large<br />

compared with those of nitrogen and phosphorus<br />

and represent excess amounts of thes e<br />

cations to the annual requirements of th e<br />

forest vegetation (table 2) . These losses are<br />

approximately four times those reported fo r<br />

the Hubbard Brook ecosystem in New Hampshire<br />

except for the potassium values whic h<br />

were nearly equal to those reported her e<br />

(Likens et al. 1970). Cole et al. (1967) also<br />

reported lower losses of potassium and calcium.<br />

The cation and silica losses (table 2) are<br />

estimates of the annual release of these chemicals<br />

by mineral weathering . In this regard, it<br />

is interesting to note that the ratio of<br />

Si :Ca:Na:Mg weathering rates for 1970 ar e<br />

8.9 :4.3 :2.1 :1 .<br />

Another mechanism for nutrient loss is th e<br />

particulate matter that is transported by the<br />

stream. This material may arise from soil<br />

erosion or from organic detritus from the<br />

forest. Losses of suspended sediment fro m<br />

this study (67 and 37 kg/ha-yr) were much<br />

lower than the 12-year annual average of 13 3<br />

kg/ha from another undisturbed watershed in<br />

the <strong>Experimental</strong> <strong>Forest</strong> (Fredriksen 1970) .<br />

Although we have not quantified the chemical<br />

loss by this means in this study, previousl y<br />

published annual losses of 0.16 kg/ha for<br />

organic nitrogen represent less than half of<br />

the dissolved component loss (table 2), still a n<br />

important part of the total nitrogen loss<br />

(Fredriksen 1971) . Cation loss was much les s<br />

than 1 percent of the loss in dissolved form .<br />

However, since erosion losses are very sporadic<br />

and dependent on extremes of storm<br />

runoff events, and the results (Fredrikse n<br />

1971) are indicative of the chemical loss during<br />

a relatively quiescent period, they shoul d<br />

be taken as minimum estimates of nutrien t<br />

loss by soil erosion processes .<br />

Biological and chemical processes of th e<br />

forest that control nutrient mobilization an d<br />

retention are undoubtedly closely linked to<br />

environmental factors. The movement of<br />

129


water is dominant because flowing water i s<br />

required to transport materials to the sites of<br />

physical-chemical and biochemical activity<br />

and to remove byproducts that appear in the<br />

outflow of experimental watersheds .<br />

Definite concentration changes in th e<br />

stream occur at the transition periods fro m<br />

drying to wetting of the forest system (figs . 3 ,<br />

4, and 5) . A key feature is either retention o f<br />

precipitation (and reaction byproducts) in soil<br />

storage or the passage of wetting front s<br />

through the soil mantle when the soils are a t<br />

field capacity . It is not surprising that, of th e<br />

annual totals, 77 percent of the calcium, 8 0<br />

percent of the organic nitrogen, and 83 per -<br />

cent of the silica were exported from th e<br />

forest in the 5 months from December 196 9<br />

through April 1970 . During this period, th e<br />

soil mantle was at field capacity and wettin g<br />

fronts from winter rainstorms regularl y<br />

flushed the entire forest system .<br />

For organic nitrogen (fig . 3), the period of<br />

maximum outflow corresponds to the tim e<br />

when wetting fronts begin to pass through the<br />

soil mantle in early December . The widening<br />

margin between input and outflow (fig. 6 )<br />

indicates an increasing contribution to outflow<br />

of organic nitrogen from the forest<br />

system. Although we do not know how muc h<br />

organic nitrogen from precipitation directl y<br />

supplements streamflow, the increased concentration<br />

and content in precipitation an d<br />

streamflow in late January suggest that<br />

sources of organic nitrogen in precipitatio n<br />

may directly augment the outflow from this<br />

experimental catchment .<br />

There is a definite change in the calcium<br />

content of streamflow as the hydrologic state<br />

of the forest system changes from winter to<br />

summer (fig. 8). The calcium outflow was<br />

closely paralleled by the content of carbonate<br />

and bicarbonate-the predominant anions of<br />

the stream . McColl and Cole (1968) demonstrated<br />

that cation mobilization is effectively<br />

controlled by these mobile anions that are<br />

formed by the carbonation of water with<br />

carbon dioxide released by respiring organisms<br />

of the forest . As yet, we cannot explain<br />

the processes that maintain a nearly constant<br />

concentration of calcium and these mobil e<br />

anions over the range of winter streamflow .<br />

Silica is nearly absent in precipitation an d<br />

is therefore entirely generated within the soil -<br />

forest system (fig. 5). Although the paren t<br />

materials of soil are the primary source o f<br />

silica, silica is cycled in generous amounts b y<br />

forests as indicated by litterfall studie s<br />

(Remezov et al. 1955) and the occurrence o f<br />

plant opal crystals in forest soils (Paeth<br />

1970). Silica in streamflow may originate<br />

from primary mineral weathering, secondary<br />

mineral dissolution, or release from decomposition<br />

of detritus . Silica may be taken out o f<br />

solution by formation of secondary mineral s<br />

within the soil . The occurrence of minimu m<br />

silica concentration in the dry and war m<br />

seasons of the year and maxima in the winte r<br />

and early spring strongly contrasts calciu m<br />

concentration-time trends (fig . 4) .<br />

Acknowledgments<br />

The work reported in this paper was sup -<br />

ported by the U .S. <strong>Forest</strong> Service in cooperation<br />

with the Coniferous <strong>Forest</strong> Biome, U.S .<br />

Analysis of Ecosystems, International Biological<br />

Program. This is Contribution No . 29 to th e<br />

Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Bollen, W. B., and K . C. Lu. 1968. Nitroge n<br />

transformations in soils beneath red alde r<br />

and conifers . In J. M. Trappe, J . F. Franklin,<br />

R . F . Tarrant, and G . M. Hansen (eds .) ,<br />

Biology of alder, p. 141-148 . Pac. North -<br />

west <strong>Forest</strong> & Range Exp . Stn., Portland ,<br />

Oreg .<br />

Bormann, F. H., and G. E. Likens . 1967 .<br />

Nutrient cycling . Science 155 : 424-429 .<br />

Cole, D . W., S. P. Gessel, and S . F . Dice .<br />

1967 . Distribution and cycling of nitrogen ,<br />

phosphorus, potassium and calcium in a<br />

second-growth Douglas-fir ecosystem . In<br />

Primary production and mineral cycling in<br />

natural ecosystems, p. 197-232. Univ.<br />

Maine Press .<br />

Fredriksen, R. L. 1969. A battery powered<br />

proportional streamwater sampler . Wate r<br />

Resour. Res . 5: 1410-1413 .<br />

130


1970. Erosion and sedimentation<br />

following road construction and<br />

timber harvest on unstable soils in three<br />

small western Oregon watersheds. USDA<br />

<strong>Forest</strong> Serv. Res. Pap. PNW-104, 15 p . ,<br />

illus. Pac. Northwest <strong>Forest</strong> & Range Exp .<br />

Stn., Portland, Oreg.<br />

1971 . Comparative water quality-natural<br />

and disturbed streams followin g<br />

logging and slash burning . In <strong>Forest</strong> lan d<br />

uses and stream environment, p . 125-137 .<br />

Oreg. State Univ., Corvallis .<br />

Likens, G . E., F . H. Bormann, N . M. Johnson ,<br />

and others . 1970. Effects of forest cuttin g<br />

and herbicide treatment on nutrien t<br />

budgets in the Hubbard Brook Watershedecosystem<br />

. Ecol. Monogr . 40: 23-47 .<br />

McColl, J. G., and D . W . Cole. 1968 . A mechanism<br />

of cation transport in a forest soil .<br />

Northwest Sci . 42 : 134-140 .<br />

Moodie, C . D. 1964 . Nutrient inputs in rain -<br />

fall at nine selected sites in Washington .<br />

Wash . Agric. Exp. Stn., Wash. State Univ.<br />

Paeth, D. C. 1970. Genetic and stability relationships<br />

of four western Cascade soils . 12 6<br />

p. Ph .D . thesis, on file at Oreg . State Univ. ,<br />

Corvallis.<br />

Remezov, N. P., L. N. Bykova, and K. M.<br />

Smirnova. 1955. Nitrogen and mineral<br />

cycles in forests . Akidemiya Nank SSSR.<br />

Trudy Instituta Lesa ., p. 167-194 .<br />

Tarrant, R . F ., K. C. Lu, C . S. Chen, and W.<br />

B. Bollen. 1968. Nitrogen content of precipitation<br />

in a coastal Oregon forest opening.<br />

Tellus XX : 554-556 .<br />

Wilson, A . T. 1959. Organic nitrogen in New<br />

Zealand snows . Nature 183 : 318-319 .<br />

Zavitkovski, J ., and M . Newton . 1968. Ecological<br />

importance of snowbrush ,<br />

Ceanothus velutinus, in the Oregon Cascades<br />

. Ecology 49 : 1134-1145 .<br />

131


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Nutrient cycling in throughf all<br />

and litterfall in 450-year-old<br />

Douglasfir stands<br />

Albert Abee<br />

an d<br />

Denis Lavender<br />

School of <strong>Forest</strong>ry<br />

Oregon State University<br />

Corvallis, Orego n<br />

Abstract<br />

Comparisons of nutrient concentrations (N, P, K+, Ca ++ , Mg++) found in canopy throughfall and litterfall<br />

were made on the H. J. <strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong>. Six old-growth Douglas-fir (Pseudotsuga menziesii )<br />

stands were studied which represented six forest communities common to the western Cascades ofOregon. These<br />

community types span a large portion of the temperature and moisture gradients present in the area. The<br />

preliminary data indicate that nutrient concentration in throughfall was highest during the summer and fall, and<br />

lowest during the winter. Nutrient input through throughfall generally followed the same trends . Nutrient<br />

return through litterfall was greatest in the needles. More amounts of N, P, and Ca++ were transferred to the soil<br />

through litterfall than through throughfall, while more K + and Mg++ were added to the soil through throughfall.<br />

Litterfall was maximum during the winter. Future studies will correlate the results from the nutrient analysis to<br />

the moisture and temperature gradients .<br />

Introduction<br />

The worldwide interest of scientists in litterfall<br />

production during the past century, has<br />

been shown by Bray and Gorham (1964) i n<br />

their review of litter production in the forest s<br />

of the world. Methodology reports ranged<br />

from utilization of randomly located collection<br />

devices of varied design, separation, oven -<br />

drying, and chemical analysis of several litter<br />

components, to merely raking up and air drying<br />

the litter on a unit area basis. In spite of<br />

the large number of papers cited in the abov e<br />

review, data of litter production from natural ,<br />

old-growth ecosystems are meager. Even less<br />

is known about litterfall in old-growth<br />

Douglas-fir (Pseudotsuga menziesii) forest<br />

types. The examination of seasonal fluctuations,<br />

nutrient concentration changes associated<br />

with defoliation, and nutrient composition<br />

of various litterfall categories are scarce<br />

(Kira and Shidei 1967) .<br />

The first published report of an investigation<br />

of litterfall in coniferous forests of the<br />

Pacific Northwest is that of Tarrant, Isaac ,<br />

and Chandler (1951). These workers collected<br />

the litter of several species for 1 year and estimated<br />

nutrient movement by multiplyin g<br />

litter weight by the percent elemental conten t<br />

of foliage collected from trees, an inexact procedure<br />

. More detailed measurements of th e<br />

nutrient cycle in Douglas-fir forests have bee n<br />

published for stands in New Zealand (Wil l<br />

1959) and the United States (Dimock 1958) .<br />

In addition, workers at both the University o f<br />

Washington (Rahman 1964) and Oregon Stat e<br />

University 1 have collected substantial dat a<br />

describing litterfall in both managed an d<br />

natural Douglas-fir stands . Riekerk and Gessel<br />

(1965) and Cole and Gessel (1968) summariz e<br />

1 D . P . Lavender, unpublished data .<br />

133


a number of very sophisticated studies o f<br />

nutrient movement through Douglas-fir ecosystems<br />

in Washington .<br />

All of the above studies, save that of Tar -<br />

rant et al ., however, were concerned with<br />

litterfall and nutrient movement through relatively<br />

young stands .<br />

Several studies (LeClerc and Breazeal e<br />

1908, Mes 1954, Tukey and Amling 1958 ,<br />

and Tukey et al . 1958) have demonstrated<br />

that rainfall may remove substantial quantities<br />

of nutrient elements from the foliage of<br />

horticultural plants . Similarly, studies of th e<br />

elemental content of precipitation under<br />

forest stands (Tamm 1951, Madgwick an d<br />

Ovington 1959, Will 1959, and Voigt 1960 )<br />

have demonstrated that rainwater which ha s<br />

passed through tree crowns ( "throughfall")<br />

contains significantly higher quantities o f<br />

many nutrient elements than rainfall collected<br />

in adjacent openings .<br />

In the Pacific Northwest, studies reporte d<br />

by Rahman (1964), Tarrant et al . (1968) and<br />

Cole and Gessel (1968) have yielded data<br />

which describe the movement of nutrient s<br />

from the atmosphere and tree crowns to th e<br />

forest floor by precipitation . Finally, unpublished<br />

data by Lavender describe the<br />

movement of nutrients from the crowns of<br />

both fertilized and control second-growth<br />

Douglas-fir stands to the forest floor by<br />

precipitation .<br />

The purpose of the present study was to<br />

measure the movement of nutrients in canop y<br />

throughfall and litterfall in several associatio n<br />

types of old-growth Douglas-fir stands . These<br />

community types were selected to represent<br />

the range of environments occurring on the H .<br />

J. <strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong> and are also<br />

indigenous to the Pacific Northwest. This<br />

paper will report on the results of our efforts<br />

to date .<br />

Study Area<br />

The H . J. <strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong> en -<br />

compasses 15,000 acres and is characterize d<br />

by steep topography with approximately one -<br />

fifth of its land area in gentle slopes o r<br />

benches. Elevations within the forest vary<br />

from 457 m to more than 1,523 m . Precipitation<br />

is heavy, varying from 226 cm per year a t<br />

lower elevations to as much as 356 cm per year<br />

along the highest ridges. A considerable snowpack<br />

develops on the higher slopes while rai n<br />

predominates at the lower elevations. Mean<br />

temperatures within the forest range fro m<br />

35°F in January to 65°F in midsummer<br />

(Berntsen and Rothacher 1959) .<br />

Methods<br />

Within the <strong>Experimental</strong> <strong>Forest</strong>, six communities<br />

were chosen (table 1), each name d<br />

for characteristic plants in both the overstory<br />

and understory . The six old-growth communities<br />

are presented in order of increasing<br />

elevation. Each of the six plots are 0 .2024<br />

hectare in size and are equipped with eight<br />

litter traps ; each is 2,601 cm 2 in area, located<br />

on a random basis in each plot . Litter was<br />

collected every 4 to 6 weeks during the snow -<br />

free months of 1970-71 . Heavy snow pack<br />

prevented litter collection during much of th e<br />

winter of 1970 . Therefore, data describing<br />

nutrient movement in the litter for this perio d<br />

are weak because : (1) the necessarily infrequent<br />

collections do not permit accurat e<br />

assessment of the rate of litterfall, and (2) lit -<br />

ter which remained in the traps for lon g<br />

periods was subjected to leaching . The following<br />

fall, three litter traps on each plot wer e<br />

equipped with a 113-liter reservoir to collec t<br />

the precipitation which passed over the litter .<br />

Analysis of this water will provide a measure<br />

of the nutrients leached from the litter .<br />

Crown and stem maps were made for each<br />

plot to aid in evaluating the variation of litter -<br />

fall between traps .<br />

After collection the litter for each trap wa s<br />

dried at 70°C, separated into classes (needles ,<br />

cones, twigs, branches, hardwoods, bark,<br />

lichens and mosses), and weighed. Prior to<br />

chemical analyses, litter from the eight litter<br />

traps per plot was composited into tw o<br />

samples, each representing four traps . In addition,<br />

some consolidation of litter collected on<br />

different dates was necessary . Each sample<br />

was analyzed to determine the levels of nitrogen,<br />

phosphorus, potassium, calcium an d<br />

134


Table 1.-Characteristics of study plot s<br />

Plot<br />

Elevation<br />

PPT/<br />

year<br />

Percent species<br />

composition )<br />

Diameter<br />

range<br />

Average<br />

d .b.h .<br />

Basa l<br />

area 2<br />

Stems /<br />

hectare<br />

meters centimeters - - - centimeters- - -<br />

1 . Pseudotsuga- 457 211 .8 Psme 89 .3 11 - 163 46 .7 83 .06 27 7<br />

Holodiscus Tsme 5 . 4<br />

Tabr 3 . 6<br />

Acci 1 . 7<br />

2 . Tsuga- 488 228 .3 Psme 17 .0 8 - 139 37 .8 97 .55 494<br />

Rhododendron- Tsme 69 . 0<br />

Berberis Tabr 8 . 0<br />

Thpl 5 . 0<br />

Conu 1 . 0<br />

3 . Tsuga- 762 230 .6 Psme 26 .5 10 - 213 68 .1 120 .87 168<br />

Polystichum Tsme 61 . 8<br />

Tabr 8 . 8<br />

Thpl 2 . 9<br />

4 . Tsuga - 610 229 .6 Psme 52 .8 8 - 157 34 .0 69 .77 45 0<br />

Rhododendron - Tsme 23 . 1<br />

Gaultheria Tabr 1 . 1<br />

Thpl 19 . 8<br />

Cach 2 . 2<br />

Conu 1 . 1<br />

5 . Tsuga-Abies- 975 Psme 16 .1 8 - 173 66 .3 129 .60 27 7<br />

Linneae Tsme 46 . 4<br />

Thpl 33 . 9<br />

Tabr 3 . 6<br />

6 . Abies- 1,311 Psme 46 .3 8 - 117 55 .4 109 .63 33 1<br />

Tiarella Tsme 16 . 4<br />

Abam<br />

Abpr<br />

37 .3<br />

1 Psme Pseudo tsuga menziesi i<br />

Tsme Tsuga mertensiana<br />

Tabr Taxus brevifolia<br />

Acci Acer circina turn<br />

Thpl Thuja plicat a<br />

Conu Cornus nutalli i<br />

Cach Castanopsis chrysophylla<br />

Abam Abies amabilis<br />

Abpr Abies procera<br />

2 Square meters/hectare .<br />

135


magnesium.<br />

In addition to the litter traps, each plot was<br />

equipped with- four 20-inch-high rain gages .<br />

Each gage was assigned to one of 20 rando m<br />

locations within the plot after each collection ,<br />

in accordance with a method described b y<br />

Wilm (1943). Higher elevation plots als o<br />

contain a rain gage on a platform 10 feet fro m<br />

ground level to provide a water sample durin g<br />

months of heavy snow cover . Water was collected<br />

and the volume measured at approximately<br />

2-week intervals . The samples wer e<br />

returned to the laboratory on the day of<br />

collection, filtered, and stored at -12°C unti l<br />

thawed for analysis. For analysis, the fou r<br />

samples per plot were combined into tw o<br />

samples and analyzed for total potassium ,<br />

calcium, magnesium, orthophosphate, an d<br />

total phosphorus. Nitrogen in the form of<br />

ammonium, nitrate, nitrite, and organic nitrogen<br />

was also determined .<br />

Precautions were taken to keep contamination<br />

of water samples to a minimum . Funnel s<br />

with glass wool stoppers were provided for<br />

each rain gage to keep organic matter fro m<br />

contaminating water samples . Mercuric chloride<br />

was added to the rain gages in the summer<br />

and fall to keep microorganism activity<br />

to a minimum . The cold temperature helpe d<br />

to reduce microorganism and insect activit y<br />

during the winter .<br />

Several techniques were investigated in an<br />

effort to arrive at a measure of crown density .<br />

Basal area and volume poorly describe intercepting<br />

crown cover in old-growth defectiv e<br />

stands as canopy development tends to re -<br />

main constant after trees reach maturity ;<br />

hence, direct estimates were used . Photo -<br />

graphs which were taken above each samplin g<br />

point with a 35-mm camera were shown over<br />

a spherical dot grid to give a means of comparing<br />

crown densities .<br />

(fig. 1). Water sample concentrations were<br />

highest during the summer when precipitatio n<br />

was minimal (table 2) . Concentrations were<br />

lowest during the winter when precipitatio n<br />

was highest. As precipitation decreased fro m<br />

winter to spring, concentration of throughfal l<br />

samples for each element increased . Throughfall<br />

concentrations were also high during th e<br />

fall when precipitation first starts .<br />

Unlike N, the average total input of P ,<br />

Mg+, Ca++, and K+ generally follows the<br />

same trend as did the concentration curves<br />

(fig. 2). The greatest amount of each element<br />

was leached out during the fall when precipitation<br />

first washes the canopy . Decreasin g<br />

amounts were leached out with increasin g<br />

precipitation . Potassium input reached a lo w<br />

point during the winter, at which time 68 per -<br />

cent of the total precipitation had fallen, an d<br />

z<br />

0<br />

I-<br />

2<br />

CC<br />

I-<br />

z<br />

w<br />

U<br />

z<br />

0<br />

U<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Results and Discussion<br />

Throughfall Results<br />

Nutrient concentration in throughfall<br />

samples for plots 1 through 4 for all elements<br />

appeared to be the same during each season<br />

FALL<br />

Figure 1 . Average concentration of plots 1 through 4<br />

for each element by season .<br />

0<br />

136


Z<br />

w<br />

2<br />

w<br />

J<br />

w<br />

3<br />

0<br />

Table 2. Average total precipitation increased sharply from winter to spring ,<br />

across all plots by season slightly decreasing from spring to summer .<br />

Season<br />

Precipitation<br />

Calcium and P reached low points during th e<br />

spring, at which time 92 percent of the tota l<br />

pre pitation had allen, and increased fro m<br />

spring to summer. Magnesium input was great -<br />

centimeters<br />

est from fall to winter and remained approximately<br />

the same from winter to summer.<br />

Fall<br />

70.89<br />

Nitrogen input slightly increased from fall to<br />

winter, decreasing from winter to sprin g<br />

Winter<br />

92 .0 5<br />

reaching a low point during the summer .<br />

There appears to be no difference in term s<br />

Spring<br />

57.9 6<br />

of net kg per hectare per year between plot s<br />

1 to 4 for each element with the exception o f<br />

Summer<br />

18 .0 8<br />

plot 3 (table 3) . Plot 3 had more K + and les s<br />

Ca++ than plots 1, 2, and 4 .<br />

Total 238 .98<br />

100<br />

8 0<br />

20<br />

Figure 2. Total average kg per hectare of plots 1<br />

through 4 for each element by season .<br />

Throughfall Discussion<br />

In general, the total nutrient input an d<br />

throughfall concentrations were highest in th e<br />

summer and fall and lowest during the winte r<br />

and spring months. This seems to indicate<br />

that each tree or canopy has a constant fraction<br />

of elements which can be removed fro m<br />

the foliage through leaching elements . Once<br />

the rains start in the fall, the majority of eac h<br />

nutrient is leached out . As the rains increase<br />

in quantity and duration during the winter<br />

• and spring months, the available fraction of<br />

nutrients is further depleted. Decreasing<br />

precipitation from spring to the end of summer<br />

allows the nutrient fraction to increase<br />

•<br />

again until the total fraction of leachabl e<br />

nutrients is reached .<br />

Variations found between plot 3 and plot s<br />

1, 2, and 4 with respect to K + and Ca++ coul d<br />

be due to differences in soil types (data no t<br />

available yet) . If soil types are different with<br />

respect to nutrient availability, the difference s<br />

between plots could be explained by luxur y<br />

consumption .<br />

Another possible source of the nitroge n<br />

found in the throughfall samples is nitrogen -<br />

fixing bacteria . Jones (1970) in his study o f<br />

nitrogen fixation by bacteria in the phyllo -<br />

sphere of Douglas-fir (Pseudotsuga douglasii)<br />

in England isolated bacteria from the leaf<br />

surfaces of Douglas-fir. He found that th e<br />

bacteria could fix atmospheric nitrogen whe n<br />

provided with a carbohydrate source . The fat e<br />

13 7


Table 3 .-Net kg per hectare per year of nutrients collected in throughfall gages<br />

++ Mg++<br />

Item N P K + Ca<br />

Input from atmosphere ) 1 .298 0 .232 0 .106 2 .085 1 .27 3<br />

Throughfall input :<br />

Plot 1<br />

3 .999 2 .308 17 .416 5 .983 2.608<br />

2 2 .979 2 .398 15 .749 4 .438 2 .08 6<br />

3 3 .729 2 .970 30 .350 2 .134 1 .45 6<br />

4 2 .710 3.283 23 .379 5 .104 2 .34 3<br />

Average 3 .354 2 .740 21 .724 4 .416 2 .123<br />

1 Fredriksen unpublished data-data collected from open area on the H . J . <strong>Andrews</strong> <strong>Experimental</strong> Fores t<br />

at 610 meters .<br />

Table 4 .-Distribution of metric tons/hectare between litter components by plo t<br />

Plot 1 2 3 4 5a 5b 1 6 Average 2<br />

Needles 2 .002 2 .246 2 .950 3 .200 2 .533 2 .533 3 .741 2 .77 7<br />

Percent of total 32 .84 35 .56 46 .44 62 .30 15 .17 55 .91 54 .10 47 .1 5<br />

Reproductiv e<br />

structures .834 1 .141 1 .284 .742 .536 .536 .518 .74 3<br />

Percent of total 13 .68 18 .06 20.22 14 .46 3 .20 11 .83 7 .49 14 .3 1<br />

Wood material 2 .280 2 .760 1 .876 1 .009 13 .479 1 .267 2 .524 1 .95 3<br />

Percent of total 37 .40 43 .68 29 .53 19 .66 80 .74 27 .96 36 .50 33.1 4<br />

Hardwoods and<br />

mosses 1 .022 .175 .401 .197 .195 .195 .206 .36 5<br />

Percent of total 16 .77 2 .77 6 .32 3 .84 1 .17 4 .30 2 .98 6 .2 0<br />

Total 6 .138 6 .317 6 .512 5 .131 16 .694 4.530 6 .916 5 .89 1<br />

Note : In plot 5, an extremely large slab of bark from a nearby snag fell into a trap causing high values for tota l<br />

tons/hectare . Over a longer period of time, this type of variation between litter components can be expected to<br />

occur randomly throughout each plot . However, due to the limited sampling time thus far recorded, the one<br />

extreme value will be temporarily ignored .<br />

1 Excluding extreme bark sample .<br />

2 Excluding 5a .<br />

138


of the nitrogen was not determined. However ,<br />

Jones suggested that it could be washed to th e<br />

ground .<br />

Litterfall Dat a<br />

Despite the differences in stand characteristics<br />

shown in table 1, little variation in total<br />

litter production was found among stands fo r<br />

the year 1970-71 (table 4) . Average yearl y<br />

litterfall production for all plots was 5 .8 9<br />

metric tons/hectare . This is approximately 1 %<br />

times the average 3 .5 metric tons per hectare<br />

reported by Bray and Gorham (1964) fo r<br />

cool, temperate forests, but closer to the yiel d<br />

they reported for a latitude comparable t o<br />

their study area (fig . 1). From worldwid e<br />

data, these authors reported that nonlea f<br />

litter averaged from 27 percent to 31 percent<br />

of total litter production . The stands reported<br />

here averaged 47 percent nonleaf (woody )<br />

litter for the 1-year period .<br />

In terms of total kg/hectare of litter, the<br />

vast majority fell during the winter (fig . 3) .<br />

This is the period when snowfall is greatest ,<br />

consequently much litter breaks under th e<br />

1600<br />

1400<br />

1200<br />

weight of the snow . Needle cast was greatest<br />

in the fall, decreasing during the winter, an d<br />

gradually increasing during spring . Hardwoo d<br />

and moss litter was greatest in the fall, de -<br />

creasing throughout the rest of the year .<br />

Woody material and cone litterfall was greatest<br />

during the winter.<br />

Nutrient concentration of litterfall components<br />

varied considerably among plots an d<br />

seasons (table 5). Plots 1 and 3 were chosen<br />

to represent the range of values that can be<br />

found in nutrient return through litterfall .<br />

There appears to be no consistent trend b y<br />

season or plot for the litter component concentrations.<br />

However, average yearly concentrations<br />

of each nutrient for each litter clas s<br />

are comparable between plots 1 and 3 . There<br />

is a substantial difference between tota l<br />

kg/hectare for N, P, K +, and Ca++ betwee n<br />

plots (table 6). Plot 1 had more kg of Ca ++<br />

per hectare than did plot 3 . Plot 3 had greate r<br />

amounts of N, P, and K + than did plot 1 .<br />

Little difference occurred among plots fo r<br />

Mg++<br />

The greatest portion of nutrient inpu t<br />

through litterfall came in the needles (tabl e<br />

7). Needle litterfall contributed about 54 per -<br />

cent of the total nutrient input . Cone litterfall<br />

accounted for 13 percent while twig litterfall<br />

accounted for 11 percent of the total . Together,<br />

the needle, cone, and twig litterfall<br />

account for 78 percent of the total nutrien t<br />

input through litterfall .<br />

Litterfall Discussio n<br />

40 0<br />

200<br />

Figure 3 . Average kg per hectare by season and litter<br />

component .<br />

Variations in nutrient concentration foun d<br />

between litterfall components among plot s<br />

and season can be expected if foliage characteristics<br />

such as age and species are no t<br />

constant . We also observed that the age of th e<br />

tissue, and when it falls, varies throughout th e<br />

year for each plot. This is primarily due to<br />

environmental parameters such as win d<br />

action, rainstorms, and snowfall . Difference s<br />

in soil types could also have affected concentrations<br />

.<br />

Differences between total nutrient inpu t<br />

through litterfall (table 6) are affected by th e<br />

distribution of litter components within th e<br />

total. Where two plots seem to produce corn -<br />

139


Table 5.-Average percent of N, P, K +, Ca ++, and Mg++ by season, plot, and litter componen t<br />

Litte r<br />

component<br />

and season<br />

Plot 1 Plot 3<br />

N P K + Ca ++ Mg++ N P K+ Ca ++ Mg+ +<br />

Needles :<br />

Fall 0 .368 0 .087 0 .109 1 .901 0 .017 0 .434 0 .120 0 .161 1 .815 0 .02 0<br />

Winter .691 .108 .210 1 .305 .018 .763 .124 .280 1 .103 .08 1<br />

Spring .462 .124 .159 1 .241 .028 .717 .119 .103 1 .112 .01 7<br />

Summer .467 .143 .200 1 .740 .034 .474 .105 .145 1 .286 .02 5<br />

Average .497 .115 .169 1 .546 .024 .597 .117 .172 1 .329 .02 0<br />

Cones :<br />

Fall .530 .084 .122 .397 .045 .518 .063. .099 .149 .01 5<br />

Winter .294 .029 .046 .146 .016 .387 .033 .051 .148 .01 1<br />

Spring .488 .049 .133 .214 .041 .544 .064 .102 .178 .01 1<br />

Summer .487 .068 .151 .369 .027 .561 .063 .159 .206 .01 6<br />

Average .449 .057 .105 .281 .032 .502 .055 .103 .170 .01 3<br />

Twigs :<br />

Fall .340 .033 .051 .852 .009 .434 .051 .124 1 .107 .01 6<br />

Winter .375 .040 .057 1 .261 .011 .398 .032 .075 1 .110 .01 2<br />

Spring .424 .072 .076 .999 .081 .358 .047 .046 .823 .00 8<br />

Summer _ .408 .043 .105 1 .034 .013 .363 .055 .102 1 .054 .01 5<br />

Average .386 .047 .072 1,036 .013 .388 .046 .086 1 .023 .01 3<br />

Branches :<br />

Fall .218 .017 .030 .713 .008 .207 .013 .025 .522 .00 6<br />

Winter .028 .017 .080 .598 .006 .296 .039 .082 .843 .01 0<br />

Sprin g<br />

Summer .102 .008 .030 .338 .00 5<br />

Average .213 .017 .055 .655 .007 .201 .020 .046 .567 .00 7<br />

Bark :<br />

Fall .332 .047 .076 .566 .014 .417 .038 .074 .517 .01 0<br />

Winter .404 .033 .050 .963 .010 .431 .035 .057 .413 .009<br />

Spring - _ .476 .080 .055 .300 .00 9<br />

Summer .320 .030 .051 .637 .011 .541 .056 .163 .474 .01 2<br />

Average .352 .036 .059 .722 .011 .466 .052 .087 .426 .01 0<br />

Hardwoods :<br />

Fall .630 .095 .167 2 .395 .037 .591 .124 .475 2 .428 .06 1<br />

Winte r<br />

Sprin g<br />

Summer .546 .092 .302 2 .127 .07 0<br />

Average .588 .094 .234 2 .261 .053 .591 .124 .475 2 .428 .06 1<br />

Mosses and lichens :<br />

Fall .417 .053 .117 .408 .016 .659 .126 .229 .413 .01 9<br />

Winter 1 .264 .100 .132 .567 .015 1 .234 .122 .350 .329 .01 7<br />

Spring - - -- _ - - - -<br />

Summer 1 .313 .100 .246 .376 .020 1 .427 .130 .284 .364 .01 8<br />

Average .998 .084 .165 .450 .017 1 .106 .126 .287 .368 .018<br />

140


Table 6. Total kg per hectare per year for plots 1 and 3 for each element by litter class<br />

Litter class<br />

Plot 1 Plot 3<br />

N P K+ Ca ++ Mg++ N P K + Ca ++ Mg+ +<br />

Needles 7 .93 2 .27 3 .16 35 .89 0 .47 15 .61 3 .82 5 .89 44 .08 0 .6 5<br />

Percent of total 36.12 58 .85 49 .54 50 .22 44 .61 47 .76 68 .20 60 .25 69 .90 59 .7 9<br />

Cones 3 .46 .45 .83 2 .06 2 .13 6 .60 .69 1 .33 2 .05 .1 7<br />

Percent of total 15 .74 11 .59 13 .03 2 .89 20 .31 20 .23 12 .40 13 .63 3 .24 15 .4 6<br />

Branches 1 .97 .16 .74 5 .09 .06 .82 .12 .29 2 .15 .0 2<br />

Percent of total 8 .98 4 .06 11 .56 7 .13 5 .28 2 .50 2 .00 2 .97 3 .40 2 .0 6<br />

Twigs 3 .52 .38 .66 12 .01 .10 3 .54 .34 .75 9 .56 .1 0<br />

Percent of total 16.07 9.86 10 .33 16 .79 9 .50 10 .82 6 .20 7 .67 15 .15 9 .2 7<br />

Bark 1 .62 .11 .07 3 .56 .30 3 .00 .28 .49 3 .18 .0 6<br />

Percent of total 7 .39 3 .17 3.08 4 .97 2 .92 9 .18 5 .00 5 .04 5 .03 5 .1 5<br />

Hardwoods 2 .33 .39 .67 12 .48 .17 .38 .06 .11 1 .51 .0 5<br />

Percent of total 10 .60 9 .17 10 .49 17 .46 16 .06 1 .16 1 .00 1 .14 2 .39 4 .1 2<br />

Moss and lichens 1 .12 .08 .12 .39 .01 2 .73 .28 .90 .54 .0 3<br />

Percent of total 5 .10 2 .11 1 .98 .54 1 .32 8 .35 5 .00 9 .16 .85 3 .0 9<br />

Total 21 .95 3 .86 6 .39 71 .49 1 .07 32 .68 5 .59 9 .77 63 .06 1 .09<br />

Table 7.-Average percent of litterfall totals for plots 1 and 3 for each element by litter class<br />

Litter class N P K + Ca ++ Mg++ Average<br />

Needles 41 .9 63.5 54 .9 60 .1 52 .2 54 .5 2<br />

Cones 18 .0 12 .0 13 .3 3 .1 17 .9 12.8 6<br />

Branches 5 .7 3 .0 7 .3 5 .3 3 .7 5 .0 0<br />

Twigs 13 .4 8 .0 9 .0 16 .0 9 .4 11 .1 6<br />

Bark 8 .3 4 .1 4 .1 5 .0 4 .0 5 .1 0<br />

Hardwoods 5 .9 5 .1 5 .8 9 .9 10 .1 7 .3 6<br />

Moss and lichens 6 .7 3 .6 5 .6 .7 2 .2 3 .7 6<br />

r<br />

14 1


parable total quantities of litter on a yearly<br />

basis, amounts of the various litter components<br />

are important . The amounts of each<br />

litter component are important becaus e<br />

concentrations of nutrient elements vary fo r<br />

each litter component (see table 5) . Table 4<br />

shows that the greatest differences betwee n<br />

plots 1 and 3 in terms of kg/hectare of litter -<br />

fall occur between needles and woody material.<br />

Plot 3 produced 947 kg/hectare more of<br />

needles than plot 1, while plot 1 produced<br />

403 kg/hectare more of woody material tha n<br />

plot 3. However, table 5 shows that th e<br />

average concentration of nitrogen, fo r<br />

example, is much higher in needles than it is<br />

in woody material ; consequently, variations<br />

such as found in table 6 are brought about .<br />

Nutrient input in both throughfall and litterfall<br />

appears to have the same trend for K + ,<br />

P, and Ca++ when comparing plots 1 and 3 .<br />

The litterfall analysis shows plot 3 havin g<br />

more K +, P, and less Ca ++ than plot 1 . Th e<br />

throughfall data for plots 1 and 3 show th e<br />

same results (table 3) . However, differences<br />

between N and Mg++ input through throughfall<br />

and litterfall for each plot do not agree .<br />

Table 3 shows little difference in N input fo r<br />

each plot, while table 6 shows N being highe r<br />

in plot 3 . A possible explanation for this ha s<br />

been given in the preceding paragraph . Tabl e<br />

3 shows that plot 1 had more Mg ++ input<br />

than plot 3 . Litterfall input for Mg ++ was<br />

about the same between plot 1 and 3 . A<br />

possible explanation for the difference between<br />

plots 1 and 3 for Mg++ in throughfall<br />

might be that the hardwood litterfall ac -<br />

counted for 10.1 percent of the total Mg ++<br />

input (table 7) . Table 1 shows that 1 .7 per -<br />

cent of the species composition in plot 1 wa s<br />

hardwoods, while plot 3 shows no hardwoods .<br />

More amounts of N, P, and Ca ++ were<br />

transferred to the soil through litterfall tha n<br />

through throughfall ; while more K + and Mg ++<br />

were added to the soil through throughfal l<br />

(table 8) .<br />

There appeared to be no relationship between<br />

nutrient concentration and elevation .<br />

Rather, concentration on each plot seemed t o<br />

be correlated with crown density . Basal are a<br />

and crown density in old-growth Douglas-fi r<br />

stands seem to be distributed randomly ove r<br />

the sites on the H . J. <strong>Andrews</strong> <strong>Experimental</strong><br />

<strong>Forest</strong>. Consequently, no real correlatio n<br />

could be seen between basal area and crow n<br />

density with moisture and elevation . Thi s<br />

could also be due to the fact that the range o f<br />

environments sampled on the H . J. <strong>Andrews</strong><br />

<strong>Experimental</strong> <strong>Forest</strong> was not great enough ,<br />

indicated by total precipitation . Table 1<br />

shows that there was no real difference i n<br />

total precipitation between plots 1 through 4 .<br />

Summary<br />

The preliminary data indicate that nutrien t<br />

concentration in throughfall varied with sea -<br />

son . Highest concentrations were found in th e<br />

summer and lowest concentrations during th e<br />

winter. Nutrient input through throughfal l<br />

generally followed the same trends as di d<br />

nutrient concentrations . Nutrient return<br />

through litterfall was greatest in the needles .<br />

Together, the needles, twigs, and cones ac -<br />

counted for 78 percent of the nutrient inpu t<br />

Table 8.-Total nutrient input in kg per hectare per yea r<br />

Input N P K + Ca ++ Mg++<br />

Average throughfall input 3 .3544 2 .740 21 .7230 4 .416 2.12 3<br />

Average litterfall input 27 .3235 4 .725 8 .0784 67 .2738 1 .08 0<br />

Total 30 .6779 7 .465 29 .8014 71 .6888 3 .203<br />

142


through litterfall . More amounts of N, P, and<br />

Ca++ were transferred to the soil throug h<br />

litterfall than through throughfall, while mor e<br />

K+ and Mg++ were added to the soil through<br />

throughfall. Litterfall was maximum during<br />

the winter .<br />

Acknowledgments<br />

The work reported in this paper was supported<br />

in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferous<br />

<strong>Forest</strong> Biome, U .S. Analysis of Ecosystems ,<br />

International Biological Program . This is Contribution<br />

No. 30 to the Coniferous Fores t<br />

Biome .<br />

Literature Cited<br />

Berntsen, C. M., and J . Rothacher . 1959 . A<br />

guide to the H . J. <strong>Andrews</strong> Experimenta l<br />

<strong>Forest</strong>. USDA <strong>Forest</strong> Serv . Pac. Northwes t<br />

<strong>Forest</strong> & Range Exp . Stn ., 21 p . Portland ,<br />

Oreg .<br />

Bray, J. Roger, and Eville Gorham . 1964 .<br />

Litter production in forests of the world .<br />

Advan. Ecol. Res . 2: 101-157 .<br />

Cole, Dale W ., and Stanley P. Gessel. 1968 .<br />

Cedar River Research . A program for<br />

studying the pathways, rates, and processes<br />

of elemental cycling in a forest ecosystem .<br />

Univ . Wash., Coll . <strong>Forest</strong> Res ., Inst. <strong>Forest</strong><br />

Prod. Contrib. No . 4, 54 p .<br />

Dimock, E . J. 1958 . Litter fall in a young<br />

stand of Douglas-fir. Northwest Sci . 32 :<br />

19-29 .<br />

Jones, K. 1970. Nitrogen fixation in th e<br />

phyllosphere of the Douglas-fir, (Pseudotsuga<br />

douglasii) . Ann. Bot. (London) 34 :<br />

239-244 .<br />

Kira, T ., and T. Shidei . 1967 . Primary production<br />

and turnover of organic matter in different<br />

forest ecosystems of the wester n<br />

Pacific. Jap . J. Ecol. 17(2) : 70-87 .<br />

Le Clerc, J. A., and J. F. Breazeale . 1908 .<br />

Plant food removed from growing plants by<br />

rain or dew . In J . A. Arnold (ed.), U .S .<br />

Department of Agriculture yearbook, p .<br />

389-402 .<br />

Madgwick, H. A. I., and J. D. Ovington . 1959 .<br />

The chemical composition of precipitatio n<br />

in adjacent forest and open plots . <strong>Forest</strong>ry<br />

32(1) : 14-22 .<br />

Mes, Margaretha G . 1954. Excretion (recretion)<br />

of phosphorus and other mineral elements<br />

by leaves under the influence of rain .<br />

S. Afr . J . Sci. 50(7) : 167-172 .<br />

Rahman, Abu Hamed Mohammed Mojibur .<br />

1964 . A study of the movement of elements<br />

from leaf crowns by natural litter -<br />

fall, stemflow and leaf wash . 119 p . M.F .<br />

thesis on file, Univ . Wash ., Seattle .<br />

Riekerk, Hans, and Stanley P . Gessel. 1965 .<br />

Mineral cycling in a Douglas-fir fores t<br />

stand . Health Phys . 11 : 1363-1369 .<br />

Rothacher, Jack. 1963 . Net precipitation<br />

under a Douglas-fir forest . <strong>Forest</strong> Sci . 9(4) :<br />

423-429 .<br />

Tamm, Carl O. 1951. Removal of plant nutrients<br />

from tree crowns by rain . Physiol .<br />

Plant. (4): 184-188 .<br />

Tarrant, R . F., Leo A. Isaac, and Robert F .<br />

Chandler, Jr . 1951. Observations on litte r<br />

fall and foliage nutrient content of som e<br />

Pacific Northwest tree species . J. For .<br />

49(12) : 914-915 .<br />

, K . C. Lu, C . S. Chen, and W . B .<br />

Bollen. 1968. Nitrogen content of precipitation<br />

in a coastal Oregon forest opening .<br />

Tellus 20(3) : 554-556 .<br />

Tukey, H . B., Jr., and H. J. Amling . 1958 .<br />

Leaching of foliage by rain and dew as an<br />

explanation of differences in the nutrien t<br />

composition of greenhouse and field-grown<br />

plants. Mich. Quart. Bull. 40(4) : 876-881 .<br />

, H. B. Tukey, and S. H. Wittwer.<br />

1958. Loss of nutrients by folia r<br />

leaching as determined by radioisotopes .<br />

Proc. Am. Soc. Hortic. Sci. 71 : 496-506 .<br />

Voigt, G. K. 1960. Alternation of the composition<br />

of rainwater by trees. Am . Midland<br />

Nat . 63(2) : 321-326 .<br />

Will, G . M. 1959. Nutrient return in litter an d<br />

rainfall under some exotic conifer stands i n<br />

New Zealand . New Zealand J. Agric. Res.<br />

2 : 719-734 .<br />

Wilm, H . G. 1943. Determining net rainfal l<br />

under a conifer forest . J. Agric. Res. 67 :<br />

501-512 .<br />

143


Estimating Biomass<br />

and Other State Variables<br />

145


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 1972 .<br />

Direct, nondestructive measurement of<br />

biomass and structure in living,<br />

old-growth Douglas-fir<br />

William C . Denison,<br />

Diane M . Tracy ,<br />

Frederick M . Rhoade s<br />

Department of Botan y<br />

and Plant Pathology<br />

Oregon State University<br />

Corvallis, Orego n<br />

Martha Sherwood<br />

Department of Biolog y<br />

University of Orego n<br />

Eugene, Orego n<br />

Abstract<br />

Previous studies of biomass and structure of Douglas-fir have examined trees less than 100 years old. This<br />

paper describes methods for measuring older trees, illustrated by data from a tree 60 m tall and 450 years old.<br />

Rock climbing techniques, modified for use on trees, are employed to climb the main trunk . A movable spa r<br />

provides access to lateral branches. The trunk is measured, the position of each branch system is located on it ,<br />

and the branch systems are scored for 10 variables related to biomass and structure. An importance value,<br />

calculated for each branch system, is used in selecting a set of branch systems for detailed measurement. The<br />

data permit diagramatic reconstruction of the tree, or estimates of the distribution or total amount of component<br />

parts .<br />

Introduction<br />

The techniques described in this paper are<br />

of two kinds ; technical climbing methods an d<br />

methods of tree description and measuremen t<br />

which depend upon technical climbing to pro -<br />

vide access to the top of the tree . In combination,<br />

these techniques are designed to provid e<br />

quantitative descriptions of the biomass, surface<br />

area, and spatial distribution of th e<br />

aboveground parts of individual trees .<br />

The climbing techniques developed from a<br />

National Science Foundation-Undergraduate<br />

Research Participation study of epiphytes o n<br />

old-growth Douglas-fir (Pseudotsuga<br />

menziesii) . Previous studies (Coleman, Muenscher,<br />

and Charles 1956; Hoffman and Kazmierski<br />

1969) were limited to epiphytes<br />

within 2 m of the ground, but this study wa s<br />

designed to study the canopy populations a s<br />

well. An initial attempt to use felled trees wa s<br />

unsuccessful because the surface of the trun k<br />

which hit the ground was destroyed and th e<br />

branch systems with their epiphytes wer e<br />

shattered and scattered . The climbing techniques<br />

were developed as an alternative an d<br />

proved to be practical, effective, and<br />

economical .<br />

Five old-growth Douglas-firs have bee n<br />

rigged and climbed to date, but only one has<br />

been subjected to the measurement and analysis<br />

described herein. Previous studies of biomass<br />

in this species (Burger 1935 ; Newbol d<br />

1967 ; Reukema 1961) have been limited t o<br />

trees less than 100 years old, but the tree s<br />

used in our study are 450 years old and large ;<br />

147


60-80 m tall, 1-1 .5-m dbh, with their lowest<br />

branches 12-25 m above the ground .<br />

Climbing trees of this height is dangerous .<br />

A fall would probably be fatal . But probably<br />

there is greater danger of injury to personnel<br />

under the tree in the event equipment or part s<br />

of branch systems are accidentally dropped<br />

upon them . We have worked through two sea -<br />

sons without injury .<br />

The climbing techniques are described in<br />

detail to permit others to adopt them . Th e<br />

rock-climbing techniques on which they are<br />

based are described in books on mountaineering<br />

(Blackshaw 1970 ; Manning 1967) but may<br />

be unfamiliar to biologists .<br />

The access technique, the methods o f<br />

description and measurement, and one application<br />

of the data, a diagramatic reconstruction<br />

of a tree, are treated in this paper .<br />

Another paper in this symposium (Pike et al .<br />

1972) describes the estimation of tree surfac e<br />

area and epiphyte biomass .<br />

Riggin g<br />

Rigging involves an initial ascent and preparation<br />

of the tree for subsequent climbing .<br />

The ascent is slow, and requires exceptiona l<br />

agility and endurance . A team of three experienced<br />

climbers should rig the tallest trees i n<br />

a day and a half .<br />

The procedure used in ascending the tre e<br />

and placing climbing and belay ropes is outlined<br />

in figures 1-4 .<br />

1<br />

Figure 1 . A lag screw is used to fasten a steel hanger<br />

to the tree .<br />

The Access Techniques<br />

Access to the tree involves three steps : rigging,<br />

climbing, and use of the spar, eac h<br />

described below. Rigging and climbing techniques<br />

are modified from direct-aid rock -<br />

climbing techniques . Basic safety procedures ,<br />

equipment, terminology, and philosophy hav e<br />

been adopted from mountaineering . Anyone<br />

without prior mountaineering experience who<br />

expects to adopt these techniques should consult<br />

one of the standard texts (Blackshaw<br />

1970; Manning 1967) . The specialized rope<br />

and hardware may be obtained from either<br />

mountaineering or yachting suppliers .<br />

Basic equipment for all climbers includes :<br />

hard hat, heavy climbing shoes, and a harnes s<br />

of nylon webbing to which a belay rope i s<br />

attached. The climber is always belayed-tha t<br />

is, she is protected by a safety rope held b y<br />

another experienced climber . Climbing an d<br />

belay ropes are 11 mm nylon "goldline" an d<br />

the webbing used in making slings, stirrups ,<br />

etc . is of nylon and 40-50 mm wide .<br />

Figure 2 . A carabiner is clipped into the hanger .<br />

148


Figure 3. Two additional carabiners are used to faste n<br />

two climbing stirrups to the first carabiner. (Only<br />

the upper ends of the webbing stirrups are shown . )<br />

The belay rope is clipped through the carabiner .<br />

5<br />

Figure 4 . The climber ascends the climbing stirrup s<br />

until the first carabiner is at her waist . Tension on<br />

the belay rope holds her in position while sh e<br />

drives the next lag screw .<br />

6<br />

Figure 5 . The climbing rope is fastened to the tree b y<br />

four lag screws, hangers, and carabiners . A loop, Figure 6 . The belay rope runs through a block hun g<br />

tied near the end of the rope, is connected to the by heavy chain with welded links. The chain i s<br />

carabiners by two loops of webbing. The free end fastened around the tree above a limb and secure d<br />

of the rope is tied around the tree using a bowline . by two 1-cm bolts with washers and lock nuts .<br />

149


Lag screws, 8 by 150 mm, are used t o<br />

attach equipment. They do not sag or "work<br />

out" of the thick, soft bark of Douglas-fir as<br />

nails do . The screws are driven with a hammer<br />

but may be tightened or removed with a<br />

wrench .<br />

A climbing rope is attached near the top of<br />

the tree and subsequent ascents are made on<br />

this fixed rope . Figure 5 illustrates the usual<br />

method of attachment.<br />

The block for the belay line is attached<br />

below the point at which the climb path i s<br />

obstructed by branches. In climbing abov e<br />

this point the belay rope is carried up throug h<br />

carabiners placed during rigging . Below th e<br />

block, on the open trunk, the climber is belayed<br />

from above through the block . Figure 6<br />

illustrates the method of attachment for th e<br />

block .<br />

Climbing<br />

Climbing requires mastery of unfamiliar<br />

skills, but no greater strength or agility tha n<br />

climbing a ladder, a 70-m ladder . An experienced<br />

climber, in good physical condition ,<br />

can climb a 75-m tree in 20 minutes or less ,<br />

depending upon the distribution of branche s<br />

and the lean of the trunk. A clumsy, middleaged<br />

man can clamber up with equal safety ,<br />

but it takes longer .<br />

The techniques involved in climbing th e<br />

fixed rope are illustrated in figures 7-13 .<br />

7<br />

8 9<br />

Figure 7 . A band of rubber cut from a motorcycle<br />

innertube is fastened around the sling and over the<br />

toe of the boot to prevent the climber from accidentally<br />

stepping out of the sling .<br />

Figure 8 . When weight is placed on the jumar, it s<br />

toothed gate grips the rope so that the jumar ca n<br />

move neither up nor down .<br />

Figure 9 . When there is no weight on the jumar, it s<br />

gate may be lowered and the jumar moved up or<br />

down the rope .<br />

Figure 10. The jumar cannot be removed from th e<br />

rope unless the safety catch is depressed to permit<br />

the gate to open fully .<br />

150


1 1<br />

Figure 11 . The climber places her weight on on e<br />

foot, thus locking the jumar on that side, while sh e<br />

moves the other jumar up or down . She ascends or<br />

descends by shifting her weight from one foot t o<br />

the other and moving the opposing jumars. Sh e<br />

stands in webbing slings which hang from th e<br />

jumars . The jumars are connected to each other by<br />

a short length of rope which passes through a carabiner<br />

clipped to her belay harness .<br />

Figures 12 and 13. At those points at which th e<br />

climbing rope is fastened by a carabiner, th e<br />

jumars must be disconnected from the rope, one at<br />

a time, and moved around the carabiner .<br />

Most of the carabiners used in rigging ar e<br />

removed once the tree is rigged, but a few ar e<br />

left in place to prevent the climbing rop e<br />

from hanging away from the trunk .<br />

Shouted instructions from climber to belayer<br />

are normally adequate during climbing ,<br />

but once the climber enters the canopy conversation<br />

is severely impeded by distance an d<br />

intervening foliage . A pair of small radi o<br />

transceivers facilitates longer communication s<br />

and permits the belayer to record data dictated<br />

by the climber .<br />

The Spar<br />

Use of a special boom, hereafter called "the<br />

spar," permits the climber to reach any point<br />

within 4 m of the trunk .<br />

The spar is specially constructed of three<br />

lengths of mast-grade spruce laminated in a<br />

"tee" cross section. It is 4 m long and together<br />

with its hardware and rigging it weighs<br />

15 kg. It is awkward to maneuver into place ,<br />

especially where there are branches close<br />

above the one to be sampled. Under normal<br />

151


14<br />

Figure 14 . The inboard end of the spar. The hinge is<br />

of steel, 8 mm thick except for the central block ,<br />

which is 36 by 36 mm . The top of the spar is a<br />

single piece of spruce, 22 mm high, 75 mm wide ,<br />

and 4 m long . The bottom is of two pieces, eac h<br />

22 mm thick, 100 mm high, and 4 m long, glued<br />

face to face .<br />

Figure 15. The outboard end of the spar. Note that<br />

the supporting ropes are not fastened to the out -<br />

board end . They pass through blocks and guides<br />

on opposite sides of the spar and are secured at th e<br />

hinge end .<br />

circumstances it should be possible to raise<br />

the spar from the ground and have it ready<br />

for use in 2 hours or less ; in difficult circumstances<br />

it may take a full day .<br />

Figures 14-17 illustrate the constructio n<br />

and operation of the spar .<br />

The ropes used to support the outboard<br />

end of the spar are of nylon, but, unlike thos e<br />

used in climbing, are of a special braid whic h<br />

minimizes stretching (Samson Yachtbraid) .<br />

This prevents the spar from sagging out of<br />

position as the weight of the climber moves<br />

away from the tree .<br />

Note that the ropes supporting the spar ar e<br />

fastened at the inboard end of the spar, not a t<br />

the outboard end . The position at which th e<br />

spar rests depends upon the relative tensio n<br />

on these ropes . Thus a climber on the spar<br />

may maneuver herself through 180° of arc ,<br />

without returning to the trunk, by pulling on<br />

one or the other of the ropes running along<br />

the spar .<br />

The climber sits in a "swing seat" suspended<br />

by loops of webbing . A climbing stirrup<br />

permits her to move along the spar b y<br />

alternately sitting in the seat while she move s<br />

the stirrup and standing in the stirrup whil e<br />

she moves the seat .<br />

In moving the spar, the ropes are coiled an d<br />

the hinge is folded back along the spar an d<br />

fastened . In placing it on the tree, it is hung ,<br />

hinge downward, and maneuvered until th e<br />

hinge is in place . The hinge is fastened to th e<br />

tree with lag screws . Then the outboard end i s<br />

lowered into place. In removing the spar the<br />

process is reversed .<br />

The use of these three techniques in combination<br />

enables the climber to work with<br />

comparative freedom and safety for hours at a<br />

time if need be .<br />

152


Figure 16. The spar in use. The suspending ropes pass<br />

upward through carabiners on opposite sides of<br />

the tree and are tied off near the inboard end of<br />

the spar . The spar is shown in the open but woul d<br />

normally be placed adjacent to a branch system .<br />

Figure 17 . Raising the spar . A "trolley" of light<br />

nylon line is stretched from a clear area on th e<br />

trunk to an open area on the ground . The spar is<br />

packaged, with its ropes coiled and hinge folded ,<br />

and clipped to the "trolley" with carabiners . A<br />

handline and pulley aid the climber in pulling th e<br />

spar up into the tree .<br />

Tree Description<br />

and Measurement<br />

Description and measurement of a tree proceeds<br />

in two steps : an initial survey, in whic h<br />

the trunk and all of the branches are de -<br />

scribed, followed by detailed measurement o f<br />

a sample set of branch systems . The tree i s<br />

arbitrarily subdivided into components : th e<br />

main trunk plus a number of branch systems .<br />

A branch system is defined as one or mor e<br />

branches, living or dead, arising from the same<br />

point on the trunk. Several branch system s<br />

may arise at the same height, but at differen t<br />

points around the circumference of the trunk .<br />

For examples of kinds of data recorded and<br />

calculated, the reader should refer to table 1<br />

while reading the following sections .<br />

Survey<br />

As the tree is climbed, the trunk is marked<br />

off vertically in meters, by use of tape. At<br />

convenient intervals, 5 to 10 m, a benchmar k<br />

is established and its height verified by transit<br />

from the ground. The diameter and inclination<br />

of the trunk are recorded at the benchmarks.<br />

It is not safe to climb above the poin t<br />

where the trunk is 10 cm or less in diameter ,<br />

so we have treated the portion above tha t<br />

point as an exceptional branch system . When ,<br />

through damage to the original leader, one or<br />

more secondary leaders have developed, the y<br />

are treated in the same fashion as the main<br />

trunk .<br />

153


Each branch system is numbered. The position<br />

of each branch system on the trunk i s<br />

recorded by height and compass quadrant .<br />

Each branch system is scored for 10 variables.<br />

Most of these variables relate to the<br />

structure and biomass of the tree, but sinc e<br />

this study began as a survey of epiphyte distribution,<br />

three variables relate to epiphyte load .<br />

1. Number of Main Axes<br />

A "main axis" is defined as a branch more<br />

than 4 cm in diameter, originating within 1<br />

dm of the trunk. Six classes are recognized :<br />

no axes (i .e., no branch more than 4 cm i n<br />

diameter), one axis, two axes, three axes, fou r<br />

axes, and five or more axes .<br />

2. Extension from Trun k<br />

The extension is measured as the horizonta l<br />

distance from the trunk to the furthest tip o f<br />

any part of the branch system . Five classes are<br />

recognized : 0-1 m, 1-3 m, 3-5 m, 5-10 m, and<br />

more than 10 m .<br />

3. Total Length of Living Axe s<br />

An "axis" is any branch more than 4 cm i n<br />

diameter. The total estimated here is of th e<br />

length of all living axes within the branc h<br />

system . Six classes are recognized : 0-1 m, 1-5<br />

m, 5-10 m, 10-15 m, and more than 20 m .<br />

4. Total Length of Dead Axe s<br />

The definition and classes are as in 3 above ,<br />

except that here the branches are dead .<br />

5. Area of Branchlets and Foliag e<br />

An estimate is made of the area of a n<br />

imaginary figure formed by connecting th e<br />

outer tips of all branchlets with the point o f<br />

attachment to the trunk and projecting th e<br />

outline on a horizontal plane . Six classes ar e<br />

recognized : 0-1 m 2 , 1-2 m 2 , 2-5 m2 , 5-10 m 2 ,<br />

10-20 m 2 , and more than 20 m 2 .<br />

6. Density of Branchlets and Foliage<br />

Within the area delimited in 5 above, a n<br />

estimate is made of the area occupied b y<br />

branchlets and foliage . Six classes, expressed<br />

as percents of the total area, are recognized : 0<br />

percent (no foliage), 0-20 percent, 20-40 percent,<br />

40-60 percent, 60-80 percent, an d<br />

80-100 percent .<br />

7. Maximum Diameter of Axi s<br />

The maximum diameter of the largest axi s<br />

is measured to the nearest centimeter .<br />

8. Lichen Cover : All Specie s<br />

The lichen cover is estimated for all axes in<br />

the branch system . Only foliose and fruticos e<br />

species are counted . The cover is expressed a s<br />

a percent of the total surface area on all axes .<br />

Six classes are recognized : 0 percent (n o<br />

lichens), 0-20 percent, 20-40 percent, 40-60<br />

percent, 60-80 percent, and 80-100 percent .<br />

9. Lichen Cover : Lobaria species<br />

The cover is estimated and expressed for<br />

Lobaria species alone, using the same classes<br />

listed in 8 above .<br />

10. Bryophyte Cover : All Specie s<br />

The cover is estimated and expressed for al l<br />

bryophytes using the same classes listed in 8<br />

above .<br />

The data from this survey are tabulated ,<br />

with the branch systems arranged by number ,<br />

in the order in which they occur on the tree .<br />

A sample group of branch systems is the n<br />

selected for detailed measurement .<br />

Sample Selectio n<br />

The number of branch systems selected fo r<br />

detailed measurement (n) depends upon th e<br />

complexity of the tree : two branch system s<br />

are selected in a small, uniform tree ; three, i n<br />

a normal tree ; four or five, in a tree with more<br />

than one leader . In our example, five branc h<br />

systems were selected and measured .<br />

An importance value (v) is calculated fo r<br />

each branch system according to the following<br />

formula :<br />

where ,<br />

v = b 2 +b 1 2 + ad + 1 + m<br />

b = class number (1-6) of total lengt h<br />

of living axes<br />

b 1 class number (1-6) of total lengt h<br />

of dead axe s<br />

a = class number (1-6) of area of<br />

branchlets and foliage<br />

d = class number (1-6) of density o f<br />

branchlets and foliage<br />

l = class number (1-6) of lichen cover ,<br />

all specie s<br />

m = class number (1-6) of bryophyte cove r<br />

The importance value is listed for each<br />

branch system, and both running totals and a<br />

grand total (V) are calculated . The grand total<br />

(2,111 in our example) is set equal to the<br />

number of branch systems to be selected (five<br />

154


in our example) and the running totals recalculated<br />

. For example, branch number 87 has<br />

an uncorrected running total of 28 . The corrected<br />

total is: 28(5/2,111)=0 .066. After thi s<br />

correction, each branch system is represented<br />

by an interval, proportional to its importance<br />

value, which is the difference between<br />

its corrected running total and that of<br />

the preceding branch system .<br />

A random number is drawn : a three-place<br />

fraction of 1 .000 . In our example, the number<br />

drawn was 0 .870. Those branch systems<br />

are selected which have intervals which include<br />

: the random number, 1 + the random<br />

number, 2 + the random number, etc . In ou r<br />

example, the numbers 0 .87, 1.870, 2 .870 ,<br />

3 .870, and 4 .870 corresponded to branc h<br />

system numbers 7, 33, 61, 93, and 116 .<br />

The branch systems selected in this way ar e<br />

measured in detail as described below .<br />

Branch System Measuremen t<br />

The measurement of each selected branc h<br />

system is carried out in 10 stages . The first<br />

four stages occur in the field and require the<br />

use of the spar ; the remaining stages occur i n<br />

the laboratory .<br />

1. Each axis is marked off in decimeter<br />

lengths and the total length of the main axe s<br />

is recorded .<br />

2. The diameter is recorded every 4 dm .<br />

3. The transition points (4-cm diam) a t<br />

which an axis becomes a branchlet are<br />

identified .<br />

4. At one-fourth of the transition point s<br />

the distal portion, consisting of branchlets<br />

and attached foliage, is cut off and lowered t o<br />

the ground for further study .<br />

5. The annual rings are counted at the cu t<br />

end of the removed branch and the age of the<br />

sample is recorded .<br />

6. The foliage-bearing portion of each<br />

sample is cut into lengths bearing a singl e<br />

year's needles .<br />

7. A subsample of 20 of each year 's<br />

needles is measured (length and width) while<br />

fresh .<br />

8. The foliage and attached fragments of<br />

branchlet are dried at 85° and weighed . Th e<br />

weight of each year's foliage is recorded, as is<br />

the weight of its associated segments o f<br />

branchlet .<br />

9. The older branchlets, those without<br />

foliage, are cut into lengths representin g<br />

approximately 1 year 's growth and their<br />

lengths and diameters are recorded .<br />

10. The segments of older branchlet are<br />

dried at 85° and weighed .<br />

Analysi s<br />

Several types of information can be summarized<br />

from data accumulated by the techniques<br />

described . Two of these, surface are a<br />

of the trunk and limbs and epiphyte biomass ,<br />

are described in a paper by Pike et al .(1972) .<br />

Analyses of wood and foliage biomass ar e<br />

being developed, but at present our results ar e<br />

incomplete since they are based on a singl e<br />

tree .<br />

Another kind of analysis results in a diagram<br />

of the tree (fig . 18) ; a graphic summarization,<br />

to scale, of the structure of an individual<br />

tree . This diagram is an east-wes t<br />

vertical section . It serves both as a map fo r<br />

those working on the tree and as a chart of th e<br />

distribution of branch systems and foliage .<br />

Reconstruction of the trunk is based upon<br />

benchmark measurements of diameter, height ,<br />

and inclination. A second leader, originatin g<br />

on the north side of the main trunk at th e<br />

40-m level, is drawn to the same scale but<br />

displaced to the right . Only those branc h<br />

systems are shown which project to the eas t<br />

or to the west. From existing data one could<br />

draw a comparable north-south section displaying<br />

the remaining branch systems . Eac h<br />

branch system is shown in its correct positio n<br />

on the trunk and the length of the longes t<br />

horizontal line indicates, to scale, the distanc e<br />

the entire branch system projected away fro m<br />

the trunk . However, the arrangement of axes<br />

and foliage within a branch system is symbolic<br />

rather than pictorial .<br />

Living branch systems are represented by a<br />

parallelogram, representing foliage, usuall y<br />

accompanied by one or more solid lines ,<br />

representing axes . Parallelograms without<br />

solid lines are branch systems in which n o<br />

branch was more than 4 cm in diameter .<br />

155


156<br />

Figure 18 . Semidiagrammatic reconstruction of Tree Number 1, Watershed 10 .<br />

Only those branch systems extending to the east or west are shown . Th e<br />

second trunk is drawn to scale at the right height, but is displaced to the right .<br />

Solid horizontal lines represent living branches (axes) or branch systems<br />

more than 4 cm in diameter . The length of the line indicates, to scale, how fa r<br />

the whole branch system extends from the trunk . Dead branches are represented<br />

by dashed lines . The parallelograms represent branchlets and foliage .<br />

The length of a parallelogram represents the horizontal area covered by th e<br />

branch system, and the height represents the density of branchlets an d<br />

foliage within that area . Parallelograms without horizontal lines are branc h<br />

systems with no branches more than 4 cm in diameter . Numbered branc h<br />

systems are those selected for detailed measurement .


Dashed lines are dead branches . The area of a<br />

foliage parallelogram is a measure of the<br />

amount of foliage within the branch system .<br />

The length of a parallelogram is a function of<br />

the area of the branch system ; the height of a<br />

parallelogram is a function of the density of<br />

foliage within the branch system .<br />

Individual branch systems are easily identified.<br />

For example, those branch systems<br />

which were selected for detailed measuremen t<br />

(numbers 7, 33, 61, 93, and 116) are identified<br />

by number on the diagram .<br />

Discussion<br />

Although the access technique is still being<br />

improved, it is now very nearly routine fo r<br />

trees with a growth pattern similar t o<br />

Douglas-fir . It could undoubtedly be modified<br />

for use in measuring large trees with different<br />

growth patterns . It would be particularly<br />

interesting to attempt to apply the method to<br />

trees of the upper canopy in lowland tropica l<br />

wet forests .<br />

We have not had adequate opportunity t o<br />

evaluate the accuracy of our methods of<br />

description and measurement . The basic pat -<br />

tern seems satisfactory, but there will b e<br />

changes made before we extend it to additional<br />

trees . We will increase the precision o f<br />

initial estimates of lengths of axes . We need<br />

better methods for estimating foliage biomass .<br />

The importance value used in selecting branc h<br />

systems combined values for foliage biomas s<br />

with values for epiphyte load . This resulted i n<br />

selection of a subset that was a poor compromise<br />

between the conflicting requirements fo r<br />

measurement of tree biomass and measurement<br />

of epiphyte biomass . In the future ,<br />

separate importance values will be calculate d<br />

for tree biomass and epiphyte biomass ; an d<br />

separate sets of branch systems will be selected<br />

and measured .<br />

Acknowledgments<br />

The authors are grateful to many people<br />

who contributed to the development of thi s<br />

project . Don Kirkpatrick taught us the ascent<br />

technique and made the first ascent . Othe r<br />

climbers (Diane Nielsen, Tom Denison, Jan e<br />

McCauley, and Karen Berliner), working wit h<br />

the authors, have helped in data taking an d<br />

have made improvements in climbin g<br />

methods. Dr. Jack Culver (Benton Boa t<br />

Works) suggested improvements in the desig n<br />

of the spar, and James Ewanowski (Arborea l<br />

Constructions) built it. Sue Carpenter helpe d<br />

with the drawings . Dr. Scott Overton suggested<br />

the basic format of the sampling pattern<br />

and has guided its development .<br />

This project began (July-August 1970) as a<br />

student project under the National Scienc e<br />

Foundation Undergraduate Research Participation<br />

Program (Grant Number GY-7641) in<br />

the Department of Botany and Plant Pathology,<br />

Oregon State University . Subsequent<br />

work has been supported by National Science<br />

Foundation Grant Number GB-20963 to the<br />

Coniferous <strong>Forest</strong> Biome, U .S . Analysis o f<br />

Ecosystems, International Biological Program .<br />

This is Contribution No . 31 from the Coniferous<br />

<strong>Forest</strong> Biome .<br />

Literature Cited<br />

Blackshaw, A . 1970 . Mountaineering . 552 p .<br />

Hammondsworth : Penguin Books .<br />

Burger, H . 1935 . Holz, Blattmenge and<br />

Zuwachs. II. Die Douglasie. Mitt. Schweiz .<br />

Anst. forstl . Vers. 19 : 21-72 .<br />

Coleman, B . B., W. C. Muenscher, and D . R .<br />

Charles . 1956 . A distributional study of th e<br />

epiphytic plants of the Olympic Peninsula ,<br />

Washington . Am. Midland Nat . 56 : 54-87 .<br />

Hartley, H. O. 1966. Systematic samplin g<br />

with unequal probability and without re -<br />

placement. J. Am. Statist . Assoc . 61 :<br />

739-748 .<br />

Hoffman, G. R., and R . G. Kazmierski . 1969 .<br />

An ecological study of the epiphytic bryophytes<br />

on Pseudotsuga menziesii on the<br />

Olympic Peninsula, Washington. I . A<br />

description of the vegetation . Bryologist<br />

72 : 1-18 .<br />

Manning, H. (ed.), 1967. Mountaineering the<br />

freedom of the hills . 2d ed . 485 p . Seattle :<br />

The Mountaineers .<br />

157


Newbold, P . J. 1967 . Methods for estimatin g<br />

the primary production of forests . I .P.P .<br />

Handbook #2 . 62 p. Oxford, England :<br />

Blackwell Sci . Publ .<br />

Pike, Lawrence H., Diane M . Tracy, Martha A .<br />

Sherwood, and Diane Nielsen . 1972. Estimates<br />

of biomass and fixed nitrogen o f<br />

epiphytes from old-growth Douglas-fir . In<br />

Jerry F. Franklin, L . J . Dempster, an d<br />

Richard Waring (eds .), Proceedings-researc h<br />

on coniferous forest ecosystems-a symposium,<br />

p. 177-187, illus . Pac . Northwest<br />

<strong>Forest</strong> & Range Exp . Stn ., Portland, Oreg .<br />

Reukema, D . L. 1961 . Crown developmen t<br />

and its effect on stem growth of six<br />

Douglas-firs . J. For. 59 : 370-371 .<br />

Table 1 .-Exemplary data for branch systems on tree #1-a Douglas-fi r<br />

on Watershed 10, H . J. <strong>Andrews</strong> <strong>Experimental</strong> Fores t<br />

c<br />

w<br />

N<br />

so<br />

S<br />

O0<br />

m<br />

m o<br />

EO sO+ .,<br />

V Q<br />

.w<br />

G<br />

o ~<br />

•d S+<br />

C F<br />

Total Lengt h<br />

of Branches<br />

E al<br />

E ~<br />

Lichen Cove r<br />

All<br />

Specie s<br />

-% -<br />

Lobari a<br />

-% -<br />

w E<br />

x<br />

o<br />

il w<br />

-m-<br />

x m<br />

Liv e<br />

-m -<br />

Dea d<br />

-m -<br />

N• m<br />

-cm-<br />

-% -<br />

w<br />

-c LL H<br />

O 0<br />

0 0<br />

WI V<br />

-%-<br />

88 59 .1 TOP<br />

87 59 .1 S<br />

86 58 .8 W<br />

85 58 .5 E<br />

84 58 .2 W<br />

83 58 .2 N<br />

82 58 .2 S<br />

81 57 .6 S<br />

22 41 .1 W<br />

21 40 .8 W<br />

20 40 .2 W<br />

19 39 .3 W<br />

18 38 .4 W<br />

17 38 .1 W<br />

2 20 .4 N<br />

1 15 .9 N<br />

128 49 .8 W<br />

127 49 .5 N<br />

93 40 .5 E<br />

92 40 .5 N<br />

91 40 .2 N<br />

90 40 .2 N<br />

89 39 .9 N<br />

1 1-3 1-5 30 2-5 10-20 40-60 40-60 15 1 5<br />

1 1-3 1-5 5 1-2 20-40 0-20 0-20 13 2 8<br />

1 1-3 1-5 6 1-2 0-20 0-20 0-20 11 3 9<br />

1 0-1 0-1 5 0-1 0-20 0-20 5 44<br />

1 1-3 1-5 5 1-2 020 0-20 0-20 11 5 5<br />

1 1-3 1-5 8 0-1 0-20 0-20 0-20 9 6 4<br />

1 0--1 0-1 5 0-1 0-20 0-20 0-20 6 7 0<br />

2 1-3 1-5 6 2-5 20-40 0-20 0-20 0-20 17 8 7<br />

1 1-3 1-5 9 0-1 40-60 0-20 0-20 11 94 9<br />

1 0-1 0-1 6 0-1 0-20 5 95 4<br />

1 1-3 1- 5<br />

8 0- 1<br />

0-20 12 96 6<br />

At this point the second trunk leaves the tree, 40m, N Compas s<br />

2 3-5 5-10 7 5-10 60-80 0-20 0-20 0-20 33 99 9<br />

1 5-10 5-10 8 2-5 20-40 20-40 0-20 22 102 1<br />

1 0-1 0-1 3 0-1 80-100 9 103 0<br />

5 3-5 10-15 8 10-20 40-60 0-20 0-20 20-40 41 137 8<br />

4 3-5 10-15 5-10 10 10-20 60-80 0-20 20-40 55 143 3<br />

Data for trunk 2 which left the tree between branch 19 & 2 0<br />

1 3-5 1-5 6 1-2 0-20 0-20 0-20 0-20 12 1445<br />

1 3-5 1-5 6 1-2 0-20 0-20 0-20 11 L45 6<br />

1 3-5 1-5 12 5-10 20-40 0-20 0-20 20-40 21 206 2<br />

1 1-3 1-5 9 0-1 20-40 0-20 20-40 11 207 3<br />

1 0-1 0-1 9 0-1 0-20 0-20 0-20 6 207 9<br />

1 3-5 5-10 11 2-5 20-40 0-20 0-20 20-40 23 210 2<br />

1 1-3 1-5 7 0-1 0-20 0-20 0-20 9 211 1<br />

158


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 1972<br />

Estimation of biomass and transpiratio n<br />

in coniferous forests using<br />

tritiated water<br />

Abstract<br />

J . R . Kline,<br />

M . L . Stewart,<br />

C . F . Jorda n<br />

Radiological Physics Divisio n<br />

Argonne National Laboratory<br />

Argonne, Illinois 60439<br />

Nondestructive measurement of biomass and transpiration rates in trees is dependent on a new application of<br />

the established theory of tracer dynamics in steady state systems. The method utilizes HTO as a tracer for H2 0<br />

in the plant. Both transpiration and biomass measurements require experimental determination of tritiu m<br />

activity in the tree as a function of time. Biomass measurement requires one additional parameter-the mean<br />

residence time of water in the plant. In this paper we examine various theoretical and experimental alternative s<br />

for determining mean residence times. Data supporting one alternative is presented. The measurement of<br />

conducting tissue biomass is also discussed .<br />

Introduction<br />

Kinetic theory and tracer distributio n<br />

methods have been used extensively in measurement<br />

of the properties of flowing system s<br />

both in biology and engineering . The theoretical<br />

aspects and the underlying assumption s<br />

governing the practical application of tracer<br />

dynamics have been thoroughly discussed by<br />

authors in many fields. Reports by Zierler<br />

(1964), Bergner (1961, 1964a, 1964b, 1965 ,<br />

1966), and Ljunggren (1967) have describe d<br />

the use of tracer dynamics and kinetic theory<br />

in the nondestructive measurement of flo w<br />

rates, mean residence times, and compartmental<br />

volumes of steady state biological an d<br />

engineering systems .<br />

It is the purpose of this paper to presen t<br />

the state of the art in the application of th e<br />

theory and tracer methods to the nondestructive<br />

measurement of tree biomass and transpiration,<br />

and to examine the experimenta l<br />

parameters that can be measured as well a s<br />

the underlying theoretical assumptions on<br />

which they are based .<br />

The nondestructive measurement of transpiration<br />

rates and biomass in trees utilize s<br />

tritiated water (HTO) as a tracer for water .<br />

Tritiated water is added to the tree water pool<br />

by injection into the trunk near ground level .<br />

The fate of the HTO tracer-labeled water in a<br />

tree is followed by monitoring tritium activity,<br />

as a function of time, in foliage and smal l<br />

branches. Both transpiration and biomas s<br />

measurements require experimental determination<br />

of tritium concentration-tim e<br />

curves. Biomass requires additionally th e<br />

mean residence time of water in the tree and<br />

the mean moisture content of the tree .<br />

Theoretical Discussion<br />

Transpiration Measurement<br />

Application of the theory of tracer dynamics<br />

to the problem of measuring transpira-<br />

159


tion rates in plants has been previously<br />

demonstrated' 2 (Kline et al . 1970) . The<br />

theory itself has been discussed extensively b y<br />

Bergner (1961, 1964a, 1964b, 1965, 1966) ,<br />

Zierler (1964), Ljunggren (1967), and Orr and<br />

Gillespie (1964) . Transpiration measurement s<br />

depend upon use of the Stewart-Hamilto n<br />

equation shown by equation 1 :<br />

M = F f 0 - f(t)dt , (1 )<br />

where M = total activity of tritium initially<br />

injected (disintegrations pe r<br />

minute, DPM) ,<br />

F = the flow rate of water throug h<br />

the tree (ml/hour x tree) ,<br />

f(t) = activity distribution of tritiu m<br />

at the points of exit from th e<br />

system (DPM/ml), and<br />

t = time (hr) .<br />

Equation 1 states simply that the product<br />

of the flow rate (F) and the total integral of<br />

the curve of activity versus time (fig. 1) is<br />

equal to the total activity of the tracer whic h<br />

was originally injected . In practice the<br />

activity-time curve is measured experimentally<br />

and the total activity injected is<br />

fixed by the experimenter . The flow rate (F )<br />

is the only unknown quantity in equation 1<br />

and is solved algebraically . The value of F is<br />

the daily flow rate which is averaged over day -<br />

time and nighttime flows. The average i s<br />

taken over the full time interval of residenc e<br />

of tritium in the tree . Shorter term resolution<br />

of transpiration is not possible with thi s<br />

method . Examples of transpiration rate s<br />

which have been obtained using equation 1<br />

with tritiated water as the tracer are given in<br />

table 1 for field-grown coniferous trees .<br />

1 J. R. Kline, C. F. Jordan, and R . C . Rose . Transpiration<br />

measurements in pines using tritiated wate r<br />

as a tracer . In D. J. Nelson (ed .), Third Nationa l<br />

Symposium on Radioecology, Proc ., May 10-12 ,<br />

1971, Oak Ridge, Tenn . (In press . )<br />

'J . R . Kline, M. L. Stewart, C . F . Jordan, an d<br />

Patricia Kovac_ Use of tritiated water for determination<br />

of plant transpiration and biomass under fiel d<br />

conditions . In Symposium on the Use of Isotopes an d<br />

Radiation in Soil-Plant Research Including Applications<br />

in <strong>Forest</strong>ry, Proc . Int . At . Energy Agency Conf .<br />

SM-151, December 13-17, 1971 .Vienna, , Austria . (I n<br />

press .)<br />

50 30 0<br />

Figure 1 . Typical activity-time response curve obtained<br />

by injecting a jack pine (Pins banksiana)<br />

tree with tritiated water and sampling twigs as a<br />

function of time for tritium content . Peak arriva l<br />

time is indicated by Tp .<br />

Biomass Measuremen t<br />

The measurement of tree biomass is base d<br />

on that part of the theory of tracer dynamic s<br />

which permits calculation of the pool size o f<br />

the compartment through which flow take s<br />

place . In trees the pool refers simply to th e<br />

average total water content of the tree . Whe n<br />

pool size has been computed, it is a simpl e<br />

matter to convert to biomass using the aver -<br />

age moisture percentage of the wood .<br />

Derivation of an expression which permit s<br />

computation of compartmental pool size i s<br />

given by Zierler (1964) . Zierler's expression i s<br />

given by equation 2 :<br />

C =F Tm , (2)<br />

where C = compartment volume (ml) ,<br />

F = flow rate through the compartment<br />

(ml/hr), and<br />

Tm = mean residence time of th e<br />

flowing substance (hr) .<br />

Equation 2 states simply that the compartment<br />

pool size is given by the product of th e<br />

160


Table 1.-Transpiration rates, mean residence times, computed biomasses, and observe d<br />

biomasses for field-grown red and jack pine tree s<br />

Red pine (Pinus resinosa) (1970 )<br />

Tree<br />

number<br />

Dbh<br />

(CM)<br />

Transpiration<br />

rate (F )<br />

(ml/hr)<br />

resi ence<br />

time (Tp )<br />

(hr)<br />

Computed<br />

dry biomas s<br />

(kg)<br />

Observed<br />

dry biomass<br />

(kg)<br />

3 14 .1 1170 50 42 .4 42 . 6<br />

4 9.5 222 110 17 .7 12 . 2<br />

5 12 .1 935 42 28 .4 26 . 3<br />

6 13 .3 1280 42 38 .9 35 . 1<br />

7 12 .3 725 49 25 .7 26 . 7<br />

8 8 .2 413 49 14 .6 12 . 1<br />

10 12 .3 254 98 18 .0 20.6<br />

Mean tree weight ± SE 26 .5 ± 4 .1 25 .1 ± 4 . 3<br />

Mean forest biomass ± SE (kg/ha) 7.7 x 10 4 ± 1 .2 7.3 x 10 4 ± 1 . 2<br />

Jack pine (Pinus banksiana) (1971 )<br />

Tree<br />

number<br />

Dbh<br />

(CM)<br />

Transpiration<br />

rate (F)<br />

(<br />

/hr)<br />

re isdeece<br />

time (T )<br />

(hr) p<br />

Computed<br />

dry biomass<br />

(kg )<br />

Observe d<br />

dry biomass<br />

( kg)<br />

1 10 .5 807 49 38.5 19 . 9<br />

2 12 .0 833 52 33.5 26 .7<br />

3 10 .0 540 48 21 .5 19 . 9<br />

4 10.0 941 27 25.6 32 . 0<br />

5 8.9 570 45 25 .4 18 . 1<br />

6 11 .0 833 25 19 .2 26 . 0<br />

7 11 .0 609 23 17 .0 32 . 3<br />

8 9.5 506 46 24 .6 24 . 6<br />

9 10 .4 492 27 18 .3 20 . 9<br />

11 11 .4 1083 26 28 .0 28 . 5<br />

12 9.4 903 24 27 .2 20 . 0<br />

Mean tree weight ± SE 23.3 ± 2 .0 24 .4 ± 1 . 5<br />

Mean forest biomass ± SE (kg/ha) 6.3 x 10 4 ± 0 .5 6.1 x 104 ± 0 .4<br />

16 1


flow rate and the mean residence time of th e<br />

system .<br />

In plants the pool size (C) is also given b y<br />

the difference between wet and dry weigh t<br />

(W-D) of the plant . Moisture fraction is conventionally<br />

calculated by equation 3 :<br />

W-D<br />

W<br />

= f ,<br />

where W = wet weight of sample (gm) ,<br />

(3)<br />

D = dry weight of sample (gm), an d<br />

f = fractional moisture content .<br />

Equation 3 holds equally well for the cas e<br />

where the determination is done on the entire<br />

plant or for a representative subsample of the<br />

plant. In the case where the entire plant is the<br />

sample, W-D can be substituted for C in equa -<br />

tion 1 resulting in equation 4, which is an<br />

expression for the moist weight of the plant :<br />

W=1F•Tm . (4)<br />

f<br />

Moist weight can be converted to dry weigh t<br />

using equation 3 . This results in equation 5<br />

which is an expression for dry biomass of th e<br />

plant :<br />

D = 11-f FTm . (5)<br />

f<br />

Equation 5 requires the experimental determination<br />

of f, F, and Tm for its solution . In<br />

the absence of a feasible method for deter -<br />

mining f on the entire tree, it is necessary t o<br />

measure it on subsamples. Ideally the subsamples<br />

should be weighted for different tre e<br />

parts such as trunk, branches, and leaves .<br />

Since there is usually no method available for<br />

measuring weighting factors, we have followed<br />

the practice of estimating moistur e<br />

content of the tree trunk since this represent s<br />

the greater portion of the biomass of the tree .<br />

In our experience, moisture content of plan t<br />

parts has not been greatly different from one<br />

another and no serious errors are introduced<br />

by following this procedure . If plants are<br />

found where unweighted estimates of f diffe r<br />

from the true weighted value, then this coul d<br />

be a significant source of error in the estimate<br />

of biomass .<br />

The value of F in equation 5 is the mean<br />

flow rate which has previously been calculated<br />

using equation 1 . This means that th e<br />

reliability of the estimate of biomass can b e<br />

no better in general than the reliability o f<br />

flow or transpiration rate . Biomass estimate s<br />

are normally expected to have lower statistical<br />

precision than transpiration estimate s<br />

since they require the use of additional measured<br />

parameters .<br />

The most difficult parameter to measure i n<br />

equation 5 is Tm, the mean residence time .<br />

There appear to be at least three possibl e<br />

approaches to obtain this quantity : (1) measure<br />

the slope of the declining branch of th e<br />

activity-time curve ; (2) compute the firs t<br />

moment of the curve ; or (3) measure th e<br />

transit time between the point of injection<br />

and the point of exit of the tracer from the<br />

system (Donato et al . 1964) .<br />

The slope method for mean residence tim e<br />

would be valid in the case where the plant is<br />

labeled to equilibrium with the tracer . If all of<br />

the water molecules of the trees were labeled<br />

equally with HTO, then the rate at which tritium<br />

activity declines in the tree would b e<br />

proportional to the amount of tritiu m<br />

present . Such systems are described by an<br />

equation of the form A = AO e- Xt where AO i s<br />

initial activity, A is activity at time t and X i s<br />

the rate constant of loss . The term X is th e<br />

slope when data described by this relationshi p<br />

are plotted on semilogarithmic coordinates .<br />

The mean residence time for such a relation -<br />

ship is simply the reciprocal of X (Tm =<br />

This is a frequently used relationship for obtaining<br />

Tm although in practice it is some -<br />

times used without verification of the assumptions.<br />

In some systems it is possible to label t o<br />

equilibrium by injecting the tracer into the<br />

system continuously ; however, in large trees<br />

this would entail larger than desirable releases<br />

of radioactivity to the environment . Therefore,<br />

it is preferable to label by the instantaneous<br />

pulse method . When the pulse method<br />

of labeling is used, the injected material<br />

moves upward in the tree while retaining its<br />

pulse shape. The pulse may undergq consider -<br />

able broadening but the system cannot be<br />

assumed to have achieved uniform labeling . In<br />

162


this case the slopes of activity-time curve s<br />

have two components, one reflecting the turn -<br />

over rate of water and the other reflecting the<br />

pulse shape. There is no method currently<br />

available for resolving these components in a n<br />

activity-time curve, and therefore the slope of<br />

the curve cannot validly be used to compute<br />

Tm .<br />

Ljunggren (1967) has described the compu -<br />

tation of mean residence times for flowin g<br />

systems using the first moment of th e<br />

activity-time curve. The first moment is th e<br />

centroidal axis of the activity-time distribution-that<br />

vertical axis which divides th e<br />

distribution into two parts having equal areas .<br />

Equation 6 indicates the method :<br />

f0`°t f(t) dt<br />

Tm* _ (6)<br />

f ff(t) dt<br />

In general, Tm* will always be the tru e<br />

mean residence time of the tracer in th e<br />

system under study . The tracer mean residence<br />

time of the system will satisfy the relationship<br />

of equation 2, however, only under<br />

the conditions of identical behavior of trace r<br />

and substance traced . In trees, the nominal<br />

mean residence time for water, Tm, is no t<br />

equal to the tracer mean residence time Tm *<br />

since the tracer apparently undergoes interactions<br />

with the conducting vessels of the<br />

tree . The relationship is given by the<br />

expression<br />

Tm *=f 1 Tm +f2 TH<br />

where TH is the residence time of the fractio n<br />

of tritium which has undergone some interaction<br />

with the wood and fl and f2 are fractions<br />

of the total tritium which pass throug h<br />

the tree without interaction and with inter -<br />

action, respectively . Possible interactions include<br />

isotopic exchange of tritium wit h<br />

hydrogen of the wood or diffusion of tritiu m<br />

into non flowing compartments of plant<br />

water. <strong>Experimental</strong>ly Tm*>Tm has been<br />

found for all trees which have been examine d<br />

to date, indicating that the term f2 TH has a<br />

nonzero value in trees . This phenomenon has<br />

been termed "holdback" by others who have<br />

examined tracer dynamics in flowing vessels<br />

(Ljunggren 1967) . Because of "holdback" the<br />

value of Tm* cannot be used without correction<br />

to solve equation 2 . The problem of finding<br />

an appropriate value for f2 TH is presentl y<br />

unsolved .<br />

In general, the mean residence time for a<br />

flowing system can always be obtained by<br />

measuring the activity distribution of a trace r<br />

in the system at two points along the flow<br />

pathway (Ljunggren 1967) . If Ti is the time<br />

of passage of the tracer at point 1 and T2 the<br />

time of passage at point 2 further down -<br />

stream, then Tm = T2 - Ti where Tm is the<br />

mean residence time between the two sampling<br />

points. Since the tracer normally undergoes<br />

peak broadening, Ti and T2 are taken as<br />

the times when the peak of the distributio n<br />

passes the sampling points . In trees we fix th e<br />

initial position of the tracer by the injectio n<br />

at time Ti = O . In this special case Tm = T2 .<br />

The value of T2 could be measured at an y<br />

point in the tree downstream to the injection<br />

point. In the special case where the downstream<br />

sampling point is tree foliage, the n<br />

Tm = Tp where Tp is simply the time of pea k<br />

arrival in the foliage .<br />

As a first approximation it can be assumed<br />

that Tp is not affected by "holdback" as was<br />

Tm* because tritium is probably remove d<br />

from free flowing forms equally over the en -<br />

tire activity distribution . That is, interactio n<br />

of tritium with conducting vessels could a s<br />

well occur with the isotope in the leadin g<br />

edge of the distribution or the trailing edge .<br />

The peak position would, therefore, not b e<br />

affected by these interactions . This assumption<br />

requires experimental verification whic h<br />

is given in the results section .<br />

Correction for Nonconducting Tissue<br />

The foregoing suggests that tritium trace r<br />

experiments can only be used to measur e<br />

actually conducting biomass in trees . Such tissues<br />

as bark, flowers, fruit, and nonconducting<br />

heartwood will not be included in the estimate.<br />

Roots are not included in the estimate<br />

since the tracer injection is normally done i n<br />

tree trunks above the roots . In practical biomass<br />

measurements for forestry purposes, th e<br />

163


most serious problem is the omission of non -<br />

conducting heartwood from the direct measurement.<br />

The theory of tracer dynamics cannot<br />

be used for direct measurement of thi s<br />

quantity, and it is, therefore, necessary t o<br />

make an approximation . Equation 7 expresse s<br />

the relationship between total biomass an d<br />

the biomass of heartwood and sapwood :<br />

where VT<br />

VTPT = VHPH + V SPS , (7)<br />

total volume of tree tissu e<br />

( cm3 ) ,<br />

weighted mean wood<br />

density (cm3 ) ,<br />

VH ; VS= volume of heartwood an d<br />

sapwood (cm 3 ), an d<br />

PH ;PS<br />

density of heartwood an d<br />

sapwood ( )<br />

cm 3<br />

Substituting the relationship s<br />

PH/PS = K<br />

and VH/V S = A<br />

into equation 7 results in equation 8 :<br />

VTPT = V SpS ( AK + 1) . ( 8 )<br />

Assuming that the volume of heartwoo d<br />

and that of the total tree can be approximated<br />

by a right circular cone, an expression<br />

for A is derived as follows :<br />

r2<br />

where r<br />

PT<br />

X = Ar (2r + Ar )<br />

= mean radius of heartwood at<br />

base (cm), and<br />

Ar = mean thickness of sapwoo d<br />

at base (cm) .<br />

Upon substituting for A in equation 8, a fina l<br />

expression for tree biomass is obtained :<br />

r 2 K<br />

VTpT = D = V S p S ( Or(2r + Dr) + 1)<br />

. (9)<br />

The numerical group V.T.I. is the total plan t<br />

biomass (D) and the group VsPs is the conducting<br />

or sapwood biomass as measured b y<br />

the tritium method. Equation 9 is not sensitive<br />

to the assumption that wood volumes ar e<br />

approximated by right circular cones. Th e<br />

same result is obtained for a right circular<br />

cylinder or for intermediate figures .<br />

Equation 9 has not yet been evaluated experimentally<br />

. It is proposed here to suggest<br />

some of the anticipated lines of research to b e<br />

undertaken in the Coniferous Biome. A principal<br />

problem for the solution of equation 9<br />

lies in accurate measurement of the quantitie s<br />

r and Ar . These quantities fundamentally refer<br />

to the radii of nonconducting and conductin g<br />

wood, respectively . In the simplest case they<br />

may be coincident with heartwood and sapwood<br />

as observed visually. Their evaluatio n<br />

could then be done by straightforward measurement<br />

of tree cores .<br />

An accurate solution also depends on th e<br />

nature of the transition zone between con -<br />

ducting and nonconducting tissue . If this is<br />

sharply defined, then r and Ar will be wel l<br />

defined and an accurate solution to equation<br />

9 can be obtained . If the conduction under -<br />

goes a gradual transition across the radius o f<br />

the tree, there may be no practical method<br />

for assigning values to r and An It is possibl e<br />

that reasonable values can be obtained b y<br />

studying tritium distribution along wood<br />

cores which have been taken from tritium -<br />

labeled trees. These considerations apply t o<br />

large trees which have appreciable volumes o f<br />

heartwood . In the trees for which we hav e<br />

experimental data, heartwood was a mino r<br />

part of the total tree volume and was no t<br />

considered .<br />

Results<br />

The mean residence time Tm is the most<br />

difficult parameter of equation 5 to obtain ,<br />

principally because it is not generally known a<br />

priori which of several possible means of computing<br />

it is the correct one . The desired value<br />

is the nominal mean residence time Tm<br />

(equation 2); however, in the usual non -<br />

destructive experiment this is not directly obtainable<br />

. Where tritium-injected trees have<br />

been harvested, however, the water pool siz e<br />

(C) is directly obtainable from biomass an d<br />

moisture measurements, and it is possible to<br />

164


solve equation 2 for Tm since the flow rat e<br />

(F) is known . The values of Tm obtained b y<br />

this direct method were compared to th e<br />

values obtained by the three indirect method s<br />

which were previously discussed .<br />

The nominal mean residence time (Tm) ha s<br />

been computed for several harvested field -<br />

grown coniferous trees and the values compared<br />

with those obtained from curve slopes ,<br />

from first moment calculations and from peak<br />

arrival times . The results show in general, that<br />

Tm > TS for slope calculations and Tm <<br />

Tm * for first moment calculations .<br />

The nominal mean residence time agree s<br />

most closely with Tp which was obtained b y<br />

the midpeak or peak arrival time method .<br />

2 0<br />

I 0<br />

o ► 1 I I_ I 1<br />

0 10 20 30 40 50 6000 110 120 13 0<br />

Tm(hr )<br />

Figure 2 . Relationship between nominal mean residence<br />

time (Tm) of plant water and mean residence<br />

time obtained by peak arrival (Tp) fro m<br />

activity-time curves .<br />

Figure 2 shows the relationship between T m<br />

and Tp for a group of harvested trees . Th e<br />

slope of the linear least squares relationshi p<br />

between the two estimates is 0 .916, and the<br />

intercept is 7 .05 as compared with the expected<br />

slope of one and expected intercept of<br />

zero (1 :1 line, fig. 2). The coefficient o f<br />

determination (r2 ) is 0.70 for the relationship.<br />

The results suggest that Tp is an unbiased<br />

estimator of Tm . This confirms that Tp<br />

is the best estimate available for Tm since<br />

other known possibilities have been examined<br />

and found to be biased either above or belo w<br />

the true values .<br />

The scatter of data points on figure 2 indicates<br />

that forest biomass determinations must<br />

C<br />

at present be approached statistically . Any<br />

particular determination of mean residenc e<br />

time of water may be substantially in error ;<br />

however, a group of determinations appear s<br />

to converge on the true Tm values of th e<br />

group. <strong>Experimental</strong> error reduction is on e<br />

objective of continuing studies in the Coniferous<br />

Biome program .<br />

Table 1 shows a comparison between tre e<br />

biomass as computed by equation 2 an d<br />

actual harvested weights. These data were obtained<br />

in a uniform age plantation of red pin e<br />

(Pinus resinosa) in 1970, and a similar stan d<br />

of jack pine (Pinus banksiana) in 1971. Th e<br />

results show individual examples of substantial<br />

error; however, the estimates of mean tre e<br />

biomass and mean forest biomass as measured<br />

by the tritium method agree well with thos e<br />

estimated by direct harvest . There is, of<br />

course, no necessity for determining biomas s<br />

of a group of trees on all individuals simultaneously<br />

. Biomass can be determined at an y<br />

time during which transpiration flow is takin g<br />

place. The results suggest that the biomass of<br />

these forests could have been reliably determined<br />

by the tritium method alone .<br />

Conclusions<br />

Experiments designed to measure transpiration<br />

rates in field-grown trees may also b e<br />

used with little extra data collection for non -<br />

destructively measuring tree biomass . The tritium<br />

method can, in general, be used only for<br />

determination of biomass which is actually<br />

transmitting water . Flowers, fruits, bark, and<br />

nonconducting heartwood are not included i n<br />

the estimate. We have proposed a method for<br />

including heartwood in the estimate but have<br />

not yet proved it experimentally .<br />

The most difficult parameter to obtain for<br />

calculation of biomass is the nominal mea n<br />

residence time of water in the plant . After<br />

examining several alternatives for obtainin g<br />

this quantity, we conclude from theory and<br />

experiment that the time of arrival of pea k<br />

tritium activity in tree foliage is the most reliable<br />

estimate of nominal mean residence time .<br />

Estimates of biomass of field-grown coniferous<br />

trees show considerable statistical varia -<br />

165


tion; however, the mean of a group of such<br />

determinations was a reliable and unbiase d<br />

estimate of the mean obtained by direct harvest.<br />

With additional research on the problem<br />

it may be possible to reduce the experimental<br />

error of the method and to improve the<br />

reliability of individual estimates .<br />

Acknowledgments<br />

The work reported in this paper was performed<br />

under the auspices of the U .S . Atomic<br />

Energy Commission and supported in part b y<br />

National Science Foundation Grant No .<br />

GB-20963 to the Coniferous <strong>Forest</strong> Biome ,<br />

U .S . Analysis of Ecosystems, Internationa l<br />

Biological Program . This is Contribution No .<br />

32 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Bergner, P .-E . E. 1961 . Tracer dynamics: I .<br />

A tentative approach and definition of<br />

fundamental concepts . J. Theor. Biol. 2 :<br />

120-140 .<br />

. 1964a . Tracer dynamics and th e<br />

determination of pool sizes and turnove r<br />

factors in metabolic systems . J. Theor .<br />

Biol . 6 : 137-158 .<br />

. 1964b. Kinetic theory : Som e<br />

aspects on the study of metabolic processes<br />

. In R. M. Kniseley and W . N. Taux e<br />

(eds.), Dynamic clinical studies with radioisotopes,<br />

p. 1-18 . U .S. At. Energy Comm .<br />

TID-7678. Germantown, Md .<br />

. 1965 . Exchangeable mass :<br />

Determination without assumption of isotopic<br />

equilibrium . Science 150 : 1048-1050 .<br />

r . 1966. Tracer theory: A review .<br />

Isotope & Radiat . Tech . 3 : 245-262 .<br />

Donato, L ., C. Giuntini, R . Bianchi, and A .<br />

Maseri . 1964 . Quantitative radiocardiography<br />

for the measurement of pulmonary<br />

blood volume . In R. M. Kniseley<br />

and W. N. Tauxe (eds .), Dynamic clinical<br />

studies with radioisotopes, p . 267-283 . U .S .<br />

At. Energy Comm . TID-7678 . German -<br />

town, Md .<br />

Kline, J . R ., J. R. Martin, C . F . Jordan, and J .<br />

J. Koranda. 1970 . Measurement of transpiration<br />

in tropical trees using tritiate d<br />

water. Ecology 5(6) : 1068-1073 .<br />

Ljunggren, K . 1967. A review of the use of<br />

radioisotope tracers for evaluating parameters<br />

pertaining to the flow of materials in<br />

plant and natural systems. Isotope &<br />

Radiat. Tech. 5(1) : 3-24 .<br />

Orr, J. S., and F . C. Gillespie . 1968. Occupancy<br />

principle for radioactive tracers i n<br />

steady state biological systems . Scienc e<br />

162 : 138-139 .<br />

Zierler, K. L. 1964. Basic aspects of kineti c<br />

theory as applied to tracer distributio n<br />

studies . In R. M. Kniseley and W . N. Taux e<br />

(eds.), Dynamic clinical studies with radioisotopes,<br />

p . 55-79 . U .S . At. Energy Comm .<br />

TID-7678. Germantown, Md .<br />

166


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Theodolite surveying fo r<br />

nondestructive biomas s<br />

sampling<br />

Eugene E . Addor, Botanist<br />

U .S . Army Corps of Engineers<br />

Waterways Experiment Statio n<br />

Vicksburg, Mississippi 93180<br />

A bstract<br />

By theodolite surveying, the relative location of points in space may be calculated by triangulation . With th e<br />

aid of computers, data gathered by theodolite surveying may provide a dimensional analysis of individual trees.<br />

Because the system is nondestructive, the rates and patterns of change in the spatial structure of trees and stand s<br />

may be monitored by repetitive surveying. This paper presents a preliminary test of the approach upon trees in a<br />

40-year-old Douglas-fir (Pseudotsuga menziesii) plantation in western Washington. From experience gained in<br />

the initial experiment, recommendations are made to increase the precision of repetitive measurements .<br />

Introduction<br />

The theodolite is an instrument used in precise<br />

surveying to locate points in space by triangulation<br />

. The use of high speed computers<br />

for converting angle measurements to poin t<br />

locations allows theodolite surveying techniques<br />

to be used for describing the physical<br />

structure of vegetation assemblages in considerable<br />

detail with relative ease . Such a surveying<br />

procedure has been used to construc t<br />

simulation models for studying the effects of<br />

vegetation on engineering activities (West e t<br />

al . 1971) . Since the method is nondestructive ,<br />

it offers the possibility of repetitive samplin g<br />

with high inherent precision. Exploratory<br />

surveys were made recently to test the applicability<br />

of the system for this purpose (West<br />

and Allen 1971) . This paper examines som e<br />

data from a Douglas-fir stand at the Coniferous<br />

Biome intensive site in Washington .<br />

Description of the<br />

Surveying System<br />

The procedure requires two theodolite s<br />

placed at an arbitrary distance apart and located<br />

conveniently to the subject trees (fig .<br />

1) . The vertical and horizontal angles fro m<br />

each instrument to every point located in th e<br />

sample space are measured with respect to a<br />

base line. The instruments can be move d<br />

about to obtain clear lines-of-site to desire d<br />

points in the sample space and every instrument<br />

location (turning point) is referenced t o<br />

the base line by conventional traverse survey -<br />

Figure 1 . Instrument set-up for surveying spatia l<br />

structures of trees. Two theodolites are in use ; th e<br />

instrument in the middle is a spotting laser .<br />

167


ieoo --<br />

~e0o -<br />

I/00 -<br />

1200 -<br />

10o0 -<br />

u<br />

X<br />

N<br />

boo -<br />

eoo --<br />

40 0<br />

200<br />

0- /s.scm -.0-IA4 CM<br />

I I I I I I I I I I I I I l l I l l l l l l l l l<br />

-300 -200 -100 0 100 200 300 -300 -200 -100 0 100 200 3,<br />

Y AXIS, C M<br />

a . APRILMEASUREMENT<br />

b. OCTOBERMEASUREMENT<br />

Figure 2 . Graphic display of point location data from a Douglas-fir tree near Seattle, Washington, on whic h<br />

branching was not surveyed in detail . Bole graphed to show diameter to scale.<br />

168


ing methods . Every point in the sample spac e<br />

is thus located with respect to any arbitraril y<br />

defined three-coordinate system . Details of<br />

the procedure are presented elsewhere (West<br />

and Allen 1971) . 1<br />

One of the theodolites is equipped with a<br />

specially designed circular reticle for measuring<br />

branch or stem diameter, employing th e<br />

principle of stadia measurement . All measure d<br />

angles and reticle readings are recorded in th e<br />

field on specially designed data forms, an d<br />

trigonometric conversions of field data t o<br />

point locations and stem or branch diameter s<br />

are made by computer . Figure 2 is an exampl e<br />

of a computer graphic of one of the trees o n<br />

the Thompson site .<br />

Description of the Sample<br />

The Douglas-fir stand, located on the A. E .<br />

Thompson Research Area in the Cedar Rive r<br />

watershed, lies some 64 km southeast of<br />

Seattle, Washington . The Research Area is<br />

described in detail by Cole and Gessel (1968) .<br />

Measurements were obtained from a group o f<br />

eight contiguous trees on each of two proximate<br />

(not contiguous) permanent research<br />

plots (designated 1 and 2 on the Researc h<br />

Area) within an even-aged 40-year-old plantation<br />

. Plot 1 received three applications o f<br />

nitrogen as ammonium sulfate (NH 4 SO 4 at<br />

the rate of 222 kg/ha in October 1963 ,<br />

October 1964, and May 1970 . Plot 2 was left<br />

untreated as a control .<br />

The eight trees selected for measuremen t<br />

on each plot were selected first by choosin g<br />

an arbitrary point within each plot, and the n<br />

taking the eight trees nearest to each point a s<br />

sampling trees. The only controlling criterio n<br />

placed upon location of the starting point o n<br />

each plot was that it should be far enough<br />

within the plot to avoid inclusion of boundary<br />

trees in the sample . Measurements on the<br />

sample trees were made in April 1970, befor e<br />

bud burst, and then again in October 1970 ,<br />

after the apparent end of the growing season .<br />

Data were taken to include the coordinate<br />

location and the diameter (outside hark) at<br />

the following points :<br />

a) At the base of the tree, defined as being<br />

at the duff line or ground line, as wel l<br />

as could be determined . Diameters were<br />

measured with a tape .<br />

b) Diameters at breast height (d .b.h . )<br />

150 cm above the duff line were<br />

marked with a ribbon for subsequent<br />

remeasurement .<br />

c) The bole at every fourth whorl where<br />

limbs were still present (diameters wer e<br />

calculated from reticle readings) .<br />

d) The base of the live crown, defined as<br />

the lowermost whorl at which more<br />

than 50 percent of the branches hel d<br />

green leaves . The location of a whorl i s<br />

defined as the approximate centroid of<br />

branch emergence ; diameter measurements<br />

on the bole are made just below<br />

the lowermost branch as well as just<br />

above the uppermost branch of th e<br />

whorl (fig. 3). (Diameters calculated<br />

from reticle readings . )<br />

e) At every fourth branch whorl within<br />

the live crown, or at least one branc h<br />

whorl within the middle one-third of<br />

the live crown .<br />

f) The topmost whorl in April and that<br />

same whorl plus the new topmos t<br />

whorl in October .<br />

g) The top of the leader, at the base of th e<br />

terminal bud whorl .<br />

1 E . E . Addor and H . W . West. A technique fo r<br />

measuring the three-dimensional geometry of standing<br />

trees . U .S. Army Waterways Experiment Station ,<br />

Vicksburg, Mississippi . Unpublished .<br />

Figure 3. Definition of branch whorl location, an d<br />

location of diameter measurements at whorl .<br />

169


h) Point locations and diameters were<br />

measured for various defined points o n<br />

crown branches but will not be discussed<br />

here .<br />

A total of 30 turning points (instrument<br />

set-ups) were established for the April survey<br />

and all were referenced to a common coordinate<br />

system . The same points of reference<br />

were used again in October, but no deliberate<br />

attempt was made to duplicate the surveyin g<br />

sequence, e .g., to site each point on the tree<br />

and to measure each diameter from the same<br />

turning point . Nonetheless, the sequence tha t<br />

was adopted for the first survey was approximated<br />

during the second survey, as a result of<br />

constraints within the stand . Thus the<br />

sampled points were mostly viewed from th e<br />

sample angles during both surveys . Since both<br />

surveys were referenced to the same coordinate<br />

system, the reported coordinate locations<br />

of surveyed points are in theory exactl y<br />

comparable, so that any difference in th e<br />

reported location of a point represents a displacement<br />

of that point by wind action ,<br />

growth, or survey error .<br />

Results of<br />

Theodolite Survey<br />

Crown Cover and Stand Densit y<br />

The crown cover was essentially closed an d<br />

the branching structure relatively dense . The<br />

ground area occupied by the eight sample d<br />

trees on each plot was determined in April b y<br />

traversing the ground points representing th e<br />

outer crown limits of the outermost trees o f<br />

the group . The crown coverage so determine d<br />

was 31 .2 m 2 on plot 2 (unfertilized) an d<br />

44 .0 m 2 on plot 1 (fertilized) representing a<br />

crown area per tree of 3 .9 m 2 and 5 .5 m 2 ,<br />

respectively, or a density of approximatel y<br />

2,567 and 1,818 trees per hectare .<br />

In this same stand, in October, 1965, Dic e<br />

(1970) destructively analyzed 10 trees from a<br />

0.0045-hectare (45 m 2 ) plot, which is 4 .5 m 2<br />

per tree, or approximately 2,222 trees per<br />

hectare . These values lie reasonably between<br />

our values for the unfertilized and fertilized<br />

trees. Unfortunately we are not certain ho w<br />

the crown boundaries of his trees relate to th e<br />

boundaries of his 45 m 2 plot .<br />

The problem of the true relation of the<br />

crown cover per tree (tree mean area) to<br />

sample plot boundaries is controversial (Greig -<br />

Smith 1964) . Supposedly, tree randomness<br />

with respect to sample-plot boundaries should<br />

balance the excluded and included portions of<br />

included and excluded trees, but the true relation<br />

is apparently complicated by both plot<br />

size and plot shape . A crown-limit traverse is<br />

relatively simple with a theodolite survey and<br />

is easily converted to area by the computer . It<br />

should therefore be worthwhile to examine<br />

whether such a procedure would resolve th e<br />

problem of the relation between sample plo t<br />

boundary and tree crown boundary in th e<br />

determination of crown cover, stand density ,<br />

or tree mean area .<br />

Patterns in Diameter Measurement s<br />

Examination of the diameter data from th e<br />

theodolite survey suggests that the unfertilized<br />

trees exhibited greater increment on th e<br />

upper portion of the bole than on the lower ,<br />

whereas the fertilized trees showed approximately<br />

equal growth pattern throughout th e<br />

length of the bole . Such patterns seem reason -<br />

able because of the difference in / stan d<br />

density. Similar patterns have been reporte d<br />

in unthinned and thinned stands of Douglas -<br />

fir surveyed with an optical dendrometer over<br />

a 2-year period (Groman and Berg 1971) .<br />

Certain sources of inaccuracies in diameter<br />

measurements with the theodolite syste m<br />

should be mentioned . First, diameters calculated<br />

from reticle readings are dependen t<br />

upon accurate measurements of the distance .<br />

Second, interpolation errors from this source<br />

may be important when estimating smal l<br />

branch diameters with the reticle . Countering<br />

these disadvantages is the possibility of<br />

measuring diameter at any point on a stem or<br />

branch regardless of the direction or angle of<br />

inclination . Other instruments such as the<br />

optical dendrometer have no reticle inscripted<br />

and thus are restricted to measuring bole<br />

diameter.<br />

170


Patterns in Point Displacemen t<br />

Measuring changes in the physical structur e<br />

of vegetation consists primarily of simply<br />

measuring the displacement of defined points<br />

over a specified time interval . Obviously, an<br />

error in determining the location of a poin t<br />

either at the beginning or at the end of th e<br />

time interval will result in an error in th e<br />

measurement of displacement .<br />

Measurement errors may stem from a<br />

variety of sources, depending upon th e<br />

methods of measurement . Since theodolite<br />

surveying is dependent upon calculation o f<br />

point locations by trigonometric relations ,<br />

both instruments must be precisely sighted o n<br />

the spot to be located . Disparities in th e<br />

assumed location of the target point wil l<br />

cause errors in the calculated location of th e<br />

point. Horizontal disparities will cause horizontal<br />

and vertical errors according to th e<br />

angle of convergence and the slope of lines-ofsite,<br />

while a vertical disparity will not locate a<br />

point at all .<br />

To reduce errors from this cause, a spottin g<br />

laser (fig. 1) can be used to project a bright<br />

orange spot a few millimeters in diameter<br />

onto the tree at the selected target point . Thi s<br />

provides a definitive target for sighting th e<br />

theodolites at one given time, but it does not<br />

resolve the problem of relocating the exact<br />

point of measurement for periodic remeasurement<br />

. The spotting laser was at the Thompson<br />

Site during both the April and October<br />

surveys, but it was inoperable much of the<br />

time. Periods of its use and nonuse may<br />

account for some of the patterns in the data .<br />

Other sources of error include the usua l<br />

reading and transcription errors by instrument<br />

men and note keepers. Errors from these<br />

various sources may or may not be critical ,<br />

depending upon their magnitude and frequency,<br />

and the special purpose for which th e<br />

survey is being made . For the purpose of<br />

monitoring subtle changes in the vegetation<br />

structure over a brief time period, even ver y<br />

small errors may be important . Gross<br />

anomalies in the data may be identified and<br />

approximately corrected during data editing<br />

and preliminary analysis, but small error s<br />

regardless of source may not be distinguish -<br />

able from true displacement.<br />

For the purpose of discussion, any difference<br />

between the calculated location of a<br />

defined point from one observation to<br />

another (specifically, for the present case ,<br />

from April to October), in any coordinat e<br />

direction, may be defined as an "apparent displacement"<br />

of that point in that direction . I t<br />

can then be defined that the apparent displacement<br />

always consists of two components<br />

: True displacement resulting fro m<br />

changes in the shapes of the trees, and errors<br />

resulting from inaccuracies and mistakes i n<br />

instrument reading, note keeping, and calcula -<br />

tions . Hereinafter, these latter will be referred<br />

to collectively as "survey error ." The question<br />

to be resolved, then, is what proportion o f<br />

apparent displacement can be attributed t o<br />

each of these two components . Data from the<br />

surveyed Douglas-fir stand provide a n<br />

opportunity to examine the theodolite<br />

surveying system with respect to thes e<br />

problems .<br />

Figure 4 is a set of graphs of the apparent<br />

displacement in the xy (horizontal) plant of<br />

defined points at various levels on the tree<br />

boles, as measured from April to October .<br />

They are : (A) at the base of the tree, (B) at<br />

the base of the live crown, (C) at an arbitrary<br />

intracrown whorl, and (D) at the whorl that<br />

was defined as the topmost whorl in Apri l<br />

(i .e., at the base of the April leader, three<br />

trees are omitted from the intracrown whorl<br />

data due to omissions in the survey) . With few<br />

exceptions, the apparent displacement in the<br />

horizontal plane at the base of the tree i s<br />

within plus or minus 3 cm, with a slight systematic<br />

bias in the positive x direction and in<br />

the negative y direction, but with the points<br />

for the unfertilized and fertilized trees reason -<br />

ably well interspersed . Since trees are<br />

anchored at the base, it may be assumed that<br />

any apparent displacement of the tree axis in<br />

the horizontal plane at that level must represent<br />

a surveying error . Therefore this sligh t<br />

systematic error at this level on these tree s<br />

may be interpreted as a horizontal error i n<br />

relocating the established origin of the coordinate<br />

system. The absence of a separation i n<br />

this plane at this level between the apparent<br />

displacement of the unfertilized and fertilize d<br />

171


+10<br />

q<br />

A<br />

0 -29 ~---o-- -<br />

D<br />

-A &A<br />

>.<br />

o"f<br />

I<br />

0 0<br />

o<br />

o FERTILIZE D<br />

q UNFERTILIZE D<br />

I I 1<br />

1<br />

q +37<br />

o<br />

-~<br />

-1 0<br />

1 q + 1 9<br />

o--><br />

of<br />

0<br />

I 1<br />

-10 0 +10 -10<br />

APPARENT DISPLACEMENT, C M<br />

0<br />

(x)<br />

+10<br />

Figure 4 . Apparent displacement of the tree boles in x and y (the horizontal plane) at various levels on th e<br />

sampled trees, from April to October : (A) at the base of the trees, (B) at the base of the live crowns, (C) a t<br />

an intracrown whorl, (D) at the base of the April leader (based on data from table 2 in West and Allen 1971) .<br />

172


trees indicates that the coordinate system wa s<br />

surveyed across the intervening distance between<br />

the two plots (about 20 m) with negligible<br />

error in this plane. If the data are adjusted<br />

for the horizontal error in relocation of<br />

the coordinate system origin, then the displacement<br />

error is quite small, and it must b e<br />

conceded that remeasurement of the horizontal<br />

angles to the trees has been achieve d<br />

with a fair degree of success, despite the 3 0<br />

turning points used in accomplishing th e<br />

survey .<br />

Two explanations may be suggested for th e<br />

increasing scatter of points in the horizontal<br />

plane with increasing height on the tree ,<br />

shown on figure 4B, C, and D. First, it may be<br />

assumed that surveying errors have increase d<br />

with increasing elevation of lines-of-site, or<br />

second, it may be assumed that the positio n<br />

of the boles are less stable in the horizontal<br />

plane at higher levels on the tree . Since point<br />

locations in the horizontal plane are calculated<br />

from horizontal angles, irrespective of<br />

angles of elevation, there is no reason t o<br />

assume that surveying errors should increas e<br />

in relation to elevation. It follows therefore<br />

that errors in the location of points in th e<br />

horizontal plane at any elevation might b e<br />

equal to, but should not exceed, the errors in<br />

location of the base of the trees in this plane .<br />

It follows in turn that the apparent horizonta l<br />

displacement of points on the upper boles of<br />

these trees must be a true displacement . The<br />

pattern is consistent with what would b e<br />

expected as a result of movement of the trees<br />

by wind .<br />

Figure 5 is a set of graphs of the apparent<br />

displacement of z (the vertical plane, or elevation)<br />

for the same defined points on the bol e<br />

that are shown on figure 4, respectively .<br />

These show a considerable scatter in th e<br />

apparent vertical displacement at the base o f<br />

the live crown, moderate scatter in apparen t<br />

vertical displacement at the intracrown whorl ,<br />

and again a relatively close clustering of<br />

apparent displacement at the base of the April<br />

leader .<br />

This pattern of variation may be attributed<br />

to relative differences in the difficulty o f<br />

identifying the defined points with respect<br />

to elevation. This difficulty is more or less<br />

inherent in the definition of the points ; that is<br />

to say that "at or near the duff or ground<br />

line" is less definitive than is "the centroid of<br />

the branch whorl," whereas the topmos t<br />

whorl (base of the leader) is the smallest an d<br />

most definitive of the defined points. The<br />

differences in vertical scatter of points for th e<br />

base of the live crowns and for the intracrow n<br />

whorl may be attributed to errors of approximating<br />

the exact elevational location of th e<br />

latter through obscuring branches and foliage .<br />

These inconsistencies of identification hav e<br />

important implications with regard to measurement<br />

of length relations, such as the tota l<br />

height of trees or the ratio of bole length t o<br />

length of live crown (although, of course, th e<br />

significance of the implication is determine d<br />

by the magnitude of the dimensions and th e<br />

special purpose for which the relations ar e<br />

being measured) .<br />

It was observed above that the apparent<br />

displacement of defined points on the unfertilized<br />

and the fertilized trees were reasonably<br />

well interspersed with respect to the horizontal<br />

plane . With respect to elevation, how -<br />

ever, the data exhibit systematic tendencie s<br />

that require explanation . Specifically, figure<br />

5B shows a seemingly strong tendency toward<br />

a positive apparent displacement of about 6<br />

or 7 cm for the base of live crowns on th e<br />

fertilized trees, while the apparent displacement<br />

of this point on the unfertilized trees<br />

appears to be randomly dispersed about zero .<br />

A possible explanation is that a vertical error<br />

was committed in extending the coordinat e<br />

system across the distance (approximatel y<br />

20 m) between the two groups of trees . However,<br />

if this were the case, then the sam e<br />

apparent vertical displacement must occur a t<br />

all other levels on the fertilized trees ; if it<br />

does not, then its absence from other level s<br />

must be accounted for . The data for the intracrown<br />

whorl (fig. 5C), though somewhat<br />

more scattered than the data for the base o f<br />

live crown, do indeed suggest a similar disparity<br />

between central tendencies for the tw o<br />

groups of trees, but a similar disparity is no t<br />

obvious at the base of the trees (fig . 5A), nor<br />

at the base of the April leaders (fig . 5D) . It<br />

may be that for the base of the tree, the disparity<br />

is simply obscured by the scatter of the<br />

173


30<br />

2 2 0<br />

V<br />

A B C D<br />

0<br />

q<br />

0 q<br />

0<br />

0<br />

0<br />

q 0<br />

A&<br />

II<br />

0<br />

A<br />

0 FERTILIZE D<br />

q UNFERTILIZE D<br />

-30<br />

Figure 5 . Apparent displacement of points on the tree boles in x and y as a function of z (elevation), from Apri l<br />

to October: (A) at the base of the trees, (B) at the base of the live crowns, (C) at an intracrown whorl, (D) a t<br />

the base of the April leader (based on data from table 2 in West and Allen 1971) .<br />

174


data; there is no obvious explanation for it s<br />

absence at the base of the April leader .<br />

Conclusions<br />

The preliminary survey of Douglas-fir tree s<br />

at the Thompson Site suggests that theodolite<br />

surveying is adaptable to the purpose o f<br />

detecting and monitoring subtle patterns o f<br />

change in vegetation structures . In terms of<br />

quantity and quality of data obtained for th e<br />

energy expended, this procedure appears t o<br />

be commensurate with any other known tre e<br />

measurement system, and it offers advantage s<br />

not offered by any other system . First, be -<br />

cause the location of every point in the<br />

sample is determined relative to an arbitrar y<br />

point defined as the origin of the coordinate<br />

system, the apparent displacement of any<br />

point on the tree is independent of th e<br />

apparent location of any other point on th e<br />

tree. Second, displacement of points in any<br />

direction can be measured with this procedure,<br />

such as, for example, the tips o f<br />

branches radially disposed about the stem an d<br />

growing at various angles from the vertical .<br />

Finally, and of considerable importance, thi s<br />

system is entirely nondestructive and is there -<br />

fore well suited to continued monitoring o f<br />

growth trends over extended time periods .<br />

Theoretically, the precision of the technique<br />

is within millimeters, since it is basicall y<br />

the same as is used for precision engineerin g<br />

surveying. There are, however, a few practica l<br />

limitations on the attainable precision .<br />

Following are a few observations about particular<br />

problems .<br />

A possible solution to the problem of<br />

target point identification would be to clim b<br />

the trees prior to the initial survey and afi x<br />

permanent sighting targets at points of interest.<br />

Also, a network of similarly smal l<br />

definitive targets could be establishe d<br />

throughout the sample area for use as permanent<br />

reference points, one of which could b e<br />

used to define the origin of the coordinate<br />

system. A series of carefully controlled<br />

experiments should be designed specifically t o<br />

determine the effects of operator error on the<br />

limits of attainable precision under different<br />

kinds of working conditions .<br />

Dense crown branches and foliage may<br />

place constraints on the usefulness or convenience<br />

of this method in some kinds o f<br />

vegetation.<br />

When precision is required, surveyin g<br />

should be avoided during periods of adverse<br />

weather conditions. Tarpaulins can be suspended<br />

over the instruments so that work<br />

may be carried on during rain or snow, but<br />

these conditions also affect visibility within<br />

the forest . Winds that are strong enough to<br />

cause movement of the trees will obviously<br />

increase the probability of meaningless<br />

apparent displacements, and should be<br />

avoided .<br />

As of this writing, the system has been use d<br />

exclusively for the dimensioning of trees . Th e<br />

principles upon which it is based, however ,<br />

are universally applicable mathematical relations.<br />

Therefore the technique should b e<br />

eminently suited to the study of a variety o f<br />

ecological problems involving relations that<br />

can be described in a three-coordinate system .<br />

These might include, for example, the structure<br />

of bird rookeries, the spatial arrangemen t<br />

of epiphyte or parasite plants, or flower and<br />

fruit distributions .<br />

Acknowledgments<br />

The work on which this paper is based wa s<br />

authorized and financed by the Recreation and<br />

Environmental Branch, Office of the Chief of<br />

Engineers, Washington, D .C., in cooperation<br />

with the Coniferous <strong>Forest</strong> Biome, U .S. Analysis<br />

of Ecosystems, International Biological Program.<br />

This is Contribution No . 33 to the<br />

Coniferous <strong>Forest</strong> Biome. I thank my referees ,<br />

whose constructive criticisms were helpful in<br />

arriving at tenable explanations for some perplexing<br />

patterns in the data .<br />

Literature Cited<br />

Cole, D . W., and S. P. Gessel . 1968 . Ceda r<br />

River research-a program for studyin g<br />

pathways, rates, and processes of elementa l<br />

cycling in a forest ecosystem . Univ . Wash .<br />

Coll. For . Res . Monogr . Contr. No . 4, 53 p .<br />

175


Dice, Steven F . 1970. The biomass and nutrient<br />

flux in a second growth Douglas-fir ecosystem<br />

(a study in quantitative ecology) .<br />

53 p. Ph.D. thesis on file, Univ . Wash . ,<br />

Seattle .<br />

Greig-Smith, P . 1964 . Quantitative plan t<br />

ecology. Ed. 2, 256 p., illus . Washington ,<br />

D.C . : Butterworths .<br />

Groman, William A ., and Alan B. Berg . 1971 .<br />

Optical dendrometer measurement of increment<br />

characteristics on the boles of<br />

Douglas-fir in thinned and unthinned<br />

stands. Northwest Sci . 45(3) : 171-177 .<br />

and H. H . Allen . 1971. A technique<br />

for quantifying forest stands fo r<br />

management evaluations . U .S . Army Eng .<br />

Waterways Exp . Stn. Tech . Rep . M-71-9 ,<br />

29 p ., illus .<br />

West, H . W ., R . R . Friesz, E . A. Dardeau, Jr . ,<br />

and others . 1971 . Environmental characterization<br />

of munitions test sites : Vol. 1 ,<br />

Techniques and analysis of data . U .S. Army<br />

Eng. Waterways Exp . Stn. Tech. Rep .<br />

M-71-3, 31 p ., illus .<br />

176


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Estimates of biomass and fixed<br />

nitrogen of epiphytes from<br />

old growth Douglas fir<br />

Lawrence H . Pike and Diane M . Tracy<br />

Department of Botan y<br />

Oregon State Universit y<br />

Corvallis, Orego n<br />

Martha A . Sherwoo d<br />

Department of Biology<br />

University of Orego n<br />

Eugene, Orego n<br />

Diane Nielse n<br />

Division of Ecology and Systematic s<br />

Cornell University<br />

Ithaca, New Yor k<br />

bstract -<br />

Epiphytes are sampled concurrently with measurements of surface area of trunk and branch systems of<br />

old-growth Douglas-fir (Pseudotsuga menziesii) . Crude predictions of epiphyte biomass in branch systems are<br />

corrected by more detailed sampling of a subset of branch systems. Nitrogen analyses enable conversion of<br />

epiphyte biomass to the total amount of nitrogen present in the epiphytes.<br />

Introduction<br />

Epiphytic lichens and mosses are a conspicuous<br />

component of forest ecosystems in th e<br />

Pacific Northwest . Because of their ability t o<br />

concentrate materials from the environment<br />

and the ability of some of them to fi x<br />

atmospheric nitrogen, their importance i n<br />

nutrient cycling within the system may be<br />

greater than their contribution to total biomass<br />

would suggest .<br />

In old-growth Douglas-fir forests, epiphyt e<br />

biomass is expected to be in a steady state ;<br />

epiphyte growth is balanced by litterfall, in<br />

situ decomposition, and consumption b y<br />

herbivores. Annual turnover of epiphytic<br />

lichens varies from 5 to 25 percent (Edward s<br />

et al. 1960, Pike 1971) .<br />

Epiphyte-fixed nitrogen accounts, at least<br />

in part, for the nitrogen needed for th e<br />

growth of these epiphytes, and may represen t<br />

a significant input to the forest ecosystem .<br />

Nitrogen is added to water flowing over tre e<br />

surfaces through decomposition of epiphyte s<br />

and leakage from nitrogen-fixing epiphytes ,<br />

and enters the soil from the epiphyte syste m<br />

via throughfall, stemflow, and litterfall .<br />

Measurements of biomass are necessary t o<br />

relate process studies of epiphytes to thei r<br />

contribution to forest mineral cycles on an<br />

ecosystem level . Estimating epiphyte biomass<br />

by felling and sampling selected trees is<br />

neither possible nor desirable in old-growt h<br />

Douglas-fir forests (where trees may approac h<br />

100 m in height) because such felling i s<br />

destructive, not only of the host tree, but also<br />

of the epiphytes one wishes to study .<br />

This paper outlines the methods and give s<br />

177


preliminary results of techniques develope d<br />

for sampling epiphytes on old-growt h<br />

Douglas-fir in the H . J. <strong>Andrews</strong> Experimenta l<br />

<strong>Forest</strong> . The relatively nondestructive samplin g<br />

is an extension of the procedure for estimating<br />

tree structure and biomass described b y<br />

Denison et al . (1972) .<br />

The Study Tree<br />

Methods<br />

Epiphyte sampling by the methods de -<br />

scribed here has been carried out on a single<br />

old-growth Douglas-fir tree . This tree i s<br />

located on a north-facing slope in watershe d<br />

10 of the H . J. <strong>Andrews</strong> <strong>Experimental</strong> <strong>Forest</strong> ,<br />

75 km east of Eugene, Oregon . The base of<br />

this 65 m tall tree is at an elevation of 500 m .<br />

The surrounding stand of old-growth Douglas -<br />

fir has an understory which includes wester n<br />

hemlock (Tsuga heterophylla) and vine mapl e<br />

(Acer circinatum) . Epiphytic lichens an d<br />

bryophytes are present on the understory<br />

trees and shrubs as well as the overstory<br />

Douglas-fir .<br />

Sampling of Epiphytes<br />

For sampling, the tree was divided into<br />

trunk(s) and branch systems . The trunk is the<br />

main vertical axis of the tree ; branch systems<br />

are sets of branches leaving the trunk at the<br />

same point and are made up of axes (>4 c m<br />

in diameter) and branchlets (


where C is the percentage cover by epiphytes ,<br />

r is the radius at the base of the axis, and h i s<br />

the length of the axis . The bulk of epiphyte<br />

biomass for these branch systems is assume d<br />

to occur on the axes; if this is not the case ,<br />

the biomass of epiphytes on the many small<br />

branch systems, which are made up entirely<br />

of branchlets, may be significantly underestimated<br />

.<br />

Since EIV is based partly on subjective estimates<br />

of length of axes and percentage cove r<br />

by epiphytes, the relationship between EI V<br />

and epiphyte biomass may be expected t o<br />

vary from worker to worker and from sampling<br />

period to sampling period, and for thi s<br />

reason will be treated separately for each tree .<br />

However, it should be possible to correct<br />

importance values so that correlations may be<br />

obtained that will hold across a set of trees .<br />

Sampling Branch Systems<br />

The basic sampling unit for epiphytes o n<br />

axes in the branch systems is a "cylindrat, "<br />

which is analagous to a quadrat but runs completely<br />

around the axis so that two edges ar e<br />

fused . Our 1 dm cylindrats include the entir e<br />

surface for a distance of 1 dm along an axis .<br />

Thus the surface area of the cylindrat varies ,<br />

depending on the diameter of the axis in th e<br />

region sampled .<br />

Cylindrats were spaced along an axis with a<br />

distance of 4 dm from the center of on e<br />

cylindrat to the center of the next . Th e<br />

distance from the trunk to the first cylindra t<br />

sampled along a main axis was varied from 0<br />

to 1, 2, and 3 dm so as to avoid errors whic h<br />

would be introduced by horizontal zonation<br />

on the axis near its origin from the trunk .<br />

Estimates of epiphyte cover were made, an d<br />

then the epiphytes were stripped from th e<br />

cylindrat and bagged . Axis diameter at th e<br />

center of each cylindrat was measured t o<br />

enable calculation of surface area of the axes .<br />

Branchlets (


anch system . Total epiphyte weight for a<br />

branch system is the sum of the weights o n<br />

axes and on branchlets .<br />

Nitrogen Analyse s<br />

Samples analyzed for nitrogen conten t<br />

were collected in watershed 10 in August an d<br />

October 1971 . Samples were air dried, an d<br />

nitrogen analyses were performed by Denni s<br />

Lavender of the <strong>Forest</strong>ry Sciences Laboratory,<br />

Oregon State University, using th e<br />

Kjeldahl method . Air drying samples analyze d<br />

for nitrogen content avoids losses of nitroge n<br />

that may occur with ovendrying . In order to<br />

enable expression of the nitrogen contents o n<br />

an ovendry-weight basis, air dried samples of<br />

epiphytes were ovendried at 100 0 C to determine<br />

weight loss on drying .<br />

Results<br />

The results presented here are preliminary<br />

results from the first tree sampled in water -<br />

shed 10 of the H . J. <strong>Andrews</strong> Experimenta l<br />

<strong>Forest</strong> . The estimates are crude . They are presented<br />

to give an idea of how the methodology<br />

is being applied and the order of magnitude<br />

of results that are being obtained . Thes e<br />

preliminary results are being used in improving<br />

and refining the sampling strategy .<br />

Biomass of Epiphytes on the Trun k<br />

Total epiphyte biomass on the trunks of<br />

tree 1 is estimated to be 4.5 kg (table 1) .<br />

Nearly 90 percent of this is bryophytes, an d<br />

nearly 50 percent is found within 8 m of th e<br />

ground. Lichens contribute the bulk of th e<br />

epiphyte biomass on the trunk from about 5 0<br />

to 60 m from the ground . Epiphyte biomas s<br />

per unit area for the trunk as a whole i s<br />

0.31 g/dm 2 , and is much higher than thi s<br />

figure only within 8 m of the ground and a t<br />

the top of the second trunk, where large<br />

patches of Lobaria oregana (Tuck.) Mull. Arg.<br />

were encountered .<br />

Biomass of Epiphytes on the Branch System s<br />

The frequency distribution of branch sys -<br />

tems by EIV is presented in figure 1 .<br />

Epiphyte biomass on axes of the five<br />

branch systems sampled in detail ranged fro m<br />

0 to 198 g (table 2) ; that on the branchlets<br />

ranged from 1 to 48 g (table 3) . The five<br />

branch systems show a relationship betwee n<br />

EIV and epiphyte biomass (fig . 2) . Values<br />

from the least-squares regression line wer e<br />

used to convert the number of branch system s<br />

in an EIV class to an estimate of the tota l<br />

epiphyte biomass represented by that clas s<br />

(fig. 3) .<br />

The low importance value classes<br />

(EIV < 10 ), although representing man y<br />

branch systems, make only a small contribution<br />

to the epiphyte totals for the tree . Abou t<br />

one-half of the epiphyte biomass is contributed<br />

by branch systems with EIV abov e<br />

35 . In this connection, the preliminary nature<br />

of these results must again be emphasized a s<br />

no branch system with an EIV higher than 3 2<br />

was selected for detailed sampling .<br />

Epiphyte Biomass for the Whole Tre e<br />

Total epiphyte biomass for the tre e<br />

sampled is estimated to be 18 .3 kg ; 13 .8 kg ,<br />

or about 75 percent of the total, is on the<br />

branch systems (table 4) . Assuming that the<br />

distribution of epiphyte biomass by specie s<br />

on the five branch systems studied is the sam e<br />

as the distribution overall, 50 percent of the<br />

total epiphyte biomass for the tree is bryophytes<br />

(table 4) . However, there is a decrease<br />

in the proportion of total epiphyte weight<br />

represented by bryophytes from larger<br />

diameter to smaller diameter stem sections .<br />

Bryophytes make up 86 percent of th e<br />

epiphyte biomass on the trunk, 47 percent of<br />

that on the axes of the branch systems<br />

(>4 cm diameter), and only 3 percent of that<br />

on the branchlets (


Table 1.-Epiphyte biomass on the trunk of tree 1 in watershed 10 '<br />

Species<br />

y<br />

Heigh t<br />

(m)<br />

Trunk<br />

surface<br />

area<br />

L<br />

tl " o<br />

rz es<br />

CI .o0<br />

o<br />

tl tl<br />

tlb0<br />

o<br />

~l o<br />

All<br />

species<br />

din '<br />

g<br />

g<br />

g/dm z<br />

TRUNK 1 :<br />

0-4 1,500 300 470 340 + 1,120 0 .75<br />

4-8 1,240 80 210 740 - 1,040 .83<br />

8-12 1,130 20 200 90 30 340 .30<br />

12-16 1,110 50 150 40 20 + 270 .24<br />

16-20 1,060 140 140 70 20 360 .34<br />

20-24 1,010 20 110 10 - 10 - - 150 .1 5<br />

24-28 960 + + 10 -- 10 .0 1<br />

28-32 900 20 60 20 90 .1 0<br />

32-36 860 + 10 10 + + 20 .0 2<br />

36-40 820 200 30 50 + + 290 .3 6<br />

40-44 750 + + + + 10 .0 1<br />

44-48 670 10 110 10 130 .1 9<br />

48-52 580 + 30 + + + 40 .0 7<br />

52-56 490 + + + 10 50 - 70 .1 3<br />

56-60 402 - 50 10 -- + 60 .1 4<br />

60-63 250 70 10 10 + 10 - 100 .3 9<br />

TRUNK 2 :<br />

41-4 5 440 10 + 10 - + - + - 10 .0 3<br />

45-49 310 60 + 20 - - 10 - 90 .28<br />

49-51 120 - + + - - 10 + 10 310 330 2 .84<br />

Total 14,610 850 1,570 1,480 120 20 10 110 20 10 10 320 4,530<br />

' Weights for each species of epiphyte are to the closest 10 g ; + indicates the presence of less than 5 g .<br />

Entries have been rounded and will not necessarily add to the total .<br />

18 1


Table 2.-Epiphyte biomass on axes of branch systems sampled '<br />

Branc h<br />

system<br />

Distance from<br />

beginning of axis<br />

Section of axi s<br />

Diamete r<br />

at base<br />

Surface<br />

area<br />

Epiphyte biomas s<br />

dm cm dm 2 g/dm 2 g<br />

7 2<br />

33 0-3 .5 9 .0 9 .62 0 0<br />

3.5-7 .5 8 .5 10 .37 0 0<br />

7.5-11 .5 8 .0 9 .74 .01 . 1<br />

11.5-15 .5 7 .5 8 .80 .02 . 2<br />

15.5-19 .5 6 .5 8 .17 .02 . 1<br />

19.5-23 .5 6 .5 7 .38 .02 . 1<br />

23.5-27 .5 5 .2 6 .28 .02 . 1<br />

27.5-31 .5 4 .8 5 .66 .02 . 1<br />

31.5-33 .5 4 .2 2 .59 .04 . 1<br />

Total 68 .6 . 9<br />

61 0-2 .5 8 .0 6 .28 .24 1 . 5<br />

2.5-4 .5 8 .0 5 .34 .30 1 . 6<br />

4.5-8 .5 9 .0 11 .00 .31 3 . 4<br />

8.5-12 .5 8 .5 10 .21 2 .50 25 . 6<br />

12.5-16 .5 7 .8 9 .27 2 .97 27 . 6<br />

16.5-18 .5 7 .0 4 .30 1 .48 6 . 4<br />

0-3 .0* 4 .5 4 .01 .25 1 . 0<br />

Total 50 .4 67 . 1<br />

93 0-1 .5 12 .0 5 .54 .52 2 . 9<br />

1.5-5 .5 11 .5 14 .14 1 .13 16 . 0<br />

5.5-9 .5 11 .0 13 .35 1 .17 15 . 7<br />

9.5-13 .5 10 .2 11 .94 .97 11 . 5<br />

13.5-17 .5 8 .8 11 .15 1 .30 14 . 4<br />

17.5-21 .5 9 .0 11 .15 1 .68 18 . 7<br />

21.5-25 .5 8 .8 10 .52 2 .03 21 . 4<br />

25.5-29 .5 8 .0 9 .11 1 .28 11 . 1<br />

29.5-33 .5 6 .5 8 .48 1 .40 11 . 8<br />

33.5-37 .5 7 .0 8 .80 2 .26 19 . 9<br />

37.5-41 .5 7 .0 7 .39 1 .38 10 . 2<br />

41.5-45 .5 4 .8 5 .97 2 .25 13 . 4<br />

45.5-49 .5 4 .8 5 .81 2 .39 13 . 9<br />

49.5-53 .5 4 .5 5 .34 .54 2 . 9<br />

53.5-56 .0 4 .0 3 .14 .36 1 . 1<br />

0-2 .0* 6 .5 4 .01 .23 . 9<br />

2.0-6.0* 6 .2 7 .38 .35 2 . 6<br />

6.0-10.0* 5 .5 6 .28 .26 1 . 7<br />

10.0-14.0* 4 .5 5 .34 .10 . 5<br />

0-4.0** 4 .5 5 .34 1 .38 7 . 3<br />

Total 160 .2 198 . 0<br />

116 0-2 .5 10 .0 7 .66 .04 . 3<br />

2.5-6 .5 9 .5 11 .50 .23 2 . 6<br />

6.5-10 .5 8 .8 10 .56 .37 3 . 9<br />

10.5-14 .5 8 .0 9 .55 .67 6 . 4<br />

14.5-18 .5 7 .2 8 .61 .69 6 . 0<br />

18.5-22 .5 6 .5 7 .67 .25 2 . 0<br />

22.5-26 .5 5 .7 6 .72 .26 1 . 7<br />

26.5-30 .5 5 .0 5 .78 .24 1 . 4<br />

30.5-32 .0 4 .2 1 .93 .12 . 2<br />

Total 70 .0 24 .4<br />

1 Asterisks (*, * *) denote secondary axes .<br />

2 No axes greater than 4 cm diameter .<br />

182


NZ. 4<br />

50 100 150<br />

EPIPHYTE IMPORTANCE VALU E<br />

Figure 1 . Frequency distribution of branch systems by epiphyte importance value . EIV is an estimate of th e<br />

total area (dm 2 ) covered by epiphytes on the axes of the branch system .<br />

Table 3 .-Number of branchlets and epiphyte weights on branchlets<br />

for branch systems sampled from tree 1 in watershed 1 0<br />

Total weight of<br />

Branch Number of epiphytes on Mean epiphyte Total number of<br />

system branchlets branchlet number weight per branchlets in<br />

number sampled - branchlet branch system<br />

1 2 3<br />

Estimate d<br />

total weight of<br />

epiphytes on<br />

branchlet s<br />

g ----- - g g<br />

7 1 0.40 0.40 3 1 . 2<br />

33 3 .01 0 .08 0 .51 .20 13.5 ± 1 .5 1 2 . 7<br />

61 1 15 .99 15 .99 3 48 . 0<br />

93 1 4 .14 4 .14 4 16 . 6<br />

116 3 .14 5 .62 1 .72 2.48 12.5 ± 1 .5' 31 .0<br />

' Estimated . Total number of branchlets not recorded .<br />

18 3


1 .5<br />

50<br />

EPIPHYTE IMPORTANCE VALU E<br />

Figure 2 . Relationship between epiphyte importance value and total epiphyte biomass for the five branc h<br />

200 -<br />

•<br />

T<br />

0 10 2 0<br />

EPIPHYTE IMPORTANCE VALU E<br />

Figure 3 . Estimates of total epiphyte biomass on the branch systems in each EIV class .<br />

184


Table 4.-Biomass estimates (kg) for epiphytes on trun k<br />

and branch systems of tree 1 in watershed 1 0<br />

Branch system s<br />

Epiphyte Trunk Total<br />

Axes<br />

Branchlets<br />

BRYOPHYTES<br />

Hypnum circinale 1 .6 2 .6 4. 2<br />

Dicranum spp . 1 .5 1 .5 3. 0<br />

Other mosses .1 0 .1 . 2<br />

Scapania bolanderi .8 .1 1 . 0<br />

Other liverworts .5 . 6<br />

Bryophyte total 3 .9 4 .9 .1 8.9<br />

LICHEN S<br />

Sphaerophorus globosus .1 1 .8 1 . 9<br />

Other lichens with green<br />

algal phycobionts .2 .4 .3 . 9<br />

Lobaria oregana .3 3.2 3 .0 6 . 5<br />

Other lichens with Nostoc<br />

phycobionts .1 . 1<br />

Lichen total .6 5 .4 3 .4 9 . 4<br />

TOTAL 4 .5 10 .3 3 .5 18 .3<br />

a much higher nitrogen content than thos e<br />

which do not (table 5). In two of these associations,<br />

Lobaria oregana and Peltigera<br />

aphthosa (L.) Willd ., Nostoc is a secondary<br />

phycobiont located in cephalodia. Levels of<br />

nitrogen in the nitrogen-fixing lichens are<br />

similar to those previously reported (Pike<br />

1971). The nitrogen contents of the mosse s<br />

and nonnitrogen-fixing lichens are comparabl e<br />

to those reported by Rodin and Bazilevic h<br />

(1967) from tundra and conifer ecosystems ,<br />

but are much lower than those reported fro m<br />

the agricultural Willamette Valley (Pik e<br />

1971) .<br />

Using these values of nitrogen content, th e<br />

biomass estimates were converted to estimate s<br />

of the total quantity of nitrogen in th e<br />

epiphytes on this one tree (table 6). Thes e<br />

results indicate that 65 percent of the tota l<br />

epiphyte nitrogen is located in lichens an d<br />

that 55 percent of the total is found in a<br />

single species, Lobaria oregana . Since one-half<br />

of the Lobaria oregana occurs on branchlets,<br />

more than 25 percent of the epiphyte nitrogen<br />

is found there .<br />

Discussion<br />

Our estimate of 18.3 kg of epiphyte biomass<br />

on the one Douglas-fir tree sampled i s<br />

considerably higher than the average 0 .3 kg<br />

per tree found in a stand of Picea engelmannii<br />

and Abies lasiocarpa in British Columbi a<br />

(Edwards et al. 1960) and the 0 .4 to 1 .3 k g<br />

per tree found in stands of Pinus banksiana<br />

and Picea mariana in Saskatchewan (Scotter<br />

1962) . This high epiphyte biomass is relate d<br />

to the tremendous size of old-growth Douglasfir.<br />

When wet, the epiphyte load on this tree<br />

is probably three to four times the dry weight<br />

and may be a significant factor affectin g<br />

branch fall. (See Barkman (1958) for wate r<br />

capacity of lichens and bryophytes . )<br />

There are approximately 60 old-growt h<br />

Douglas-fir trees per hectare of forest in<br />

18 5


Table 5 .-Nitrogen contents of common epiphytes in water -<br />

shed 10. Analyses were performed on air-dry material;<br />

results are expressed on an ovendry-weight basis<br />

Epiphyte<br />

Percent nitroge n<br />

Lichens with green algal phycobionts :<br />

Alectoria sarmentosa 0 .49<br />

Hypogymnia enteromorpha .50<br />

Hypogymnia imshaugii .66<br />

Platismatia glauca .4 1<br />

Platismatia herrei .50<br />

Platismatia stenophylla .52<br />

Sphaerophorus globosus .4 2<br />

Lichens with Nostoc phycobionts:<br />

Lobaria oregana 1 .9 3<br />

Peltigera aphthose 2.84<br />

Pseudocyphellaria anomala 3.07<br />

Pseudocyphellaria anthraspis 2 .6 2<br />

Sticta weigelii 3 .78<br />

Bryophytes :<br />

Dicranum scoparium .87<br />

Hypnum circinale .95<br />

Isothecium spiculiferum 1 .1 0<br />

Table 6 .-Estimates of the total quantity of nitrogen containe d<br />

in the epiphytes on tree 1 in watershed 1 0<br />

Epiphyte<br />

Nitrogen<br />

g<br />

BRYOPHYTES<br />

Hypnum circinale 40<br />

Dicranum spp . 2 6<br />

Other bryophytes 1 7<br />

Bryophyte total 8 2<br />

LICHEN S<br />

Sphaerophorus globosus 8<br />

Other lichens with green algal phycobionts 4<br />

Lobaria oregana 12 7<br />

Other lichens with Nostoc phycobionts 4<br />

Lichen total 14 3<br />

TOTAL 225<br />

186


watershed 10 (C . T. Dyrness, personal communication).<br />

Our 18.3 kg of epiphytes, per<br />

tree would then correspond to 1 .1 metric tons<br />

per hectare, a figure well within the range of<br />

values reported from northern conifer forests<br />

(Edwards et al . 1960, Scotter 1962, Rodin<br />

and Bazilevich 1967). Our estimate is only for<br />

the overstory Douglas-fir trees, however, an d<br />

does not take into account the considerabl e<br />

biomass of epiphytes that is located on understory<br />

trees and shrubs .<br />

For comparison, our estimate of the tota l<br />

dry weight of needle biomass on the tree<br />

sampled is 84 kg, a figure 4 .6 times as high as<br />

the estimate of epiphyte biomass .<br />

Our finding, that a large proportion of<br />

epiphyte biomass is present on small branch -<br />

lets, indicates the importance of adequate<br />

sampling of this part of the tree and demonstrates<br />

the importance of nondestructiv e<br />

sampling because the branchlets are particularly<br />

likely to be destroyed when a large, old -<br />

growth Douglas-fir tree is felled . Since th e<br />

contribution from the branchlets is significant,<br />

EIV should be modified to include a<br />

component from the branchlets to avoi d<br />

underestimating the portion of epiphyte biomass<br />

located on small branch systems .<br />

Our methodology does not include estimates<br />

of biomass of crustose lichens or free -<br />

living algae. We have observed, even on twig s<br />

smaller than 1 cm in diameter, that cover of<br />

crustose lichens is regularly greater than 5 0<br />

percent of the total surface area of the twigs .<br />

Our estimates must be considered underestimates<br />

since they include only bryophytes an d<br />

foliose and fruticose lichens .<br />

Acknowledgments<br />

We would like to acknowledge the assistance<br />

of Fred Rhoades, Jane McCauley, and<br />

Tom Denison in the field sampling and o f<br />

Mary Vreeland, Karen Barrett, and Kare n<br />

Berliner in sorting and weighing the epiphytes .<br />

This project began (June-August 1970) as a<br />

student project supported by the National<br />

Science Foundation Undergraduate Researc h<br />

Participation Program (Grant No. GY-7641 )<br />

to the Department of Botany, Oregon Stat e<br />

University . Subsequent work was supporte d<br />

by National Science Foundation Grant No .<br />

GB-20963 to the Coniferous <strong>Forest</strong> Biome ,<br />

U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution No .<br />

34 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Barkman, J. J. 1958 . Phytosociology and<br />

ecology of cryptogamic epiphytes . 628 p . ,<br />

illus. Assen, Netherlands : Van Gorcum.<br />

Denison, William C., Diane M. Tracy ,<br />

Frederick M. Rhoades, and Martha Sherwood<br />

. 1972 . Direct, nondestructiv e<br />

measurement of biomass and structure i n<br />

living old-growth Douglas-fir . In Jerry F .<br />

Franklin, L. J. Dempster, and Richard H .<br />

Waring (eds.), Proceedings-research on<br />

coniferous forest ecosystems-a symposium,<br />

p. 147-158, illus. Pac. Northwest<br />

<strong>Forest</strong> & Range Exp . Stn., Portland, Oreg.<br />

Edwards, R. Y., J. Soos, and R . W. Ritcey .<br />

1960 . Quantitative observations o n<br />

epidendric lichens used as food by caribou .<br />

Ecology 41 : 425-431 .<br />

Pike, L . H. 1971. The role of epiphytic<br />

lichens and mosses in production and nutri -<br />

ent cycling of an oak forest. 172 p ., illus .<br />

Ph.D. thesis, on file at Univ. Oreg ., Eugene .<br />

Rodin, L. E., and N . I. Bazilevich . 1967 . Production<br />

and mineral cycling in terrestrial<br />

vegetation . [Trans. from Russian .] 288 p . ,<br />

illus. Edinburgh : Oliver and Boyd .<br />

Scotter, G . W. 1962. Productivity of arborea l<br />

lichens and their possible importance t o<br />

barren-ground caribou (Rangifer arcticus) .<br />

Arch . Soc . Bot. Fenn . " Vanamo" 16 :<br />

155-161 .<br />

187


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Litter, foliage, branch, and ste m<br />

production in contrasting lodgepole<br />

pine habitats of the Colorado Front Rang e<br />

William H . Moir '<br />

Colorado State University<br />

A bstrac t<br />

Harvest data are presented from a 70-year-old naturally thinned stand of lodgepole pine (Pins contorta<br />

Dougl.) in the subalpine P . contorta/Vaccinium myrtillus habitat and from five stands of similar-aged lodgepole<br />

pine in the drier P . contorta/Geranium fremontii habitat. These latter stands include both naturally thinned and<br />

artificially uniform-thinned treatments . Stand structure, natural mortality, biomass components, growt h<br />

increments of stem wood, branches, and foliage, and net primary production of these materials are given in<br />

tables. Litter production over a 4-year period is also reported for each stand ; sources of variation in litte r<br />

production are discussed.<br />

Introduction<br />

<strong>Forest</strong>s of lodgepole pine (Pinus contorta<br />

Dougl.) are extensive in the central and northern<br />

Rocky Mountains and portions of the<br />

Cascade Range . Studies on net primary production<br />

are lacking, however, in virtually al l<br />

habitats in which this pine is the major re -<br />

source. Biomass studies in certain lodgepole<br />

pine habitats in Alberta have been publishe d<br />

(Kiil 1968, Johnstone 1971), and in Colorado,<br />

the biomass of forest floor humus and<br />

pine foliage has been reported (Moir and Grier<br />

1969, Moir and Francis 1972) . These studies<br />

were not extended to net primary productivity<br />

. Our rather extensive knowledge of net<br />

stem wood production in the central Rocky<br />

Mountains (Myers 1967) does not reveal th e<br />

relationship between productivity and th e<br />

lodgepole pine habitats or include branch an d<br />

foliage production. I report below, therefore ,<br />

results of a 4-year study in Colorado concerning<br />

aspects of net primary productivity i n<br />

mostly 70-year-old natural stands of lodgepole<br />

pine .<br />

Methods<br />

Six stands on the east slopes of the Front<br />

Range in Boulder County, Colorado, wer e<br />

studied . Stand LH2 occurs in the subalpine P .<br />

contorta/Vaccinium myrtillus habitat; the<br />

others occur in the drier, montane P .<br />

contorta/Geranium fremontii habitat. Habitat<br />

features and pine population statistics ar e<br />

given by Moir (1969), Moir and Francis<br />

(1972), and table 1 . In each stand a rectangular<br />

study plot was established in an area<br />

where the forest structure appeared homogeneous.<br />

Plot size varied with stem density<br />

(table 1). Annual litter fall was periodicall y<br />

measured from randomized 0 .25 m 2 micro -<br />

plots permanently located within each plo t<br />

(Moir and Grier 1969) . All surface organi c<br />

material in the microplots was collected a t<br />

4-month intervals in 1969, and in late August,<br />

the remaining years . Mineral contamination a t<br />

collection was minimized by computing litter<br />

'Present address: Box 31, Rodeo, New Mexico<br />

88056 .<br />

189


Table 1 .-Some population characteristics of Pinus con torta in stand s<br />

on the east slopes of the Colorado Front Range<br />

Stand<br />

and<br />

year<br />

Age'<br />

Plot<br />

area<br />

Number of stems by d.b.h. classes<br />

0 .2 .5 2 .5-5 5-7 .5 7 .5-10 10.15 15-20 20-2 5<br />

Density<br />

Mortalit y<br />

Years m 2 cm Stems/ha Stems/ha x yr<br />

2 .2 71 8 4<br />

1966 130 136 27 2 35,700<br />

1970 82 135 25 4 29,300 1,60 0<br />

4 .1 72 12 6<br />

1966 3 35 56 38 5 10,900<br />

1970 1 24 49 39 8 9,600 32 5<br />

1 .1 72 12 6<br />

1966 1 26 46 30 10 9,000<br />

1970 24 35 36 12 1 8,600 100<br />

1 .3 103 15 0<br />

1966 2 7 5 25 8 3,10 0<br />

1970 5 5 23 12 3,000 2 5<br />

4 .3 71 37 5<br />

1966 29 1 1 12 45 2 1,600 2<br />

1970 25 2 7 45 8 1,650 0<br />

LH2 77 15 0<br />

1966 2 4 10 18 19 4 3,80 0<br />

1970 2 4 9 18 20 4 3,800 0<br />

' Reference year 1970 . Stand age is based upon the oldest trees .<br />

2 This stand was thinned in 1934-35 . Density refers to stems over 2 .5 cm d .b .h .<br />

fall as loss-on-ignition (Moir and Grier 1969) ,<br />

and the tendency of debris of the forest floo r<br />

to "drift" into the microplots was offset b y<br />

slightly increasing the effective area of th e<br />

microplot during computation .<br />

In each of plots 1 .1, 4.3, and LH2 of table<br />

1, the stems of lodgepole pine were placed<br />

into five diameter (d .b.h.) classes, with an<br />

equal number of stems in each class. In September<br />

1969, the tree at the midpoint of eac h<br />

d .b .h. class was harvested (Ovington, Forrest ,<br />

and Armstrong 1967), giving five trees fro m<br />

each plot. Current and second year foliage<br />

and twigs were separated from each other an d<br />

from older material. Branches with any livin g<br />

foliage whatever were classified as living an d<br />

separated from dead branches at harvest time .<br />

From each stem replicate disks were obtained<br />

at the stem base, at 3 to 4 dm ., and just belo w<br />

the lowest living branch (crown base) . From<br />

each disk I measured diameter inside bark ,<br />

diameter outside bark (d .o .b.), wood density<br />

(volume determined by water immersion) ,<br />

bark density at 102° C., and average radial<br />

growth during each of growing seasons 196 8<br />

and 1969. In addition, measurements of<br />

19 0


d.o .b. were made at periodic lengths along th e<br />

felled stems. Since this harvest procedure i s<br />

not effective- for estimating cone productio n<br />

(Ovington, Forrest, and Armstrong 1967 ,<br />

Lotan and Jensen 1970), this is not include d<br />

in the study . Biomass values are reported at<br />

700 C. ovendry weights except for bole materials<br />

(stem wood and bark) which are reporte d<br />

at 102° C. ovendry weights .<br />

Stem biomass was calculated by the Smalian<br />

equation, multiplying each volume segment<br />

by the appropriate wood density observed<br />

from the disks . Stem biomass increments wer e<br />

computed as the difference between Smalian<br />

biomass for the years 1969 and 1968 . Similarly,<br />

bark biomass for each bole was deter -<br />

mined by the Smalian difference betwee n<br />

diameters inside and outside bark, using bark<br />

densities for each volume segment . The increment<br />

of living branches was calculated using<br />

estimative ratios (Whittaker and Woodwel l<br />

1968), assuming that relative branch increments<br />

are approximated by relative ste m<br />

wood increments . Branch mortality was computed<br />

by a canopy displacement model<br />

(Madgwick 1968). The computation assume s<br />

a canopy steady state in which the upward<br />

displacement of the envelope of green foliag e<br />

within a stand leaves a wake of dead branche s<br />

that were alive the previous year . The average<br />

displacement of foliage was further assume d<br />

to be the mean height growth of each stand .<br />

Stem mortality was computed from the mortalities<br />

of table 1, in which the biomass o f<br />

dead stems during the mortality period wa s<br />

determined from the basic allometric relation -<br />

ship between stem biomass and d .b.h .<br />

Calculations were based on the equation s<br />

below :<br />

Biomass and growth increments .<br />

H = total stem length (ft )<br />

T = stem length to crown<br />

(lowest living whorl )<br />

d i , d2 , d 3<br />

= diameter inside bark at ground<br />

level, 1 ft, and lowest living<br />

whorl respectively (inch) .<br />

P 1, P2, P3 = wood densities at ground level ,<br />

1 ft, and lowest living whorl<br />

(g/cc) .<br />

k = conversion constant t o<br />

metric = 0 .0772 2<br />

Ar' 1 , Ore, Ara = mean radial growth (inch) at<br />

ground, 1 ft, and crown base.<br />

Ws = biomass (kg) at 102° C<br />

II. Living branches.<br />

AWb = Wb ~ s<br />

Ws<br />

+ new twig biomass<br />

Branch mortality (per tree) = Wb(AH )<br />

T '<br />

where AH . is average height growth for<br />

stand, T' is stem length within the crown<br />

over which branch mortality is observed ,<br />

and Wb is the live branch biomass along<br />

stem portion T' , Wb is total live branc h<br />

biomass of the crown (along stem lengt h<br />

H-T) .<br />

III. Basic allometric equation .<br />

W = k 1 D k 2 , for biomass W and DBH of D ,<br />

and regression constants ,<br />

k 1 and k 2 .<br />

In the case of plot 4 .3 where a clear trend of<br />

stem wood increment with d .b.h . exists, the<br />

plot increment of stem wood was computed<br />

by linear regression over the independent<br />

variable, d .b.h .<br />

I. Stem .<br />

Ws = k[(di + di )P1 + (di + 43) (T - 1)p2<br />

+(d3 + .75 2 ) (H-T-3)ps ]<br />

AWs = W s - Ws- 1 , where Ws- 1 is obtaine d<br />

by substituting<br />

di - 2Ori for di (i = 1,2,3) .<br />

Litter Productio n<br />

Results<br />

Tables 2 and 3 give annual litter productio n<br />

in the six stands over the 4-year period . The<br />

mean annual litter production was 0 .4 6<br />

191


Table 2 .-Annual litter fall in Colorado stands of Pinus contorta .<br />

Values are given as loss on ignition weights .<br />

Year<br />

Stand<br />

2 .2 4 .1 1 .1 1.3 4 .3 LH2 Mean<br />

T otal<br />

Number<br />

of samples<br />

1967 0 .46 0 .47 17 1<br />

1968 0.37 0 .47 .36 0 .63 0 .30 0 .38 .40 5 2<br />

1969 .41 .56 .50 .88 .52 .69 .59 5 2<br />

1970 .25 .39 .37 .45 .30 .54 .37 52<br />

Mean .34 .47 .42 .65 .37 .52 .46<br />

1 A fire in our drying oven caused the loss of many samples . Only the samples from stand 1 .1 entirel y<br />

escaped the fire and can be used for comparison with other years.<br />

Table 3 .-Constituents of the 1967-68 annual litter fall, as los s<br />

on ignition in Pinus contorta stands in Colorado<br />

Stand<br />

Numbe r<br />

of samples<br />

Needles Branches 9 Cones d Cones Bark<br />

% kg/m 2 % kg/m 2 % kg/m 2 % kg/m 2 % kg/m 2<br />

2 .2 7 66 0 .24(18) 1 18 0.07(65) 12 0.05(59) 1 0 .00(58) 2 3 0 .01(20 )<br />

4 .1 7 64 .30(10) 13 .06(105) 18 .08(98) 1 .00(90) 5 .02(28 )<br />

1 .1 7 68 .24(11) 10 .04(29) 15 .05(52) 1 .00(45) 6 .02(36 )<br />

1 .3 8 54 .34(28) 6 .04(60) 34 .21(103) 2 .01(44) 5 .03(59 )<br />

4 .3 15 83 .25(17) 2 .01(105) 9 .03(72) 3 .01(35) 2 .01(70 )<br />

LH2 8 72 .26(14) 10 .04(55) 10 .04(94) 4 .02(21) 4 .01(56 )<br />

Mean 52 68 .27 10 .04 16 .07 2 .01 4 .0 2<br />

1<br />

Coefficient of variation in parentheses .<br />

2 Trace quantities are less than 0 .005 kg/m2 . Rounding errors may cause the percentages for each stand to<br />

total around 100 . Only trace quantities of nonpine debris are found in each stand .<br />

192


kg./m 2 . Both tables show, however, that<br />

numerous sources of variation in litter fal l<br />

occur in lodgepole pine forests. Variatio n<br />

within plots is evident (table 3) . Generally<br />

needle fall is less variable from microplot to<br />

microplot (CV from 10 to 18 percent except<br />

in low density stand 1 .3) than other components<br />

of the annual litter production .<br />

Branches and female cones exhibit very high<br />

spatial variability, and the latter is furthe r<br />

enhanced (beyond the values of table 3 )<br />

through activities of the common pine squirrel ,<br />

Tam ias c i u r u s h udsonicus . Generally th e<br />

percentages of needles, branches, etc . are<br />

rather uniform within and between stands ,<br />

but marked departures from mean percent -<br />

ages are found in the low density stands 1 . 3<br />

and 4.3 . These departures probably resul t<br />

from the lessened sloughing of branches a t<br />

lower portions of the bole . Thus, branches<br />

comprised only 2 percent of the annual litte r<br />

fall in stand 4 .3, but 18 percent in high<br />

density stand 2 .2. The very high cone accumulation<br />

in stand 1 .3 is partially explained by<br />

fire thinning around 1900 which remove d<br />

many smaller, noncone-bearing trees, and<br />

partly by the chance distribution of collectio n<br />

sites near prominent cone-bearing trees . The<br />

mean annual litter fall at the bottom of tabl e<br />

2 indicates that stands with 'least canopy mas s<br />

(2 .2 and 4 .3) have lowest litter fall, and thos e<br />

with greatest canopy mass (1 .3 and LH2) have<br />

the highest litter fall . This relationship is no t<br />

clear, however, when only a single year ' s<br />

annual litter fall is considered .<br />

In all stands, 1969 was the maximum litter<br />

fall year. Nevertheless, year-to-year differences<br />

are not consistent from stand to stand .<br />

Thus, 1970 was the year of least litter production<br />

in stand 2 .2 but not LH2 . Seasonal variations<br />

are also significant. The inclusive perio d<br />

from early June to late August gave 38 per -<br />

cent of the total annual litter production i n<br />

1969, but only about 26 percent of the total<br />

from the same stands the following year .<br />

While some materials such as branches may b e<br />

shed somewhat uniformly through the year ,<br />

others such as needles showed weak seasonal<br />

peaks from June through October . In only<br />

one year was there a conspicuous "peak sea -<br />

son" of needle fall in late August and September;<br />

in general, each stand was shedding<br />

needles continuously throughout the year .<br />

Biomass and Net Growth Increment s<br />

The live components of the abovegroun d<br />

pine crop are given in table 4. From the data<br />

the coefficients, k l and k 2 , of allometric<br />

equations of the form, W = k l D k 2 , were computed<br />

by regression (W is biomass in kg and D<br />

is d .b.h. in cm from columns of table 4) .<br />

These allometric equations and the entire<br />

population of d .b.h. values in each of plots<br />

4.3, 1 .1, and LH2 were used to compute th e<br />

stand biomass values of table 5 . This biomas s<br />

and the mean increments of table 4 were use d<br />

to compute the "new growth" values of tabl e<br />

6. The mean values of table 2 and the appropriate<br />

percentages of litter constituents i n<br />

table 3 were multiplied to give the litter fal l<br />

results of table 6. Stand LH2 in the P.<br />

contorta/V. myrtillus habitat is clearly th e<br />

most productive . In this stand the proportion<br />

of green needles to living shoot parts is very<br />

high in average to dominant tree sizes, as<br />

shown in the last column of table 4 . By contrast<br />

stand 1 .1 in the P. contorta/G. fremonti i<br />

habitat (Moir 1969) has in general a lowe r<br />

ratio of green needles to living shoot parts i n<br />

average to dominant trees . Current year<br />

fascicles comprise a greater end-of-season proportion<br />

of the green canopy (25 to 31 per -<br />

cent) in stand 1 .1 than in stand LH2 (17 to<br />

27 percent) . Most trees in the latter stand retain<br />

fascicles for several years longer than<br />

stand 1 .1 . Stand 4 .3 occurs in the same habitat<br />

as stand 1 .1 but was thinned in 1934-3 5<br />

to about 2 .5 x 2.5 m stem spacing . The consequence<br />

was to reduce light competition an d<br />

encourage lateral branch growth . Surviving<br />

stems retained a high proportion of needle s<br />

(11 to 17 percent of the shoot weight), an d<br />

abscission would not occur until about th e<br />

5th year. By 1969 this stand had age<br />

reached a "closed" canopy condition ; exces s<br />

needles were abscissing as suggested by th e<br />

high needle percentage of table 5 and by th e<br />

negative net needle production in table 6 .<br />

Further discussion of foliage characteristics of<br />

these stands is given by Moir and Franci s<br />

(1972) .<br />

193


Table 4. Biomass and growth increments from harvested Pinus contorta trees<br />

Stand and tree D .b .h .<br />

Total<br />

shoot<br />

Bole 2<br />

Stem wood Living branches Green needles<br />

Ws<br />

p /Ws Wb WWb Wn L\/Wn Wn/S 3<br />

cm • kg kg % kg % kg %+ %<br />

LH2 5 19 .3 145 111 .2 103 .1 2.2 12 .4 4 .9 12 .3 22 8 . 9<br />

4 16 .5 91 68 .0 63 .1 1 .4 7 .8 3 .3 5 .8 18 7 . 1<br />

3 12 .4 49 42 .0 38 .8 2.2 2 .7 4 .8 2 .7 17 5 . 7<br />

2 10 .0 27 23 .4 21.0 1.5 1 .1 3 .6 1 .0 27 3 .9<br />

1 7 .6 11 9 .2 8 .5 .7 .4 3 .9<br />

Mean 1 .8 4 .2 2 1<br />

1 .1 5 10 .0 23 17 .9 16.1 1 .9 1 .4 5 .5 1 .4 25 6 . 3<br />

4 7 .6 13 9 .6 8.6 1 .3 .8 4 .5 .8 30 6 .7<br />

3 7 .6 13 10 .8 9 .8 2.2 .7 5 .7 .8 28 6 .4<br />

2 5 .1 4 3 .5 3 .1 .8 .13 4 .6 .16 31 4 . 1<br />

1 3 .6 2 1 .8 1 .6 .05 .02 1 . 1<br />

Mean 1 .6 5 .1 2 8<br />

4 .3 5 14 .0 44 24 .2 22 .5 1 .7 6 .5 5 .1 4 .9 25 12 .6<br />

4 12 .7 30 19 .3 18 .1 1 .8 4 .2 4 .3 3 .0 24 10 .6<br />

3 11 .7 25 17 .4 15 .8 1 .9 3 .2 5 .3 2 .6 27 10 . 8<br />

2 10 .7 18 10 .7 9 .7 2 .2 2 .5 4 .8 2 .6 20 15 .5<br />

1 8 .9 15 9 .0 8.6 2 .3 2 .0 7 .5 2 .4 26 17 . 1<br />

Mean 5 .4 24<br />

Includes bole, living and dead branches, cones, and needles .<br />

2 Stem wood plus stem bark .<br />

3s refers to total shoot less dead branches .<br />

Table 6 suggests that organic production<br />

and turnover in the three stands is approximately<br />

steady (Kira and Shidei 1967) . The<br />

only meaningful net annual accumulation is<br />

the stem wood component (and the unmeasured<br />

root wood). Of course, my branch mortality<br />

model assumed a steady state condition<br />

for these stands, but the assumption is not<br />

disturbed in table 6 where new growth and<br />

branch mortality are balanced by the observed<br />

contribution of branches and needle s<br />

to the litter fall .<br />

The inherent photosynthetic capacity o f<br />

stand LH2 is high by virtue of the great lea f<br />

biomass and surface display of current-year<br />

foliage (Moir and Francis 1972) . The tw o<br />

stands of the drier habitat have considerably<br />

less foliage. Stand 4.3 has only half the increment<br />

of current-year foliage as stand 1 . 1<br />

(table 6). Since current-year foliage has<br />

greater photosynthetic capacity than older<br />

pine foliage (e .g., Larson and Gordon 1969) ,<br />

this difference may account for the reduce d<br />

amount of net stem wood production in stand<br />

194


Table 5 .Pinus contorta stand biomass, 1970, from harvest s<br />

on the east slopes of the Colorado Front Rang e<br />

Compartment<br />

Stan d<br />

4 .3 1 .1 LH 2<br />

kg/m 2 96 kg/m 2 kg/m 2 96<br />

Live pine crop :<br />

Bole' 3 .24 46 9 .44 57 21 .68 5 8<br />

Live branches .75 11 .83 5 2 .11 6<br />

Roots 1 .782 25 2 4 .28 26 9 .68 2 6<br />

Green needles .50 7 .84 5 1 .74 5<br />

Cones .35 5 .57 3 .24 1<br />

Dead branches .39 6 .57 4 1 .69 4<br />

Total 7 .01 100 16 .53 100 37 .14 10 0<br />

Standing dead timber 3 0 1 .06 .1 5<br />

<strong>Forest</strong> floor humus4 2 .85 3 .16 3 .8 6<br />

Ground flora s .01 0 .005<br />

Total 9 .87 20 .75 41 .18<br />

' Stem wood plus bark .<br />

2 Biomass of roots with diameters over about 5 mm were computed on the basis of nine excavated roo t<br />

systems . Percentages and totals in stands 1 .1 and LH2 assume root biomass to be about 25 percent of the liv e<br />

pine crop (Rodin and Bazilevich 1967) .<br />

3 Above ground parts only (bole plus branches) .<br />

4 After Moir and Grier 1969 .<br />

5 Shoot parts only, from harvests in August 1967 .<br />

4.3 despite both stands having identical leaf<br />

area indices of 4 .5 m 2 /m 2 (Moir and Francis<br />

1972) .<br />

Conclusions<br />

Pine stands in the Lodgepole Pine Zone of<br />

Colorado (Moir 1969) are not very productive<br />

per unit of current-year leaf biomass . It is instructive<br />

to compare the net primary productivity<br />

of stand LH2 (the most productive in<br />

this study) with pine stands having simila r<br />

steady state quantities of annual foliage production.<br />

Unfortunately, there are very few<br />

published studies of net primary productivit y<br />

for older natural stands of Pinus (Art and<br />

Marks 1971, Kira and Shidei 1967) . However ,<br />

several studies within young plantations o r<br />

natural stands do permit comparison, as give n<br />

in table 7 . It is clear that annual stem an d<br />

branch production in the P. contorta stand i s<br />

well below that of the three younger pine<br />

types. Numerous studies within developin g<br />

pine plantations have led workers to conclude<br />

that within a few years after canop y<br />

closure gross productivity levels off and i n<br />

following years net production declines a s<br />

stand respiration continues to increase. Forrest<br />

(1969) concluded that the weight of<br />

foliage in developing P. radiata plantation s<br />

(the oldest of his series of stands is given in<br />

table 7) is probably constant after about 1 2<br />

years. However, I am not convinced on th e<br />

19 5


Table 6.-Annual turnover from certain shoot component s<br />

of Pinus contorta stands in Colorado '<br />

Component<br />

Stand<br />

4 .3 I 1 .1 LH 2<br />

------------------ kg/m2<br />

Bol e<br />

New growth 0 .06 0 .15 0.3 9<br />

Stem mortality 0 - .02 0<br />

Net .06 .13 .3 9<br />

Branches<br />

New growth .04 .04 .0 8<br />

Branch mortality -.03 - .01 - .0 3<br />

Litter fall - .01 - .04 - .0 5<br />

Net 0 - .01 0<br />

Needles<br />

New growth .12 .24 .3 7<br />

Litter fall - .32 - .29 - .3 8<br />

Net - .20 - .05 - .01<br />

' Litter fall is determined by the equation :<br />

Li = fiL, where L is the 3- or 4-year mean (table 2) and fi is the fractional<br />

component of needles or branches from table 3 .<br />

Table 7.-Some comparisons of annual increments in shoo t<br />

components of contrasting pine stand s<br />

Stand Age Density<br />

Annual production<br />

Foliage Stem Branches<br />

References<br />

yr sterns/ha metric tons/ha<br />

P. contorta 77 3,800 3 .7 3.9 0 .8 This study<br />

P. radiata 12 1,560 3 .1-3 .3 10 .8 2 .9 Forrest 196 9<br />

P. virginiana 17 5,750 4 .3 5 .8 3 .6 Madgwick 196 8<br />

P. densiflora 33 2,340 3 .4 8.7 2 .1 Hatiya et al . 196 5<br />

196


asis of the structure of the different pin e<br />

crops (tables 1 and 5 and comparable descriptions<br />

from references in table 7) that the<br />

three lodgepole pine stands of this study carry<br />

appreciably greater burdens of respiratory<br />

biomass (Yoda et al. 1965). The major limitation<br />

of pine productivity in the Central and<br />

Northern Rocky Mountains may be the generally<br />

infertile soils and inimical continental<br />

climates. Under these environmental conditions<br />

not even the thinning treatment (stand<br />

4.3) and subsequent surge of productivity i n<br />

surviving stems gave increased yield on an are a<br />

basis (table 6) .<br />

Management decisions in the Colorad o<br />

Lodgepole Pine Zone must recognize th e<br />

inherent low productivity of P. contorta .<br />

Where wood production must continue t o<br />

have management priority, efficiency o f<br />

utilization can be increased at least 10 percent<br />

if live branches can be economically include d<br />

in the harvest (table 5) (Young 1968) . Management<br />

can also be directed toward uneven -<br />

aged stands for purposes of esthetics, enhanced<br />

groundcover production, or better<br />

game utilization. Small, irregularly shaped<br />

clearcut areas are also an attractive management<br />

possibility for long-term planning in this<br />

low productivity region . This study suggests<br />

strongly that because of its very low productivity,<br />

especially when compared against th e<br />

productivity of intensively managed pin e<br />

plantations, lodgepole pine stands in environments<br />

of site index of 90 or less should b e<br />

regarded as only minor resources of woo d<br />

harvest. A greater percentage of such land i n<br />

the lodgepole pine region of North Americ a<br />

might be devoted to recreational usage .<br />

A much neglected management tool in<br />

lodgepole pine forests is prescribed burning .<br />

Periodic fires were an important feature of<br />

lodgepole forests (Moir 1969) . This study and<br />

others (Kiil 1968) reveal that high quantities<br />

of slowly decomposable materials build up<br />

within natural stands . Fairly high quantities<br />

of nutrients become locked up in the forest<br />

floor humus (Moir and Grier 1969) ; this<br />

humus together with the shaded condition of<br />

the forest floor in closed stands has adverse<br />

effect upon the ground flora (Basile and<br />

Jensen 1971, Moir 1966). Controlled fire has<br />

at least four possible beneficial effects in<br />

lodgepole pine forests : (1) The reduction of<br />

fuel and wildfire probability, (2) stimulatio n<br />

of ground vegetation, (3) a nutrient "pulse"<br />

effect stimulating tree production, (4) tree<br />

thinning and conversion to uneven-age d<br />

stands. The possible use of controlled surfac e<br />

fires in P. contorta should be given<br />

considerably more attention .<br />

Acknowledgments<br />

Work reported in this paper was supporte d<br />

by the National Science Foundation, Grant No .<br />

B020357 in cooperation with the Coniferou s<br />

<strong>Forest</strong> Biome, U .S. Analysis of Ecosystems ,<br />

International Biological Program . This is Contribution<br />

No. 35 to the Coniferous <strong>Forest</strong><br />

Biome .<br />

Literature Cited<br />

Art, H. W., and P. L. Marks. 1971. A summary<br />

table of biomass and net annual primary<br />

production in forest ecosystems o f<br />

the world . In <strong>Forest</strong> biomass studies, 15th<br />

IUFRO Congr., p. 3-32 . Orono : Univ.<br />

Maine Press .<br />

Basile, J. V., and C. E. Jensen . 1971 . Grazing<br />

potential on lodgepole pine clearcuts i n<br />

Montana. USDA <strong>Forest</strong> Serv . Res. Pap.<br />

INT-98, 11 p . Intermountain <strong>Forest</strong> &<br />

Range Exp . Stn ., Ogden, Utah .<br />

Forrest, W. G. 1969. Variations in the accumulation,<br />

distribution and movement of<br />

mineral materials in radiata pine plantations.<br />

276 p. Ph .D. thesis on file at<br />

Australian Nat . Univ., Canberra .<br />

Hatiya, K., K. Doi, and R. Kobayashi. 1965 .<br />

Analysis of the growth in Japanese red pin e<br />

(Pinus densiflora) stands. A report on th e<br />

matured plantation in Iwate Prefecture .<br />

Govt. For. Exp. Stn. Bull. (Tokyo) 176 :<br />

75-88 .<br />

Johnstone, W . D. 1971 . Total standing cro p<br />

and tree component distributions in thre e<br />

stands of 100-year-old lodgepole pine . In<br />

<strong>Forest</strong> biomass studies, 15th IUFR O<br />

Congr., p. 81-89 . Orono : Univ. Maine Press .<br />

197


Kiil, A. D. 1968. Weight of the fuel complex<br />

in 70-year-old lodgepole pine stands of different<br />

densities . Can. Dep. For. & Rural<br />

Dev., For. Br. Dep . Publ . 1228, 9 p .<br />

Kira, T., and T. Shidei . 1967 . Primary production<br />

and turnover of organic matter in different<br />

forest ecosystems of the wester n<br />

Pacific . Jap . J. Ecol . 17: 70-87 .<br />

Larson, P. R., and J . C. Gordon . 1969 . Photo -<br />

synthesis and wood yield . Agric. Sci. Rev .<br />

7 : 7-14 .<br />

Lotan, J . E., and C. E. Jensen . 1970. Estimating<br />

seed stored in serotinous cones of<br />

lodgepole pine. USDA <strong>Forest</strong> Serv . Res .<br />

Pap. INT-83, 10 p . Intermountain <strong>Forest</strong> &<br />

Range Exp . Stn ., Ogden, Utah .<br />

Madgwick, H. A. I. 1968. Seasonal changes in<br />

biomass and annual production of an old -<br />

field Pinus virginiana stand. Ecology 49 :<br />

149-152 .<br />

Moir, W . H . 1966. Influence of ponderosa<br />

pine on herbaceous vegetation . Ecology 47 :<br />

1045-1048 .<br />

. 1969. The lodgepole pine zon e<br />

in Colorado. Am. Midland Nat. 81 : 87-98 .<br />

and R . Francis . 1972. Foliage<br />

biomass and surface area in three Pinus<br />

contorta plots in Colorado . <strong>Forest</strong> Sci. 18 :<br />

41-45, illus .<br />

and H. Grier . 1969. Weight an d<br />

nitrogen, phosphorus, potassium, and cal -<br />

cium content of forest floor humus o f<br />

lodgepole pine stands in Colorado . Soil .<br />

Sci. Soc. Am. Proc . 33 : 137-140 .<br />

Myers, C. A. 1967. Yield tables for manage d<br />

stands of lodgepole pine in Colorado an d<br />

Wyoming. USDA <strong>Forest</strong> Serv . Res. Pap .<br />

RM-26, 20 p . Rocky Mountain <strong>Forest</strong> &<br />

Range Exp . Stn ., Fort Collins, Colo .<br />

Ovington, J. D., W . G . Forrest, and J . S. Armstrong.<br />

1967 . Tree biomass estimation . In<br />

Primary productivity and mineral cycling i n<br />

natural ecosystems-symp . Am. Assoc .<br />

Advan. Sci. 13th Annu . Meet . & Ecol . Soc .<br />

Am. (New York) Proc., p. 4-31 . Orono :<br />

Univ . Maine Press .<br />

Rodin, L . E., and N . I. Bazilevich . 1967 . [Production<br />

and mineral cycling in terrestrial<br />

vegetation.] 288 p. London : Oliver &<br />

Boyd . (English translation .)<br />

Whittaker, R . H ., and G. M. Woodwell . 1968 .<br />

Dimension and production relations of<br />

trees and shrubs in the Brookhaven <strong>Forest</strong> ,<br />

New York. J . Ecol . 56: 1-25 .<br />

Y o d a, K., K. Shinozaki, H. Ogawa, K .<br />

Hozumi, and T . Kira . 1965. Estimation of<br />

the total amount of respiration in wood y<br />

organs of trees and forest communities . J .<br />

Biol . (Osaka) 16 : 15-26 .<br />

Young, H. E. 1968 . Challenge of complete<br />

tree utilization . <strong>Forest</strong> Prod. J. 18 : 83-86 .<br />

198


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Small mammal and bird populations<br />

on Thompson site, Cedar River :<br />

parameters for modeling<br />

Sterling Miller, Curtis W . Erickson,<br />

Richard D . Taber, and Carl H . Nellis<br />

College of <strong>Forest</strong> Resource s<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

Preliminary estimates of small mammal and bird populations on the Thompson site in the Cedar Rive r<br />

watershed were made from 1969 to 1971 . Mammal populations were estimated through a kill-trap grid, and bird<br />

populations through systematic direct observations . The small mammal fauna, notably rich in insectivorou s<br />

forms, has as its most abundant members Trowbridge shrew (Sorex trowbridgei), vagrant shrew (Sorex vagrans) ,<br />

shrew-mole (Neurotrichus gibbsi), Oregon vole (Microtus oregoni), and deer mouse (Peromyscus maniculatus) .<br />

Bird populations differed between summer and winter. In the summer the most abundant species wer e<br />

Swainson's thrush (Hylocichla ustulata), winter wren (Troglodytes troglodytes), Oregon junco (Junco oreganus) ,<br />

black-throated gray warbler (Dendroica nigrescens), chestnut-backed chickadee (Parus rufescens), brown-heade d<br />

cowbird (Molothrus ater), and MacGillivray's warbler (Oporornis tolmiei) . In winter only two birds were<br />

common: winter wren and golden-crowned kinglet (Regulus satrapa). For these species estimates were obtaine d<br />

for abundance, biomass and, for the mammals, energy flow .<br />

Introduction<br />

This paper represents a first step in providin g<br />

the information on terrestrial vertebrate s<br />

which will be needed in the development o f<br />

ecosystem models for the Western Coniferou s<br />

<strong>Forest</strong> Biome . It covers the initial investigations<br />

on small mammal and bird populations ,<br />

which have centered on the Thompson site of<br />

the Cedar River watershed, King County, Wash -<br />

ington . Consideration is given to methodology ,<br />

results in terms of species populations, ecological<br />

role, biomass and energy, and the desirable<br />

directions for further investigations .<br />

The Thompson site consists of a second -<br />

growth Douglas-fir (Pseudotsuga menziesii)<br />

forest about 60-70 years old, on Barnesto n<br />

soils formed from glacial outwashes . Elevations<br />

vary from 210-310 meters . There are<br />

small variations in ecological conditions with -<br />

in the forest, as indicated by differences i n<br />

the understory vegetation .<br />

Every biotic community is composed o f<br />

individual species, and these must be identified<br />

and studied individually in building u p<br />

knowledge concerning the whole . Schwarz<br />

(1967, p. 225-226) outlines our task and it s<br />

difficulties as follows :<br />

In order to determine the role of the species<br />

in the energy metabolism of the ecosystem,<br />

we must, first of all, determin e<br />

with the necessary reliability the following<br />

parameters of the population : th e<br />

absolute numbers of animals ; their biomass;<br />

the structure of the populatio n<br />

(because the intensity of the matter an d<br />

energy metabolism of animals of different<br />

weight, sex, age, physiologica l<br />

199


condition differ) ; the turnover rate o f<br />

the population, and the intensity o f<br />

metabolism of different groups of animals<br />

. . . Whilst it is not difficult to estimate<br />

the intensity of metabolism of certain<br />

species in a calorimetric chamber, i t<br />

is practically impossible to estimate the<br />

loss of energy by a bat, a mole, or a<br />

dolphin in the process of their natura l<br />

life activity .<br />

In the light of these considerations we wil l<br />

examine our preliminary data and, through a<br />

comparison of actual and needful information,<br />

delineate the work still to be done .<br />

Small Mammal Studies<br />

Since small mammals are generally inconspicuous,<br />

and do not make readily detectable<br />

signs, the determination of species presence<br />

and especially population density is widely<br />

recognized as a difficult problem .<br />

The most commonly used method fo r<br />

estimating population densities in smal l<br />

mammals is to capture, mark, and release a<br />

sample, and then capture a second sampl e<br />

containing both marked and unmarked individuals<br />

. Several underlying assumption s<br />

(Leslie 1952) are : all individuals have equa l<br />

probabilities of capture, released individual s<br />

disperse randomly into the populations, an d<br />

there is no mortality, natality, or significan t<br />

ingress or egress during the experimenta l<br />

period .<br />

Since these assumptions were not war -<br />

ranted in our case, and since we needed to<br />

collect specimens for data on weight, foo d<br />

habits, and reproduction, we considered intensive<br />

removal methods . Grodzinski et al .<br />

(1966) proposed such a method for IBP smal l<br />

mammal studies. They suggested intensive<br />

kill-trapping following a prebaiting period on<br />

a defined grid of locations . This provided data<br />

for an estimation procedure in which a regression<br />

line, plotted for the number of animals<br />

caught each day (on the ordinate axis) against<br />

the cumulative number previously caught ,<br />

would provide an estimate of populatio n<br />

density (the point where the regression line<br />

intersected the abscissa) . Earlier work on this<br />

approach was that of DeLury (1947), Hayn e<br />

(1949), and Zippin (1958) .<br />

The major flaw in the method of Grodzinski<br />

et al. (1966) is that it gives no accurat e<br />

estimate of the area sampled by the grid, be -<br />

cause the small mammals move some unknown<br />

distance to the trapping point . Sinc e<br />

the area sampled is not defined, populatio n<br />

densities cannot be determined .<br />

Adamczyk and Ryszkowski (1968) suggested<br />

that the sample grid be surrounded o n<br />

each side by an external belt of trapping locations<br />

to catch animals moving toward th e<br />

inner grid before they could reach it, thereb y<br />

controlling the "periphery effect ."<br />

The data provided by the external trappin g<br />

belt cannot be used, because the distance<br />

travelled by each animal before being caught<br />

is highly variable (Miller 1970) .<br />

Adamczyk and Ryszkowski (1968) recommend<br />

a 5-day prebaiting period to accusto m<br />

the mammals to the trap locations, followe d<br />

by a 5-day period of removal trapping . Their<br />

basic assumptions are : (1) all residents of the<br />

inner grid are captured, (2) prebaiting does<br />

not increase the resident density, and (3) individuals<br />

not resident on the inner grid are no t<br />

captured on the inner grid ; immigrants and<br />

individuals resident in the outer grid will b e<br />

captured, if at all, in the outer grid .<br />

Of these three assumptions, we could tes t<br />

only the second . Adopting a 5-day prebaiting<br />

period, we ran pairs of trap-lines-one pre -<br />

baited and one control-in each case . Total<br />

catches over 5 days were not different,<br />

though prebaited lines took a higher proportion<br />

of the catch on the first day (Mille r<br />

1970) . This finding corroborates that o f<br />

Babinska and Bock (1969), who showed tha t<br />

prebaiting did not significantly increase th e<br />

density of resident small mammals .<br />

Our trapping procedure was as follows : a<br />

5 .94-ha (12 .4-acre) sample area was divide d<br />

into 256 stations, 16 rows by 16 lines, wit h<br />

15 .2-meter spacing . Anchored at each statio n<br />

was a paper plate which was baited with oatmeal,<br />

millet, sunflower seeds, oats, and wheat<br />

for 5 days . Then the remaining bait was removed<br />

and two mouse traps were placed on<br />

each plate, one trap baited with a peanut<br />

butter-bacon grease mixture, and the other<br />

200


with a birdseed-Crisco mixture . Traps were<br />

run for 5 full days .<br />

Two such grids were established to sampl e<br />

the range of variability within the Thompso n<br />

site, and each was trapped twice, once in summer<br />

and once in winter, as follows : Grid I :<br />

July 22 to 27, 1969, and March 22 to 27 ,<br />

1970 ; Grid II : September 6 to 11, 1969, and<br />

February 18 to 23, 1970 . The results (table 1 )<br />

show several patterns . There is a difference<br />

between summer and winter populations i n<br />

almost all cases as well as a number of differences<br />

between plots . From other work on<br />

small mammal populations we would predict<br />

that there would also be variations from year<br />

to year .<br />

Since this method assumes that all th e<br />

mammals on the inner grid are caught, it i s<br />

important that trapping be highly efficient .<br />

Study of the trapping data (Miller 1970) suggests<br />

that winter populations may not be as<br />

trappable as summer populations, so winte r<br />

estimates in table 1 may be low . We have<br />

planned further work to increase trapping<br />

efficiency and so improve the accuracy of<br />

these estimates .<br />

Meanwhile, it is possible with our presen t<br />

data to obtain at least preliminary values for<br />

small mammal biomass (table 2) . Note that<br />

the differences between the two plots in tota l<br />

biomass are rather small, and that the dro p<br />

from summer to winter, in total biomass, is<br />

Table 1 .-Small mammal density estimates from grid-trapping (number per hectare )<br />

Species<br />

Grid I Gri I I<br />

July March September February<br />

Sorex trowbridgei 16 .8 3.0 8.2 3 .0<br />

S. vagrans 4 .3 2.6 1 .3 . 4<br />

Neurotrichus gibbsi 3 .5 .9 3.6 1 . 3<br />

Microtus oregoni 1 .3 .9 1 .7 1 . 7<br />

Peromyscus maniculatus 1 .7 1.3 2.2 1 . 3<br />

Total 27.6 8.7 17 .0 7 .7<br />

Table 2.-Small mammal biomass (grams per hectare)<br />

Species<br />

Grid I<br />

Gri d If<br />

July March September February<br />

Sorex trowbridgei 82 .3 18.5 41 .0 16 . 5<br />

S. vagrans 27 .4 17 .8 6.7 2 . 7<br />

Neurotrichus gibbsi 27 .9 7.7 51 .4 11 . 1<br />

Microtus oregoni 19 .8 14 .1 28.6 27 . 9<br />

Peromyscus maniculatus 27.7 22.0 36.6 22 . 2<br />

Total 185 .1 80 .1 164 .3 80.4<br />

20 1


proportionally about the same for each plot .<br />

Future studies will tell us what fractions of<br />

this biomass are attributable to variou s<br />

nutrient substances ; that is, what the chemical<br />

constitution of these small mammal bodies is .<br />

We can already make preliminary estimate s<br />

of energy flow through these small mamma l<br />

populations. The initial energy source for ecosystems<br />

is sunlight. It is fixed by green plants<br />

and may be passed to populations of plant -<br />

eating animals. Within animal population s<br />

energy is used in respiration and tissue production<br />

and may also be passed on to othe r<br />

(predatory) animal populations . Much less<br />

energy is expended in tissue production tha n<br />

respiration .<br />

We can combine published data on the<br />

caloric value of animal tissue and on smal l<br />

mammal metabolic rates with our estimates o f<br />

biomass and information on populatio n<br />

dynamics to estimate the energy consumed i n<br />

tissue production and respiration for ou r<br />

small mammal populations .<br />

The energy incorporated in tissue production,<br />

for small mammals, has been estimate d<br />

to average 1 .5 kcal/gm (Gorecki 1965) . Th e<br />

energy expended by the animal in its dail y<br />

activity-the metabolic rate-varies by species .<br />

The values used in our calculations are given<br />

in table 3 .<br />

We do not as yet have a detailed knowledg e<br />

of the dynamics of these small mammal populations.<br />

However, we now know enough t o<br />

establish upper and lower limits of energ y<br />

Table 3 .-Metabolic rates used in calculating the energy<br />

expended in respiration by small mammal s<br />

Species Cal/gm/day Authority<br />

Sorex trowbridgei 806 Pearson (1948 )<br />

S. vagrans 920 Pearson (1948), Gebczyski (1965 )<br />

Neurotrichus gibbsi 800 Authors' estimate<br />

Microtus oregoni 500 Estimate based on value fo r<br />

red-backed vole (Pearson 1947)<br />

Peromyscus maniculatus 440 McNab (1963)<br />

Table 4.-Estimated energy flow (respiration + tissue production )<br />

through small mammal population s<br />

(kcal per hectare per year )<br />

Species Minimum Maximu m<br />

Sorex trowbridgei 5,250 15,75 0<br />

S_ vagrans 3,470 7,42 0<br />

Neurotrichus gibbsi 2,760 6,81 0<br />

Microtus oregoni 4,300 6,43 0<br />

Peromyscus maniculatus 4,450 6,86 0<br />

Total 20,230 43,27 0<br />

202


flow for these species by using data on reproduction,<br />

seasonal population density and<br />

growth of young (Miller 1970) . These estimates<br />

are given in table 4 . If necessary, we<br />

can strive for more accuracy in future estimates<br />

through a more detailed knowledge -o f<br />

the dynamics of these populations .<br />

The role and foraging stratum of mammal s<br />

in the forest ecosystem can be roughly categorized<br />

as shown in table 5. Secondary consumers<br />

are relatively numerous in the small<br />

mammal populations pointing to the abundance<br />

of their invertebrate foods in the litter<br />

and soil of the forest .<br />

In categorizing such roles, we must recognize<br />

that the primary consumers do on occasion<br />

eat other animals, and that the secondar y<br />

consumers do eat plants . It remains for us t o<br />

delineate in more detail the nature of th e<br />

seasonal diet for each species, as well as the<br />

rates of consumption, and the amounts an d<br />

composition of the excretory products .<br />

Table 5 .-Names, foraging strata, and consumer roles of mammals o f<br />

the Thompson site, Cedar River watershed, Washington<br />

Scientific name<br />

Common name<br />

Foraging<br />

stratum<br />

Consumer<br />

role'<br />

Canis latrans Coyote G II °<br />

Cervus canadensis Elk (wapiti) G,S I °<br />

Chiroptera Bats C II°<br />

Eutamias townsendi Townsend's chipmunk G,S I°<br />

Lepus americanus Snowshoe hare G,S I°<br />

Lynx rufus Bobcat G II°<br />

Microtus oregoni Oregon vole G I°<br />

Mustela erminea Shorttail weasel G II °<br />

Mustela frenata Longtail weasel G II °<br />

Neotoma cinerea Bushytail woodrat G,S I °<br />

Neurotrichus gibbsi Shrew-mole B,L II °<br />

Odocoileus hemionus Black-tailed deer G,S I °<br />

Peromyscus maniculatus Deer mouse L,G I °<br />

Sorex trowbridgei Trowbridge shrew L II°<br />

Sorex vagrans Vagrant shrew L II°<br />

Tamiasciurus douglasi Chickaree G,S,C I°<br />

Ursus americanus Black bear L,G,S I°<br />

Zapus trinotatus Jumping mouse G I°<br />

1 B = soil layer ; L = litter layer ; G = ground layer, under 1 foot ; S = shrub layer, 1 to 6 feet ; C = crown layer ,<br />

area occupied by living crowns of forest overstory .<br />

21° = primary consumer, eats mostly plant material ; II° = secondary consumer, eats mostly animal matter .<br />

20 3


Bird Studies<br />

The forest birds were sampled both summer<br />

and winter on a 6 .0-ha plot chosen t o<br />

represent the Thompson site . This plot was<br />

divided into fifteen 0 .4-ha sample units on a<br />

map of the whole plot . A series of transect s<br />

was made through the whole plot on each<br />

sample day by Erickson who noted each bird<br />

observation by location and species . Durin g<br />

the summer sample period (May and June) ,<br />

many breeding males were singing, whic h<br />

facilitated their detection . During the winter<br />

observation period (December), there had t o<br />

be more dependence on sight than sound .<br />

Five days of observation were spent at eac h<br />

season, and all observations took place during<br />

mornings when there was no heavy rain. Th e<br />

results for the two seasons are given in table<br />

6 .<br />

Some less common, but in some case s<br />

rather large birds, are also found on th e<br />

Thompson site . During the coming year we<br />

will be able to make population estimates of<br />

these species, adding substantially to our total<br />

estimate of bird biomass .<br />

The contrast between summer and winte r<br />

populations is much sharper for birds than fo r<br />

small mammals . This has implications for ecosystem<br />

modeling, since some birds reproduc e<br />

on the Thompson site but winter elsewhere ,<br />

presumably suffering some mortality, an d<br />

thus constituting a one-way movement o f<br />

energy and nutrients out of the Thompso n<br />

site . From a knowledge of body weights w e<br />

have calculated the biomass of common bird s<br />

on the Thompson site by season (table 6) .<br />

Energy-flow estimates for these bird populations<br />

cannot yet be made because neithe r<br />

population estimates of the larger birds no r<br />

annual cycles of abundance of each bird species<br />

are yet available for the Thompson site .<br />

This must be one of our next topics of study .<br />

The role and foraging stratum of birds is<br />

roughly categorized in table 7. As in the case<br />

of small mammals, some of these birds shif t<br />

seasonally in their consumer-roles . A diet of<br />

invertebrates is essential for the young of al l<br />

species. Also, shifts in the relative abundanc e<br />

of available food will presumably be reflecte d<br />

in food habits . These are also topics on which<br />

more work must be done .<br />

Table 6.-Abundance (number per 100 hectares) and biomass (grams per hectare)<br />

of the most common birds by species and seaso n<br />

Species<br />

Abundance<br />

Biom as s<br />

Summer Winter Summer Winter<br />

Molothrus ater 26 0 11 .4 0<br />

Junco oreganus 43 0 9.9 0<br />

Troglodytes troglodytes 36 33 3.5 3 . 1<br />

Hylocichla ustulata 16 0 5.8 0<br />

Dendroica nigrescens 43 0 4.5 0<br />

Parus rufescens 23 0 2.3 0<br />

Regulus satrapa 0 33 0 2 . 0<br />

Oporornis tolmiei 26 0 1 .6 0<br />

Total 213 66 39.0 5 .1<br />

204


Table 7.-Names, foraging strata, and consumer roles of birds of th e<br />

Thompson site, Cedar River watershed, Washingto n<br />

Scientific name<br />

Common name<br />

Foraging Consumer<br />

stratum' role 2<br />

Bombycilla cedrorum Cedar waxwing C I°<br />

Bonasa umbellus Ruffed grouse L,G,S,C I°<br />

Bubo virginianus Great homed owl G II°<br />

Colaptes cater Red-shafted flicker G,C II°<br />

Columba fasciata Band-tailed pigeon L,S,C I°<br />

Corvus brachyrhynchos Common crow G II°<br />

Corvus corax Common raven G II °<br />

Dendragopus obscurus Blue grouse L,G,S,C I°<br />

Dendrocopos villosus Hairy woodpecker S,C II°<br />

Dendroica nigrescens Black-throated gray warbler S,C II°<br />

Empidonax spp . Empidonax flycatchers S,C II°<br />

Hylocichla ustulata Swainson's thrush L,G II°<br />

Ixereus naevius Varied thrush L,G II°<br />

Junco oreganus Oregon junco L,G I °<br />

Loxia curvirostra Red crossbill C I°<br />

Molothrus ater Brown-headed cowbird G I°<br />

Oporornis tolmiei MacGillivray's warbler S,C II°<br />

Parus rufescens Chestnut-backed chickadee S,C II°<br />

Passerella iliaca Fox sparrow G I°<br />

Perisoreus canadensis Gray jay G,C II °<br />

Pipilo erythrophthalmus Rufous-sided towhee G I °<br />

Piranga ludoviciana Western tanager S,C I °<br />

Regulus calendula Ruby-crowned kinglet C II °<br />

Regulus satrapa Golden-crowned kinglet C II °<br />

Selasphorus rufus Rufous hummingbird G,S I °<br />

Sitta canadensis Red-breasted nuthatch S,C II °<br />

Sphyrapicus varius Yellow-bellied sapsucker S,C I °<br />

Spinus pinus Pine siskin C I °<br />

Troglodytes troglodytes Winter wren G,S II °<br />

Turdus migratorius Robin L,G II °<br />

Vermivora celata Orange-crowned warbler S,C II °<br />

Vireo gilvus Warbling vireo S,C II°<br />

B = soil layer ; L = litter layer ; G = ground layer, under 1 foot ; S = shrub layer, 1 to 6 feet ; C = crown layer ,<br />

area occupied by living crowns of forest overstory .<br />

2 1° = primary consumer, eats mostly plant material ; II° = secondary consumer, eats mostly animal matter .<br />

20 5


Further Studies<br />

The data provided for small mammal populations<br />

will be more useful if we are able t o<br />

make more accurate population density estimates<br />

and if we determine the yearly, seasonal,<br />

and geographical fluctuations in<br />

density. Further, a more detailed understanding<br />

of population dynamics will permit a<br />

more accurate estimate of energy flow, and<br />

set the stage for studies of uptake, storage ,<br />

and loss of chemical materials by smal l<br />

mammal populations .<br />

Of course, the same sort of informatio n<br />

must be obtained for the large mammals ,<br />

which include several important primary consumers.<br />

For large mammals the role of movement<br />

must be studied, since seasonal movements<br />

entail a translocation of materials, an d<br />

patterns of mortality provide loci of nutrien t<br />

release .<br />

For birds the study of movements will also<br />

be important, since long-distance annua l<br />

migrations provide a potential mechanism for<br />

the annual loss of materials from the ecosystem<br />

.<br />

A prime consideration in all such studie s<br />

will be the establishment of confidence limits<br />

for the data, and estimates of additiona l<br />

sampling necessary for results of a stipulate d<br />

level of accuracy . Then we will be in a position<br />

to respond to the needs of modeling in a<br />

realistic way with regard to the actual field o r<br />

laboratory work which provision of any particular<br />

piece of information will require .<br />

In addition to these inventories of vertebrate<br />

animals in the forest ecosystem, we will<br />

seek an understanding of the effects that<br />

animal populations may have on the populations<br />

of plants which constitute their food ,<br />

i.e ., the control function exerted by animal s<br />

over plant populations. A knowledge of suc h<br />

relations, properly quantified, will help pro -<br />

vide the basis for models of the dynamics o f<br />

plant populations .<br />

Acknowledgments<br />

This study was supported by the Department<br />

of Game, State of Washington ; Institute<br />

of <strong>Forest</strong> Products, University of Washington ;<br />

Ecology Training Grant from the Nationa l<br />

Science Foundation to the Zoology Department,<br />

University of Washington ; and by<br />

National Science Foundation Grant No .<br />

GB-20963 to the Coniferous <strong>Forest</strong> Biome ,<br />

U .S. Analysis of Ecosystems, Internationa l<br />

Biological Program . This is Contribution No .<br />

36 to the Coniferous <strong>Forest</strong> Biome, IBP . We<br />

also wish to thank the personnel of the Cedar<br />

River watershed and the Seattle Water Department<br />

for their assistance and cooperatio n<br />

throughout the study .<br />

Literature Cited<br />

Adamczyk, K ., and L . Ryszkowski. 1968 .<br />

Estimation of the density of a rodent popu -<br />

lation using stained bait. Acta Ther .<br />

13(17) : 295-311 .<br />

Babinska, J ., and E . Bock. 1969 . The effect of<br />

prebaiting on capture of rodents . Act a<br />

Ther. 14(19) : 267-270 .<br />

De Lury, D . B . 1947 . On the estimation o f<br />

biological populations. Biometrics 3 :<br />

145-167 .<br />

Gebczynski, M . 1965 . Seasonal and age<br />

changes in the metabolism and activity o f<br />

Sorex araneus L . Acta Ther. 10(22) :<br />

303-331 .<br />

Gorecki, A. 1965 . Energy values of body in<br />

small mammals . Acta Ther . 10(23) :<br />

333-352 .<br />

Grodzinski, W ., Z. Pucek, and L . Ryszkowski .<br />

1966. Estimation of rodent numbers b y<br />

means of prebaiting and intensive removal .<br />

Acta Ther . 11(10) : 297-314 .<br />

Hayne, D . W. 1949 . Two methods for estimating<br />

population from trapping records . J .<br />

Mammalogy 30(4) : 399-411 .<br />

Leslie, P. H. 1952. The estimation of population<br />

parameters from data obtained by<br />

means of the capture-recapture method. II .<br />

The estimation of total numbers . Biometrika<br />

39(3&4): 363-388 .<br />

McNab, B . K. 1963 . A model of the energy<br />

budget of a wild mouse . Ecology 44(3) :<br />

521-532 .<br />

206


Miller, S. 1970. Small mammal populations i n<br />

a Douglas-fir forest : Cedar River, Washington.<br />

102 p. M .S. thesis on file at Univer .<br />

Wash ., Seattle .<br />

Pearson, O . 1947 . The rate of metabolism o f<br />

some small mammals. Ecology 28(7) :<br />

127-145 .<br />

1948 . Metabolism of small mammals<br />

with remarks on the lower limit of<br />

mammalian size . Science 108(2794) : 44 .<br />

Schwarz, S. S. 1967 . Indirect methods of estimating<br />

field metabolism of mammals . In K .<br />

Petrusewicz (ed.), Secondary productivity<br />

of terrestrial ecosystems, p. 225-239 .<br />

Warsaw, Poland : Panstwowe Wydawnictwo<br />

Naukowe .<br />

Zippin, C. 1958. The removal method of<br />

population estimation . J. Wildlife Manage .<br />

22(1) : 82-90 .<br />

207


Terrestrial Process Studies<br />

209


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Terrestrial process studies i n<br />

conifers: a review<br />

Abstract<br />

R . B . Walker, Department of Botan y<br />

D . R . M . Scott, College of <strong>Forest</strong> Resource s<br />

D . J. Salo, Department of Botan y<br />

aqd<br />

K . L . Reed, College of <strong>Forest</strong> Resources<br />

University of Washingto n<br />

Seattle, Washingto n<br />

A few studies on the physiological processes of conifers go back as much as a century, but most informatio n<br />

has been accumulated during the past 40 years . These efforts have involved most of the widely distribute d<br />

species to some extent, but limited information is available even on those species most studied, and very little<br />

on the remainder. With Douglas-fir (Pseudotsuga menziesii), the most studied species of our region, considerable<br />

information is available, but insufficient for process modeling . Of other species of the Western United State s<br />

much less is known . Thus intensive studies ofphysiological processes will be necessary in the Coniferous Biom e<br />

effort on each of the species of principal importance . In this paper, the current status of information on th e<br />

following topics is reviewed: CO 2 assimilation and respiration; transpiration, water conduction and wate r<br />

deficits; translocation of photosynthates; and mineral nutrition . Stomatal behavior and leaf resistance are<br />

considered with respect to the various gas exchanges . Energy budgets, nutrient cycling, growth, and modeling o f<br />

processes are closely related to the foregoing topics, but are covered elsewhere in this symposium.<br />

Introduction<br />

Central in the goals of the IBP is the measurement<br />

of productivity and production, and<br />

particularly the understanding and predictio n<br />

of these for vegetation and ecosystems . It<br />

follows that a basic knowledge of the physiological<br />

and related processes involved is<br />

essential to reaching these goals . In the Coniferous<br />

<strong>Forest</strong> Biome our concern is naturally<br />

centered around the conifers, although associated<br />

deciduous and herbaceous species are<br />

also important. However, this review concentrates<br />

on the conifers as the species demanding<br />

our principal efforts .<br />

Experimentation with conifers, which began<br />

about a century ago and was taken up by<br />

various workers periodically, has been included<br />

in several general reviews in the Encyclopedia<br />

of Plant Physiology (Leyton 1958,<br />

Huber 1956, Stalfelt 1956, Pisek 1960), an d<br />

discussed in detail by Kramer and Kozlowski<br />

(1960) . Since then, a number of reviews concerning<br />

woody plants have covered various<br />

aspects: viz. food relations (Kozlowski an d<br />

Keller 1966), mineral nutrition (Baule an d<br />

Fricker 1967), and translocation (Zimmermann<br />

and Brown 1971) . Of course, the results<br />

of many experiments with broadleafed wood y<br />

plants are pertinent to conifers . Further, general<br />

principles established with herbaceou s<br />

plants may be applicable to conifers, although<br />

this carryover is more tenuous and in each<br />

case usually demands experimental verification .<br />

This review will concentrate on findings o f<br />

the past decade which, together with the<br />

work discussed in the reviews mentione d<br />

above, serves as the background on which th e<br />

studies of the Coniferous Biome are based .<br />

Included is considerable unpublished material<br />

211


from theses. The principal processes an d<br />

subjects which will be covered in sequence are<br />

(A) assimilation of CO 2 and respiration ,<br />

(B) transpiration, water conduction, an d<br />

water deficits, (C) translocation of photosynthates,<br />

and (D) mineral nutrition . Stomatal<br />

status and leaf resistance are considere d<br />

in relation to various aspects of gas exchange .<br />

Energy budgets, nutrient cycling, growth, and<br />

modeling of processes are considered b y<br />

other authors in this symposium, so are no t<br />

emphasized in this paper .<br />

Assimilation of CO 2<br />

and Respiration<br />

The principal emphasis in the IBP on assessment<br />

and understanding of dry matter<br />

production naturally focuses attention o n<br />

photosynthesis and the conditions and factors<br />

which influence its rate . Likewise the magnitude<br />

of respiratory losses is of complementary<br />

interest. Both short-term and long-term influences<br />

on the rates of these processes are o f<br />

importance in estimating seasonal and annual<br />

totals . Further, both inter- and intra-specifi c<br />

variability must be taken into account . Regardless<br />

of the factor under study, methods of<br />

individual measurement assume marked<br />

importance, and carrying over these measurements<br />

commonly made on portions of the<br />

plant to the entire forest stand is even mor e<br />

important. The latter aspect is considered i n<br />

the papers of this symposium on terrestrial<br />

modeling.<br />

Measuremen t<br />

The appearance in late 1971 of the detailed<br />

and authoritative book, Plant Photosyntheti c<br />

Production, Manual of Methods (Sestak,<br />

6atsky, and Jarvis, eds .), makes a detailed<br />

consideration of principles and techniques o f<br />

measurement unnecessary here . Thus a general<br />

consideration of methods and their applicability<br />

in the Coniferous Biome wil l<br />

suffice .<br />

Many useful studies have been performe d<br />

under controlled-environmental conditions ,<br />

using assimilation chambers of various types<br />

(Jarvis et al . 1971) . Such studies may be criticized<br />

on the basis that plants grown or hel d<br />

for study in controlled-environment rooms or<br />

boxes may behave differently from materia l<br />

growing under natural conditions . Thus a<br />

number of studies have been devoted to ga s<br />

exchange of conifers in the field (e .g., Botkin ,<br />

Woodwell, and Tempel 1970 ; Gentle 1963 ;<br />

Helms 1965, 1970 ; Hodges 1967 ; Hodges and<br />

Scott 1968; Kunstle 1971; Ungerson and<br />

Scherdin 1965; and Woodman 1971) . Also<br />

the extensive unpublished studies on Picea<br />

abies (L.) Karst of W . Koch' and of O . L .<br />

Lange and E . -0. Schulze 2 should be mentioned.<br />

To eliminate the influence of greenhouse<br />

or controlled-environment on the<br />

plants, several workers have used seedlings o r<br />

saplings brought in from the field or garden ,<br />

or made measurements on excised branches<br />

from field growing trees (Brix and Ebell 1969 ;<br />

Keller 1971 ; Parker 1963; Pisek, Larcher ,<br />

Moser, and Pack 1969 ; Pisek and Kemnitzer<br />

1968 ; and Poskuta 1968) .<br />

All of the measurements in the studies<br />

cited in the previous paragraph were made using<br />

assimilation chambers of various kinds .<br />

Difficulties and potential errors associate d<br />

with use of such enclosures are well covere d<br />

by Larcher (1969a) and Jarvis et al . (1971) .<br />

The major concerns pertain to excessive<br />

boundary layer resistances if stirring is inadequate,<br />

to significant departures of irradiatio n<br />

and energy balances from those of unenclose d<br />

foliage, and to errors inherent in the measuring<br />

itself .<br />

Techniques other than those using assimilation<br />

chambers are available (Larcher 1969a ,<br />

Denmead and Mcllroy 1971) . Harvest techniques<br />

give direct measurements of dry matte r<br />

accumulation, but must pertain with conifer s<br />

to time intervals of several weeks or months .<br />

Thus they cannot give information on responses<br />

to various controlling factors, nor d o<br />

they give information on respiratory losses .<br />

However, apart from their intrinsic value ,<br />

they may serve as an independent check on<br />

I <strong>Forest</strong> Botanical Institute, University of Munich ,<br />

Germany .<br />

2 Botanical Institute II, University of Wiirzburg ,<br />

Germany .<br />

212


the validity of extrapolating from assimilatio n<br />

chamber measurements to forest stands . Thus<br />

Botkin, Woodwell, and Tempel (1970) noted<br />

that values of production calculated fro m<br />

assimilation chamber data were about 30 per -<br />

cent above harvest values for the Brookhave n<br />

oak-pine forest . Meteorological techniques<br />

have been applied to conifer stands in a few<br />

cases (Baumgartner 1969, Denmead 1969 ,<br />

Kinerson 1971) . These can give values for en -<br />

tire stands without the influence of enclosures,<br />

but require large even stands, level<br />

topography, and uniform atmospheric conditions,<br />

thus measurement of either short or<br />

long term influences of factors affecting<br />

assimilation is very difficult. The use of<br />

meteorological data in combination wit h<br />

physiological information has potential fo r<br />

mathematical modeling and prediction of<br />

CO 2 assimilation (Reed and Webb 1972) .<br />

Difficulties arise in defining and measurin g<br />

the surfaces active in reception of light and i n<br />

CO 2 exchange. The general problems an d<br />

techniques for these assessments, particularl y<br />

in relation to broad-leafed plants and grasses ,<br />

are covered in the review of Kvet and Marshal l<br />

(1971) . The needle leaves of conifers usuall y<br />

have longitudinal rows of stomata on th e<br />

lower surface or on the interior surfaces i n<br />

fascicled species . Thus it is common to consider<br />

the "one-side" area as accounting fo r<br />

most of the gas exchange . Areas can be attained<br />

by tedious linear measuring and counting<br />

methods (Kvet and Marshall 1971 ,<br />

Madgwick 1964), but often dry weights o f<br />

foliage only have been reported to avoid thi s<br />

time-consuming effort. The introduction of<br />

the glass-bead technique (Thompson an d<br />

Leyton 1971) should encourage the determination<br />

of surface areas as well as dr y<br />

weights so that data can be compared between<br />

species . Further, because of the mathematical<br />

relations of gas diffusion and leaf<br />

resistances, most models of photosynthesi s<br />

require that values of photosynthesis b e<br />

expressed on a unit leaf area basis (Reed and<br />

Webb 1972) . The foliar distribution in crown s<br />

and stands must be obtained for extrapolation<br />

of gas exchange data, and several description s<br />

of such methods are available (Stephens 1969 ,<br />

Kinerson 1971) .<br />

Factors Affecting CO 2 Assimilation<br />

and Respiration<br />

In general, the principle of limiting factor s<br />

will apply to the relative importance of th e<br />

environmental and endogenous variables<br />

which control photosynthetic and respirator y<br />

rates. Thus in consideration of an individua l<br />

variable or factor, its effect may be masked at<br />

particular times by the limiting influence o f<br />

another factor. In the discussions of any individual<br />

factor which follows, it is implicit that<br />

the influences described are those of the<br />

factor when it is believed to be limiting the<br />

metabolic process (Stalfelt 1960) .<br />

Light Intensity<br />

Numerous studies in controlled-environmental<br />

systems and using excised branche s<br />

have shown that conifers exhibit typical<br />

limiting-factor response to light, with saturation<br />

dependent on the species and growing<br />

conditions (Pisek 1960) . For those species of<br />

principal interest in the Coniferous Biome a<br />

number of values for saturating light intensity<br />

have been reported (table 1) . These values are<br />

in general agreement with the range of<br />

1,850-3,200 ft-c stated by Polster (1967b) t o<br />

give maximal net photosynthesis in bot h<br />

broadleafed and coniferous trees, except for<br />

the value quoted by Kiinstle (1971) . Further<br />

verification of light responses under field conditions<br />

are needed for adequate modeling<br />

even of Douglas-fir, Pseudotsuga menziesii<br />

(Mirb.) Franco. Western hemlock, Tsuga<br />

he te rophylla (Rafinesque) Sargent, Pin us<br />

ponderosa Laws., and the Abies species are<br />

much in need of further study with respect t o<br />

light .<br />

Recent studies on photorespiration and its<br />

enhancement with increasing light intensity i n<br />

spruce species (Poskuta 1968, Ludlow an d<br />

Jarvis 1971b) indicate the need to understand<br />

the variations of this process in the species o f<br />

special interest in the Coniferous Biome . This<br />

information will not be necessary for our<br />

earlier empirical models, but will become<br />

essential for more theoretically based model s<br />

in the future as well as being necessary for<br />

estimates of gross productivity .<br />

213


Table 1 .-Saturating light intensities for net photosynthesis in various species of conifers<br />

Description of plan t<br />

material<br />

Saturating ligh t<br />

intensity<br />

Author<br />

- - - foot-candle'-- -<br />

Douglas-fir (Pseudotsuga menziesii (Mirb .) Franco) :<br />

65-day-old seedlings grown with 1,100 ft-c 3,000 Krueger and Ferrell ,<br />

long day ; temp 19°C day, 9°C night 196 5<br />

100-day-old seedlings grown at 18°C and 1,000 ft-c 2,500 Brix, 196 7<br />

1-year-old seedlings-outdoor grown ca . 2,300 Krueger and Ruth, 196 9<br />

4-year-old seedlings grown in cans in the garden 2,000-3,000 Fry, 196 5<br />

2-year-old seedlings-nursery grown 1,600 Hodges, 196 5<br />

Current year foliage from 10-year-old ca. 4,600 Kunstle, 197 1<br />

trees (upper crown July-Aug . )<br />

Western hemlock (Tsuga heterophylla (Rafinesque) Sargent) :<br />

2-year-old seedlings-nursery grown 1,400 Hodges, 1965<br />

1-year-old seedlings-outdoor grown Krueger and Ruth, 196 9<br />

in light shade 2,30 0<br />

in heavy shade 1,25 0<br />

Noble fir (Abies procera Rehd .) :<br />

2-year-old seedlings-nursery grown 2,000 Hodges, 196 5<br />

Sitka spruce (Picea sitchensis (Bong .) Carr .) :<br />

1-year-old seedlings-outdoor grown Krueger and Ruth, 196 9<br />

in light shade 2,30 0<br />

in heavy shade 1,25 0<br />

2-year-old seedlings-nursery grown 1,400 Hodges, 196 5<br />

3- to 4-year-old seedlings and branches 1,450-1,930 Ludlow and Jarvis,<br />

from 10-year-old forest trees<br />

1971b<br />

1 1 foot-candle (ft-c) = 10 .8 lux ; 6,750 ft-c = 1 cal cm-2 min-e ; 1 Wm-2 = 1 .43 x 10-3 cal cm-2 min-l .<br />

Temperature<br />

The concomitant responses of respiratio n<br />

and gross photosynthesis to temperature ,<br />

modification by light and moisture regime s<br />

and adaptation, as well as special influences o f<br />

low and high temperatures, all make it difficult<br />

to clearly indicate the influence of<br />

temperature alone on these processes (Krame r<br />

and Kozlowski 1960, Kozlowski and Kelle r<br />

1966, Larcher 1969b, Polster 1967b) . Thu s<br />

Pisek and his collaborators (1969) quote d<br />

rather considerable ranges of optimal temperatures<br />

at 10,000 Lux for a list of Europea n<br />

species. The total ranges for the conifers i n<br />

this list extended from 9° to 22°C, with the<br />

means showing a much narrower range of 12°<br />

to 15°C . Larcher (1969b) assembled data o n<br />

summer temperature optima for some 1 8<br />

conifers . These varied from the low range of<br />

12°-16°C for Pinus cembra L. to the hig h<br />

range of 20°-22°C for Abies balsamea (Mill.) .<br />

Data for species of our special interest i n<br />

the Coniferous Biome are not abundant . For<br />

Douglas-fir seedlings values of 20°-25°C a t<br />

saturating light (Krueger and Ferrell 1965) ,<br />

and 18°-20°C at 1,000 ft-c (Brix 1967) hav e<br />

been reported . For 10-year-old saplings of thi s<br />

species growing in the garden we noted a<br />

somewhat lower range of 12°-17°C . 3 This<br />

3 R . B . Walker and D . J . Salo . Unpublished data .<br />

214


may be explained by the observation of Brix<br />

and Ebell (1969) that dry matter production<br />

of Douglas-fir shoots in bud dormancy had a<br />

broad optimal range of 12° to 24°C, with<br />

maximum increase in stem diameter and i n<br />

total plant dry weight at 18°C .<br />

Respiration is believed to follow the van't<br />

Hoff Q 1 0 relationship up to about 60°C ,<br />

where enzyme inactivation ensues . This is supported<br />

by measurements of dark respiration ,<br />

and by the fall off in net photosynthesi s<br />

above 20°-25°C (Pisek 1960, Larcher 1969b) .<br />

Studying current-year needles of 10-year-ol d<br />

Douglas-fir, Kiinstle (1971) found the Q1 0 of<br />

respiration of fully expanded needles to be<br />

about 2.2, but to be about 4 .0 in buds in late<br />

spring. Also Brix and Ebell (1969) noted a<br />

steady decline in dark respiration of Douglas -<br />

fir needles with years of age. However, more<br />

intensive study of respiration not only o f<br />

needles but of non-green parts will be neede d<br />

in our Biome effort for modeling production .<br />

This is even more pertinent for the specie s<br />

other than Douglas-fir .<br />

Since the evergreens may function through -<br />

out the year, special interest has long existe d<br />

in their responses to low temperatures (Pise k<br />

1960) . The low-temperature compensatio n<br />

points for conifers in the winter are in th e<br />

range of -5° to -8°C, which is also the expected<br />

range of freezing of the needle s<br />

(Larcher 1969b) . Again considering species of<br />

special interest in the Coniferous Biome, th e<br />

low-temperature compensation point fo r<br />

Sitka spruce proved to be -5°C (Ludlow an d<br />

Jarvis 1971b) . Respiration is very low a t<br />

temperatures below 5°C, and in our climat e<br />

such temperatures usually predominate onl y<br />

when light intensities are very low . Thus with<br />

light often limiting, temperature influences o n<br />

photosynthesis are small, and net assimilatio n<br />

in Douglas-fir was near the maximum fo r<br />

short times even at 0°C if illumination wa s<br />

less than 500 ft-c (see footnote 3) . However ,<br />

the assimilation rate declined if a temperatur e<br />

under 2°C was maintained for an hour or<br />

more .<br />

Wintertime depressions in photosyntheti c<br />

capacity have been studied frequently in th e<br />

past (Pisek 1960) . These declines have bee n<br />

attributed to a combination of stomatal<br />

closure and a probable biochemical facto r<br />

(Parker 1963), to stomatal closure in Picea<br />

abies, Pinus sylvestris L., and Juniperus cornmunis<br />

L. in northern Sweden (Ungerson an d<br />

Scherdin 1965), to a breakdown of the photo -<br />

synthetic apparatus in Pinus taeda L. (Perry<br />

and Baldwin 1966), to a reversible effect<br />

associated with freezing of the needles an d<br />

presumed dehydration in Abies alba Mill .<br />

(Pisek and Kemnitzer 1968), to adverse needle<br />

water relations (Zelawski and Kucharska<br />

1967), and to a typical seasonal behavior in<br />

pines (McGregor and Kramer 1963) . Seasonal<br />

variations will be considered in more detail i n<br />

a subsequent section .<br />

There is some information on high temperature<br />

compensation points of photosynthesis ,<br />

although nothing pertaining particularly to<br />

species of Western United States . The si x<br />

European conifers studied at Innsbruck (Pisek<br />

et al . 1968) varied from 46° to 50°C in heat<br />

resistance, and from 37° to 41°C in the maximum<br />

temperature for net photosynthesis . If<br />

our species fall into similar ranges, as seem s<br />

likely, there is little reason to be concerned<br />

with this aspect because only very occasionally<br />

would air temperatures reach the<br />

compensation value for photosynthesis in our<br />

region. Only if leaf temperatures should<br />

greatly exceed air temperatures on hot day s<br />

would this be a matter of interest.<br />

Carbon Dioxide<br />

Under favorable conditions of light intensity,<br />

temperature and moisture, carbo n<br />

dioxide concentration may sharply limi t<br />

photosynthesis . Thus Koch (1969) reported<br />

some threefold increase in yield of severa l<br />

conifers from raising CO 2 concentration to<br />

5X the normal atmospheric level in controlle d<br />

environment experiments. Also Fry (1965 )<br />

recorded a measurable increase in net assimila -<br />

tion of Douglas-fir saplings when he increase d<br />

CO 2 from ambient (0 .035 percent) to 0 .0 5<br />

percent if light intensity was above 1,600 ft-c .<br />

Likewise Ludlow and Jarvis (1971b) foun d<br />

that net photosynthesis of Sitka spruce seedlings<br />

increased almost linearly with CO 2 concentration<br />

as it was varied from 0 to 0 .0 4<br />

percent, then continued to rise less steeply t o<br />

215


at least 0 .06 percent CO 2 . However, in th e<br />

forest it is not feasible to increase CO 2 concentration<br />

except for the natural contributions<br />

from the soil and litter . It is important<br />

to make periodic checks in the field using in -<br />

creased CO 2 in assimilation chambers to<br />

detect if the ambient CO 2 concentration is<br />

limiting, and to minimize the "draw down" of<br />

CO 2 in enclosures to avoid excessive deviations<br />

from assimilation rates of the unenclosed<br />

foliage. Also the atmospheric CO 2 concentration<br />

is known to be variable, and this<br />

should be monitored for taking into account<br />

any limiting effect on photosynthesis .<br />

Water Defici t<br />

A substantial literature documents the reductions<br />

in CO2 assimilation in various species<br />

caused by even moderate water deficits<br />

(Kozlowski and Keller 1966) . Some reductio n<br />

may occur in the afternoon with high transpirational<br />

rates even on well-watered soils ,<br />

but lowered soil moisture is the principal<br />

cause of severe reductions . Using 4-yearold<br />

Douglas-fir saplings in containers, Fry<br />

(1965) found that depletion of bulk soil<br />

water to a potential of -5 bars reduced ne t<br />

photosynthesis to zero (compensation point )<br />

and reduced respiration to about 10 percen t<br />

of normal. Since needle water potential<br />

dropped to about -30 bars, soil water potential<br />

at the root surfaces was probably considerably<br />

below -5 bars. Hodges (1965, 1967) ,<br />

studying 2-year-old Douglas-fir seedlings in<br />

the field, found net assimilation closely correlated<br />

with leaf water potential . Using 7- to<br />

9-year-old saplings of about 0 .5-0.75-m height ,<br />

Hinckley (1971) found net assimilation in<br />

Abies amabilis (Dougl.) Forbes and Abie s<br />

procera Rehd. to drop sharply with reduced<br />

water potential (pressure chamber method) o f<br />

the needles, reaching 10 percent of maximu m<br />

assimilation at about -10 bars needle water<br />

potential in the A . procera, and at about -1 3<br />

bars in the A. amabilis .<br />

Such declines in net assimilation caused b y<br />

reduced soil moisture and lowered leaf wate r<br />

potential are well recognized to be caused b y<br />

increases in stomatal and other leaf resistance s<br />

(Fry 1965, Slatyer 1967, Slavik 1971, Jarvis<br />

1971) . Considerable information has bee n<br />

accumulated on stomatal behavior in Douglas -<br />

fir (Fry 1965, Reed 1968, Phillips 1967) ,<br />

ponderosa pine4 (Lopushinsky 1969), an d<br />

true firs (Lopushinsky 1969, Hinckley an d<br />

Ritchie 1970). However, meager informatio n<br />

is available on actual values for stomatal an d<br />

other leaf resistances in species of specia l<br />

interest in the Coniferous Biome . Fry (1965 )<br />

calculated stomatal resistance in Douglas-fi r<br />

saplings from stomatal infiltration pressures .<br />

From these pressures ranging from 0 .25 to<br />

1.65 atm he calculated stomatal resistances o f<br />

2.3 to 32 sec cm-1 for CO 2 , and additional<br />

leaf resistances of 3 .7 to 83 sec cm-1 . Ree d<br />

(1972) found general agreement with thes e<br />

values. Ludlow and Jarvis (1971b) reported<br />

stomatal resistances in Sitka spruce seedling s<br />

in excess of 20 sec cm-1 at very low irradiance<br />

or at very low temperatures, with normal<br />

values of about 4 sec c m ' . Their data showed<br />

mesophyll resistances of 20 to 30 sec cm-1 or<br />

more at very low irradiances and at very lo w<br />

temperatures, with normal values of about 6<br />

to 8 sec cm- ' . Leaf resistance is such an<br />

important factor that understanding stomatal<br />

and other components of resistance in all species<br />

of special interest in the Biome is necessary<br />

for the development of reasonable<br />

models of photosynthesis and transpiration .<br />

Mineral Nutrition<br />

A number of experiments have established<br />

that mineral deficiency may limit photosynthesis<br />

(Kozlowski and Keller 1966, Baule and<br />

Fricker 1967) . In the Pacific Northwest,<br />

nitrogen deficiency is common in forests, thu s<br />

making determination of the influences o f<br />

this element very important in the Coniferou s<br />

Biome program. The stands at the Cedar Rive r<br />

Watershed show incipient N deficiency, s o<br />

care will be taken to assess the influence o f<br />

this element. Keller (1971) recently reporte d<br />

a 50 percent increase in CO2 assimilation i n<br />

needles of Picea abies seedlings with increas e<br />

in N content from 0 .41 to 1 .11 percent . Also<br />

Brix (1971) recently measured increases i n<br />

4 A . P . Drew, L. G . Drew, and H . C . Fritts . Unpublished<br />

data .<br />

216


net assimilation in excised branches fro m<br />

24-year-old Douglas-fir trees which had been<br />

fertilized with N . However, the increases wer e<br />

noted only at light intensities of 2,000 ft- c<br />

and above .<br />

Diurnal and Seasonal Pattern s<br />

(including adaptation )<br />

Diurnal cycles or fluctuations in light, temperature,<br />

water vapor pressure gradients, an d<br />

other external factors are interactive with<br />

internal factors such as endogenous rhythm o f<br />

stomatal behavior, recovery from water deficit ,<br />

and translocation. Therefore the response of a<br />

tree day by day is very complex .<br />

Gentle (1963) and Helms (1965) studie d<br />

diurnal responses of 38-year-old Douglas-fir i n<br />

detail over a 4-year period . Superimposed o n<br />

the daily responses to light and darkness were<br />

short-term and longer term fluctuations attributable<br />

to varying temperature, radiation and<br />

cloud cover, relative humidity, water deficit,<br />

and undetermined influences . Midday depressions<br />

were common in the summer but also<br />

occurred in cool autumn weather . Similar detailed<br />

studies of diurnal behavior of the other<br />

species important in the Biome are lacking .<br />

The diurnal endogenous changes in stomatal<br />

aperture have been studied in several<br />

species. Fry (1965) noted that saplings grow n<br />

from seed gathered at Pack Demonstratio n<br />

<strong>Forest</strong>, La Grande, Washington, did no t<br />

exhibit any closing of the stomata during th e<br />

day or night if water status was favorable .<br />

This was largely confirmed by Reed (1968) ,<br />

although he observed a slight closing tendenc y<br />

at night. Phillips (1967) studied 38-year-old<br />

Douglas-fir trees at Pack Demonstration <strong>Forest</strong><br />

and found that the stomata of leaves i n<br />

the lower crown closed at night. A geographic<br />

or provenance difference in stomatal response<br />

of Douglas-fir was demonstrated in the field<br />

in southern Oregon, where stomata were open<br />

at night in the spring and early summer, but<br />

closed at night from July through September<br />

(Reed 1972). Drew, Drew, and Fritts (see<br />

footnote 4) and Lopushinsky (1969) independently<br />

observed that the stomata of Pinus<br />

ponderosa are closed at night . Abies amabilis<br />

growing in the field was observed to hav e<br />

more open stomata in early morning than did<br />

Abies procera, indicating that the stomata of<br />

A. amabilis were probably more open durin g<br />

the night (Hinckley 1971). The stomatal<br />

behavior of western hemlock is unknown an d<br />

needs attention .<br />

The influence of season of the year ha s<br />

already been touched on in connection with<br />

winter depression of assimilation . In a more<br />

general context, seasonal patterns of CO 2<br />

assimilation were studied by Gentle (1963)<br />

and by Helms (1964, 1965) in the 38-year-ol d<br />

Douglas-fir stand mentioned above . As would<br />

be expected, winter assimilation was low and<br />

quite variable, depending on the light an d<br />

temperature prevailing. Spring and autum n<br />

performances were good, although somewha t<br />

below the summer . Higher temperatures presumably<br />

made respiration a good deal highe r<br />

in the summer than in spring and autumn and .<br />

particularly than in the winter . Also Wood -<br />

man (1968, 1.971) studied net assimilation in<br />

one of the trees of this lame stand during th e<br />

growing season of 1967. A somewhat different<br />

pattern can be expected in conifers<br />

growing in regions with very dry summers ,<br />

with much Of the yearly total of photosynthesis<br />

occurring during the winter and spring .<br />

Such studies of seasonal patterns are imperative<br />

for adequate models of production in<br />

forest stands . This is particularly true becaus e<br />

a considerable number of variables in addition<br />

to those of the physical environment have<br />

marked influences (Kozlowski and, Kelle r<br />

1966). These include morphological variatio n<br />

(sun and shade foliage, stomatal distribution ,<br />

cuticular thickness, and others), adaptatio n<br />

with respect to light intensity and temperature<br />

in par icular, interspecific and intraspecific<br />

variations, shading, dormancy, age of<br />

foliage, position in the crown, and the natur e<br />

of past seasons . Some information on these<br />

features is available for Douglas-fir. However<br />

very limited knowledge, gained mostly fro m<br />

studies with seedlings, is available in this connection<br />

for the other species of special concern<br />

in the Coniferous Biome programwestern<br />

hemlock, ponderosa pine, and one of<br />

the true firs l Hodges and Scott 1968, Kruege r<br />

and Ruth 1969, Ludlow and Jarvis 1971b t<br />

217


Pharis et al. 1967). Again Helms (1970) ha s<br />

studied CO 2 assimilation in ponderosa pine<br />

during the summer in the natural environment,<br />

and envisions extending this study t o<br />

all seasons of the year. Intensive work on CO 2<br />

assimilation of all of the other species named<br />

above from the standpoints of environmenta l<br />

factors and the other influences on their behavior<br />

will be needed for realistic modeling o f<br />

their photosynthetic production .<br />

Water Relationships<br />

Although water deficit was taken up in th e<br />

previous section as a potential limiting facto r<br />

in CO 2 assimilation and respiration, the general<br />

subject of water relationships is of concern<br />

in water balance, nutrient cycling, an d<br />

other soil-plant-atmosphere relations, as wel l<br />

as in its effects on plant metabolism . Also th e<br />

balance between water absorption and transpiration<br />

determines any water deficit in th e<br />

plant. Water relationships of woody plants<br />

were thoroughly reviewed by Polster (1967a) .<br />

Water Absorptio n<br />

The factors affecting water uptake in<br />

woody plants were discussed in detail i n<br />

Kramer and Kozlowski (1960) . In the usuall y<br />

well-drained and well-aerated soils of the<br />

western forests, temperature of the soil an d<br />

roots and the level of soil moisture are th e<br />

principal factors affecting water uptake in th e<br />

root zone. These factors are being regularl y<br />

monitored at the intensive study sites of th e<br />

Biome . Their influence will be of most concern<br />

in the magnitude of water deficits . Sinc e<br />

water uptake is wholly passive during times o f<br />

high water use, increased resistance to uptake<br />

in the root zone can be expected to enhanc e<br />

water deficits . Further, rate of water uptake<br />

and conduction may exert some effect o n<br />

mineral uptake (Slatyer 1967) . Investigation s<br />

of water uptake will probably not be a part o f<br />

the Biome studies, because of the difficultie s<br />

of measuring this quantity in large soil-roote d<br />

plants, and the near equivalence of absorptio n<br />

rates to transpiration rates .<br />

Water Conduction<br />

For woody plants, this aspect has bee n<br />

thoroughly reviewed, including special features<br />

of conifers, by several authors (Hube r<br />

1956, Kramer and Kozlowski 1960, Zimmermann<br />

and Brown 1971) . With reference to th e<br />

objectives of the Coniferous Biome, interes t<br />

in conduction pertains to the magnitude o f<br />

negative sap pressures (as a measure of water<br />

deficit), to its use as an indicator of transpirational<br />

fluctuations where actual transpirational<br />

measurements can not be made readily ,<br />

to intra-tree water adjustments includin g<br />

potential storage in stems and branches, an d<br />

to its influence on rate of translocation of<br />

mineral nutrients . Nonetheless, rates of conduction<br />

will be inferred of necessity for ou r<br />

models from transpirational rates . However ,<br />

some indication of the nature of potential<br />

storage of water in stems or branches may b e<br />

attained from measurements using sap velocity<br />

flow meters and by study of intra-crown<br />

variation in the sap pressure (Scholander technique)<br />

(Waring and Cleary 1967) of stems an d<br />

needles (Hinckley 1971, Ritchie 1971) .<br />

Transpiratio n<br />

The predominant interest in water relation -<br />

ships is with transpiration both from a hydro -<br />

logic viewpoint concerned with total plan t<br />

use, and from its influence on water deficit s<br />

and on resistance to water vapor and CO 2<br />

transfer. This major interest has resulted i n<br />

substantial work on the transpiration of conifers,<br />

and considerable attention to species o f<br />

interest in the Coniferous Biome . Thes e<br />

studies will be considered below under th e<br />

headings of measurement, influence of environmental<br />

factors, effects of leaf temperature<br />

and resistances, and the significance of<br />

water deficits .<br />

Measuremen t<br />

Although there are various methods fo r<br />

measurement of transpiration, in the fiel d<br />

with large plants the air-flow method using<br />

leaf enclosures, lysimeter systems, meteorological<br />

techniques, and tritiated water<br />

methods are most feasible. The air-flo w<br />

218


method, using psychrometers or infra-red ga s<br />

analyzers as the detecting instruments, is commonly<br />

applied to plant material enclosed i n<br />

assimilation chambers . The possible artifact s<br />

of the systems pointed out for CO 2 assimilation<br />

measurements (Jarvis et al . 1971) are<br />

even more critical in the case of transpirational<br />

assessments . This is true because of th e<br />

role of energy balance and resulting lea f<br />

temperature in determining the vapor pressur e<br />

gradient from leaf to atmosphere . Also<br />

adsorption and condensation problems ofte n<br />

make measurements difficult and less accurat e<br />

during cool moist periods either diurnal or seasonal,<br />

although rates are characteristically low<br />

under such conditions . The lysimeter technique<br />

is suited to diurnal and seasonal studie s<br />

without the artificiality of enclosing th e<br />

foliage (Fritschen 1972) . Root disturbance<br />

occurs in installation, but this should hav e<br />

little effect on transpiration and passive up -<br />

take of water . Meteorological methods, whic h<br />

measure evapotranspiration, also maintai n<br />

natural conditions around the foliage, bu t<br />

necessitate sizable stands of homogenous<br />

vegetation and uniform air conditions for bes t<br />

measurements. The tritiated water metho d<br />

(Kline et al . 1972) yields information on transpirational<br />

activity only over periods of hour s<br />

and days, but its advantage over the other<br />

methods lies in its applicability to very larg e<br />

trees, such as the 75-m-tall old-growt h<br />

Douglas-firs in the H . J. <strong>Andrews</strong> <strong>Experimental</strong><br />

<strong>Forest</strong>. During certain periods o f<br />

1972, all four of these methods of transpirational<br />

measurement will be compared at th e<br />

Thompson Research Center (Cedar River) .<br />

The lysimeter study will be continuou s<br />

throughout the 1972 spring through autum n<br />

season, and the simultaneous measurement o f<br />

transpired water in the gas stream from the<br />

assimilation cuvettes will permit estimation of<br />

leaf resistance as well as giving data on transpirational<br />

rates .<br />

Influence of Environmental Factors<br />

General discussions of these factors an d<br />

their influences have been given by Stalfelt<br />

(1956), Kramer and Kozlowski (1960), an d<br />

Kozlowski and Keller (1966) . Emphasis here<br />

will be on more recent studies with conifers<br />

and especially with species of special concern<br />

in the Coniferous Biome .<br />

The dominant influence of the vapor pressure<br />

gradient from leaf to air on transpirational<br />

rate when stomata are open has long<br />

been recognized. This was verified by Ritchie<br />

(1971) in a study with Abies amabilis and A .<br />

procera of about 14-m height, using vapor<br />

pressure gradient and also saturation deficit of<br />

the atmosphere . On a statistical basis, th e<br />

latter value accounted for 56 percent of th e<br />

variation in transpirational rate on a seasona l<br />

basis .<br />

Depletion of soil moisture has also lon g<br />

been recognized as a major factor in the control<br />

of transpiration . This was verified by<br />

Mullerstael (1968) using three species of pine ,<br />

as well as other evergreens, and he als o<br />

showed marked species differences . Hinckley<br />

(1971) showed great reduction in transpiration<br />

(sap velocity) in Abies amabilis and A .<br />

procera with soil water depletion. Reed<br />

(1972) developed a computer simulation o f<br />

transpiration in the field . He showed that lo w<br />

vapor pressure gradients limited transpiratio n<br />

in the spring, but increased stomatal resistance<br />

resulting from depleted soil moisture an d<br />

water deficits limited transpiration during th e<br />

summer. He also showed that the availabl e<br />

soil moisture can vary greatly from year t o<br />

year, which can cause considerable seasonal<br />

and annual differences in total transpiration .<br />

Although moderate air movement is commonly<br />

recognized as a factor which increase s<br />

transpiration because boundary layers are reduced,<br />

few studies have been conducted with<br />

high air velocities. Tranquillini (1969) varied<br />

wind velocity from 0 .5 to 20 m sec-1 , and<br />

observed that net assimilation of Pinu s<br />

cembra increased somewhat up to about 4 m<br />

sec-t , then declined . In Picea abies the decline<br />

started at 1 .5 m sec-t . Caldwell (1970), working<br />

with the same Pinus cembra plants ,<br />

showed that stomatal aperture and transpiration<br />

rate were only slightly reduced by hig h<br />

wind speeds, but photosynthesis was reduce d<br />

considerably because of changes in needle display<br />

to the light. High wind would be expected<br />

to be of importance in the Coniferou s<br />

Biome only in extreme habitats .<br />

219


Effects of Leaf Temperature an d<br />

Leaf Resistances<br />

Leaf temperature establishes the vapo r<br />

pressure of water in the intercellular spaces o f<br />

the leaf, thus it is imperative to have goo d<br />

measurements of this value to accompan y<br />

transpirational (and net assimilation) estimations<br />

. Methods for determination of lea f<br />

temperature have been recently described i n<br />

detail (Perrier 1971) .<br />

Although transpirational rates may approach<br />

potential evaporation with favorabl e<br />

soil and leaf water status, commonly leaf<br />

resistances markedly impede water loss . This<br />

was well demonstrated by Waggoner an d<br />

Turner (1971) in a study of transpiration in<br />

Pinus resinosa Ait. with both natural and<br />

artificially induced stomatal closure markedly<br />

reducing transpiration . The separation of stomatal<br />

and other leaf resistance is very desirable<br />

when feasible (Jarvis 1971) . The influence<br />

of wax in the stomatal pores in reducin g<br />

transpiration in Sitka spruce was noted by<br />

Jeffree et al . (1971) . Similar wax accumulations<br />

increasing with age of needles i n<br />

Douglas-fir have recently been observed o n<br />

material collected from the 35-year-old tree s<br />

at the Thompson Research Center (Cedar<br />

River) . 5 In the Coniferous Biome studies ,<br />

more information is needed on these resistances<br />

in all species for successful productio n<br />

models .<br />

Significance of Water Deficit s<br />

Conifer species differ in their ability to<br />

maintain photosynthetic production in th e<br />

presence of water deficits, and in their abilit y<br />

to control water loss under similar environmental<br />

conditions (Hodges and Scott 1968) .<br />

The status of woody plants can be assesse d<br />

over the season by periodic tests of the pre -<br />

dawn pressure chamber value (Waring an d<br />

Cleary 1967), and this used in establishin g<br />

moisture gradients and ecological tolerance s<br />

and distributions. In the Coniferous Biom e<br />

such assessments are needed throughout th e<br />

growing season over wide geographic areas<br />

5 P. Machno and K . L. Reed. Unpublished data<br />

with all species of primary interest, and this<br />

program is already well under way (Waring et<br />

al. 1972) .<br />

Translocation<br />

of Photosynthates<br />

and Relations to Growth<br />

The translocation of photosynthate in<br />

woody plants was thoroughly reviewed b y<br />

Kozlowski and Keller (1966), and very recently<br />

discussed in detail by Zimmerman n<br />

and Brown (1971) .<br />

Among the conifers the pines, especiall y<br />

Pinus resinosa, have received the greatest attention<br />

(Rangnekar et al . 1969, Dickman n<br />

and Kozlowski 1970) . The former group ,<br />

working with 15-year-old trees in a plantation,<br />

found that each branch appeared to b e<br />

self-supporting, contributing more liberally t o<br />

its own growth than to the tree leader, whic h<br />

probably expands in part from reserve carbohydrates<br />

. Further, they suggested that th e<br />

products of cambial growth are largely de -<br />

rived from current photosynthate . The results<br />

of Dickmann and Kozlowski (1970), wh o<br />

labeled the 1-year-old needles of second-orde r<br />

branches of 20-year-old trees, are in general<br />

agreement with this, as they found that recovery<br />

from the various sinks was high and i n<br />

the order : 2d-year cones > terminal needles ><br />

lateral needles > terminal internode > latera l<br />

internodes > 1-year-old wood .<br />

Ross (1972) labeled different-aged branc h<br />

segments of 9-year-old Douglas-fir trees o f<br />

about 6-m height . He likewise was able t o<br />

discern source-sink relationships. The proportion<br />

of 14 CO 2 exported from needles was a<br />

function of their age, of the attractive "pull"<br />

of external sinks, and inversely of water<br />

stress . One-year-old needles exported a large r<br />

proportion of their photosynthate than di d<br />

current-year needles even late in the growin g<br />

season . The attractive "pull" of the sink s<br />

decreased in the order: elongating ne w<br />

1st-order internodes and needles, elongatin g<br />

new 2d-order internodes and needles, an d<br />

stem . The fact that 1-year-old needles mainly<br />

supplied adjacent new shoots and thus ex -<br />

220


ported little photosynthate basipetally in contrast<br />

with the preferential translocation of<br />

photosynthate toward the stem by 2-year-ol d<br />

needles may be explained by the ability of a<br />

sink such as the stem to mobilize photosynthate<br />

falling off with distance in a n<br />

approximately logarithmic manner .<br />

These studies give clues to the relationships<br />

between CO 2 assimilation in the leaves an d<br />

the formation and differentiation of new cells<br />

in apical and cambial growth . These relationships<br />

are not simple, and care must be take n<br />

to avoid oversimplified connections . For<br />

example, net assimilation rates can give an<br />

index of amounts of photosynthate availabl e<br />

for growth only if position and age of th e<br />

foliage studied are taken into account (Ros s<br />

1972). With caution, however, net assimilation<br />

rates taken in different parts of th e<br />

crown can be used to estimate daily mean<br />

assimilation and the total photosynthate production,<br />

on the basis of the whorls o f<br />

branches of different heights (Woodman<br />

1968, 1971) .<br />

In the Coniferous Biome studies, effort s<br />

must be made to bring together net assimilation,<br />

translocation, and terminal and lateral<br />

growth into a workable model . It may be<br />

necessary in such efforts to take into accoun t<br />

growth inhibitors and promoters in such a<br />

model, since Lavender and Hermann (1970 )<br />

have pointed out the importance of thes e<br />

substances in their studies of light and photo -<br />

period effects in Douglas-fir . Although there<br />

is information to build on with respect to re d<br />

pine and Douglas-fir (see above), further wor k<br />

will be necessary with Douglas-fir to utiliz e<br />

the information effectively . Clearly, specifi c<br />

studies of translocation and growth in wester n<br />

hemlock, ponderosa pine, and a true fir ar e<br />

needed .<br />

Mineral Nutrition<br />

A number of books cover the general field<br />

of mineral nutrition and specifically that of<br />

forest tree species (Baule and Fricker 1967 ,<br />

Bengtson 1968, Epstein 1972) . The aspects of<br />

particular concern in the Coniferous Biom e<br />

are mineral uptake and cycling, and the role<br />

of mineral elements in metabolic processes .<br />

Mineral cycling studies in second-growt h<br />

Douglas-fir forests have been carried out for<br />

over a decade (Grier and Cole 1972) . Mineral<br />

cycling is important in the nutrition of th e<br />

trees since limited nutrient capital is available .<br />

These studies need to be extended to th e<br />

other species of special importance in th e<br />

Coniferous Biome program .<br />

The importance of nitrogen in chlorophyll<br />

production and the close correlation between<br />

chlorophyll contents and CO 2 uptake were<br />

pointed out by Keller and Wehrmann (1967 )<br />

with reference to Picea abies and Pinus sylvestris<br />

. The widespread occurrence of nitrogen<br />

deficiency in Douglas-fir in western Washington<br />

and Oregon (Gessel et at . 1965) give s<br />

reason for careful attention to this element i n<br />

studies of coastal Douglas-fir in the Coniferous<br />

Biome . Also it will be wise to make a<br />

sufficiently broad spectrum of mineral analyses<br />

to detect low levels of other element s<br />

which might be limiting photosynthesis or<br />

growth .<br />

The foregoing indicates that a large back -<br />

ground of information on terrestrial production<br />

processes in woody plants is already<br />

available. With respect to Douglas-fir, the<br />

species currently receiving most emphasis i n<br />

the Coniferous Biome, the background i s<br />

appreciable and useful, but often lacking i n<br />

specific data needed for the construction of<br />

mathematical models . These deficiencies must<br />

be made up for adequate modeling of processes<br />

in this species. With other species of<br />

special concern in the Biome-western hemlock,<br />

ponderosa pine, and true firs-muc h<br />

more background information as well as<br />

specific data is needed in order to develo p<br />

effective production models .<br />

A cknowledgments<br />

The work reported in this paper was sup -<br />

ported in part by the University of Washingto n<br />

and in part by National Science Foundatio n<br />

Grant No. GB-20963 to the Coniferous Fores t<br />

Biome, U .S. Analysis of Ecosystems, Inter -<br />

national Biological Program . This is Contribution<br />

No . 37 to the Coniferous <strong>Forest</strong> Biome .<br />

221


Literature Cited<br />

Baule, H., and C. Fricker. 1967 . The fertilizer<br />

treatment of forest trees . (Transl . from<br />

German .) 259 p. Munich : BLV Verlagsges .<br />

Baumgartner, A. 1969. Meteorological approach<br />

to the exchange of CO 2 betwee n<br />

the atmosphere and vegetation, particularl y<br />

forest stands . Photosynthetica 3(2) :<br />

127-249 .<br />

Bengtson, G . W . (ed .) . 1968 . <strong>Forest</strong> fertilization<br />

: theory and practice . 306 p . Muscl e<br />

Shoals, Ala . : Tenn. Valley Authority .<br />

Botkin, D., G. M . Woodwell, and Neal<br />

Tempel. 1970. <strong>Forest</strong> productivity estimated<br />

from carbon dioxide uptake .<br />

Ecology 51 : 1057-1060 .<br />

Brix, H . 1967 . An analysis of dry matter production<br />

of Douglas-fir seedlings in relatio n<br />

to temperature and light intensity . Can . J .<br />

Bot. 45 : 2063-2072 .<br />

1971 . Effects of nitrogen fertilization<br />

on photosynthesis and respiration in<br />

Douglas-fir. <strong>Forest</strong> Sci . 17 : 407-414 .<br />

and L . F. Ebell . 1969 . Effects of<br />

nitrogen fertilization on growth, leaf area ,<br />

and photosynthesis rate in Douglas-fir . <strong>Forest</strong><br />

Sci. 15 : 189-195 .<br />

Caldwell, M. M . 1970. Plant gas exchange a t<br />

high wind speeds . Plant Physiol. 46(4) :<br />

535-537 .<br />

Denmead, O . T. 1969. Comparative micrometeorology<br />

of a wheat field and a forest<br />

of Pin us radiata . Agric. Meteorol. 6 :<br />

357-371 .<br />

and I. C. Mcllroy. 1971. Measurement<br />

of carbon dioxide exchange in th e<br />

field . In Z . Sestak, J . 6atsky, and P . G .<br />

Jarvis (eds.), Plant photosynthetic production,<br />

manual of methods, p . 467-516 . Th e<br />

Hague : Dr . W . Junk N .V . Publishers .<br />

Dickmann, D . I ., and T. T. Kozlowski. 1970 .<br />

Mobilization and incorporation of photo -<br />

assimilated 14 C by growing vegetation an d<br />

reproductive tissues of adult Pinus resinosa<br />

trees. Plant Physiol . 45 : 284-288 .<br />

Epstein, E . 1972. Mineral nutrition of plants :<br />

principles and perspectives, 412 p . New<br />

York: John Wiley and Sons .<br />

Fritschen, L . J. 1972. The lysimeter installation<br />

on the Cedar River watershed . In<br />

Proceedings-research on coniferous forest<br />

ecosystems-a symposium, p . 255-260, illus.<br />

Pac. Northwest <strong>Forest</strong> & Range Exp . Stn . ,<br />

Portland, Oreg .<br />

Fry, K . E. 1965. A study of transpiration an d<br />

photosynthesis in relation to the stomatal<br />

resistance and internal water potential i n<br />

Douglas-fir. 192 p. Ph.D. thesis on file ,<br />

Univ . Wash ., Seattle .<br />

Gentle, S . W. 1963 . The effect of local weather<br />

conditions on the assimilation of carbo n<br />

dioxide by Douglas-fir in western Washing -<br />

ton. 122 p . Ph .D. thesis on file, Univ .<br />

Wash., Seattle .<br />

Gessel, S . P., T. N . Stoate, and K . J. Turnbull ,<br />

1965 . The growth behavior of Douglas-fi r<br />

with nitrogenous fertilizer in western Washington<br />

. Res . Bull. No. 1, 204 p . Coll. For . ,<br />

Univ . Wash., Seattle .<br />

Grier, Charles C ., and Dale W . Cole. 1972 .<br />

Elemental transport changes occurring during<br />

development of a second-growth Douglas-fir<br />

ecosystem . In Proceedings-research<br />

on coniferous forest ecosystems-a symposium,<br />

p . 103-113, illus . Pac. Northwest<br />

<strong>Forest</strong> & Range Exp . Stn., Portland, Oreg .<br />

Helms, J. A. 1964. Apparent photosynthesi s<br />

of Douglas fir in relation to silvicultural<br />

treatment . <strong>Forest</strong> Sci . 10: 432-442 .<br />

1965. Diurnal and seasonal patterns<br />

of net assimilation in Douglas-fir, a s<br />

influenced by environment . Ecology 46 :<br />

698-708 .<br />

1970. Summer net photo -<br />

synthesis of ponderosa pine in its natura l<br />

habitat. Photosynthetica 4(3) : 243-253 .<br />

Hinckley, T . M. 1971 . Plant water stress an d<br />

its effect on ecological patterns of behavior<br />

in several Pacific Northwest conifer species ,<br />

177 p . Ph .D . thesis on file, Univ . Wash . ,<br />

Seattle .<br />

and G . A. Ritchie . 1970 . Within -<br />

crown patterns of transpiration, wate r<br />

stress, and stomatal activity in Abies<br />

amabilis . <strong>Forest</strong> Sci . 16: 490-492 .<br />

222


Hodges, J . D. 1965. Photosynthesis in forest<br />

tree seedlings of the Pacific Northwest<br />

under natural environmental conditions .<br />

177 p. Ph.D. thesis on file, Univ . Wash . ,<br />

Seattle .<br />

1967. Patterns of photosynthesis<br />

under natural conditions . Ecology<br />

48: 234-242 .<br />

and D. R. M. Scott. 1968 .<br />

Photosynthesis in seedlings of six conifer<br />

species under natural environmental conditions.<br />

Ecology 49 : 973-980 .<br />

Huber, B . 1956. Die Gefassleitung. In W .<br />

Ruhland (ed .), Handb . Pflanzenphysiol . III :<br />

541-582 .<br />

Jarvis, P. G. 1971 . The estimation of resistances<br />

to carbon dioxide transfer . In Z .<br />

Sestak, J. Jatsky, and P . G. Jarvis (eds .) ,<br />

Plant photosynthetic production, manua l<br />

of methods, p. 566-631 . The Hague : Dr. W .<br />

Junk N .V. Publishers .<br />

J. Catsky, F. E. Eckardt, W .<br />

Koch, and D. Koller. 1971 . General principles<br />

of gasometric methods and the mai n<br />

aspects of installation design . In Z . Sestak ,<br />

J. Catsky, and P . G. Jarvis (eds.), Plant<br />

photosynthetic production, manual of<br />

methods, p . 49-110 . The Hague : Dr. W .<br />

Junk N .V . Publishers .<br />

Jeffree, C. E ., R. P. C. Johnson, and P . G .<br />

Jarvis. 1971 . Epicuticular wax in the stomatal<br />

antechamber of Sitka spruce and its<br />

effects on the diffusion of water vapour<br />

and carbon dioxide. Planta 98 : 1-10 .<br />

Keller, Theo. 1971. Der Einfluss der Stickstoffernahrung<br />

auf den Gaswechsel der<br />

Fichte. Allg. Forst- u . Jagdzt. 142(4) :<br />

89-93 .<br />

and J. Wehrmann . 1963 . C02<br />

assimilation, Wurzelatmung and Ertrag vo n<br />

Fichten and Kiefersamlingen bei unterschiedlichen<br />

Mineralstoffernahrung. Mitt .<br />

Schweiz . Anst. Forstl. Versuchsw . 39(4) :<br />

217-242 .<br />

Kinerson, R .,Jr. 1971 . Selected aspects of gas<br />

and energy exchange in a young, matur e<br />

Douglas-fir stand . 91 p. Ph.D. thesis on file ,<br />

Univ . Wash ., Seattle .<br />

Kline, J . R ., M . L. Stewart, and C . F. Jordan .<br />

1972 . Estimation of biomass and transpiration<br />

in coniferous forests using tritiated<br />

water . In Proceedings-research on coniferous<br />

forest ecosystems-a symposium, p .<br />

159-166, illus. Pac. Northwest <strong>Forest</strong> &<br />

Range Exp. Stn ., Portland, Oreg .<br />

Koch, W. 1969 . Untersuchungen caber di e<br />

Wirkung von CO 2 auf die Photosynthese<br />

einiger Holzgewachse unter Laboratoriumsbedingungen.<br />

Flora 158 : 402-428 .<br />

Kozlowski, T., and Theo. Keller . 1966. Food<br />

relations of woody plants . Bot. Rev . 32(4) :<br />

293-382 .<br />

Kramer, P . J., and T . Kozlowski . 1960 . Physiology<br />

of trees. 642 p . New York : McGraw-<br />

Hill Co .<br />

Krueger, K . W., and W . K. Ferrell . 1965 . Comparative<br />

photosynthetic and respiratory<br />

responses to temperature and light by Pseudotsuga<br />

menziesii var. menziesii and var .<br />

glauca seedlings . Ecology 46 : 794-801 .<br />

and R . H. Ruth. 1969 . Comparative<br />

photosynthesis of red alder, Douglas -<br />

fir, Sitka spruce, and western hemloc k<br />

seedlings . Can . J. Bot . 47: 519-528 .<br />

Kunstle, E . 1971. Der Jahresgang de s<br />

CO 2 -Gaswechsel von einj ihrigen Douglasien-trieben<br />

in einem 20 ja"hrige m<br />

Bestand. Allg. Forst- u. Jagdzt. 142(4) :<br />

105-108 .<br />

Kvet, J ., and J. K. Marshall. 1971 . Assessment<br />

of leaf area and other assimilating surfaces .<br />

In Z . Sestak, J. Catsky, and P. G . Jarvis<br />

(eds.), Plant photosynthetic production ,<br />

manual of methods, p . 517-555 . The<br />

Hague : Dr . W. Junk N .V. Publishers .<br />

Larcher, W. 1969a. Physiological approache s<br />

to the measurement of photosynthesis i n<br />

relation to dry matter production by trees .<br />

Photosynthetica 3(2) : 150-166 .<br />

1969b . The effect of environmental<br />

and physiological variables on the<br />

carbon dioxide gas exchange of trees .<br />

Photosynthetica 3(2) : 167-198 .<br />

Lavender, D. P., and R. K. Hermann. 1970 .<br />

Regulation of the growth potential of<br />

Douglas-fir seedlings during dormancy .<br />

New Phytol . 69 : 675-694 .<br />

Leyton, L. 1958 . The mineral requirements of<br />

forest plants . In W . Ruhland (ed .), Handb .<br />

Pflanzenphysiol . IV : 1026-1039 .<br />

223


Lopushinsky, W . 1969. Stomatal closure i n<br />

conifer seedlings in response to leaf moisture<br />

stress . Bot . Gaz . 130 : 258-263 .<br />

Ludlow, M . M ., and P . J. Jarvis. 1971a . Methods<br />

for measuring photorespiration i n<br />

leaves . In Z . Sestak, J. Catsky, and P . G .<br />

Jarvis (eds .), Plant photosynthetic production,<br />

manual of methods, p . 294-315 . Th e<br />

Hague : Dr. W . Junk N .V . Publishers .<br />

and P . J. Jarvis. 1971b . Photo -<br />

synthesis in Sitka spruce . 1 . General characteristics<br />

. J . Appl . Ecol . 8(3) : 925-953 .<br />

McGregor, W . H. D ., and P . J. Kramer . 1963 .<br />

Seasonal trends in rates of photosynthesi s<br />

and respiration of loblolly pine and whit e<br />

pine seedlings . Am . J. Bot. 50(8) : 760-765 .<br />

Madgwick, H . A. 1964 . Estimation of surface<br />

area of pine needles with special reference<br />

to Pinus resinosa . J. For . 62 : 636 .<br />

Miillerstael, H . 1968 . Untersuchungen uber<br />

den Gaswechsel zweijahriger Holzpflanze n<br />

bei fortschreitender Bodenaustrocknung .<br />

Beitr. Biol. Pflanz. 44(3) : 319-341 .<br />

Parker, J. 1963. Causes of the winter decline<br />

in transpiration and photosynthesis in some<br />

evergreens. <strong>Forest</strong> Sci . 9 : 148-166 .<br />

Perrier, A. 1971 Leaf temperature measurement<br />

. In Z . Sestak, J . Catsky, and P . G .<br />

Jarvis (eds.), Plant photosynthetic production,<br />

manual of methods, p . 632-671 . The<br />

Hague : Dr . W. Junk N .V. Publishers .<br />

Perry, T. O., and G . W . Baldwin. 1966 . Winter<br />

breakdown of the photosynthetic apparatus<br />

of evergreen species . <strong>Forest</strong> Sci. 12 :<br />

298-300 .<br />

Pharis, R. P., H. Hellmers, and E . Schuurmans.<br />

1967. Kinetics of the daily rate of<br />

photosynthesis at low temperature for two<br />

conifers. Plant Physiol. 42(4) : 525-531 .<br />

Phillips, R. A. 1967. Stomatal characteristic s<br />

throughout a tree crown . 67 p . M .S . thesis<br />

on file, Univ . Wash., Seattle .<br />

P i s e k, A . 1960. Die photosynthetische n<br />

Leistungen von Pflanzen besonder e<br />

Standorte : immergrune Pflanzen . In W .<br />

Ruhland (ed.), Handb . Pflanzenphysiol .<br />

V 2 : 415-459 .<br />

and R. Kemnitzer . 1968 . Der<br />

Einfluss von Frost auf die Photosynthes e<br />

der Weisstanne (Abies alba) . Flora 157 B :<br />

314-326 .<br />

, W . Larcher, W . Moser, and I .<br />

Pack . 1969 . Temperaturabhangigkeit an d<br />

optimaler Temperaturbereich der Netto -<br />

Photosynthese . Flora 158 B: 608-630 .<br />

, W . Larcher, I . Pack, and R .<br />

Unterholzner . 1968 . Temperaturmaximu m<br />

der Netto-Photosynthese and Hitzeresistenz<br />

der Blatter. Flora 158 B: 110-128 .<br />

Polster, H . 1967a . Wasserhaushalt . In H. Lyr ,<br />

H. Polster, and H . -J . Fiedler, Geholzphysiologie,<br />

p . 147-196 . Jena : G . Fischer Verlag .<br />

1967b. Photosynthese, Atmung ,<br />

and Stoffproduktion . In H . Lyr, H . Polster ,<br />

and H. -J . Fiedler, Geholzphysiologie, p .<br />

197-246 . Jena : G . Fischer Verlag .<br />

Poskuta, J . 1968 . Photosynthesis, photorespiration<br />

and respiration of detached<br />

spruce twigs as influenced by oxygen concentration<br />

and light intensity . Physiol .<br />

Plant . 21 : 1129-1136 .<br />

Rangnekar, P . V ., D . F . Forward, and N . J .<br />

Nolan. 1969 . Foliar nutrition and wood<br />

growth in red pine : the distribution of<br />

radiocarbon photoassimilated by individual<br />

branches of young trees . Can . J. Bot. 47 :<br />

1701-1711 .<br />

Reed, K. L. 1968 . The effects of sub-zero<br />

temperatures on the stomata of Douglas-fir .<br />

79 p . M .S. thesis on file, Univ . Wash . ,<br />

Seattle .<br />

1972 . A computer simulation<br />

model of seasonal transpiration in Douglas -<br />

fir based on a model of stomatal resistance .<br />

133 p. Ph.D. thesis on file, Oreg . State<br />

Univ ., Corvallis .<br />

Reed, Kenneth L ., and Warren L. Webb .<br />

1972 . Criteria for selecting an optima l<br />

model : terrestrial photosynthesis . In<br />

Proceedings-research on coniferous forest<br />

ecosystems-a symposium, p . 227-236 ,<br />

illus. Pac . Northwest <strong>Forest</strong> & Range Exp .<br />

Stn., Portland, Oreg .<br />

Ritchie, G. A. 1971 . Transpiration, water<br />

potential and stomatal activity in relatio n<br />

to microclimate in Abies amabilis an d<br />

Abies procera in a natural environment .<br />

226 p. Ph.D. thesis on file, Univ . Wash . ,<br />

Seattle .<br />

Ross, Stephen D . 1972. The seasonal an d<br />

diurnal source-sink relationships for photo -<br />

assimilated 14 CO 2 in the Douglas-fi r<br />

224


anch. 170 p . Ph .D . thesis on file, Univ .<br />

Wash., Seattle .<br />

Sestak, Z ., J . Catsky, and P . G . Jarvis (eds .) .<br />

1971 . Plant photosynthetic production ,<br />

manual of methods . 818 p . The Hague : Dr .<br />

W. Junk N .V . Publishers .<br />

Slatyer, R . O . 1967 . Plant-water relationships .<br />

366 p . London : Acad . Press .<br />

Slavik, B . 1971. Determination of stomatal<br />

aperture . In Z . Sestak, J . Catsky, and P . G .<br />

Jarvis (eds .), Plant photosynthetic production,<br />

manual of methods, p . 556-565 . The<br />

Hague : Dr . W . Junk N .V . Publishers .<br />

Stalfelt, M . G . 1956 . Die stomatare Transpiration<br />

and die Physiologie der Spaltoffnungen<br />

. In W . Ruhland (ed .), Handb . Pflanzenphysiol.<br />

III : 351-426 .<br />

1960 . Die Frage der begrenzenden<br />

Faktoren . In W . Ruhland (ed .), Handb .<br />

Pflanzenphysiol . V 2 : 213-254 .<br />

Stephens, G . R . 1969 . Productivity of red<br />

pine. 1 . Foliage distribution in tree crowns<br />

and stand canopy . Agric . Meteorol . 6 :<br />

275-282 .<br />

Thompson, R . B., and L . Leyton . 1971 . Method<br />

for measuring the leaf surface area of<br />

complex shoots . Nature 229 : 572 .<br />

Tranquillini, W. 1969 . Photosynthese and<br />

Transpiration einiger Holzarten bei verschieden<br />

starkem Wind . Centralbl. fur da s<br />

Gesamte Forstw . (Wien) 86(1) : 35-48 .<br />

Ungerson, J ., and G . Scherdin . 1965 . Untersuchungen<br />

uber Photosynthese and<br />

Atmung unter nattirlichen Bedingungen<br />

wa`hrend des Winterhalbjahres bei Pin u s<br />

silvestris, Picea excelsa, and Juniperus cornmunis<br />

. Planta 67 : 136-167 .<br />

Waggoner, P . E ., and N . C . Turner . 1971 .<br />

Transpiration and its control by stomata i n<br />

a pine forest . Bull . Conn . Agric. Exp. Stn .<br />

726, 87 p .<br />

Waring, R . H ., and B . D . Cleary . 1967 . Plant<br />

moisture stress : evaluation by pressur e<br />

bomb . Science 155(3767) : 1248-1254 .<br />

, K. L . Reed, and W. H. Emmingham<br />

. 1972 . An environmental grid fo r<br />

classifying coniferous forest ecosystems . In<br />

Proceedings-research on coniferous fores t<br />

ecosystems-a symposium, p . 79-91, illus .<br />

Pac . Northwest <strong>Forest</strong> & Range Exp . Stn . ,<br />

Portland, Oreg .<br />

Woodman, James N . 1968 . The relationshi p<br />

of net photosynthesis to environment with -<br />

in the crown of a large Douglas-fir tree . 18 8<br />

p . Ph .D . thesis on file, Univ . Wash ., Seattle .<br />

1971 . Variation of net photo -<br />

synthesis within the crown of a large forest -<br />

grown conifer . Photosynthetica 5(1) :<br />

50-54 .<br />

Zelawski, W ., and J . Kucharska . 1967 . Winter<br />

depression of photosynthetic activity i n<br />

seedlings of Scots pine (Pinus silvestris, L .) .<br />

Photosynthetica 1(3-4) : 207-213 .<br />

Zimmermann, M . H ., and C . L. Brown . 1971 .<br />

Trees : structure and function . 336 p . New<br />

York : Springer-Verlag Inc .<br />

225


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium.<br />

Bellingham, Washington-March 23-24, 197 2<br />

Criteria for selecting an optimal model:<br />

terrestrial photosynthesis<br />

Kenneth L . Ree d<br />

College of <strong>Forest</strong> Resources<br />

University of Washingto n<br />

Seattle, Washington 98196<br />

an d<br />

Warren L . Web b<br />

School of <strong>Forest</strong>r y<br />

Oregon State University<br />

Corvallis, Oregon 9733 1<br />

A bs tract<br />

In theory, there exists an infinite number of models of a given system. These models differ in resolution,<br />

scope, descriptive and predictive power. Because the models described in the literature generally reflect th e<br />

personal perspectives and goals of the researchers who developed the models, it is helpful to be able to evaluat e<br />

the potential of a given model to provide the information needed in another research problem . Several criteria<br />

for selection of models or modeling approach are suggested in this paper. The use of the criteria is illustrated by<br />

discussion of several mathematical models of photosynthesis taken from the literature . The models discussed<br />

include a linear regression model, an energy budget model, and models based on theory of enzyme kinetics and<br />

gas exchange .<br />

Introduction<br />

A great many models are available that de -<br />

scribe photosynthesis, some of which will be<br />

discussed below . Because of the many models<br />

now available, we had hoped that we woul d<br />

not have to develop a new model-that w e<br />

need only choose or modify an existing model .<br />

Thus it was necessary to consider certain criteria<br />

by which to judge the applicability of a<br />

given model to the purposes of the IBP .<br />

Several of the criteria will be discussed be -<br />

low. This list is probably not complete, but i s<br />

suggested as a guide to model selection, modeling<br />

approach, or both . Because most scientists<br />

are modelers in a broad sense, this discussion<br />

is intended primarily for those scientists<br />

who are least familiar with formal aspects<br />

of modeling .<br />

Criteria for Model Selectio n<br />

Because plant growth and presumably community<br />

composition are greatly influenced b y<br />

photosynthetic capacity, that phenomeno n<br />

has been the object of a great deal of study .<br />

Photosynthesis is only one component of the<br />

biochemical system resulting in plant growth .<br />

Failure to consider other factors in growth, a s<br />

well as the fact that photosynthesis can be<br />

uncoupled from growth, results in the general<br />

failure of attempts to predict growth from<br />

photosynthesis. However, a model of photosynthesis<br />

is an important component of a<br />

growth model and because the process o f<br />

photosynthesis is a physical-chemical system ,<br />

attempts to model photosynthesis should b e<br />

consistent with some basic tenets of systems<br />

theory as well as physical reality .<br />

227


Systems<br />

A system, as defined by Klir (1969), is imposed<br />

upon an object-a segment of nature<br />

(the earth, a community, a tree, an automobile,<br />

a computer)-by the observer from a distinct<br />

point of view. Everything that does not<br />

belong to the object is the environment . The<br />

boundary between the object and its environment<br />

cannot be clearly defined ; thus th e<br />

delimitation of the object is somewhat arbitrary<br />

and reflects the personal perspectives o f<br />

the modeler . Because we usually cannot stud y<br />

an object in its entirety (because of its complexity),<br />

we observe or measure values of certain<br />

quantities . The choice of quantities to<br />

measure depends on what we consider to b e<br />

of interest or important to the given purpos e<br />

(Klir 1969) .<br />

Most scientists are familiar with the concept<br />

of a system, especially as it is applied i n<br />

thermodynamics. For example, the laws of<br />

thermodynamics were derived by studyin g<br />

idealized closed systems in equilibrium . In<br />

dealing with such closed systems it is possible<br />

to study independent state variables such as<br />

pressure, temperature, and volume, which de -<br />

fine the state of the system. A closed thermodynamic<br />

system can exchange heat and work<br />

but not matter with its environment (Daniels<br />

and Alberty 1967) .<br />

As Ludvig von Bertalanffy (1969) pointe d<br />

out, biological systems are open system s<br />

whose structure is maintained by energy, in -<br />

formation, and matter flow through the systems.<br />

The behavior of the system is the resul t<br />

of the interactions of the elements of th e<br />

system and the flow through the system . Because<br />

the thermodynamic system is closed<br />

and idealized, and because the state variable s<br />

are independent (i .e., one can be varied with -<br />

out affecting another), it is convenient and<br />

correct to describe the system in terms o f<br />

total differentials; for example, the relation<br />

of internal free energy, E, of a system to<br />

temperature and volume, T and V, can be expressed<br />

(Daniels and Alberty 1967) :<br />

E = f(T, V)<br />

dE _<br />

(3T<br />

) V dT + () T dv (1)<br />

Equation 1 implies that the variables E, T,<br />

and V are independent ; one can be varied<br />

while the others are held constant, but mor e<br />

importantly, equation 1 implies summativity<br />

in the system where the behavior of a summa -<br />

tive system is the physical sum of the behavior<br />

of the parts . Because this approach is useful<br />

in thermodynamics and other special<br />

areas, biologists often attempt to explai n<br />

observed phenomena by assuming that the<br />

elements of all systems are independent (vo n<br />

Bertalanffy 1969) . We in plant physiology<br />

have accepted as a standard procedure th e<br />

isolation of a plant in a growth chamber, an d<br />

the study of the response of the plant to on e<br />

factor while holding the others constant . This<br />

approach is epitomized by Cleary (1970) who<br />

derived the following model of photosynthesis :<br />

the n<br />

P = f(M,T,L,N,Pr)<br />

dP= (3P) dM y {('r,) dT<br />

am T,L,N P dT M .L.N.P<br />

+ .<br />

. + dpr~ dP r<br />

1, 7'. L,N<br />

(2 )<br />

where P = photosynthesis, M,T,L,N, and Pr<br />

are moisture, temperature, light, nutrition an d<br />

preconditioning effect, respectively . This<br />

model assumes a summative system analogou s<br />

to a closed thermodynamic system .<br />

This assumption is not totally valid in a<br />

biological system like photosynthesis . Light ,<br />

for example, has an effect not only on the<br />

photochemical reactions within the leaf, but<br />

also is converted to heat, which influences<br />

leaf temperature. Likewise, leaf temperature<br />

is affected by transpiration rates, which is in<br />

turn affected by temperature, moisture status<br />

of the leaf and air, and so on, but equation 2<br />

does not account for these interactions. While<br />

this model is inadequate in that respect ,<br />

Cleary (1970) does recognize that photo -<br />

synthesis is not a simple light-and-temperatur e<br />

related phenomenon but represents a complex<br />

interaction of diverse factors, all of whic h<br />

should be taken into account in order to full y<br />

understand the process . Some of the factor s<br />

listed by Cleary (1970) probably are independent,<br />

namely, Pr and N.<br />

228


Von Bertalanffy (1969) calls nonsummative<br />

systems Gestalten ; for example, the nonadditivity<br />

of mixing equal volumes of concentrated<br />

sulfuric acid and water. Von Bertalanffy<br />

provides another simple example where he<br />

points out that the voltage of three isolated<br />

conductors would be different from that i f<br />

they were interconnected . Biological quantities<br />

often act very much like the charge i n<br />

von Bertalanffy's example in that their values<br />

in concert are different from their isolated<br />

values . Thus, it is essential for a biological re -<br />

searcher to understand the distinction between<br />

summative systems and Gestalten . It is<br />

advisable to consider a biological system to b e<br />

a Gestalt unless it can be proved otherwise .<br />

We do not believe that the concepts of systems<br />

can be overemphasized . It is not necessary<br />

for each researcher to be adept at formulating<br />

systems models, but he should be awar e<br />

that few things in a biological system operat e<br />

independently, and that it is difficult if no t<br />

impossible to understand biological phenomena<br />

if we simply take the one factor-on e<br />

response approach. As a result, in modelin g<br />

biological systems, the single most importan t<br />

concept is that the model must be a syste m<br />

model, consistent with the concepts of<br />

systems theory .<br />

Theoretical Validit y<br />

We view a model as a tool to aid in under -<br />

standing and prediction . Because there are, i n<br />

theory at least, an infinite number of model s<br />

that can apply to a given system, we must<br />

select that model which provides the greatest<br />

understanding and predictive power withi n<br />

certain limitations . If our system of interest<br />

can best be explained by physical principles ,<br />

then the model should have physical validity ,<br />

but there are systems which are not explain -<br />

able from the physical paradigm, for example ,<br />

animal behavior . In other cases, the physica l<br />

theory is inadequate and the approach may b e<br />

fruitless .<br />

Given a physical-chemical system, like<br />

photosynthesis, a physically valid model is usu -<br />

ally a good choice . As an almost trivial example,<br />

photosynthesis, P, has been expressed b y<br />

Chartier (1966) as a function of light :<br />

P 1 bL (3 )<br />

where a/b is P at light saturation, and a is the<br />

slope of the curve (fig . 1) at zero light since<br />

a = dP/dL(1 + bL) . Lommen and coworkers<br />

(1971) express the same phenomenon as :<br />

PmL<br />

P m<br />

(L) = 1 + K L /L<br />

where PmL is photosynthesis at light and C O 2<br />

saturation and KL is that light intensity at<br />

which Pm(L) = 1/2 PmL (fig. 2) .<br />

a/s<br />

P<br />

Figure 1 . Photosynthesis as a function of light . From<br />

Chartier (1966) A = slope of line O,X . Descriptio n<br />

in text .<br />

PM (L)<br />

PML<br />

------------------ -<br />

K<br />

(4)<br />

Figure 2 . Photosynthesis as a function of light . Fro m<br />

Lommen et al.(1971) . Description in text .<br />

L<br />

L<br />

229


It is obvious that in essence equation 4 i s<br />

identical to equation 3, where PmL = a/b an d<br />

K = 1/b . Both models describe the light dependence<br />

of photosynthesis equally well ; on e<br />

model is not more complex than the other .<br />

Yet, we would choose equation 4 because th e<br />

parameters have more physical meaning . Th e<br />

curve could also be described by :<br />

P=00 + a 1L+ 32 L 2 + 03 L3 . . . (5)<br />

which has even less appeal because it is nearl y<br />

impossible to determine the physical meanin g<br />

of the parameters ((3 i).<br />

It could be argued that equation 4 is littl e<br />

more valid than equation 3. Equation 4 is<br />

derived from the Michaelis-Menton equatio n<br />

that describes the rate of a single enzymati c<br />

reaction as a function of a substrate concentration<br />

. Photosynthesis is not a true Michaelis-<br />

Menton case because it is an integration o f<br />

photochemistry and a chain of enzymati c<br />

reactions . The reductionist might argue that<br />

we should model photosynthesis as a functio n<br />

of the photochemical and the enzymatic reaction<br />

rates . Further, it is obvious that temperature,<br />

substrate availability, and several othe r<br />

factors are important in photosynthesis. Also ,<br />

since equation 4 considers only light an d<br />

(indirectly) CO 2 effects, it falls short of ou r<br />

first criterion, that the model be a system<br />

model insofar as possible. These points brin g<br />

up our third criterion : resolution level .<br />

Resolution<br />

According to Klir (1969) every quantity w e<br />

observe must be determined in space . That is ,<br />

CO 2 concentration, incident radiant energy ,<br />

and temperature may be specified at som e<br />

point in space . This specification may be irrelevent<br />

for some studies (e .g., the location of<br />

a laboratory may be trivial with respect to th e<br />

study) but when concerned with a weathe r<br />

forecast, it would be important to identif y<br />

the spot on earth where temperature is measured.<br />

We may also specify the interval and<br />

accuracy of our measurements, giving the<br />

space-time resolution level (Klir 1969) .<br />

It may be convenient to think of the spatial<br />

resolution level as the hierarchical level of<br />

resolution (von Bertalanffy 1969) . Biological<br />

systems can be placed in hierarchies corresponding<br />

with levels of organization (table 1) .<br />

0 . ?mo<br />

Table 1 .-Levels of organization<br />

1. Subatomi c<br />

2. Atomi c<br />

3. Molecular<br />

4. Subcellular<br />

5. Cellular<br />

6. Organ or tissue<br />

Thus a system can be thought of as a component<br />

in a larger system, and composed of sub -<br />

systems (von Bertalanffy 1969) . Each level of<br />

organization seems to have certain properties<br />

unique to that level, making prediction o f<br />

system behavior from the viewpoint of th e<br />

lower systems somewhat hazardous . This concept<br />

is implicit in the idea of nonsummativit y<br />

of systems. The reductionist might argu e<br />

against this view, but the limits of reductionism<br />

are obvious . It is impossible to predict<br />

all the behavior of a tree from the molecular<br />

level . If we hope to discover general ecosystem<br />

principles comparable to the fundamental<br />

principles of physics, probably we wil l<br />

have to emphasize study from the ecosyste m<br />

level. The ideal gas law, for example, was<br />

derived from observations of the behavior o f<br />

10 23 particles acting in concert. In ecosyste m<br />

study, we are among the particles; a holistic<br />

view is needed .<br />

There are three levels of resolution under<br />

consideration for the photosynthesis model :<br />

the leaf, tree, and stand levels . Each level will<br />

require a different model and approach ; we<br />

cannot extrapolate the leaf level model to a<br />

tree by multiplying the results by the numbe r<br />

of leaves in the crown . The temporal resolution<br />

of the models will be somewhat flexibl e<br />

at the leaf level, but will probably be on a<br />

daily basis for the stand model .<br />

Practical Considerations<br />

7. Organism<br />

8. Community<br />

9. Global<br />

10. Solar Syste m<br />

11. Galactic<br />

12. Universe<br />

13. ? o<br />

There are several practical consideration s<br />

230


which may be of importance in model selection<br />

or approach . Some nonlinear systems of<br />

differential equations are extremely difficult<br />

to solve and their solution may consume a<br />

great deal of computer time . Parameter estimation<br />

is often a significant problem in man y<br />

models. Some models require highly precis e<br />

measurement of input variables, reducin g<br />

their practical applications, but their theoretical<br />

contributions may be important . On the<br />

other hand, we sometimes settle for use of a<br />

linear regression model when the system is<br />

poorly understood and it is much easier t o<br />

measure the variables than to determine their<br />

exact interrelations . Such models are useful in<br />

science, but have some built-in dangers, e .g . ,<br />

the hazard of extrapolation of linear regression<br />

models .<br />

One danger of reliance upon regressio n<br />

models is that the confidence in the model i s<br />

greatest at the mean and diminishes at th e<br />

extremes. In fact, it is possible to fit a curvilinear<br />

function to data of a given range, only<br />

to have the model become totally inadequate<br />

at the extremes .<br />

For example, net photosynthesis as a function<br />

of temperature can be described as a<br />

symmetrical quadratic (Pisek and Winkle r<br />

1958, Webb 1972) (fig. 3) .<br />

values. Thus equation 6 is valid only in th e<br />

range of temperature within which the<br />

parameters were estimated .<br />

The failure of a model to predict syste m<br />

behavior at extremes is not necessarily fatal ;<br />

such models are common in physics, e .g., the<br />

ideal gas law and Newtonian physics . It i s<br />

necessary to understand the limitations of<br />

one 's models to know when the system deviates<br />

from the model .<br />

In choosing a model, it is important t o<br />

understand the assumptions and limitations of<br />

the model . For example, it is questionable to<br />

describe a biological system with a model tha t<br />

assumes a closed reversible system . Th e<br />

assumptions implicit in the model should be<br />

compatible with present knowledge of th e<br />

system of interest .<br />

This is not to say that models of different<br />

systems cannot be used to model another<br />

system. The idea of isomorphism of system s<br />

models is central to general systems theory<br />

(von Bertalanffy 1969) . Thus thermodynami c<br />

models, for example, may be of great utility<br />

in certain biological systems models . It re -<br />

mains for the researcher to be sure that the<br />

systems are isomorphic so that such model s<br />

are applicable .<br />

Pn = X30 + R1 T - 32 T2 ( 6)<br />

But extrapolation of the curve in either direction<br />

will lead to progressively more negativ e<br />

G(T)<br />

Examination of<br />

Some Models<br />

of Photosynthesis<br />

The discussion above dealt with criteria fo r<br />

selection of models . It would be useful to discuss<br />

some models described in the literature .<br />

Leaf Environment Mode l<br />

a<br />

Figure 3 . Net photosynthesis, G(T), as a quadrati c<br />

function of temperature .<br />

T<br />

Botkin (1969) simulated photosynthesis in<br />

an open oak-pine forest near Brookhave n<br />

National Laboratory . He developed a linea r<br />

model of net photosynthesis :<br />

Pn=c 0 + 01 T+ (3 2 ln S<br />

+ Q3 (ln s) 2 + Q4 T2 + (3 5 T 1n S (7)<br />

231


where T = leaf temperature, ° K<br />

S = solar radiation, gcal cm-2 min<br />

Pn = net photosynthesis rate ,<br />

mg CO 2 (g dry wt)-1 hr 1<br />

(3i = parameters to be estimate d<br />

Equation 7 was derived from equation 8 and<br />

equation 9 :<br />

P(L)=a+b 1nS (8)<br />

which had been used to fit photosynthesis dat a<br />

from several herbaceous species (Blackman an d<br />

Rutter 1946, Blackman and Wilson 1951) .<br />

Equation 8 was coupled with equation 9 suggested<br />

by data of Pisek and Winkler (1958) ,<br />

Krueger and Ferrell (1965), and others :<br />

G(T) = a + bT + cT2 (9 )<br />

Data from two species of oak were obtained<br />

from infrared gas analysis and wer e<br />

used to estimate the parameters of equation 7<br />

by stepwise multiple regression analysis . Som e<br />

of the terms of equation 7 were nonsignificant,<br />

and were thus discarded . The final<br />

model was :<br />

Pn =Qo +(3 1 T+(3 2 In S+(3 5 T In S (10 )<br />

Having estimated the parameters of equation<br />

10, the model was used to predict photo -<br />

synthesis of oak in the field . Their model gav e<br />

fair to good agreement with subsequentl y<br />

measured net photosynthesis .<br />

In terms of the criteria discussed above ,<br />

equation 10 is generally inadequate . It doe s<br />

allow for interaction between two variables but<br />

fails to take into consideration other factor s<br />

that affect photosynthetic rate (stomatal behavior,<br />

micrometeorological conditions, plan t<br />

nutrition) . The model does satisfy the requirement<br />

that the Pn model be solved as a functio n<br />

of systems variables, that is, Pn = f [S, T] .<br />

In terms of the other criteria, the model fall s<br />

short of having a great deal of theoretica l<br />

validity in that the function is one of convenience<br />

rather than having general physicalchemical<br />

meaning . It is unnecessary to discuss<br />

in detail Botkin's model with respect to the<br />

other criteria . Like most photosynthesis models,<br />

it is specified at the leaf level, where th e<br />

inputs to the leaf are measured, and the<br />

relations between the leaf subsystems are<br />

inferred .<br />

While his model has many failings, it doe s<br />

have one virtue. If one were interested in<br />

comparing Pn = f[T,L] in two species under<br />

identical conditions, this model may be adequate,<br />

given that some other factor, e .g . ,<br />

stomatal resistance, is not limiting . The fact<br />

that the parameters of the model were estimated<br />

by least-squares allows the use o f<br />

statistical tests useful in comparing such data .<br />

Energy Budget Mode l<br />

Idso and Baker (1968) based their model of<br />

photosynthesis and their earlier model (Ids o<br />

and Baker 1967) on that of Gates (1965 )<br />

which is an energy budget model where precis e<br />

measurement of incoming radiant energy is<br />

equated with outgoing energy . Thus, at equilibrium,<br />

the amount of energy leaving a leaf i s<br />

equal to that coming in . Gates ' (1965) leaf<br />

energy budget model is :<br />

where<br />

(1 +r) (S + s) (<strong>Rg</strong>' +R a)<br />

as 2 +a t 2 (11 )<br />

-e taTZ ± C±LE= 0<br />

a s = mean total absorbance of plant to<br />

sunlight and skyligh t<br />

r = reflectance of underlying ground<br />

or plane surface to sunlight and<br />

skylight, S and s<br />

at = absorbance of plant to thermal<br />

radiation, <strong>Rg</strong> and Ra<br />

e t = emissivity of plant to thermal<br />

radiatio n<br />

Tl = leaf temperature ° K<br />

C = energy gained or lost by convection<br />

cal cm-2 min- 1<br />

L = latent heat of evaporation ca l<br />

cm-2 miri- l<br />

E = transpiration rate of leaf gm cal - 2<br />

min- 1<br />

S, s, <strong>Rg</strong>, Ra = various short and lon g<br />

wave radiation inputs ca l<br />

cm-2 min 1<br />

232


The first three terms of equation 11 represent<br />

the radiant energy available to the plant .<br />

By elaborating equation 11 and solving for Tl ,<br />

Gates was able to predict leaf temperature as<br />

a function of the incoming energy less th e<br />

outgoing energy . From data of other worker s<br />

he developed a family of curves for photo -<br />

synthesis with respect to temperature and radiant<br />

energy . Given these curves and by calculating<br />

Tl, Gates predicted photosynthesis of<br />

several species .<br />

Idso and Baker (1967) constructed a similar<br />

family of curves for sorghum. In a subsequent<br />

paper (Idso and Baker 1968), they used<br />

these curves to predict Pn on four different<br />

types of days based on input from their<br />

energy budget calculations . They did not<br />

measure Pn directly, so there was no direct<br />

verification of the models .<br />

In terms of our suggested criteria fo r<br />

analysis of models, it is obvious that th e<br />

energy budget model is consistent with th e<br />

systems approach . However, the relation of Tl<br />

and absorbed energy to Pn is empirical an d<br />

incomplete, and constitutes the weak point o f<br />

this model . Thus, from a standpoint o f<br />

theoretical validity, the Pn model suggeste d<br />

by Idso and Baker (1967, 1968) suffers fro m<br />

drawbacks identical to those of Botki n<br />

(1969). Should a more satisfactory model o f<br />

Pn as a function of light and temperatur e<br />

become available, however, it can be incorporated<br />

into the leaf energy budget model . The<br />

energy budget model itself is sound and has<br />

great promise . It is probably sufficiently<br />

general to apply to any plant, and the variou s<br />

terms can be refined where more specifi c<br />

component models are needed .<br />

Unfortunately, from a practical standpoint ,<br />

the input data necessary for such a model ar e<br />

very difficult to obtain, requiring a great deal<br />

of sensitive instrumentation. This limitatio n<br />

would preclude the applicability of the mode l<br />

to study of some ecosystems where suc h<br />

measurements would be most difficult o r<br />

impossible . Further, the resolution of th e<br />

model is great, and such variables as inciden t<br />

solar angle and leaf geometry are important .<br />

Extrapolation of this model to a tree or to a<br />

stand would require prodigious volumes o f<br />

input data and highly sophisticated subsystem<br />

models .<br />

In summary, the high-resolution energ y<br />

budget approach, while having considerabl e<br />

appeal from the systems and theoretica l<br />

standpoint, is probably limited in its utility t o<br />

theoretical studies and studies of simple r<br />

systems . The basic approach could be modified<br />

by altering the resolution level of the<br />

components and by perhaps adding som e<br />

stochastic models . In this way, the energy<br />

budget approach could be used in modeling a t<br />

a lower level of resolution . Such a model ,<br />

being simpler, could then be used at the tre e<br />

and stand levels of organization .<br />

While the work of Gates (1965) and hi s<br />

colleagues has contributed greatly to th e<br />

understanding of the energetics of the leaf, it<br />

still remained to develop a model of the<br />

process of photosynthesis itself that could<br />

satisfy our second criterion, that of theoretical<br />

validity . This problem was attacked by<br />

Brown (1969) .<br />

Process Models<br />

Brown assumed that if Pn is a first-orde r<br />

reaction, then equation 12 would hold :<br />

where P = rate of CO 2<br />

leaf area<br />

P = KI[CO 2 ]g (12 )<br />

exchange per uni t<br />

K = capacity of acceptor site in th e<br />

leaf to fix CO 2 at the site of<br />

photosynthesis<br />

I = incident radiatio n<br />

[C0 2]g = concentration of CO 2 at the fixation<br />

site<br />

Brown also assumed that I follows Beer' s<br />

law within the leaf ,<br />

x = Ie" ax (13 )<br />

where I is incident radiation at the leaf<br />

surface, a is the extinction coefficient, x is<br />

distance from the leaf surface, and Ix is<br />

radiation at x . Brown incorporated equation<br />

12 into Gaastra's (1959) derivation from<br />

Fick 's law of diffusion representing steadystate<br />

diffusion of CO 2 from the air to the<br />

leaf :<br />

233


where<br />

giving<br />

P = C a Cc<br />

Er<br />

(14 )<br />

Ca = concentration of atmospheri c<br />

CO 2 g cm- 3<br />

Cc = concentration of CO 2 at the<br />

chloroplast g cm 3<br />

Er = resistance to gas diffusion<br />

sec cm- 1<br />

D[CO 2] a<br />

P= 1 + D/KI<br />

where D is the integral exchange coefficien t<br />

D= 1 = 1<br />

ra +rs +rm E r<br />

where ra is boundary layer resistance, rs is<br />

stomatal resistance, and rm is mesophyll<br />

resistance to CO 2 diffusion in sec cm-2 .<br />

Equation 15 is gross photosynthesis, not Pn ,<br />

and thus equation 15 should be corrected for<br />

respiration :<br />

Pn = K[CO 2 /al - R<br />

(16)<br />

I + K1/D<br />

where R is respiration .<br />

These models assume that resistance t o<br />

transfer of CO 2 from respiration site to the<br />

photosynthesis site is negligible . The variable<br />

nature of rs must also be considered, requirin g<br />

an additional model of stomatal behavior .<br />

Brown (1969) notes that the complexity of<br />

equation 16 becomes great when other components<br />

are added, and states that it is<br />

necessary to solve equation 16 by arriving at<br />

the functional relationships of the components<br />

by independent means and substituting<br />

them into the general model .<br />

If the components are truly independent ,<br />

such a tactic is valid . If not, error is introduced<br />

by failing to consider changes in th e<br />

values of the parameters due to interaction, a<br />

violation of our systems criterion .<br />

Lommen and coworkers (1971) took an<br />

approach similar to that of Brown (1969) .<br />

They began with Gaastra 's (1959) derivatio n<br />

from Fick's law of diffusion now familiar to<br />

us. The biochemical relations are described b y<br />

a Michaelis-Menton type equation that de -<br />

scribes the rate of a single enzymatic reaction .<br />

Their model for photosynthesis as a function<br />

(15)<br />

of chloroplast concentration of CO 2 is :<br />

Pm<br />

P 1 + K/Cc<br />

where P = photosynthesis g cm-2 sec- 1<br />

Pm = photosynthesis at saturating CO 2<br />

g cm-2 sec- s<br />

K = concentration of CO 2 at the<br />

chloroplast when P = 1/*Pm<br />

Cc = concentration of CO 2 at the<br />

chloroplast g cm- 3<br />

Equation 14 is solved for Cc and substitute d<br />

into equation 17 giving photosynthesis as a<br />

function of atmospheric CO 2 concentratio n<br />

and stomatal resistance :<br />

(Ca + K + E rPm)<br />

P=<br />

2Er<br />

(18)<br />

- [(Ca + K + ErPm)2 - 4CaErPm] lie<br />

21 r<br />

They also derived a similar though mor e<br />

complex model giving photosynthesis as a<br />

function of Ca, K, Pm, W, and two series of<br />

resistances, S 1 , S 2 ; W is respiration .<br />

They also derived submodels of photosynthesis<br />

as a function of light and temperature ,<br />

Pm(L,T) =<br />

PmLT G(T)<br />

1 + KL/L<br />

(17 )<br />

(19 )<br />

similar to equation 4 but incorporating a term<br />

G(T) representing the temperature dependence<br />

of photosynthesis . The authors implie d<br />

that equation 19 could be incorporated int o<br />

equation 18 and their other model, but the y<br />

did not do so, and dimensional analysis o f<br />

equation 18 with PmLT and KL substituted<br />

for Pm and K, shows that such a substitutio n<br />

is physically unrealistic. Hence the model o f<br />

Lomrnen et al . (1971) is inadequate for<br />

predicting photosynthesis as a function o f<br />

light and temperature as well as CO 2 concentration<br />

and stomatal resistance .<br />

Chartier (1966, 1969, 1970) also developed<br />

a complex model of net assimilation derive d<br />

along the lines of that by Lommen et al .<br />

(1971) . Beginning with the fundamental for m<br />

(Chartier 1970) :<br />

234


P = F+ R<br />

where F = net assimilation rate<br />

per unit leaf area<br />

R = respiration rate<br />

and by analogy to Gaastra's (1959) Fick's la w<br />

function, he derived :<br />

C-F(ra +rs +rm)-nRrm<br />

F + R = 1<br />

aE C-F(rQ +rs +rm)-nRrm +rx<br />

(20)<br />

The only terms unfamiliar to us are rx, n, aE,<br />

and R . Here R is respiration in light, and n is<br />

the fraction of respiratory flux that is mixe d<br />

in the intercellular spaces (n < 1) . The term<br />

aE represents conversion of light to photosynthate<br />

where E is incident light energy, an d<br />

a is the efficiency of light energy conversion .<br />

The term rx represents resistance to carboxyla -<br />

tion, a parameter which includes biochemical<br />

restraints caused by mineral nutrition, age o f<br />

the leaf, etc .<br />

Chartier's (1970) model differs from tha t<br />

of Lommen et al. (1971) in that the effect of<br />

light is incorporated into the model . Th e<br />

effect of temperature must be included, perhaps<br />

by a multiplicative term (Lommen et al .<br />

1971, Webb 1972) . Chartier's model gives a<br />

quadratic solution for F, as in Lommen et al ,<br />

but many of the terms in both models ar e<br />

difficult if not impossible to measure in a<br />

field study . Further, respiration is represente d<br />

by a single term ; an oversimplification requiring<br />

further work .<br />

Conclusions<br />

In order to be consistent with the systems<br />

viewpoint, photosynthesis must be treated as<br />

a Gestalt, or nonsummative system. The<br />

models described above violate this criterion<br />

to some extent, most often by failing to<br />

incorporate an important factor in the model .<br />

We had hoped to be able to use one of the<br />

models by Lommen et al. (1971) or by<br />

Chartier (1970) as a tool in our field research ,<br />

but two considerations prohibit this : (1) both<br />

models have unmeasurable (in the field )<br />

terms, (2) neither model is complete, i .e . ,<br />

neither expresses photosynthesis as a function<br />

of all the known important factors .<br />

Consequently, it will be necessary to develop<br />

a model of photosynthesis as a function<br />

of light, temperature, leaf resistance and<br />

ambient CO 2 concentration with some simplifications<br />

from the above models which will<br />

increase the utility of the models. The<br />

parameters in this model will be estimate d<br />

from data after the manner of Webb (1972 )<br />

by nonlinear least-squares . The final model<br />

will have much the same utility as a regressio n<br />

model, but will not be linear and th e<br />

parameters where possible will have physical<br />

meaning. Thus, the models will be develope d<br />

with certain specific goals in mind, necessitating<br />

development of different models for th e<br />

tree and stand levels of resolution .<br />

A cknowledgments<br />

The work reported in this paper was<br />

supported by National Science Foundatio n<br />

Grant No. GB-20963 to the Coniferous Fores t<br />

Biome, U .S . Analysis of Ecosystems, International<br />

Biological Program . This is Contributio n<br />

No. 38 to the Coniferous <strong>Forest</strong> Biome, IBP .<br />

Literature Cited<br />

Blackman, G . E ., and A. J. Rutter. 1946 .<br />

Physiological and ecological studies in th e<br />

analysis of plant environment . I. The light<br />

factor and the distribution of the bluebel l<br />

(Scilla non-scripta) in woodland communities.<br />

Ann. Bot., N.S. 10: 361-390 .<br />

and G . L. Wilson . 1951. Physiological<br />

and ecological studies in the analysi s<br />

of plant environment . VI. The constancy<br />

for different species of a logarithmic relationship<br />

between the net assimilation rat e<br />

and light intensity and its ecological significance.<br />

Ann . Bot., N.S. 15: 63-94 .<br />

Botkin, D . R. 1969 . Prediction of net photo -<br />

synthesis of trees from light intensity and<br />

temperature. Ecology 50: 854-858 .<br />

Brown, K . W. 1969 . A model of the photosyn -<br />

thesizing leaf . Physiol . Plant 22 : 620-637 .<br />

Chartier, P. 1966. Etude theorique de 1 'assimi -<br />

lation brute de la feuille . Ann . Physiol. veg .<br />

8: 167-196 .<br />

235


1969. Assimilation nette d'un e<br />

culture couvrante_ II. La reponse de l'unite<br />

de surface de feuille . Ann. Physiol. Veg.<br />

11 : 221-263 .<br />

1970. A model of CO 2 assimilation<br />

in the leaf . In Prediction and measurement<br />

of photosynthetic productivity, 63 2<br />

p . Wageningen, The Netherlands : Centre<br />

for Agricultural Publishing and Documentation<br />

.<br />

Cleary, B . D. 1970. The effect of plant<br />

moisture stress on physiology and establishment<br />

of planted Douglas-fir ponderosa pin e<br />

seedlings. 85 p. Ph .D. thesis on file, Oreg.<br />

State Univ., Corvallis .<br />

Daniels, F ., and R . A. Alberty . 1967 . Physical<br />

chemistry . Ed . 3, 767 p. New York : Wiley .<br />

Gaastra, P. 1959. Photosynthesis of cro p<br />

plants as influenced by light, C0 2 , temperature,<br />

and stomatal diffusion resistance .<br />

Med. Landb . Hogesch . Wageningen, Th e<br />

Netherlands 59 : 1-68 .<br />

Gates, D. M . 1965. Energy, plants, an d<br />

ecology . Ecology 46 : 1-13 .<br />

Idso, S . B ., and D. G. Baker . 1967. Metho d<br />

for calculating the photosynthetic respons e<br />

of a crop to light intensity and lea f<br />

temperature by an energy flow analysis o f<br />

the meteorological parameters . Agron. J .<br />

59 : 13-21 .<br />

and D . G . Baker. 1968. The<br />

naturally varying energy environment and<br />

its effects on photosynthesis . Ecology 49 :<br />

311-316 .<br />

Klir, G. J. 1969. An approach to general<br />

systems theory . 323 p . New York: Van<br />

Nostrand-Reinhold Co .<br />

Krueger, K. W., and W . K . Ferrell. 1965 .<br />

Comparative photosynthetic and respiratory<br />

responses to temperature and light by<br />

Pseudotsuga menziesii var . menziesii and<br />

var . glauca seedlings . Ecology 46 : 794-801 .<br />

Lommen, P. W ., C. R . Schwintzer, C. S .<br />

Yocum, and D . M. Gates . 1971 . A model<br />

describing photosynthesis in terms of gas<br />

diffusion and enzyme kinetics. Planta 98 :<br />

195-220 .<br />

Pisek, A., and E . Winkler . 1958 . Assimilationsvermogen<br />

and Respiration der Ficht e<br />

(Picea excelsa Link) in verschiedene r<br />

Hohenlage and der Zirbe (Pinus cembra L .)<br />

and der alpinen Waltlgranze. Planta 51 :<br />

518-543 .<br />

von Bertalanffy, L . 1969 . General syste m<br />

theory . 289 p. New York: Geo. Braziller .<br />

Webb, W. L. 1972. A model of light and temperature<br />

controlled net photosyntheti c<br />

rates for terrestrial plants . In Jerry F .<br />

Franklin, L. J. Dempster, and Richard H.<br />

Waring (eds.), Proceedings-research o n<br />

coniferous forest ecosystems-a symposium,<br />

p. 237-242, illus. Pac. Northwest<br />

<strong>Forest</strong> & Range Exp . Stn ., Portland, Oreg.<br />

236


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, WashingtonMarch 23-24, 197 2<br />

A model of light and temperature<br />

controlled net photosynthetic rates<br />

for terrestrial plants<br />

Warren L . Web b<br />

<strong>Forest</strong> Research Laboratory<br />

School of <strong>Forest</strong>r y<br />

Oregon State University<br />

Corvallis, Orego n<br />

Abstract<br />

Steady-state relative net CO2 exchange was modeled in terms of a temperature-dependent respiration function<br />

and light- and temperature-dependent photosynthesis function. The parameters of the model were<br />

evaluated using the laboratory CO 2 exchange data of a group of 40 red alder seedlings (Alnus rubra Bong.). Th e<br />

model is continuous and well-behaved in the temperature region of 0-50°C for light energy between 0 .0 and 1 . 0<br />

ly/min total short-wave radiation.<br />

Introduction<br />

Consumer populations depend upon the<br />

chemical energy that is converted from sola r<br />

energy by plants in the ecosystem . The rate of<br />

conversion, or net photosynthesis, is in tur n<br />

dependent upon the genetic information avail -<br />

able to each plant and its immediate environment.<br />

Many factors influence net photosynthesis<br />

but, as Schulze (1970) found from hi s<br />

work in a beech stand, when water stress is<br />

not appreciable, the radiation and temperature<br />

regimes of the plant largely regulate net<br />

photosynthesis . This paper presents an empirical<br />

model of steady-state net CO 2 exchange i n<br />

terms of a light (L) and temperature (T) con -<br />

trolled gross photosynthesis function (Ps) an d<br />

a temperature-controlled dark respiration (Rs )<br />

function .<br />

Net CO 2 flux entering the leaf, or ne t<br />

photosynthesis (Psn), is conceptualized as th e<br />

difference between carbon fixed in the photo -<br />

synthetic process and that lost during respiration,<br />

Psn = Ps - Rs (Larcher 1969) . The two<br />

terms, net photosynthesis and net CO2 exchange,<br />

are used synonymously in this paper .<br />

Although the former is less general in that it<br />

usually applies only to CO 2 exchanges occurring<br />

in the light, it is more mechanically<br />

viable .<br />

Fluxes of CO 2 are easily measured and<br />

many investigators have reported on the ne t<br />

photosynthetic response of plants to temperature<br />

and light (Heath 1969, Rabinowitc h<br />

1969, Milner and Hiesey 1969) . Figure 1 illustrates<br />

the net photosynthetic response of<br />

small plants or individual leaves to increasin g<br />

temperature or light . The linear portion of the<br />

light curve represents the rate of the photo -<br />

chemical reaction at the chloroplast and is<br />

largely invariant among species (Rabinowitc h<br />

1969) . This linear slope can be extrapolated<br />

through zero CO2 flux, called the light compensation<br />

point, and the negative absciss a<br />

intercept interpreted as dark respiration<br />

(Chartier 1969) . At high light levels, photo -<br />

synthesis becomes saturated and its rate i s<br />

dependent upon factors such as temperatur e<br />

237


Figure 2. Effect of temperature on dark respiration .<br />

Figure 1 . Leaf CO2 exchange as related to light an d<br />

temperature .<br />

and water potential that influence the carbon -<br />

fixing enzymatic reactions .<br />

Net photosynthetic response to temperature,<br />

also shown in figure 1, is generally a<br />

symmetrical form (Pisek et al . 1969, Moone y<br />

and Harrison 1969). Pisek's data have show n<br />

that the slope of the curve and the point of<br />

maximum shifts depending upon species an d<br />

season but the response remains symmetrical .<br />

The responses to light and temperatur e<br />

seem well-defined. A model using light an d<br />

temperature as independent variables shoul d<br />

conserve the known responses to each as wel l<br />

as include any interaction . To be useful, th e<br />

model should also predict CO 2 losses that<br />

occur below the light-compensation point a s<br />

well as respiration losses occurring durin g<br />

darkness . This requires that a respiratio n<br />

response be explicitly introduced into th e<br />

model.<br />

Figure 2 represents a generalized dark respiration<br />

response to temperature for steady -<br />

state conditions. The response at low an d<br />

intermediate temperatures is exponential i n<br />

keeping with the van 't Hoff Q 1 o rule for<br />

chemical reactions (Forward 1960). Abov e<br />

some high temperature, a decline resultin g<br />

from enzyme denaturation occurs which<br />

tends to become irreversible as the time o f<br />

exposure to high temperature increases (For -<br />

ward 1960, Longridge 1963) .<br />

Net CO 2 uptake, or net photosynthesis ,<br />

can now be expressed in terms of temperature-controlled<br />

respiration and light- an d<br />

temperature-controlled photosynthesis . For<br />

purposes of this model, net CO 2 fluxes int o<br />

the leaf are assigned a positive value whil e<br />

CO 2 losses are considered negative .<br />

Psn(L,T) = Ps(L,T)-Rs(T) (1)<br />

The functional form of the model can b e<br />

expanded around the following exponential<br />

expression that is representative of the ligh t<br />

curve in figure 1 .<br />

Psn(L,T) = Bo(T) (1 - exp(B 1 L)) - Rs(T) (2 )<br />

Rs(T) is negative and represents dark respiration<br />

. Bo(T) represents the light-saturate d<br />

asymptote as a function of temperature . Th e<br />

exponential coefficient (B 1 ) determine s<br />

response behavior at low light levels and is<br />

characteristically independent of temperature .<br />

Bo (T) can be represented by a quadratic function<br />

of the following form :<br />

Where :<br />

B o (T) = Bo + B i (T-B ) 2 (3 )<br />

Bo : maximum photosynthesi s<br />

B ; : slope coefficient ; negative algebraic sign<br />

B2 : temperature of maximum photosynthesi s<br />

Respiration as a function of temperature<br />

(Rs(T)) can be modeled with the followin g<br />

expression :<br />

Rs(T) = Boexp(BI (T-B.) - exp(BI (T-B . )) )<br />

Bo /exp(1) : maximum respiratio n<br />

B1 : slope coefficient ; positive sign<br />

B 2 : temperature of maximum respiratio n<br />

238


This function is asymptotic to zero respiration<br />

at both low and high temperatures an d<br />

exhibits an exponential increase in respiratio n<br />

followed by a rapid decline at high<br />

temperatures .<br />

The completed functional expression fo r<br />

Psn(L,T) is :<br />

Psn(L,T) = -Boexp[B 1 (T-B2 )-exp(B 1 (T-B2 ))]<br />

+ [ B 2 + B3 ( T -B4 ) 2 ] [1 - exp(B s L ) ]<br />

(4)<br />

The intent of this model is to predict ne t<br />

photosynthesis in light and CO 2 losses in<br />

darkness. The model is not useful for predicting<br />

gross photosynthesis without including<br />

some function for respiration in the light .<br />

Until the experimental techniques for obtaining<br />

such measurements are developed, th e<br />

light respiration term must remain implicit i n<br />

the gross photosynthesis function . This limit s<br />

the use of the model although most C 3 plant s<br />

will respond similarly . A different model may<br />

be required for C 4 plants such as grasses that<br />

require a much higher light level to saturat e<br />

photosynthesis .<br />

Methods and Materials<br />

Values for the six parameters in the model<br />

were determined with a least squares fit of th e<br />

data using an algorithm developed by Marquardt<br />

(1963) . Net photosynthetic data were<br />

taken with a controlled environment syste m<br />

developed by Webb (1971) . The test organis m<br />

for the model was red alder (Alnus rubra<br />

Bong.) .<br />

Forty red alder seedlings, 1- to 2-years-old ,<br />

were removed from the field and propagated<br />

in a nutrient culture before transferring them<br />

to the gas-tight controlled environment chamber.<br />

The root system of the seedlings was en -<br />

closed in a nutrient flow system with temperature<br />

controlled at 11°C ± 1°C . The plants<br />

were maintained in this system for 1 month at<br />

conditions of light and temperature consisten t<br />

with those in the greenhouse .<br />

Carbon dioxide absorption rates were measured<br />

by monitoring the CO 2 depletion in the<br />

gas-tight environment chamber with a Beckman<br />

IR gas analyzer. Atmospheric CO 2 levels<br />

in the chamber were maintained between 32 0<br />

and 335 p .p.m. Light energy was varied from<br />

0.06 to 0 .68 ly/min (total shortwave radiation)<br />

at air temperatures from 5 to 30°C . For<br />

each of five light levels, CO 2 uptake was<br />

measured while temperature was changed a t<br />

the rate of 1° per 5 minutes beginning at<br />

15°C and proceeding upscale to approximately<br />

30°C . Measurements were then made<br />

while decreasing temperature at the same rat e<br />

until near 5°C at which time temperature wa s<br />

increased again up to 16°C . All CO 2 uptak e<br />

measurements were made during 3 successiv e<br />

days. Relative humidity varied between 6 5<br />

percent at low temperatures to 75 percent at<br />

high temperatures .<br />

Results<br />

A portion of the data is plotted in figures 3<br />

and 4 . The photosynthetic response to radiation<br />

in figure 3 is characterized by a linea r<br />

response at low light followed by the usual<br />

light-saturated response at high radiation . Thi s<br />

is consistent with the findings of many other<br />

investigators . The saturation value for ne t<br />

photosynthesis increases with temperature ,<br />

but the increase is not linear .<br />

Figure 4 shows the net photosynthetic<br />

response to temperature for constant radiation<br />

of 0 .19 ly/min . A quadratic function wa s<br />

fit to the data and the parameters with their<br />

respective standard errors are listed in figure<br />

4 . At 0.19 ly/min, net photosynthesis has a<br />

maximum of 20 .5°C. Although these data<br />

represent the temperature range between 5 .5 °<br />

and 27°, the data of Pisek et al . (1969) indicate<br />

a nearly symmetrical response betwee n<br />

5° and 40°C. Extrapolation of the quadratic<br />

beyond this temperature range should b e<br />

done cautiously although Phillips an d<br />

McWilliams (1971) have measured zero CO 2<br />

fluxes at high temperatures . It may be that<br />

high temperatures increase respiration until it<br />

exceeds the photosynthetic capacity .<br />

Figure 5 illustrates the response surface<br />

generated by expression (4) which is the net<br />

239


w 10 0<br />

0<br />

U)<br />

w 80<br />

1-<br />

z<br />

>-<br />

u)<br />

0 60<br />

I--<br />

0<br />

22 °<br />

i6 °<br />

Psn t Rs (TI + Bo (T1 (I - e B2y<br />

°<br />

0<br />

t- 40<br />

w<br />

z<br />

w ><br />

2 0<br />

J<br />

w<br />

cc<br />

0<br />

1<br />

0 .2 0 .4 0.6 0 .8<br />

TOTAL SHORTWAVE RADIATIO N<br />

Figure 3 . Light effects on net photosynthesis of red alder at three temperatures .<br />

0 O<br />

w 4 0<br />

z<br />

w<br />

>_<br />

1- 20<br />

J<br />

w<br />

cc<br />

I<br />

10 20 30 40<br />

TEM<strong>PE</strong>RATURE, ° C<br />

Figure 4 . Net photosynthetic response of red alder to temperature at 0 .19 ly/min total s.w . radiation .<br />

240


The important interaction between light<br />

and temperature can best be shown by numerical<br />

example . Table 1 shows the predicted<br />

temperature of maximum net photosynthesis ,<br />

or optimum temperature, and the peaks of<br />

net photosynthesis for each of five levels of<br />

light. Both the optimum temperature and th e<br />

peaks of net photosynthesis increase with<br />

light. The latter is exemplary of photosynthetic<br />

light saturation .<br />

Table 1. Predicted effect of light on<br />

maximum net photosynthesis<br />

Light Relative Temperature of<br />

ly/min maximum Psn maximum Psn<br />

Figure 5 . Response surface of net CO 2 exchange generated<br />

by Psn(L,T) = -513.3 exp[ .088(T-47.7) -<br />

exp .088(T-47.71] + [187 .3 - .105(T-41 .41 2 1<br />

[1 - exp(-9.59L)] .<br />

photosynthetic response to both light an d<br />

temperature . The functional form of the surface<br />

is included in figure 5 along with th e<br />

values of the 6 parameters obtained from a<br />

least-squares fit of the CO 2 exchange data.<br />

The average deviation of the function fro m<br />

the data was ±9 .1-pct . and the r 2 was 0 .97 .<br />

Note that the light saturation phenomeno n<br />

of photosynthesis is conserved in this model<br />

as well as the symmetrical response to temperature.<br />

Although there are no data below 5 °<br />

and above 30°, the model seems well-behave d<br />

in the regions outside the data. As light decreases<br />

the gross photosynthesis functio n<br />

tends toward 0, and the dark respiration function<br />

begins to dominate . Below the light<br />

compensation point CO 2 fluxes become negative.<br />

At zero light only respiration occurs, an d<br />

because the function goes asymptotically t o<br />

zero at both low and high temperatures, the<br />

model never predicts an uptake of CO 2 in<br />

darkness .<br />

0 .07 27 .8 1 6<br />

.15 65 .5 2 0<br />

.19 76 .2 2 1<br />

.25 90 .8 2 2<br />

.68 100 .0 23<br />

The increased temperature of maximu m<br />

net photosynthesis associated with additiona l<br />

light can be explained as an interaction between<br />

the gross photosynthesis function an d<br />

dark respiration . At low light, photorespiration<br />

is minimal and photosynthesis is largely<br />

independent of the temperature increases that<br />

step-up dark respiration . Therefore, dark<br />

respiratory CO 2 losses from cellular maintenance<br />

are not compensated by photosynthesis<br />

as temperatures rise and maximum net photo -<br />

synthesis (photosnythesis minus respiration )<br />

occurs at a relatively low temperature . As<br />

light increases, photosynthesis and its depend -<br />

ence on temperature increases . This allows<br />

photosynthesis to keep pace with temperature-controlled<br />

dark respiration and maxi -<br />

mum net photosynthesis then shifts to a<br />

higher temperature .<br />

Although the model is empirical, it has<br />

several noteworthy features . The model is<br />

continuous over a wide range of independent<br />

variables and is therefore more mathemati-<br />

24 1


cally tractable than discontinuous or piece -<br />

wise continuous models . This continuous feature<br />

should facilitate linking this model t o<br />

others .<br />

The general net photosynthetic response of<br />

plants to light and to temperature has been<br />

demonstrated with many different specie s<br />

under various preconditioning treatments .<br />

The general responses characterized in figur e<br />

1 predominate with various shifts in the pea k<br />

and magnitude of the response curves . These<br />

shifts can be described mathematically by<br />

determining CO 2 exchange at certain point s<br />

and using these data to define parameters in<br />

the basic model .<br />

In the Biome program, CO 2 exchange<br />

studies are being conducted on several species<br />

and these data will further test the model pro -<br />

posed here. A necessary addition to the model<br />

is an explicit evaluation of light respiratio n<br />

from which techniques for data acquisitio n<br />

are now being contemplated .<br />

Acknowledgments<br />

Thanks are due Michael Newton and Duan e<br />

Starr for their contribution to the model an d<br />

to Jack Colby for his programming assistance .<br />

The work reported in this paper was sup -<br />

ported by the PHS grant 2P10-Es00210 i n<br />

cooperation with the Coniferous <strong>Forest</strong><br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution<br />

No . 39 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Chartier, P . 1969. A model of CO 2 assimilation<br />

in the leaf . In Prediction and measurement<br />

of photosynthetic productivity, p .<br />

307-315 . US/IBP Tech. Meet. Proc . ,<br />

Trebon .<br />

Forward, Dorothy F. 1960 . Effect of temperature<br />

on respiration . In W . Ruhlan d<br />

(ed .), Encyclopedia of plant physiology, p .<br />

234-258 . Berlin : Springer-Verlag.<br />

Heath, O .V .S. 1969 . The physiological aspect s<br />

of photosynthesis . 309 p. London : Heinemann<br />

Educ . Books Ltd .<br />

Larcher, W. 1969 . Physiological approaches t o<br />

the measurement of photosynthesis in relation<br />

to dry matter production by trees .<br />

Photosynthetica 3(2) : 150-166 .<br />

Longridge, J. 1963. Biochemical aspects of<br />

temperature response. Annu. Rev. Plant<br />

Physiol. 14 : 441-462 .<br />

Marquardt, Donald W . 1963 . An algorithm for<br />

least-squares estimation of nonlinear<br />

parameters. J. Soc. Ind. Appl. Math . 11(2) :<br />

431 .<br />

Milner, Harold W., and William H . Hiesey .<br />

1963. Photosynthesis in climatic races o f<br />

Mimulus . I. Effect of light intensive and<br />

temperature on rate . Plant Physiol . 39(2) :<br />

208-213 .<br />

Mooney, H. A., and A. T. Harrison. 1969 . The<br />

influence of conditioning temperature on<br />

subsequent temperature-related photosynthetic<br />

capacity in higher plants . In Prediction<br />

and measurement of photosynthetic<br />

productivity, p. 411-417 . US/IBP Tech .<br />

Meet. Proc., Trebon .<br />

Phillips, V . R., and J. R. McWilliams . 1971 .<br />

Thermal responses of primary carb o<br />

enzymes from C 3 and C 4 plant adapted to<br />

contrasting temperature environments . In<br />

M . D. Hatch, C . B . Osmund, and R . O .<br />

Slatyer (eds .), Photosynthesis and photorespiration,<br />

p . 97-109. Conf., Canberra,<br />

Australia, Nov. 1970 (Proc .). New York :<br />

Wiley-Interscience .<br />

Pisek, A., W. Larcher, W. Moser, and Ida<br />

Pack. 1969. Kardinale Temperaturbereiche<br />

der Photosynthese and Grenz-temperaturen<br />

des L e b ens der Blatter verschiedener<br />

S permato-phyten. III. Temperaturabhangegkeit<br />

and optimaler Temperaturbereich<br />

der Netto-Photosynthese. Flora 158 :<br />

608-630 .<br />

Rabinowitch, E . 1969 . Photosynthesis . 273 p .<br />

New York : John Wiley & Sons, Inc .<br />

Schulze, Ernst-Detlef. 1970. Der C02 -<br />

Gaswechsel der Buche (Fagus siluatica L . )<br />

in Abhangrgkeit von den Klimafaktoren i m<br />

Freiland. Flora 159 : 177-232 .<br />

Webb, Warren L. 1971 . Photosynthetic response<br />

models for a terrestrial plant community,<br />

79 p. Ph.D. thesis on file, Oreg.<br />

State Univ., Corvallis .<br />

242


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium.<br />

Bellingham, Washington-March 23-24, 197 2<br />

Energy flux studies in a<br />

coniferous forest ecosystem<br />

Abstract<br />

Lloyd W. Ga y<br />

Associate Professor of <strong>Forest</strong> Climatolog y<br />

Department of <strong>Forest</strong> Engineerin g<br />

Oregon State University<br />

Corvallis, Orego n<br />

The fluxes of thermal energy between the atmosphere and a young Douglas-fir forest were measured durin g<br />

two contrasting summer days, one cloudless and one overcast. The energy budget components were evaluated<br />

by the Bowen ratio method, with ceramic-wick psychrometers at the 26 .16 m, 28.16 m, or 31 .16 m levels. The<br />

maximum height of the tallest trees was 28 m, and the general level at the top of the closed canopy was abou t<br />

22 m. Daily totals of the energy budget components (cal/cm 2 ) under cloudless skies on July 29, 1971, were :<br />

solar radiation, 584; net radiation, 410 ; change in storage, 5 ; convection, -135; and latent energy, -280. The<br />

albedo was 0.09 on both the clear and the overcast day . Analysis of the overcast conditions of July 31, 1971 ,<br />

yielded the following values: solar radiation, 171 ; net radiation, 134 ; change in storage, 6 ; convection, -39; and<br />

latent energy, -102.<br />

Problems of measurement and analysis are discussed. These include the storage term in the biomass, and th e<br />

small gradients of potential temperature and vapor pressure above the canopy . Clear day gradients at noon, for<br />

example, were in the order of -0.03°C m-' and -0.03 mb m-' . Techniques are presented for minimizing measurement<br />

errors.<br />

Introduction<br />

The level of biological activity at the surface<br />

of the earth is closely associated wit h<br />

cycles of energy and mass . The cycles of mas s<br />

and of energy are virtually interchangeabl e<br />

concepts. Indeed, the transpiration an d<br />

photosynthetic components of the mass cycl e<br />

can be studied through examination of th e<br />

cycle of energy, with the energy required t o<br />

change the phase of H2 0 and CO 2 serving as<br />

the connecting link .<br />

The magnitudes and phase relationships of<br />

the mass and energy cycles are affected by th e<br />

characteristics of the surface, and by the stat e<br />

of the atmosphere . The properties of vegetation,<br />

particularly of low, cultivated ecosystems,<br />

have been investigated thoroughly i n<br />

a variety of studies that have clearly demonstrated<br />

many advantages for energy budge t<br />

evaluations of transpiration and photosynthesis<br />

(Baumgartner 1965) . The advantage s<br />

include sensitivity, mobility, and the benefit s<br />

to be gained by use of a nondestructive technique<br />

. The application of these techniques t o<br />

the forest ecosystem appears feasible and useful.<br />

A number of studies have already been<br />

reported, but in general the effects of forest s<br />

upon the cycles are not yet well known<br />

(Baumgartner 1971, Tajchman 1971) . Coniferous<br />

forests, as a class, are good absorbers of<br />

solar radiation. The roughness of coniferou s<br />

crowns also appears to effectively enhanc e<br />

mixing in the atmosphere near the top of th e<br />

canopy. These factors, combined with the<br />

large surface area of canopies, make forests<br />

into very efficient exchange surfaces for water<br />

vapor, carbon dioxide, and energy .<br />

Studies of the fundamental cycles of energ y<br />

and mass have begun at the Cedar River site i n<br />

the Coniferous <strong>Forest</strong> Biome. A variety of<br />

interrelated studies are planned in coopera -<br />

243


tion with physiologists, soil scientists, hydrologists,<br />

and meteorologists . The wor k<br />

reported here is of preliminary research int o<br />

the exchange of thermal energy between a<br />

young Douglas-fir forest and the atmosphere .<br />

The energy exchange processes have bee n<br />

evaluated during two contrasting conditions :<br />

the first during clear, warm weather characterized<br />

by a large energy input to the forest ,<br />

and the second during cool, overcast conditions<br />

with a relatively small energy input to<br />

the forest. The objectives are to define th e<br />

energy transfer characteristics of the youn g<br />

forest under conditions of both high and lo w<br />

rates of transpiration . Study of these contrasting<br />

conditions will help us to understand the<br />

processes that control the exchange of energy<br />

and water vapor between the forest canop y<br />

and the atmosphere .<br />

<strong>Experimental</strong> Methods<br />

The evaluation of thermal energy exchange<br />

in the forest is simple in concept, requirin g<br />

that appropriate boundaries be defined abou t<br />

the forest site, that the quantities of energ y<br />

crossing the boundaries be identified, and that<br />

appropriate measurements be made at periodic<br />

intervals . The periodic samples can the n<br />

be combined to yield estimates of energy ex -<br />

change for the desired time interval . Th e<br />

selection of an appropriate model is an important<br />

step in evaluating the thermal energy<br />

exchange .<br />

Energy Transfer Models<br />

Several review articles have discussed th e<br />

application of various energy transfer model s<br />

to the problem of evaluating the exchange between<br />

the forest and the atmosphere<br />

(Baumgartner 1965, Federer 1970, Fritsche n<br />

1970) .<br />

The basic problem concerns the transformation<br />

of energy by the forest from radiant<br />

to nonradiant forms . The total energy thus<br />

transformed is called net radiation, Q* ; this<br />

equals the quantity of radiation of all wavelengths<br />

absorbed at the surface, minus th e<br />

radiation lost by reflection and emission . The<br />

net radiation is transformed into either a<br />

change in stored heat in the soil and biomass ,<br />

G; a flow of convective (sensible) energy between<br />

the forest and the air, H; or a flow of<br />

latent heat, AE, that is associated with a flux<br />

of water vapor E . The amount of energy<br />

transformed in photosynthesis, P, is of great<br />

importance in productivity . However, since i t<br />

amounts to only a few percent in terms of the<br />

net radiation flux over the forest (Denmead<br />

1969), P will be neglected in this study .<br />

The rate and direction of energy transfer<br />

depends upon the relative energy states of th e<br />

canopy and the atmosphere and upon th e<br />

availability of radiant energy which is derived<br />

primarily from the sun . The state of energy is<br />

determined by temperature and by vapor concentrations.<br />

Motion of the atmosphere may<br />

also enhance energy transfer . Energy transfer<br />

models thus use measurements of temperature,<br />

vapor concentration, wind, and radian t<br />

energy in order to determine the flow of<br />

energy between the forest and the atmosphere .<br />

The "Bowen ratio" model was selected for<br />

use in this study because of relative simplicity<br />

in analysis and application, and because of its<br />

general acceptance based upon tests over<br />

other types of vegetation . The method was<br />

first derived by Bowen (1926), and has been<br />

adapted for energy transfer studies by a<br />

number of workers (Fritschen 1965, Tanner<br />

1960) .<br />

The Bowen ratio model has been thoroughly<br />

described elsewhere, but a short discussion<br />

here will help to place its applicatio n<br />

into perspective . First, we must sum th e<br />

thermal energy fluxes active in the forest i n<br />

accordance with the principle of conservatio n<br />

of energy, to obtain<br />

Q*+G+H+XE+P=0 . (1)<br />

The polarity convention considers fluxes t o<br />

the surface as positive . After neglectin g<br />

photosynthesis, equation 1 can be solved for<br />

latent energy to yield<br />

XE = - (Q* + G) / (1 + (3) ( 2)<br />

where i3 is the Bowen ratio of convective hea t<br />

to latent heat (H/XE). The Bowen ratio can be<br />

written<br />

244


(3 = H/AE=yo$/De (3 )<br />

where y is the psychrometric constan t<br />

(y ^ 0 .66 mb/°C at sea level) and Da and<br />

De represent the measured differences i n<br />

potential temperature and in vapor pressure at<br />

two levels in the atmosphere above, but near ,<br />

the surface .<br />

The Bowen ratio model thus provides a<br />

rationale for partitioning the measured suppl y<br />

of thermal energy into convection and laten t<br />

heat, based upon measurements of temperature<br />

and vapor concentration at only two<br />

levels above the canopy surface. Wind measurements<br />

are not required for the analysis ,<br />

although they are often useful in interpretation<br />

of the results . The supply of thermal<br />

energy (Q* + G) is measured, so limits ar e<br />

placed on the estimates of convection and<br />

latent heat with this method .<br />

The disadvantages of the model are primarily<br />

associated with instrumentation ;<br />

accurate measurements are required in order<br />

to measure the small gradients in temperatur e<br />

and vapor found near the forest canopy . In<br />

addition, the basic relationship in equation 2<br />

becomes undefined whenever 0=-1 . This<br />

normally occurs infrequently, and ordinarily<br />

only for short periods near dawn or dusk<br />

when the amount of available energy i s<br />

limited. The magnitude of H and XE wil l<br />

normally be small at such times .<br />

Site Conditions<br />

The energy exchange studies were carrie d<br />

out in the broad, flat valley of the Ceda r<br />

River, near Seattle, Washington. The soil and<br />

stand conditions have been described by<br />

Fritschen (1972) . The stand is second-growth<br />

Douglas-fir approximately 35 years old, with<br />

an average of 570 trees per ha . The stand<br />

canopy is relatively level ; the height of the<br />

tips of the tallest trees in the vicinity of th e<br />

experimental site is about 28 m . The site is<br />

adjacent to the lysimeter tree described by<br />

Fritschen (1972) . A series of energy transfer<br />

studies are planned at this site in the future ,<br />

using the lysimeter tree, meteorological<br />

towers, and a variety of models for evaluatin g<br />

energy exchange processes .<br />

Field Measurement s<br />

Net radiation, stored heat, and temperature<br />

and vapor pressure measurements were obtained<br />

with a mobile data acquisition that ha s<br />

been described by Gay . l The truck-mounted<br />

system includes sensors and supports, cabling ,<br />

and a digital data logger with resolution of<br />

0.001 percent (0 .1 microvolt on a 10 millivol t<br />

scale). Ceramic wick, wet-bulb psychrometers<br />

of the basic design of Lourence and Pruit t<br />

(1969), as modified by Gay (n .d.), were used<br />

for the temperature and vapor pressur e<br />

measurements.<br />

The sensors were mounted on a 33 .5 m<br />

(110 ft) tall, triangular TV tower about 0 .3 m<br />

(1 foot) in width. Wind, temperature, and<br />

vapor pressure measurements were made a t<br />

six levels, respectively, 26 .16, 27 .16, 28 .16 ,<br />

29 .16, 30 .16, and 31 .16 m above the fores t<br />

floor. The radiation budget components were<br />

measured from a height of 30 .66 m above th e<br />

floor. The tip of the tallest tree in the vicinity<br />

of the tower extended to 28 m, though th e<br />

bulk of the crowns were below 24 m, and th e<br />

general level of crown closure was about 20 t o<br />

22 m above the floor . Five soil heat flux disks<br />

were installed at the -2 cm level, just beneat h<br />

the surface of mineral soil on the forest floor .<br />

Observations began on July 27 and continued<br />

through August 1. The sensors were<br />

sampled at 5-minute intervals during the day ,<br />

and at 10-minute intervals at night . The observation<br />

period spanned a range of weather conditions<br />

that included one clear and one completely<br />

overcast day. The data acquisitio n<br />

system performed well during this period .<br />

Problems in <strong>Forest</strong> Energy<br />

Budget Analyses<br />

A variety of measurement problems are en -<br />

countered in energy budget studies. In addition,<br />

forests have unique characteristics o f<br />

scale and mass that affect the application o f<br />

I L . W. Gay. An environmental data acquisitio n<br />

system . National Conference on the <strong>Forest</strong>, Weather ,<br />

and Associated Environment, Atlanta, Georgia, May<br />

18-19, 1971 . Mimeo., 8 p . Abstracted : Bull. Am .<br />

Meteor . Soc . 52 : 202-203 .<br />

245


the basic energy budget model given in equation<br />

2 . The gradients of temperature an d<br />

vapor above the rough, porous forest canop y<br />

are very slight, and measurement difficultie s<br />

increase with small gradients . Another major<br />

problem is related to the difficulty of measuring<br />

changes in stored energy in the biomass of<br />

the forest . These problems will be placed int o<br />

perspective at the Cedar River site, for the y<br />

affect the analyses and the subsequent interpretation<br />

of the results .<br />

Evaluation of Gradients<br />

Gradients of temperature, vapor concentration,<br />

and wind are small above fores t<br />

canopies, even though the transfer of energy<br />

and mass may be proceeding at high rates .<br />

The small gradients result from mixing induced<br />

by the mechanical turbulence create d<br />

by the rough canopy surface . In addition, the<br />

exchange surfaces are distributed through a<br />

considerable canopy depth so that the source s<br />

of heat and vapor are diffuse, rather tha n<br />

being concentrated as in the dense canopies of<br />

crops or other low vegetation . The smal l<br />

gradients above the forest require tha t<br />

extreme care be taken in the development o f<br />

suitable instrumentation, and in the experimental<br />

design controlling the deployment o f<br />

sensors .<br />

The Bowen ratio model assumed tha t<br />

steady state conditions prevail, i .e., the value s<br />

of the variables do not change with respect t o<br />

time during the period of analysis . This i s<br />

partially satisfied by averaging the values ove r<br />

the period of an hour before applying the<br />

model . Integration into hourly means als o<br />

reduces the small random component of error<br />

associated with the measurements. It does<br />

not, however, reduce biases that are introduced<br />

by small differences among sensors .<br />

Such biases can be a source of serious error ,<br />

particularly with small gradients that exis t<br />

above the forest . Bias errors must be handle d<br />

by techniques other than averaging .<br />

Two approaches have been used to cope<br />

with the problems involved in the measurement<br />

of small gradients above forests . In the<br />

first approach, the two levels of measuremen t<br />

required for the Bowen ratio are separated by<br />

a relatively large distance (10 m) in order to<br />

increase the differences that are being measured<br />

to a level commensurate with the sensitivity<br />

of the data system (Galoux et al . 1967 ,<br />

Storr et al. 1970) . In the second, reversing<br />

sensors have been used to cancel the effect of<br />

any small biases that may exist between<br />

sensors (Black and McNaughton 1971) .<br />

A large vertical separation of the sensors<br />

introduces questions of the representativenes s<br />

of the measured gradients which should represent<br />

the effect of the underlying surface . It is<br />

quite possible that a sensor placed 10 m abov e<br />

the canopy may be measuring properties o f<br />

the atmosphere that are derived from a surface<br />

other than the one under investigation .<br />

In contrast, a pair of reversing sensors, place d<br />

near the canopy, offers an excellent means fo r<br />

evaluating small gradients . However, the<br />

reversing mechanism introduces new problem s<br />

of design for operation and support in tal l<br />

forests .<br />

A graphical approach was used in this study<br />

to minimize sensor errors in the gradien t<br />

measurements used for the Bowen ratio<br />

analysis . Initially, the mean hourly values o f<br />

potential temperature were plotted agains t<br />

the associated values of vapor pressure t o<br />

yield 24 plots (1 per hour) for each day, with<br />

each plot containing six points (one for eac h<br />

measurement level) . The plots will be straigh t<br />

lines if similarity exists between the gradient s<br />

of potential temperature and vapor concentration,<br />

providing there are no errors of measurement<br />

(Tanner 1963). Since similarity is an<br />

assumption of the Bowen ratio method, thes e<br />

plots were used to separate out the instrument<br />

levels that exhibited small offset s<br />

throughout the day and to identify thos e<br />

levels appropriate for use in the Bowen rati o<br />

analysis .<br />

The similarity plots for July 29 are shown<br />

for levels 1, 4, 6 in figure IA, and for levels 1 ,<br />

2, 3 on July 31 in figure 1B . The potentia l<br />

temperature (e-) scale on the ordinate differ s<br />

between the 2 days. The vapor pressure (e )<br />

scale is shown in the legend . Note that the plo t<br />

for each hour has been normalized by subtracting<br />

the & and e value at the bottom level fro m<br />

the observations at each of the other levels .<br />

Therefore each plot actually shows the incre -<br />

246


0.4<br />

0.2<br />

v 0<br />

0<br />

w<br />

IX -0.2<br />

I-<br />

cr -0.4<br />

w<br />

-0.6<br />

~ 0.04<br />

A ..29<br />

JULY<br />

HOUR (PDT) -->- y y g i<br />

0 I 2 3 4 5 6 7 1I I 1 1! I r 17 .19 202122 23 2 4<br />

9 10 II 12 13 14 15 16 o 18<br />

B. 31 JULY<br />

r<br />

a<br />

I<br />

2 3 4 5 6 7 8 9<br />

HOUR (PDT)----4 .-<br />

10<br />

I I<br />

2.03.122! J<br />

1213 14 15 16 17 18 19 ° 1 23 24<br />

0- I<br />

u<br />

-0.04 • MB<br />

VAPOR PRESSURE, M B<br />

Figure 1 . Similarity between gradients of potential temperature and vapor pressure . A. Clear weather on Jul y<br />

29, 1971 . Levels 1, 4, and 6 . B . Overcast weather on July 31, 1971 . Levels 1, 2, and 3 .<br />

ment of a and e with respect to the firs t<br />

measurement level near the canopy .<br />

Though several points can be deduced fro m<br />

such similarity plots, the most important conclusion<br />

concerns adequacy of data . The linearity<br />

at the selected levels confirms their<br />

acceptability for the Bowen ratio model ,<br />

although an unexplained offset is evident at<br />

level 2 during the 1000-1200 hours on Jul y<br />

31 .<br />

Once the data is judged acceptable, on e<br />

notes that the slope of the lines is A&/Ae ; this<br />

is directly proportional to 0, the Bowen ratio ,<br />

as shown in equation 3 . Thus the relative<br />

slope of the similarity plots is an index to th e<br />

way that the surface is partitioning the net<br />

energy supply into convection and evaporation.<br />

Further, the quadrant of each hourl y<br />

plot indicates the sign of 0, and the direction<br />

of the H and XE fluxes .<br />

For example, as plotted in figure 1A and<br />

1B, both fluxes will have the same sense in<br />

quadrants I and III where the slope is positiv e<br />

((3>0). In quadrant I, H and XE are both directed<br />

away from the surface and are negative<br />

by the sign convention adopted earlier . In<br />

quadrant III, H and XE are both positive, a s<br />

their sense is toward the surface . The slop e<br />

and the (3 coefficient are negative in quadrant s<br />

II and IV. In quadrant II, H is toward th e<br />

surface (advection) while XE is directed away ,<br />

while the fluxes in quadrant IV have th e<br />

opposite sense . The similarity plots can thu s<br />

provide three things : a ready indication of the<br />

magnitude of the Bowen ratio ; an indicatio n<br />

of the direction of the H and E fluxes ; and an<br />

identification of levels suited for the Bowe n<br />

ratio analysis .<br />

247


Evaluation of Stored Hea t<br />

Application of the Bowen ratio model t o<br />

forest measurements reveals a problem i n<br />

measuring the stored heat term in equation 2<br />

that is a direct consequence of the nature of<br />

the forested surface . The scale of the forest<br />

elements makes the change in heat storage i n<br />

the biomass difficult to measure, and man y<br />

studies (Baumgartner 1956) have treate d<br />

these changes as negligible . This is certainly<br />

true on a daily basis, but an appreciable<br />

amount of energy appears to move into<br />

storage (-G) in the biomass in the early morning,<br />

and out again (+G) in the early evening .<br />

These amounts ordinarily will balance when<br />

totaled over the day, but comparisons amon g<br />

hourly totals may be in error unless the<br />

energy storage changes are estimated for th e<br />

biomass as well as for the soil .<br />

The estimates of biomass storage change<br />

were derived here by an indirect method that<br />

may prove useful in other situations as well .<br />

First, application of the Bowen ratio model to<br />

data collected during the early evening hours<br />

frequently resulted in positive estimates o f<br />

latent energy (condensation) when the vapo r<br />

gradient clearly indicated that evaporatio n<br />

was taking place . This apparent anomaly in<br />

the Bowen ratio estimate of XE could occur<br />

as a consequence of underestimating the<br />

storage term G . In order to obtain the proper<br />

sign on XE under these conditions, the estimate<br />

of G must be increased to a positiv e<br />

value that exceeds the absolute value of th e<br />

negative net radiation .<br />

Estimates of the minimum probable value<br />

of G were obtained in this manner for th e<br />

hours between sunset and the time near mid -<br />

night when the vapor gradient changed direction,<br />

indicating the beginning of condensation<br />

. As a second step, the crossover points<br />

between release (+G) and uptake (-G) o f<br />

stored energy were then estimated from m y<br />

experience and that of others (Grulois 1968 )<br />

as being about 1 hour after sunrise and 3<br />

hours before sunset . The release of store d<br />

heat (+G) is then defined by the two cross -<br />

over points and by the magnitude during th e<br />

early evening hours . The third and final ste p<br />

was to estimate the magnitude of the gain in<br />

storage (-G) during the morning hours so tha t<br />

the daily integral approximately balanced .<br />

The final result is a much better estimate o f<br />

hourly changes in stored energy than could be<br />

obtained by direct measurement of the soil<br />

storage component alone. The magnitude of<br />

the stored energy change will be investigate d<br />

further during future studies .<br />

The major problems in application of th e<br />

Bowen ratio model to the forest appear to b e<br />

associated with adequate precision of th e<br />

measurements. The similarity tests demonstrated<br />

here appear useful in eliminatin g<br />

errors from the data . Further, the changes in<br />

stored energy can be estimated by an indirect<br />

method, based upon gradient measuremen t<br />

and knowledge of the time at which th e<br />

stored heat flux reverses sign .<br />

The <strong>Forest</strong> Energy Budget<br />

Energy budget analyses were developed fo r<br />

2 days that represent quite different amount s<br />

of available energy . The solar energy input to<br />

the forest was large on July 29, 1971, a da y<br />

characterized by clear skies and a warm mea n<br />

temperature (24.4°C). In contrast, July 31 ,<br />

1971, was completely overcast with reduced<br />

levels of incoming solar radiation and a relatively<br />

cool mean temperature (17 .4°C) .<br />

Examination of the energy budgets unde r<br />

such contrasting conditions will improve our<br />

understanding of the basic processes tha t<br />

govern energy transfer between the forest and<br />

the atmosphere .<br />

The diurnal energy budget will first b e<br />

examined with respect to daily totals; a discussion<br />

of the relationships among hourly<br />

values of the energy budget components wil l<br />

then follow .<br />

The Diurnal Energy Budget<br />

The daily energy budget totals are presented<br />

for the separate periods of daylight (1 3<br />

hours) and night (11 hours) and for 24-hou r<br />

totals in table 1 . Dividing the daily totals into<br />

daylight and night portions enhances futur e<br />

comparisons that may be made with dat a<br />

collected under different daylengths at Ceda r<br />

River or elsewhere .<br />

248


Table L-Diurnal energy budget components '<br />

Period<br />

July 29-clear<br />

July 31-overcast<br />

Q* G H XE @* G H X E<br />

cal/cm2<br />

Daylight 454 -43 -143 -263 141 -7 -38 -9 6<br />

Night -44 48 8 -17 -7 13 -1 - 6<br />

Daily total 410 5 -135 -280 134 6 -39 -102<br />

1 Totals are given for the daylight hours, 0630-1930 PDT ; night hours, 1930-0630 PDT; and the full day ,<br />

0000-2400 PDT .<br />

There is a large difference in the radiatio n<br />

supply on the 2 days, as net radiation totale d<br />

410 cal/cm 2 under the clear skies of July 29 ,<br />

and only 134 cal/cm 2 for the overcast conditions<br />

of July 31 . These totals include a stead y<br />

net loss of radiation at night, amounting t o<br />

-44 cal/cm 2 under clear skies, and -7 cal/cm 2<br />

under the overcast conditions .<br />

The net radiation term represents the<br />

energy converted from radiative to nonradiative<br />

forms by the forested surface . The shortwave<br />

radiation from the sun makes up th e<br />

largest component of the net radiation . During<br />

the 13 hours of daylight on the clear day ,<br />

the forest received 584 cal/cm 2 of solar radiation<br />

and reflected 55 cal/cm 2 . Under overcast<br />

skies, the forest received 171, and reflected<br />

16, calories/cm 2 . The albedo was 0 .09 on<br />

both days .<br />

Since the gain and loss of longwave radiation<br />

also enters into the supply of radian t<br />

energy, it is not helpful to calculate a shortwave/net<br />

radiation ratio as an index of efficiency<br />

of conversion . However, the lo w<br />

albedo value (0.09) emphasizes the efficienc y<br />

with which the Douglas-fir canopy absorbs<br />

solar radiation. This low reflectivity is similar<br />

to values reported for other coniferous canopies<br />

(Stewart 1971), and is much lower tha n<br />

the 0 .2-0.25 albedo values that commonl y<br />

prevail over crops and other low vegetatio n<br />

(Monteith and Szeicz 1961) . As noted by<br />

Baumgartner (1971), forests are effectiv e<br />

absorbers of solar radiation .<br />

The changes in stored energy (G) tabulate d<br />

in table 1 for the 24-hour period are nea r<br />

zero, which is in accord with observation s<br />

reported elsewhere . This term represents the<br />

changes in heat storage of both the biomass<br />

and the soil. Most of the storage changes ar e<br />

attributed to the biomass ; the indirect<br />

methods used to estimate the changes in<br />

storage have been described in an earlier section<br />

.<br />

I estimate that -43 cal/cm 2 went into storage<br />

during the daylight hours on the clear da y<br />

and that 48 cal/cm 2 came out of storage during<br />

the night. The storage changes on the<br />

cloudy day proceeded in a similar direction ,<br />

but the magnitudes were much smaller .<br />

The storage term at Cedar River appear s<br />

large because of the large quantity of the biomass<br />

there . The biomass is as yet unmeasured ,<br />

however. Attempts will be made to measure<br />

the storage flux directly in future experiments .<br />

The convective flux for the clear day<br />

totaled -135 cal/cm 2 , directed away from the<br />

forest into the atmosphere . A slightly larger<br />

amount, -143 cal/cm2 , was lost during daylight,<br />

but 8 cal/cm2 was gained by the canopy<br />

at night when the canopy temperature s<br />

dropped below that of the air. Under overcas t<br />

sky conditions, -38 cal/cm 2 were lost over the<br />

full day .<br />

24 9


Latent energy was the largest dissipation<br />

term on each of the days, totaling -28 0<br />

cal/cm 2 for July 29 and -102 cal/cm 2 on July<br />

31 . There was a net loss of latent energy b y<br />

night, as well as by daylight, for both days .<br />

The evaporation equivalent of the laten t<br />

energy total was about 0 .5 cm on July 29 ,<br />

and 0 .18 cm on July 31 .<br />

The Bowen ratio ((3 =H/AE) is a measure of<br />

how the surface partitions the energy suppl y<br />

between sensible and latent heat . The mean<br />

daily value of 13 was 0.48 for the clear day ,<br />

and 0 .38 for the overcast day. The differenc e<br />

in (3 between days is not large, but it suggest s<br />

that the forest partitioned more of the energy<br />

supply into convection on the clear day tha n<br />

on the overcast day . From another viewpoint,<br />

the ratio of XE/Q* was 0 .67 on the sunn y<br />

day, and 0 .76 on the overcast day . This is i n<br />

the direction that one might expect for a<br />

stand of vegetation that receives a large inpu t<br />

of energy .<br />

Hourly Energy Budgets<br />

The phase relationships among the energy<br />

budget components can be examined with the<br />

aid of figure 2 which shows the hourly value s<br />

on July 29, and figure 3 which shows th e<br />

hourly values on July 31 . Each plotted poin t<br />

represents the midpoint of an hourly mean .<br />

The daytime, night and daily totals in table 1<br />

were integrated from the rates shown in thes e<br />

two figures . The 2 days exhibit different characteristics,<br />

so they will be discussed separately .<br />

The symmetry of the bell-shaped net radiation<br />

curve on July 29 confirms that the skie s<br />

were cloudless on that day . The maximu m<br />

intensity occurred during the hour centere d<br />

on 1300 hours PDT, which closely corresponded<br />

with solar noon . The net radiation<br />

values became positive about 1 hour after sun -<br />

rise and remained positive until shortly befor e<br />

sunset, indicating the hours in which there<br />

was a surplus of energy that might be dissipated<br />

through the other energy budget components.<br />

The net radiation was negative<br />

throughout the night, as the surface lost<br />

energy to the atmosphere. The greatest net<br />

radiation loss occurred at 2200, about 1 hou r<br />

after sunset at a time when the forest wes<br />

rapidly losing the absorbed solar radiatio n<br />

that had been stored during the day .<br />

The phase of the fluxes is also of interest ,<br />

4 8 10 12 14 16 18 20 22 24<br />

TIME, HR (PDT )<br />

Figure 2. Energy budget components under clear skies . Symbols: net radiation, Q* ; change in heat storage of<br />

soil and biomass, G ; convection, H ; latent energy, AE .<br />

250


4 8 10 12 14 16 18 20 22 24<br />

TIME, HR (PDT )<br />

Figure 3 . Energy budget components under overcast skies . Symbols : net radiation, Q* ; change in heat storage<br />

of soil and biomass, G ; convection, H ; latent energy, XE .<br />

as G, H, and XE all lag behind Q* . Let us<br />

consider the stored heat flux first . It reache s<br />

its peak flow into the biomass and soil (-G) i n<br />

midmorning, and reverses to flow out of th e<br />

biomass (+G) in the late afternoon and earl y<br />

evening. The change in stored heat appears t o<br />

provide a significant source of energy to th e<br />

surface throughout the night .<br />

The sensible heat flux, H, reaches its maxi -<br />

mum about two hours after G, but still a n<br />

hour before solar noon . Sensible heat i s<br />

directed away from the surface during day -<br />

light (-H), but reverses in direction as convection<br />

begins to provide energy (+H) to th e<br />

surface during the night . During this period ,<br />

the canopy cools below air temperature du e<br />

to longwave emission .<br />

The latent energy flux reached its maxi -<br />

mum about two hours after solar noon .<br />

Evaporation continued well into the night ;<br />

only during the early morning hours did a<br />

rather small amount of condensation tak e<br />

place .<br />

The marked phase shift between sensible<br />

and latent energy is of interest, as many<br />

studies have shown these two fluxes to be i n<br />

phase with net radiation (Baumgartner 1956 ,<br />

Denmead 1969, Rauner 1960) . The Douglas -<br />

fir forest, in contrast, partitioned the energy<br />

available at the surface into sensible and<br />

latent energy on a preferential basis . This<br />

partition was on a 1 :1 basis during the morning,<br />

but latent energy was apparently favored<br />

at the expense of sensible energy during th e<br />

afternoon . A similar pattern is evident i n<br />

measurements over a young Douglas-fir forest<br />

near Vancouver, B .C. (Black and McNaughto n<br />

1971), and over a mixed hardwood forest<br />

(Grulois 1968) .<br />

The phase shift in latent energy into the<br />

afternoon is probably related to a vapor pressure<br />

deficit which has an afternoon maximu m<br />

on clear days . Stewart and Thom 2 have concluded<br />

that the latent energy flux from thei r<br />

pine forest site in England is controlled more<br />

by the vapor pressure deficit than by th e<br />

supply of available energy . This conclusion is<br />

based upon their evaluation of the interplay<br />

between the relatively large internal resistanc e<br />

to transfer and a small external resistance ; the<br />

ratio for the pine site was in the order o f<br />

20 :1 .<br />

Conclusions<br />

The observations reported here are a n<br />

initial contribution toward the problem of<br />

evaluating the flow of energy and mass between<br />

the atmosphere and the young Douglas -<br />

fir forest at the Cedar River site .<br />

2 J . B. Stewart and A . S. Thom . Energy budgets i n<br />

pine forest. Institute of Hydrology, Wallingford ,<br />

Berkshire, U . K . Unpublished manuscript, 1972 .<br />

251


<strong>Forest</strong>ed surfaces are generally considere d<br />

to be effective energy exchange surfaces . Th e<br />

results confirm that this young stand has a<br />

high absorptivity for solar radiation, with a n<br />

albedo of 0 .09 for both clear and overcas t<br />

conditions . This high absorptivity contribute s<br />

to the large net radiation values that wer e<br />

measured under clear skies .<br />

The role of the forest in dissipating the net<br />

radiation is of particular interest. The porous ,<br />

aerodynamically rough canopy is effective i n<br />

transferring sensible and latent energy into<br />

the atmosphere . The large quantity of forest<br />

biomass may also involve an amount of store d<br />

thermal energy that is of significance durin g<br />

short periods, even though the daily totals are<br />

small . Summed over a 24-hour period, evapotranspiration<br />

was about 280 cal/cm 2 min (0 . 5<br />

cm water equivalent) or about two-thirds o f<br />

the net radiation that was transformed unde r<br />

clear skies . Evapotranspiration was relativel y<br />

larger on an overcast day, about three -<br />

quarters of net radiation, although the tota l<br />

amount of latent energy (102 cal/cm2 min, or<br />

0.18 cm water equivalent) was considerably<br />

lower .<br />

These results provide initial estimates o f<br />

the amounts of energy transferred during extreme<br />

conditions under cloudless and unde r<br />

overcast skies . The exchange of energy an d<br />

mass depends not only upon the energy inpu t<br />

to the forest, however, but also upon the<br />

physiological response of the vegetation . No w<br />

that the instrumentation system and the<br />

analysis model have been tested at this site ,<br />

subsequent research will include a range of<br />

environmental conditions . Instrumentation<br />

development and model testing will continue<br />

in cooperation with the lysimeter installatio n<br />

and eddy flux system of cooperating investigators<br />

. Ultimately, analysis and interpretatio n<br />

of the energy transfer studies must include<br />

physiological as well as physical characteristics<br />

of the stand .<br />

Acknowledgments<br />

The work upon which this publication i s<br />

based was supported in part by funds provided<br />

by the U .S. Department of Interior,<br />

Office of Water Resources Research, as<br />

authorized under the Water Resources Re -<br />

search Act of 1964, and administered by th e<br />

Water Resources Research Institute, Orego n<br />

State University ; and the National Scienc e<br />

Foundation Grant No . GB-20963 to the<br />

Coniferous <strong>Forest</strong> Biome, U .S. Analysis o f<br />

Ecosystems, International Biological Program .<br />

This is Contribution No . 40 to the Coniferous<br />

<strong>Forest</strong> Biome and Paper 844, <strong>Forest</strong> Researc h<br />

Laboratory, School of <strong>Forest</strong>ry, Oregon State<br />

University .<br />

Literature Cited<br />

Baumgartner, A. 1956 . Investigations on th e<br />

heat- and water economy of a young forest.<br />

Translated from Ber . Deut . Wetterdienst 5 :<br />

4-53 . Commonwealth Sci. & Ind. Res .<br />

Organ . Translation 3760, Melbourne ,<br />

Australia, 1958 .<br />

1965. The heat, water and car -<br />

bon dioxide budget of plant cover : methods<br />

and measurements . In F. E . Eckardt<br />

(ed.), Methodology of plant eco-physiology:<br />

Proceedings of the Montpellie r<br />

Symposium, p. 495-512 . Paris: UNESCO .<br />

1971 . Wald als Austauschfakto r<br />

in der Grenzschicht Erde/Atmosphare .<br />

Forstwiss. Cbl. 3: 174-182 .<br />

Black, T. A., and K. G. McNaughton . 1971 .<br />

Psychrometric apparatus for Bowen-ratio<br />

determination over forests . Bound . Laye r<br />

Meteorol . 2 : 246-254 .<br />

Bowen, I. S. 1926 . The ratio of heat losses by<br />

conduction and by evaporation from an y<br />

water surface . Phys. Rev . 27 : 779-787 .<br />

Denmead, O. T. 1969 . Comparative micro -<br />

meteorology of a wheat field and a fores t<br />

of Pinus radiata . Agric. Meteorol. 6 :<br />

357-371 .<br />

Federer, C. A. 1970. Measuring forest evapotranspiration-theory<br />

and problems . USDA<br />

<strong>Forest</strong> Serv . Res . Pap . NE-165, 25 p .<br />

Northeast. <strong>Forest</strong> Exp . Stn ., Upper Darby ,<br />

Pa .<br />

Fritschen, L . J. 1965 . Accuracy of evapotranspiration<br />

determinations by the Bowe n<br />

ratio method . Bull. Int. Assoc . Sci. Hydrol .<br />

10: 38-48 .<br />

252


1970. Evapotranspiration an d<br />

meteorological methods of estimation as<br />

applied to forests . In J . M. Powell and C .<br />

F. Nolasco (eds .), Proceedings of the Third<br />

<strong>Forest</strong> Microclimate Symposium, p. 8-27 .<br />

Can. For. Serv ., Calgary, Alberta.<br />

1972. The lysimeter installation<br />

on the Cedar River Watershed . In Jerry F .<br />

Franklin, L. J. Dempster, and Richard H .<br />

Waring (eds.), Proceedings-research on<br />

coniferous forest ecosystems-a symposium,<br />

p. 255-260, illus . Pac. Northwes t<br />

<strong>Forest</strong> & Range Exp . Stn ., Portland, Oreg.<br />

Galoux, A., G . Schnock, and J . Grulois . 1967 .<br />

Les installations ecoclimatologiques .<br />

Travaux Stat . Rech . Eaux et Forets ,<br />

Groenendaal-Hoeilaart, S4rie A, 12, 52 p .<br />

Gay, L. W. (n .d.) On the construction and use<br />

of ceramic-wick psychrometers . In R. W .<br />

Brown and B . P. Van Haveren (eds .) ,<br />

Psychrometry in water relations research .<br />

Utah Agric . Exp. Stn. (In press . )<br />

Grulois, J. 1968 . Flux thermiques et evapotranspiration<br />

au tours d 'une journ4e<br />

sereine . Bull. Soc . r. Botanique Belg . 102 :<br />

27-41 .<br />

Lourence, F. J ., and W. O. Pruitt . 1969 . A<br />

psychrometer system for micrometeoro<br />

l o gy profile determination. J. Appl .<br />

Meteorol . 8: 492-498 .<br />

Monteith, J. L., and G. Szeicz. 1961. Th e<br />

radiation balance of bare soil and vegetation.<br />

Quart. J. Roy. Meteorol. Soc. 87 :<br />

159-170 .<br />

Rauner, Yu. L. 1960 . The heat balance of forests.<br />

Izvestiya Akademii Nauk SSSR, Seriya<br />

Geograficheskaya. 1 : 49-59 . Translated by<br />

Israel Prog. Sci. Transl. No. 1342 . U .S .<br />

Dep. Commerce, Washington, D .C .<br />

Stewart, J . B. 1971. The albedo of a pine forest.<br />

Quart. J. Roy . Meteorol. Soc. 97 :<br />

561-564 .<br />

Storr, D., J . Tomlain, H . F . Cork, and R . E .<br />

Munn. 1970. An energy budget study<br />

above the forest canopy at Marmot Creek ,<br />

Alberta, 1967 . Water Resour . Res. 6 :<br />

705-716 .<br />

Tanner, C . B. 1960. Energy balance approac h<br />

to evapotranspiration from crops . Soil Sci .<br />

Soc . Am . Proc . 24: 1-9 .<br />

1963 . Basic instrumentation an d<br />

measurements for plant environment an d<br />

micrometeorology . Soils Bull . 6, various<br />

pagination . Madison : Univ . Wis .<br />

Tajchman, S . J. 1971 . Evapotranspiration an d<br />

energy balances of forest and field. Water<br />

Resour. Res . 7: 511-523 .<br />

253


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

The lysimeter installation on<br />

the Cedar River Watershed<br />

Leo J . Fritsche n<br />

Associate Professo r<br />

College of <strong>Forest</strong> Resource s<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

A lysimeter was built around the root ball of a 28 m Douglas-fir (Pseudotsuga menziesii) tree. The container ,<br />

tree, and soil weigh 28,900 kg. The sensitivity of the weighing mechanism is 630g which is equivalent t o<br />

0.06 mm of water. The installation will be used to study evapotranspiration and volume changes in relation t o<br />

soil water potential and atmospheric demand; to test cuvette and meteorological methods ; determine canop y<br />

interception ; and to assess the effects of irrigation and fertilization .<br />

Introduction<br />

Understanding the process and quantifying<br />

the rate of water transfer within a forest ecosystem<br />

has become increasingly more important<br />

in recent years . This concern has originated<br />

from both applied and basic directions .<br />

For example, it has been estimated that by<br />

1980, six major regions will have no wate r<br />

reserve. These regions are Colorado River ,<br />

South Pacific, Great Basin, Upper Ri o<br />

Grande, Pecos River, and Upper Missour i<br />

River (Colorado River Association 1966) .<br />

Furthermore, interbasin transfer is bein g<br />

practiced in at least two areas at the presen t<br />

time .<br />

Increasing need for water conservation, th e<br />

possibility of interbasin water transfer, and<br />

the treatment of watersheds to increase yiel d<br />

or change quality has demonstrated the nee d<br />

for additional research on water use by various<br />

types of vegetative cover. For example ,<br />

clearcutting a north-facing Coweeta watershe d<br />

resulted in a 1st year increase streamflow o f<br />

40.2 cm and a stabilized increase of 23 .8 cm .<br />

Although the 1st year 's increase from a south -<br />

facing watershed was only 15 cm which de -<br />

creased to insignificance by the 3d year<br />

(Swift, unpublished data), it is believed that<br />

the different microclimatic influences upon<br />

the vegetation and resulting evapotranspiration<br />

will account for the discrepancy in wate r<br />

yield. To understand fully the processes involved<br />

and to predict results from futur e<br />

cuttings require detailed investigations relating<br />

water use to its availability and to atmospheric<br />

conditions .<br />

Short-period (1 hour or less) determinations<br />

of evapotranspiration are necessary t o<br />

study the complex soil-plant-atmospheric relations.<br />

Many methods or combination o f<br />

methods have been employed to determin e<br />

evapotranspiration. The major ones includ e<br />

watershed runoff ; measurement of soi l<br />

moisture depletion by either gravimetri c<br />

sampling, with resistance blocks, or with neutron<br />

soil moisture meters ; use of weighin g<br />

lysimeters ; and application of meteorologica l<br />

models . Other methods have employe d<br />

percolation tension plates, stemflow measurements,<br />

and enclosures . Of these methods only<br />

weighing lysimeters, enclosures, and meteorological<br />

models are capable of yielding hourl y<br />

results . Enclosures alter the microclimate an d<br />

255


may bias the results . Meteorological method s<br />

must be tested in the areas of anticipated us e<br />

(Slatyer and Mcllroy 1961) . Such testing has<br />

not been accomplished in forested areas .<br />

Thus, weighing lysimetry appears to be th e<br />

only method capable of yielding the neede d<br />

short period accurate rates .<br />

Weighing lysimeters have been used to determine<br />

evapotranspiration of agriculture<br />

crops (McIlroy and Angus 1963, Pruitt and<br />

Angus 1960, Harrold and Dreibelbis 1951, Va n<br />

Bavel and Meyers 1962), but they haven't<br />

been used for natural vegetation, such a s<br />

brush and trees, which have extensive root<br />

systems .<br />

The use of weighing lysimeters to deter -<br />

mine evapotranspiration of trees is feasibl e<br />

only if the trees are uniformly spaced such as<br />

in a plantation . Installation is easier when the<br />

root system is naturally restricted, thereby<br />

reducing the size of the lysimeter and it s<br />

deadweight. Under these conditions, it would<br />

be possible to obtain accurate short period<br />

evapotranspiration rates, thus enabling a<br />

mechanistic examination of evapotranspiration<br />

in relation to the determining meteorological<br />

and soil moisture conditions .<br />

Establishment of evapotranspiration from a<br />

single tree or group of trees in a specific location<br />

is important but does not yield answer s<br />

for other types of vegetation in different<br />

climatic zones . Installation of lysimeters in<br />

many locations is not possible and may not be<br />

feasible because of the cost . However, meteorological<br />

methods if tested can be employed to<br />

establish evapotranspiration rates wher e<br />

lysimeters are not feasible. Results fro m<br />

meteorological methods can be tested against<br />

the results from a properly installed weighing<br />

lysimeter. This has been accomplished for<br />

agriculture crops (Tanner 1967, Pruitt 1963 ,<br />

Fritschen and Van Bavel 1963, Fritsche n<br />

1965) . However, differences in scale factor s<br />

such as canopy height, density, and roughnes s<br />

demand additional testing before these<br />

methods can be used over natural vegetation .<br />

The purpose of this paper is to discuss th e<br />

establishment of the weighing lysimeter in a<br />

Douglas-fir forest for the primary purpose of<br />

studying the complex relation between evapotranspiration<br />

rates and the determining<br />

meteorological and soil moisture conditions .<br />

In later stages the installation will be used t o<br />

evaluate meteorological methods for determining<br />

evapotranspiration to be used i n<br />

other areas .<br />

The Site<br />

The lysimeter installation is located on the<br />

Cedar River Watershed near Seattle, Washing -<br />

ton. The soil is a Barneston, gravelly, loamy<br />

sand originating from glacial outwash laid<br />

down at the end of the Vashon glacial period<br />

(Poulson and Miller 1952) and generally restricts<br />

the root system above the 3-foot depth<br />

(Gessel and Cole 1965) . The lateral extent of<br />

the root system is largely restricted to th e<br />

basic spacing of the trees . The area is relatively<br />

level and has a uniform canopy density<br />

making it desirable for micrometeorological<br />

investigations .<br />

The trees are 35-year-old Douglas-fir (Pseudotsuga<br />

menziesii) which regenerated naturally<br />

after logging. The average tree spacing i s<br />

5.8 m, resulting in 231 trees per 4,041 m 2<br />

consisting mostly of Douglas-fir with a fe w<br />

hemlock (Tsuga heterophylla) and maple .<br />

Trees in the 5 to 30 cm d.b.h. classes occur<br />

with the greatest frequency . Ground vegetation<br />

consists of bracken fern (Pteridiu m<br />

aquilinum), salal (Gaultheria shallon), red<br />

huckleberry (Vaccinium parvifolium), an d<br />

mosses, mainly (Eurhynchium oregonum) .<br />

The Lysimeter<br />

The lysimeter was constructed around th e<br />

root ball of a 35-year-old, 28 m tall an d<br />

38 cm diameter Douglas-fir tree . The lysimeter<br />

consists of two right cylinder containers,<br />

one located within the other . The inner -<br />

most container in which the tree is located i s<br />

366 cm in diameter and 122 cm deep, havin g<br />

a surface area of 10 .5 m 2 (fig. 1). The container<br />

at "field capacity " with tree and soi l<br />

weighs 28,900 kg (fig . 2) .<br />

The inner container is resting on a hydraulic<br />

transducer located on the bottom of th e<br />

outer container. The hydraulic transduce r<br />

consists of eleven 15 m lengths of 6 .35 c m<br />

256


Figure 1 . Walls of the inner container being fitted together<br />

prior to lowering down around the root ball .<br />

Figure 3. Butyl rubber tubing used for weighing coiled<br />

on the bottom of the outer container .<br />

Figure 2 . Completed inner container with soil and tre e<br />

shown raised 90 cm for installation of the hydrauli c<br />

transducer .<br />

butyl rubber tubing with a valve stem vulcanized<br />

30 cm from one end . The tubing was<br />

filled with air-free water and was coiled o n<br />

the bottom of the outer container starting at<br />

the center (fig . 3). Holes were cut in the bottom<br />

of the outer container to allow the valv e<br />

stems to pass through . Thick-walled tubing<br />

(9.5 mm I .D.) was attached to each valve stem<br />

and extended beneath the bottom of th e<br />

outer container to a shutoff manifold . The<br />

manifold is connected to a vertical standpipe ;<br />

thus, the weight of the tree, soil, and inne r<br />

container is reflected in the height of th e<br />

water column in the stand pipe . A water-fille d<br />

dummy standpipe is located adjacent to the<br />

active standpipe for the purpose of temperature<br />

compensation and counter balance .<br />

Weight changes are detected by a ±6 .9 mb differential<br />

pressure transducer located betwee n<br />

the active and the dummy standpipes . The<br />

signal is recorded on an automatic data logging<br />

system .<br />

To provide for drainage of excess water ,<br />

eight filter candles were installed around th e<br />

periphery of the inner container about 10 c m<br />

from the bottom. The filter candles were<br />

5 cm in diameter by 20 cm ceramic tube s<br />

with a bubbling pressure in excess of 25 0<br />

millibars. The filter candles are connecte d<br />

through a manifold system to an evacuate d<br />

reservoir. The outflow from the drainage<br />

system is measured with a modified typ e<br />

tipping bucket rain gauge .<br />

To prevent the tree from blowing or falling<br />

over during and after construction, it wa s<br />

guyed from a yoke located at 10 m to the<br />

base of adjacent trees. After construction four<br />

climbable (30 cm triangular) towers were<br />

built around the lysimeter and the tree guye d<br />

to the towers with horizontal cables. The<br />

cables were loose enough to allow 15 c m<br />

motion at 10 m .<br />

The sensitivity of the lysimeter is 630 g<br />

which is equivalent to 0 .06 mm of water o r<br />

22 ppm. The lysimeter installation at Tempe ,<br />

Arizona (Van Bavel and Meyers 1962), has a<br />

sensitivity of 20 g or 0 .02 mm of water o r<br />

257


7.4 ppm . The lysimeters at Coshocton, Ohi o<br />

(Harrold and Dreibelbis 1951), have a sensitivity<br />

of 2,268 g or 0 .25 mm of water or<br />

38 ppm and the lysimeter at Davis, Californi a<br />

(Pruitt and Angus 1960), has a sensitivity o f<br />

907 g or 0 .03 mm of water or 20 ppm .<br />

The Weather Station<br />

Located adjacent to the lysimeter tree an d<br />

on a tower 33.5 m in the air are the meteorological<br />

sensors of the weather station . Thes e<br />

consist of a solarimeter, a net radiometer, air<br />

temperature sensor, vapor pressure sensor ,<br />

wind direction and speed sensors, and a<br />

tipping bucket rain gauge . In addition to thes e<br />

parameters, soil temperature is measured a t<br />

four depths inside and outside of the lysimeter.<br />

The signal from these sensors is re -<br />

corded automatically on a digital magneti c<br />

tape data logging system. At the present time ,<br />

five of the signals are being integrated continuously<br />

: solar radiation, net radiation, rain -<br />

fall, windspeed, and the drainage from th e<br />

lysimeter. The integrals of these signals an d<br />

the other parameters are recorded on the<br />

magnetic tape at hourly intervals to conserv e<br />

the magnetic tape's supply . With hourly<br />

records, the tape supply will last for 30 days .<br />

The magnetic tape (6 .4 mm) is converted t o<br />

12.7 mm computer compatible tape and the n<br />

analyzed at the University of Washingto n<br />

Computer Center with the Burroughs 550 0<br />

computer .<br />

The Proposed Uses of the<br />

Lysimeter Facility<br />

Since the lysimeter was installed during th e<br />

summer of 1970 and the completed facility<br />

was not completely checked out unti l<br />

recently, very little data is available for discussion.<br />

However, it may be enlightening to<br />

discuss the proposed uses of the facility .<br />

An Evapotranspirimeter<br />

The lysimeter installation is expected to<br />

yield short period rates of evapotranspiratio n<br />

from the 28 m Douglas-fir tree . The accurac y<br />

of the measured evapotranspiration is withou t<br />

much question . How representative the dat a<br />

are is questionable unless the water potential<br />

of the root mass is kept equal to the wate r<br />

potential of the root mass of adjacent trees .<br />

Since the root mass is restricted in a 366 c m<br />

diameter container, the amount of soil avail -<br />

able for water and nutrient extraction by thi s<br />

tree is slightly less than adjacent trees . Durin g<br />

the summer months, water should be with -<br />

drawn from the lysimeter container at a faster<br />

rate than from the adjacent soil . Therefore ,<br />

water will have to be added to the lysimeter<br />

container to maintain a water supply equal t o<br />

that in the adjacent area. During the winter<br />

months, the opposite will be true . The bottom<br />

of the lysimeter container inhibits dee p<br />

percolation . When rainfall exceeds evapotranspiration,<br />

water will have to be removed fro m<br />

the container in order to prevent a buildup of<br />

water in the bottom of the container .<br />

Evapotranspiration as a Function of Potentia l<br />

In order to study the mechanism of transpiration<br />

from a Douglas-fir tree, the evapotranspiration<br />

rates will be determined in relation<br />

to the soil-water potential and the atmosph<br />

eric evaporative demand . During thes e<br />

studies, soil moisture potential will be deter -<br />

mined with a series of thermocouple psychrometers<br />

installed in the lysimeter container .<br />

The atmospheric evaporative demand will b e<br />

calculated from meteorological parameters .<br />

Tissue Volume Change s<br />

Dendrometer bands will be installed at various<br />

locations on the tree to study the swellin g<br />

and shrinking of the tree in relation to water<br />

potential and evapotranspiration . At the sam e<br />

time, stomata aperture and plant stress will b e<br />

determined using leaf resistance meters an d<br />

thermocouple psychrometers or Scholande r<br />

bombs.<br />

Test Cuvette Technique<br />

While the detailed studies of evapotran -<br />

258


spiration are being conducted, cuvettes will b e<br />

installed at various locations within the crow n<br />

to monitor the spacial variation of transpiration.<br />

In addition, the rates of transpiration ,<br />

determined with the cuvettes, will be compared<br />

with the lysimetric transpiration rate t o<br />

determine how representative the cuvett e<br />

technique is for determining transpiration o f<br />

the tree .<br />

Test Meteorological Methods<br />

Meteorological methods such as energy balance,<br />

aerodynamic, and eddy correlation techniques<br />

have been used to determine evapotranspiration<br />

from agriculture crops. The<br />

methods need further testing in forestry be -<br />

fore they can be utilized on a wide scale . Th e<br />

evapotranspiration rates from the lysimeter<br />

tree, if representative of other trees, can b e<br />

utilized for testing the meteorologica l<br />

methods .<br />

Profiles of meteorological parameters suc h<br />

as radiation, temperature, humidity, carbo n<br />

dioxide, and windspeed will be monitored .<br />

These parameters will be utilized to calculat e<br />

the transfer coefficients (eddy diffusivities) o f<br />

sensible heat, latent heat, momentum, an d<br />

CO 2 . In addition, transpiration and photo -<br />

synthesis by layers will be calculated fro m<br />

these parameters .<br />

Canopy Interception of Precipitatio n<br />

Interception of precipitation either as rain ,<br />

dew, or snow will be represented as a weigh t<br />

increase by the lysimeter . The amount of<br />

precipitation intercepted by the crown of th e<br />

tree can be studied by providing a waterproof<br />

cover for the soil surface to prevent th e<br />

throughfall from being recorded as a weigh t<br />

increase . The throughfall will. be measure d<br />

separately . If stemflow is measured, then th e<br />

rest of the weight increase can be attribute d<br />

to canopy interception . It is obvious that thi s<br />

method can be utilized to determine th e<br />

amount of rainfall and snowfall interceptio n<br />

in relation to intensity, duration, and previou s<br />

wettings .<br />

However, this technique may be more useful<br />

in determining the amount of moisture<br />

extracted from the atmosphere as dewfall-at<br />

present, an unknown quantity. During the<br />

summer months, energy budget calculations<br />

demonstrate that Douglas-fir trees, under the<br />

conditions of the site, would be stresse d<br />

within 2 to 3 weeks after the last rainfall if<br />

they transpired at a potential rate, unless transpiration<br />

is supplemented by evaporation of<br />

dewfall. Dewfall may be a very important<br />

parameter in the survival of these trees .<br />

Fertilization and Irrigation<br />

At the present time considerable interest i s<br />

being expressed in using forested areas for disposal<br />

of sewage sludge . The benefit of irrigation,<br />

sludge, or other fertilizers to tree growth<br />

could be tested with the lysimeter installatio n<br />

by comparing the photosynthesis and transpiration<br />

of the tree being studied with th e<br />

adjacent trees .<br />

Summary<br />

During the summer of 1970, a lysimeter<br />

container (366 cm in diameter and 122 c m<br />

deep) was built around the root mass of a<br />

28 m Douglas-fir tree at the Cedar Rive r<br />

Watershed near Seattle, Washington . This container<br />

was placed on top of the hydraulic<br />

transducer which is capable of weighing th e<br />

tree, soil, and container, a mass of 28,900 kg ,<br />

with a sensitivity of 630 g or 0 .06 mm of<br />

water . A weather station was located on a<br />

33 .5 m tower adjacent to the lysimeter tree .<br />

The proposed uses of this installation are :<br />

(1) as an evapotranspirimeter, (2) as a standard<br />

for testing cuvettes, (3) as a standard for<br />

testing meteorological methods, (4) to stud y<br />

evapotranspiration in relation to soil moisture<br />

potential and atmospheric evaporative demands,<br />

(5) to study the shrinking and swellin g<br />

of the plant tissue in relation to evapotranspiration,<br />

(6) to determine interception of<br />

rainfall, snow, and dewfall by the crown, an d<br />

(7) to study the effects of irrigation and fertilizer<br />

upon the growth of the tree .<br />

259


Acknowledgments<br />

The work reported in this paper was sup -<br />

ported by National Science Foundation Gran t<br />

No . GB-20963 to the Coniferous <strong>Forest</strong> Biome ,<br />

U .S. Analysis of Ecosystems, International<br />

Biological Program. This is Contribution No .<br />

41 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Colorado River Association. 1966 . Newsletter .<br />

Colo. River Assoc . Los Angeles, Calif. Sept. ,<br />

p. 4 .<br />

Fritschen, Leo J . 1965 . Accuracy of evapotranspiration<br />

determination by the Bowe n<br />

ratio method. Bull. Int. Assoc. Hydrol .<br />

10(2) : 38-48 .<br />

and C . H . M . Van Bavel. 1963 .<br />

<strong>Experimental</strong> evaluation of models of<br />

latent and sensible heat transport over irrigated<br />

surfaces. Int . Assoc. Sci. Hydrol . Pub .<br />

62 : 159-171 .<br />

Gessel, Stanley P ., and Dale W. Cole. 1965 .<br />

Influence of removal of forest cover on<br />

movement of water and associated elements<br />

through soil . J. Soc. Am. Water<br />

Works Assoc . 57 : 1301-1310 .<br />

Harrold, L . L ., and F . R. Dreibelbis . 1951 .<br />

Agricultural hydrology as evaluated by<br />

monolith lysimeters . U .S. Dep. Agr . Tech .<br />

Bull. 1050 : 135 .<br />

Mcllroy, I . C., and D. E. Angus . 1963. Th e<br />

A s p e n d ale multiple weighed lysimeter<br />

installation . Australian Commonwealth Sci .<br />

& Ind. Res. Organ., Div. Meteorol. Phys .<br />

Tech. Pap. 14, p. 30 .<br />

Poulson, E . N., and J. T. Miller. 1952. Soi l<br />

survey of King County . USDA Soil Ser.<br />

1939, No. 31, p. 106 .<br />

Pruitt, W. O. 1963. Application of several<br />

energy balance and aerodynamic evaporation<br />

equations under a wide range of stability.<br />

Final Rep., USAEPG No . DA-36 -<br />

0390SC-80334, p . 107-124 .<br />

and D. E. Angus . 1960. Large<br />

weighing lysimeter for measuring evapotranspiration.<br />

Trans. Am. Assoc. Agric .<br />

Eng. 3(2) : 13-18 .<br />

Slatyer, R. O., and I . C. Mcllroy . 1961 . Practical<br />

microclimatology . Australian Commonwealth<br />

Sci. & Ind. Res. Organ . Various<br />

pagination .<br />

Tanner, C . B. 1967. Measurement of evapotranspiration<br />

. R. M. Hagan, H. R. Haise ,<br />

and T. W. Edminster (eds .), In Irrigation of<br />

agricultural lands. Am. Soc . Agron .<br />

Monogr. 11 : 534-574 .<br />

Van Bavel, C . H. M., and L . E . Meyers . 1962 .<br />

An automatic weighing lysimeter . Agr . Eng .<br />

43 : 580-583, 586-588 .<br />

260


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Initial steps in decomposition<br />

of Douglasfir needles<br />

under forest conditions<br />

Abstract<br />

P. L . Minyar d<br />

Graduate Student i n<br />

<strong>Forest</strong> Pathology<br />

Douglas-fir (Pseudotsuga menziesii) needle decomposition was evaluated after 6 and 12 months exposure t o<br />

forest floor conditions by characterizing solubility changes of saccharides, waxes and oils, and cellulose. Th e<br />

results lead to the hypothesis that primary processes of decomposition are initiated on the outside of the needl e<br />

with the solubilization of waxes and, as the waxes are depleted, the cellular constituents are solubilized.<br />

and<br />

C . H . Drive r<br />

Professor of<br />

<strong>Forest</strong> Pathology<br />

College of <strong>Forest</strong> Resources<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Introduction<br />

Approximately 4 percent of the biomass of<br />

the Douglas-fir consists of needles (Dic e<br />

1970) containing a portion of the carbon to<br />

be cycled within the ecosystem . Some genera<br />

of the Moniliales have been reported that are<br />

capable of utilizing cellulose and breakin g<br />

down needle waxes to release carbo n<br />

(Macauley and Thrower 1966 ; Martin and<br />

Juniper 1970) . The decomposition of needl e<br />

litter and the release of carbon in a forest is<br />

important in the nutrient cycling processe s<br />

within a conifer ecosystem . In an effort to<br />

study carbon cycling pathways and the par t<br />

played by the processes of the decompositio n<br />

of Douglas-fir needles the following investigation<br />

was conducted .<br />

Primary decomposition is defined a s<br />

changes in solubility of the needle constituents.<br />

Therefore to study decompositio n<br />

techniques were used to detect changes i n<br />

solubility . Simplistically, needles consist of an<br />

outer layer of cutin and waxes and an inne r<br />

cellular layer made up of cellulose and protoplasmic<br />

contents. Needle constituents wer e<br />

divided into three classes : (1) compound s<br />

extractable with hexane (primarily cuticl e<br />

waxes) (Martin and Juniper 1970), (2) those<br />

extractable with hot water (saccharides)<br />

(Campbell 1952), and (3) cellular components<br />

soluble in 1 percent HaOH (hemicellulose )<br />

(Campbell 1952) .<br />

Materials and Methods<br />

Needles used in this study were collected i n<br />

the following manner . Screen traps wer e<br />

attached to a tower at random intervals i n<br />

height from just below the crown of the tree s<br />

to ground level . Needles were collected fro m<br />

the traps aseptically, placed in sterile containers,<br />

and transported to the laboratory for<br />

analysis and further study. In addition needle s<br />

from the traps were placed in nylon mes h<br />

bags on the forest floor and allowed to de -<br />

compose. After exposure for 6 and 12 months<br />

the needle samples were analyzed in th e<br />

following manner .<br />

Samples were oven dried for 24 hr . at<br />

105°C and stored in airtight containers unti l<br />

analyzed .<br />

Three 8-g replications of each treatmen t<br />

sample of needles were extracted with boiling<br />

n-hexane in a Soxhlet extractor for 6 hr . and<br />

oven dried for 24 hr. at 105°C to remove the<br />

hexane and weighed . The amount of material<br />

261


extracted was determined by the weight differences<br />

before and after extraction .<br />

Water soluble materials were determined in<br />

a similar manner using boiling water .<br />

As employed by Cowling (1961), a 1 per -<br />

cent NaOH solubility test was used to determine<br />

rate of cellulose decomposition .<br />

In an effort to isolate fungi involved in the<br />

decomposition processes needles were cultured<br />

in two ways : A portion of each sample<br />

was plated on 2 percent malt extract agar and<br />

onto an antibiotic medium consisting of 2<br />

percent malt extract agar plus 1 :30,000 Rose<br />

Bengal and 50 p .p.m. streptomycin. A second<br />

portion of each sample was washed in sterile<br />

water for 2 minutes and the needles plate d<br />

onto these same agars . The plates were incubated<br />

at 26°C for 7 to 14 days. Isolation s<br />

were made from the resulting fungi .<br />

Results<br />

Table 1.-Frequency of occurrence of<br />

genera of fungi isolated from<br />

needle sample s<br />

Months<br />

Genera<br />

0 Phomopsis sp .<br />

Mucor sp .<br />

Penicillium spp .<br />

Trichoderma spp .<br />

Botrytis sp .<br />

Graphium sp .<br />

6 Phomopsis sp . 2<br />

Mucor sp . 3<br />

Penicillium spp . 4<br />

Trichoderma spp . 4<br />

Aspergillus spp . 3<br />

12 Penicillium 4<br />

Trichoderma spp . 4<br />

Mucor sp . 4<br />

1 1 = Very infrequent<br />

2 = Infrequen t<br />

3 = Frequent<br />

4 = Very frequent<br />

Frequency o f<br />

isolation'<br />

Discussion<br />

Several genera of fungi have been reporte d<br />

to exhibit the capability to utilize waxe s<br />

(Martin and Juniper 1970) . From the data i n<br />

figure 1 it appears that fungal action on th e<br />

needle waxes causes these materials to be -<br />

come more soluble in hexane after a period o f<br />

time. There is a large weight loss in hexan e<br />

extractables after a six month period and a<br />

smaller weight loss after 12 months . I t<br />

appears from this trend that during the first<br />

6-month period, waxes are acted on initiall y<br />

during primary processes of decomposition .<br />

This was followed during the second 6-mont h<br />

period by the processes of increasing solubilization<br />

of cellulose and saccharides .<br />

Dice (1970) reported that annual production<br />

of organic matter for Douglas-fi r<br />

litter is 359 kg/ha . The density data in figure<br />

1 shows that at 12 months there is a 60 per -<br />

cent reduction of organic matter or approximately<br />

213 kg/ha .<br />

The fungi listed in table 1 are being studie d<br />

to determine what role they play in th e<br />

physiochemical processes illustrated in figure<br />

1 .<br />

1 . 6<br />

1 . 5<br />

1 . 4<br />

1 . 3<br />

1 . 2<br />

1 . 1<br />

1 . 0<br />

.9-<br />

.8 /<br />

.7 /<br />

. 6<br />

. 5<br />

.4<br />

.3<br />

. 2<br />

. 1<br />

0<br />

6 1 2<br />

M<strong>ON</strong>THS<br />

- DENSITY<br />

- HEXANE ELTRACTAN.ES<br />

- I3T WATER EXTRACTABLES<br />

1% MOM EXTRACTAHLES<br />

1 . 6<br />

1 . 5<br />

1 . h<br />

1 . 3<br />

1 . 2<br />

1 . 1<br />

1 . o<br />

Figure 1 . Density and weight loss of extractables of<br />

Douglas-fir needles after decomposition .<br />

. 9<br />

. 8<br />

. 7<br />

.6<br />

.5<br />

.4<br />

.3<br />

. 2<br />

.1<br />

0<br />

3<br />

262


These preliminary results suggest the following<br />

hypothesis . Initially the primary<br />

processes of decomposition occur on the surface<br />

of the needle with the decomposition o f<br />

waxes, and as the waxes are depleted, the cell -<br />

ular constituents are acted upon .<br />

Acknowledgments<br />

The work reported in this paper was sup -<br />

ported in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferou s<br />

<strong>Forest</strong> Biome, U .S. Analysis of Ecosystems ,<br />

International Biological Program . This is Contribution<br />

No . 42 from the Coniferous Fores t<br />

Biome .<br />

Literature Cited<br />

Campbell, W. G. 1952 . Wood chemistry . L. E .<br />

Wise and E . C . Jahn (eds), 2 : 1343 p . Ne w<br />

York: Reinhold Publ. Corp .<br />

Cowling, E . B. 1961 . Comparative biochemistry<br />

of the decay of sweetgum sap -<br />

wood by white rot and brown rot fungi .<br />

U .S. Dep. Agric . Tech . Bull. 1258 : 2 .<br />

Dice, Steven F . 1970. The biomass and nutrient<br />

flux in a second growth Douglas-fir ecosystem.<br />

165 p ., illus. Ph .D . thesis, on file at<br />

Univ . Wash ., Seattle .<br />

Macauley, B . J., and L. B. Thrower . 1966 .<br />

Succession of fungi in leaf litter o f<br />

Eucalyptus regans . Brit. Mycol . Soc. Trans .<br />

49 : 509-520 .<br />

Martin, J. T., and B . E . Juniper. 1970. Th e<br />

cuticles of plants . 347 p ., illus. New York :<br />

St. Martin's Press Inc .<br />

263


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Seasonal and diurnal patterns of<br />

water status in Acer circinatum<br />

James P. Lassoie<br />

Research Assistant<br />

and<br />

David R . M. Scott<br />

Professor of <strong>Forest</strong>ry<br />

College of <strong>Forest</strong> Resources<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

Diurnal patterns of water status were monitored in Acer circinatum on clear days during the summer of<br />

1971 . Sap velocity, branch water potential, and relative leaf resistance to water vapor diffusion were used t o<br />

characterize the water status of a typical tree clump . Within the clump studied, air temperature and relative<br />

humidity were measured and used to calculate atmospheric vapor pressure deficits . Daily potential evaporation<br />

and soil moisture were calculated from U.S. Weather Bureau precipitation and air temperature data using th e<br />

Thornthwaite method. Results suggest that the seasonal course of plant water status is dependent upon changes<br />

in soil moisture levels, possible changes in leaf morphology, and differences in diurnal patterns of water statu s<br />

resulting from differences in atmospheric moisture demand .<br />

Introduction<br />

Vine maple, Acer circinatum Pursh, is a<br />

major subordinate species within the Coniferous<br />

Biome ranging from the southern mainland<br />

coast of British Columbia to the coasta l<br />

section of northern California . It is seldo m<br />

found growing with a single stem ; rather, it<br />

grows in clumps, forming a bushy mass . It is<br />

very shade-tolerant and is usually found unde r<br />

a tree canopy (Lyons 1964, Anderson 1968) .<br />

Despite the prevalence of vine maple in this<br />

area, it seems that little is known of it s<br />

physiological processes .<br />

Numerous authors have illustrated the con -<br />

trolling influence of internal water regimes o n<br />

physiological processes such as photosynthesis,<br />

respiration, translocation, and growt h<br />

of forest trees and herbaceous plants (re -<br />

viewed by Crafts 1968, Gates 1968, and<br />

Zahner 1968) . Also, transpiration, resulting i n<br />

water loss from the soil, influences the total<br />

hydrologic cycle and water balance of forest<br />

stands (Rutter 1968) . Therefore, understanding<br />

the water relations of a major subordinat e<br />

species, such as vine maple, may be quit e<br />

important to the study of productivity an d<br />

water balance of forests within the Coniferou s<br />

Biome. This study investigated diurnal pat -<br />

terns of internal water status in vine maple<br />

during the summer of 1971 . Besides supplying<br />

general information on the water relations of<br />

vine maple, the study also supplied basic data<br />

which will be used in the development of a<br />

plant water status submodel for the Coniferous<br />

Biome . The data presented are also<br />

representative of some of the types of data<br />

presently being collected for conifers within<br />

the Biome (Walker et al . 1972) .<br />

265


Methods<br />

The study was conducted at the Allen E .<br />

Thompson Research Center located about 5 5<br />

km southeast of Seattle at an elevation o f<br />

about 120 m in the foothills of the Washing -<br />

ton Cascades . The Research Center is abou t<br />

24 ha in size and is located on the western<br />

portion of the Cedar River watershed, in a<br />

40-year-old second-growth Douglas-fir, Pseudotsuga<br />

menziesii (Mirb.) Franco, plantation .<br />

Dominant trees are about 24 m in height .<br />

Principal subordinate species, besides vin e<br />

maple, are salal, Gaultheria shallon Pursh ,<br />

Oregon grape, Berberis nervosa (Pursh) Nutt ,<br />

sword fern, Polystichum munitum (Kaulf.)<br />

Presl., bracken fern, Pteridium aquilinum (L .)<br />

Kuhn var . pubescens Underw., and variou s<br />

species of moss . The climate is typical of lower<br />

elevations in western Washington . The aver -<br />

age temperature in July is 16°C and in January,<br />

3°C ; the average annual precipitation i s<br />

about 145 cm, with the majority falling during<br />

the winter months . The soil is a gravelly ,<br />

sandy loam derived from Pleistocene glacial<br />

outwash. The vegetation, climate, geology ,<br />

and soils in this area have been more completely<br />

described by Cole and Gessel (1968) .<br />

In the vine maple clump investigated, ther e<br />

were eight live stems, averaging about 3 .7 cm<br />

in diameter at breast height. Since the stems<br />

were quite supple, a spreading crown form resulted<br />

with stem lean in all directions . This<br />

straggly, crooked form is common for vin e<br />

maple growing in the shade (Lyon 1964) . The<br />

tallest erect stem in the clump was about<br />

10 m high. The clump was a moderately<br />

heavy Douglas-fir canopy .<br />

The water status of the clump was characterized<br />

by measurements of sap velocity ,<br />

branch water potential, and leaf diffusio n<br />

resistance . Sap velocity of one major stem in<br />

the clump was estimated using a heat puls e<br />

velocity (HPV) meter and techniques described<br />

by Hinckley (1971) . Numerous investigators<br />

have shown that HPV is a valid<br />

indicator of relative changes in the transpiration<br />

rate of forest trees (Skau and Swanso n<br />

1963, Wendt et al . 1967, Hinckley 1971) .<br />

Readings were taken automatically every half<br />

hour and sap velocities calculated as described<br />

by Marshall (1958) . Branch water potentia l<br />

(>Vb) was estimated using the pressure chamber<br />

apparatus and techniques described b y<br />

Scholander et al. (1965) . Many investigators<br />

have successfully used this technique to estimate<br />

VI) of trees in the field (Waring an d<br />

Cleary 1967, Klepper 1968, Hinckley an d<br />

Scott 1971) . In this study, pressure chambe r<br />

readings of at least three twig samples were<br />

taken and averaged for each observation period<br />

during the day. Samples were obtaine d<br />

from shaded portions of the crowns betwee n<br />

the heights of about 0 .5 and 2.5 meters . Eac h<br />

sample had at least two leaves attached to th e<br />

twig. Leaf resistances (RL) to water vapor diffusion<br />

(transpiration) were measured with a<br />

leaf resistance meter and techniques describe d<br />

by Van Bavel et al . (1965) . Resistances wer e<br />

calculated using the meter sensitivity to temperature<br />

relationship described by Van Bave l<br />

et al. Therefore, though the readings prove d<br />

to be adequate in indicating relative diurna l<br />

differences, their absolute magnitudes may b e<br />

in error by a constant amount . Resistances fo r<br />

two or three intact vine maple leaves (obtained<br />

from similar locations as OD samples)<br />

were measured and their values average d<br />

periodically during the day .<br />

Early morning and evening readings wer e<br />

usually impossible because of the very high<br />

resistances involved causing serious errors in<br />

measurement (Van Bavel et al . 1965). Probably,<br />

these high resistances are primarily due<br />

to nearly complete stomatal closure durin g<br />

these periods . Readings of HPV, 1'b, and R L<br />

were taken between about 0600 and 160 0<br />

Pacific Standard Time (PST) .<br />

Within the clump studied, relative humidity<br />

and air temperature (Ta) were measured i n<br />

the shade with a sling psychrometer approximately<br />

every hour between about 0600 and<br />

1600 PST . These data were used to calculat e<br />

the vapor pressure deficit (VPD) of the atmosphere.<br />

Daily potential evaporation (<strong>PE</strong>) and<br />

daily soil moisture (SM) in the top 90 cm of<br />

soil were calculated for July, August, an d<br />

September from precipitation (P) and air temperature<br />

data by the Thornthwaite metho d<br />

(Thornthwaite and Mather 1957) as modifie d<br />

by Machno (1966) . The data used in these<br />

calculations were taken from the clima -<br />

266


I<br />

- 5<br />

-l o<br />

-1 5<br />

HPV<br />

1<br />

10 . 0<br />

I . 6<br />

E<br />

U<br />

7 . 5<br />

5 . 0<br />

1 . 2<br />

0 . 8<br />

2 . 5<br />

.<br />

0 . 4<br />

3 0<br />

Ta<br />

VPO<br />

I 1 L<br />

3 0<br />

2 0<br />

10<br />

1 0<br />

- I<br />

0800 1200 1600 0800 1200 1600 0800 1200 600<br />

Time of Day (PST )<br />

Figure 1 . Daylight patterns of branch water potential (lib), heat pulse velocity (HPV), leaf resistance (RL), ai r<br />

temperature (Ta), and atmospheric vapor pressure deficit (VPD) for July 21, August 22, and September 22 ,<br />

1971, at the Thompson Research Center, Washington .<br />

tological records of the U .S . Weather Bureau<br />

Station at Landsburg, Washington, located<br />

about 3 km west of the Research Center a t<br />

approximately the same elevation .<br />

Results and Discussion<br />

Water status and environment data examined<br />

in this paper represent typical data collected<br />

on clear days during the summer of<br />

1971 . The days presented as examples are<br />

July 21, August 24, and September 22 (fig .<br />

1) . An expected overall pattern for plant<br />

water status on these days is discernible in the<br />

results. During the early morning "predawn,"<br />

the air was cool and fairly humid ; RL was<br />

high suggesting that stomata were relatively<br />

closed ; 'b was high while HPV was quite low .<br />

As the day progressed, Ta increased, thus in -<br />

creasing VPD and intensifying the evaporative<br />

demand . As stomata began to open, probabl y<br />

in response to increasing light levels, R L<br />

values began to decrease . Since VPD had in -<br />

creased and stomata had begun to open, th e<br />

tree began to transpire causing an increase i n<br />

HPV readings. Continued water loss resulted<br />

in a decrease in t/b until about midday . In th e<br />

afternoon, RL increased because of stomatal<br />

closure probably due to decreasing lib . With<br />

this came a corresponding decrease in HP V<br />

and an increase in branch water potential .<br />

The summer was marked by an abnormall y<br />

wet June, with more than 10 cm of precipitation<br />

having fallen. A rainless period followe d<br />

during most of July and August ; only on 1<br />

day, August 21, did P surpass <strong>PE</strong> (fig . 2). Th e<br />

drought ended with a rainy period during th e<br />

first 2 weeks of September . This was followed<br />

by 2 weeks without rain before another stor m<br />

occurred near the end of September . The dry<br />

period during July and August produced a<br />

26 7


1 5<br />

FIGURE 2<br />

1 0<br />

18 0<br />

FIGURE 3<br />

u<br />

160<br />

ILO<br />

-,<br />

- 0<br />

= 120<br />

I00<br />

SM<br />

- 3 . 0<br />

July 6 July 26<br />

Aug 1 5<br />

Time of year (1971 )<br />

Sept L Sept 2 L<br />

Figure 2 . Precipitation minus potential evaporation (P-<strong>PE</strong>) calculated by the Thornthwaite method for July ,<br />

August, and September 1971 at the Thompson Research Center, Washington .<br />

Figure 3 . Soil moisture (SM) in the top 90 cm of soil calculated by the Thornthwaite method and "predawn "<br />

branch water potential (*1y ) ) for July, August, and September 1971 at the Thompson Research Center ,<br />

Washington .<br />

steady drying of the soil from field capacit y<br />

to a water potential well below -1 .0 bar (fig .<br />

3). The rain in early September recharged th e<br />

soil to a point near field capacity . For the top<br />

90 cm of this soil, field capacity is near 17 . 7<br />

cm of water while the -1 .0 bar point is abou t<br />

14.2 cm (Knutsen 1965) .<br />

Since plants obtain water primarily fro m<br />

the soil and lose water to the atmosphere ,<br />

their internal water status will depend greatly<br />

upon the SM supply and the atmospheric demand.<br />

The relationship of plant water status<br />

to SM is illustrated by examining "predawn "<br />

branch water potential (*fib) periodically<br />

throughout the study period . These measurements<br />

were taken just prior to sunrise whe n<br />

'b is usually the highest of the day and<br />

should most accurately reflect soil water po -<br />

tential (Slayter 1967, Klepper 1968, Waring<br />

1969). Therefore, a continuous decrease in<br />

*lPb should have occurred during the rainles s<br />

period in July and August because of the<br />

steadily decreasing SM (i .e ., decreasing soil and<br />

water potential) during this period . Quite the<br />

opposite was observed (fig . 3) . During the<br />

period, "predawn" branch water potential decreased<br />

steadily to a minimum on August 3<br />

and then continually increased through<br />

August . This was surprising as *fib seemed to<br />

respond quickly to any change in SM, as illustrated<br />

by the immediate decrease in I1b during<br />

the temporary rainless period in September.<br />

Also, maximal and minimal points withi n<br />

this general trend proved to be significantl y<br />

different at the 1-percent level (table 1) . The<br />

apparent anomaly is probably explainable i f<br />

268


Table 1 .- Tukey's w Multiple Comparison Test for seasonal "predawn" branc h<br />

water potentials (*fib) showing significant differences at th e<br />

1-percent level. Values underlined are not significantly different.<br />

Date 7/15 9/15 8/24 8/31 8/17 8/22 7/21 7/27 8/ 3<br />

*>!ib 0.8 0.8 1.2 1.0 1 .4 1 .4 1 .4 1 .7 2 . 2<br />

certain plant and environmental changes<br />

which occurred through the summer an d<br />

which decreased daily transpirational wate r<br />

losses are taken into account .<br />

The differences in daily plant water status<br />

which occurred on clear days during the stud y<br />

period would help explain the *0b trend observed<br />

(fig. 1). These differences seemed primarily<br />

due to a general decrease in the atmospheric<br />

demand through the study period (fig .<br />

2) . Generally, higher average Ta's and VPD' s<br />

and higher <strong>PE</strong>'s occurred earlier in the summer,<br />

despite higher SM levels . A higher evaporative<br />

demand produced higher average HPV' s<br />

and greater water loss, thus developing lowe r<br />

branch water potentials . The later part of<br />

August was marked by relatively lower Ta's ,<br />

VPD's and <strong>PE</strong>'s, even though SM had de -<br />

creased. Hence, lower transpirational losse s<br />

occurred and higher 4/b developed . Durin g<br />

this period, a general increase in diffusion resistance<br />

was also evident .<br />

It is also possible that changes in leaf morphology<br />

during the summer could have helpe d<br />

produce a trend of increasing *fib during a<br />

period of continually decreasing SM . Chlorosi s<br />

in leaves, leading to necrosis before abscission ,<br />

has been shown to be induced by drought ,<br />

vascular disease, and normal senescence in<br />

woody and herbaceous plants (Kramer and<br />

Kozlowski 1960, Talboys 1968) . Such manifestations,<br />

characteristic of "drought hardening"<br />

in plants, can produce decreased transpiration<br />

and photosynthesis rates due to persistent<br />

stomatal closure even while leaves are fully<br />

turgid (Talboys 1968, Turner 1969) . By the<br />

middle of August, the leaves of the vine mapl e<br />

clump investigated had developed numerou s<br />

chlorotic spots; necrosis became evident during<br />

September. Therefore, increasing *1Pb during<br />

August could have been influenced by th e<br />

systematic reduction in transpiration in portions<br />

of the leaves .<br />

Conclusion<br />

The diurnal patterns of water status in vin e<br />

maple were examined for clear days during<br />

the summer of 1971 . As expected, the seasonal<br />

pattern of plant water status was de -<br />

pendent upon a combination of certain soil ,<br />

plant, and atmospheric factors which affecte d<br />

water uptake and loss by the plants . Soi l<br />

water potential, which steadily decreased during<br />

the rainless periods, should have progressively<br />

limited water availability and resulte d<br />

in the development of lower "predawn" plan t<br />

water potentials. However, during a continuous<br />

drying cycle, changes in certain plant an d<br />

atmospheric factors seemed to influenc e<br />

water loss greatly . Early in the summer ,<br />

greater water losses occurred due to highe r<br />

evaporative demands, and the plants were<br />

unable to recharge fully during the night eve n<br />

though adequate soil moisture was probably<br />

available . The development of chlorotic spots<br />

suggested that during a prolonged drought, o r<br />

as the leaves approached abscission, the ability<br />

of the entire clump to transpire decreased .<br />

Also, as the atmospheric demand lessened<br />

later in the summer, less water loss occurre d<br />

269


during the day. Therefore, despite low soil<br />

water supply, greater nighttime recharging<br />

was possible because the deficits developed<br />

were less .<br />

Acknowledgments<br />

The work reported in this paper was sup -<br />

ported by National Science Foundation Gran t<br />

No . GB-20963 to the Coniferous <strong>Forest</strong><br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution<br />

No . 43 to the Coniferous <strong>Forest</strong> Biome ,<br />

IBP .<br />

The authors are also indebted to Mr . Peter<br />

K. McCracken and Mr . Peter S . Machno for<br />

their assistance in various phases of this study .<br />

Literature Cited<br />

Anderson, H . C. 1968. Growth from and distribution<br />

of vine maple (Acer circinatum)<br />

on Marys Peak, western Oregon . Ecology<br />

50(1): 127-130 .<br />

Cole, D . W., and S . P. Gessel. 1968. A pro -<br />

gram for studying the pathways, rates an d<br />

processes of elemental cycling in a fores t<br />

ecosystem . Coll. of <strong>Forest</strong> Resour . Contrib .<br />

No. 4, 54 p ., Univ. Wash., Seattle .<br />

Crafts, A . S . 1968. Water deficits and physiological<br />

processes . In T. T . Kozlowski (ed .) ,<br />

Water deficits and plant growth . Vol. 2 .<br />

Plant water consumption and response, p .<br />

85-133 . New York : Academic Press .<br />

Gates, C . T. 1968. Water deficits and growth<br />

of herbaceous plants . In T . T . Kozlowsk i<br />

(ed.), Water deficits and plant growth . Vol .<br />

2 . Plant water consumption and response ,<br />

p. 135-190 . New York : Academic Press .<br />

Hinckley, T. M . 1971 . Estimate of water flo w<br />

in Douglas-fir seedlings . Ecology 52(3) :<br />

525-528 .<br />

and D . R. M. Scott . 1971. Estimates<br />

of water loss and its relation to environmental<br />

parameters in Douglas-fi r<br />

saplings . Ecology 52(3) : 520-524 .<br />

Klepper, B . 1968 . Diurnal patterns of water<br />

potential in woody plants . Plant Physiol .<br />

43: 1931-1934 .<br />

Knutsen, S . K. 1965. Hydrologic processes in<br />

30- to 35-year-old stands of Douglas-fir an d<br />

alder in western Washington . 167 p .,illus .<br />

M .S. thesis, on file at Univ . Wash., Seattle .<br />

Kramer, P. J., and T. T. Kozlowski . 1960 .<br />

Physiology of trees. 642 p., illus. New<br />

York: McGraw-Hill .<br />

Lyons, C . P . 1964. Trees, shrubs, and flower s<br />

to know in Washington . 211 p., illus .<br />

Toronto, Can . : J. M. Dent and Sons .<br />

Machno, P . S. 1966. Water balance and soi l<br />

moisture studies in western Massachusetts .<br />

136 p., illus. M .S. thesis, on file at Univ .<br />

Mass ., Amherst .<br />

Marshall, D . C . 1958. Measurement of sap<br />

flow in conifers by heat transport . Plan t<br />

Physiol . 33 : 385-396 .<br />

Rutter, A. J . 1968. Water consumption b y<br />

forests. In T . T . Kozlowski (ed.), Water<br />

deficits and plant growth . Vol. 2. Plant<br />

water consumption and response, p . 23-84 .<br />

New York : Academic Press .<br />

Scholander, P . F., E . D . Hammel, E . D. Bradstreet,<br />

and E . A. Hemmingsen. 1965. Sa p<br />

pressure in vascular plants . Science 143 :<br />

339-346 .<br />

Skau, C . M., and R . H . Swanson . 1963 . An<br />

improved heat pulse velocity meter as a n<br />

indicator of sap speed and transpiration . J .<br />

Geophys . Res. 68 : 4743-4749 .<br />

Slayter, R . O. 1967 . Plant-water relationships .<br />

366 p ., illus. New York : Academic Press .<br />

Talboys, P . W . 1968 . Water deficits and vascular<br />

disease . In T . T . Kozlowski (ed .), Water<br />

deficits and plant growth . Vol. 2 . Plant<br />

water consumption and response, p .<br />

255-311. New York : Academic Press .<br />

Thornthwaite, C . W ., and J . R. Mather . 1957 .<br />

Instructions and tables for computing potential<br />

evapotranspiration and the wate r<br />

balance . Climatol. Drexel Inst . Technol .<br />

Publ. 10(3) : 1-311 .<br />

Turner, N. C. 1969. Stomatal resistance to<br />

transpiration in three contrasting canopies .<br />

Crop Sci. 9(3) : 303-307 .<br />

Van Bavel, C . H. M ., F. S . Nakayama, and W .<br />

L. Ehrler. 1965 . Measuring transpiration resistance<br />

of leaves. Plant Physiol . 40 :<br />

535-540 .<br />

Walker, R. B., D. R. M . Scott, D . J. Salo and<br />

K . L. Reed. 1972 . Terrestrial process<br />

270


studies in conifers-a review . In Jerry F .<br />

Franklin, L . J . Dempster, and Richard H .<br />

Waring (eds.), Proceedings-Research o n<br />

coniferous forest ecosystems-A symposium,<br />

p. 211-225, illus . Pac. Northwes t<br />

<strong>Forest</strong> & Range Exp . Stn ., Portland, Oreg.<br />

Waring, R. H . 1969. <strong>Forest</strong> plants of the east -<br />

ern Siskiyous: their environmental an d<br />

vegetative distribution . Northwest Sci .<br />

43(1) : 1-17 .<br />

and B. D. Cleary. 1967. Plan t<br />

moisture stress : evaluation by pressure<br />

bomb . Science 155 : 1248-1254 .<br />

Wendt, C. W., J . R. Runkles, and R . H. Haas .<br />

1967. The measurement of water loss by<br />

mesquite (Prosopis glandulosa var . glandulosa<br />

Torr.) using the thermoelectric method<br />

. Soil Sci. Soc. Am. Proc. 31(2) :<br />

161-164 .<br />

Zahner, R . 1968 . Water deficits and growth o f<br />

trees . In T . T. Kozlowski (ed.), Water deficits<br />

and plant growth . Vol. 2. Plant wate r<br />

consumption and response, p . 191-254 .<br />

New York : Academic Press .<br />

271


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Development and testing of a n<br />

inexpensive thermoelectrically<br />

cooled cuvette<br />

David J . Salo, Department of Botan y<br />

John A . Ringo, Department of Electrical Engineerin g<br />

James H . Nishitani, Department of Botany<br />

an d<br />

Richard B . Walker, Department of Botany<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

An economical temperature-controlled cuvette system has been developed to monitor CO 2 assimilation an d<br />

dark respiration in the crowns of mature Douglas-fir (Pseudotsuga menziesii) trees at the Cedar River Thompso n<br />

site. Temperatures in the Plexiglas assimilation chambers are controlled with thermoelectric coolers, and CO 2<br />

concentration changes are measured with a differential infrared gas analyzer. Preliminary field data sugges t<br />

cuvette temperatures can be maintained within ±1 °C of ambient even under high insolation conditions .<br />

Introduction<br />

The "cuvette technique" has been used fo r<br />

field assessment of carbon dioxide assimilation<br />

since the 1930's (Bosian 1933), but the<br />

findings of these studies have often bee n<br />

criticized. These criticisms of cuvette methods<br />

have usually resulted from the difficulties of<br />

maintaining a plant-chamber environmen t<br />

closely approximating that outside the en -<br />

closure .<br />

Cuvette Problems<br />

Problems encountered have included th e<br />

development of "deep" boundary layers alon g<br />

leaf surfaces with the concomitant establishment<br />

of abnormal C0 2 , vapor pressure, an d<br />

temperature gradients . Additionally, CO 2<br />

concentrations deviating markedly fro m<br />

ambient levels have frequently occurred an d<br />

water condensation in chambers and in ai r<br />

conducting lines has been of particular con -<br />

cern at night, during cold weather periods ,<br />

and during periods of intense thermoelectri c<br />

cooling of cuvette bases . But probably th e<br />

most difficult problem to overcome has bee n<br />

that of chamber heating during periods o f<br />

high insolation . In efforts to achieve temperature<br />

control under these conditions, a variet y<br />

of chamber designs, fabrication procedure s<br />

and materials, and cooling techniques have<br />

been utilized .<br />

A common practice employed by many investigators<br />

has been the use of various thin<br />

plastic film "skins" to cover cuvette frames .<br />

Bourdeau and Woodwell (1965) used 8 mi l<br />

polyvinyl chloride (PVC), Ritchie (1969) use d<br />

2 mil PVC, and Hodges (1965) used 5 mil<br />

polypropylene . These efforts have met with<br />

varying degrees of success because plastics<br />

differ markedly in CO 2 permeability, as well as<br />

infrared and visible light transmittance . As an<br />

example, though the IR transmissivity o f<br />

polyethylene is relatively good its visible light<br />

transmittance is less than that of either Plexiglas<br />

(methyl methacrylate) or PVC . It is also<br />

273


permeable to both CO 2 and water vapor. On<br />

the other hand, visible transmittance of Plexiglas<br />

and PVC is very good while PVC infrared<br />

transmissivity is better than Plexiglas and glass ,<br />

though not as good as that of polyethylene .<br />

Other methods of reducing temperature<br />

buildups have included the rapid conductio n<br />

(700 - 1,500 1 hr ') of air through assimilation<br />

chambers (Lerch 1965) and the circulation<br />

of water, water-ethanol, or waterethylene<br />

glycol solutions through transparen t<br />

jackets surrounding them .<br />

Lange's (1962) and Ritchie's (1969 )<br />

" Klapp-Kuvettes" were attempts to reduc e<br />

temperature increases by alternately openin g<br />

and closing the chambers during experimenta l<br />

periods. Ritchie observed air temperatur e<br />

buildups of 5° to 7° C and leaf temperatur e<br />

increases of up to 17° C when his cuvette s<br />

were closed. By alternately opening and closing<br />

them, these overtemperatures could b e<br />

reduced by more than 50 percent .<br />

Currently, the most convenient method o f<br />

achieving cuvette temperature control appear s<br />

to be through the incorporation of thermoelectric<br />

coolers into chamber designs . Usin g<br />

the Peltier principle to regulate temperature ,<br />

Siemens Corporation (Erlangen, Germany )<br />

has manufactured a highly reliable and sensitive<br />

temperature and humidity controlle d<br />

system whose utility and flexibility have bee n<br />

demonstrated by Schulze (1970) . However ,<br />

cost per unit prohibits Biome acquisition of a<br />

sufficiently large number for simultaneou s<br />

sampling of CO 2 exchange in different crow n<br />

locations, on different age classes of foliage ,<br />

or on different plants . Therefore, routin e<br />

monitoring must be accomplished with a les s<br />

expensive system, reserving the Siemen s<br />

equipment available for "factors studies ."<br />

Prototype Development<br />

To complete a thorough daily and seasonal<br />

sampling program which will provide sufficient<br />

data to describe assimilation, dar k<br />

respiration, and transpiration in representativ e<br />

individual Douglas-fir (Pseudotsuga menziesii)<br />

at the Cedar River Thompson site, a syste m<br />

composed of several reasonably priced cooled<br />

cuvettes is considered highly desirable . As the<br />

first step in producing a functional system, a<br />

single prototype "two cuvette-power supply -<br />

temperature controller-fan controller "<br />

package (fig . 1) was fabricated from materials<br />

costing approximately $675 .<br />

Figure 1 . Prototype cuvette, power supply, temperature<br />

controller, fan controller package .<br />

Two 5-liter chambers constructed of<br />

3/8-inch Plexiglas are fitted with 1/8-inc h<br />

aluminum bases, removable tops, and smal l<br />

0-6 volt d .c . fans to mix the air and preven t<br />

boundary layer buildups . The aluminum floor<br />

of each chamber is securely bolted to the col d<br />

plate of an 80 watt Cambion (Cambridge ,<br />

Mass .) Model 7250-1 "Forced Convectio n<br />

Thermoelectric Assembly" (TEA) . The Mode l<br />

7250-1 coolers operate in a range between 0<br />

and 18 volts and 0 to 6 amps (larger TEA' s<br />

are also available) and are powered by d .c .<br />

supplies each of which provides power fo r<br />

two cuvettes, two fans, two temperature<br />

controllers, and two thermoelectric coolers .<br />

Each control circuit has been develope d<br />

around a temperature sensing silicon resisto r<br />

(sensistor) bridge (originally diodes fashione d<br />

from transistors were used as temperature<br />

sensors), the imbalance in which determine s<br />

the current output to the thermoelectri c<br />

device .<br />

274


<strong>Experimental</strong> System<br />

Summer field trials for the cuvette wer e<br />

conducted at the Allen Thompson Researc h<br />

Center during 1971 . The open system (fig . 2)<br />

utilized included an air intake approximately<br />

25 m above the ground on a tower located<br />

3 m from that used as an access to study<br />

trees. Air was drawn to the ground, through a<br />

20-liter "mixing reservoir," to assure uniformity<br />

and prevent "noisy" carbon dioxid e<br />

analyzer output, and then passed through a<br />

manifold for distribution to two Model 506 R<br />

Reciprotor (Copenhagen) pumps . One of<br />

these pumped a comparison airstream to th e<br />

URAS II (Hartmann and Braun, Frankfurt )<br />

infrared gas analyzer housed 46 m away in th e<br />

permanent site building and the other move d<br />

air to the cuvette located at 17 .2 m, back to<br />

the ground, through a flow meter, and then t o<br />

the differential analyzer. The attenuating<br />

reservoir, manifold, pumps, and flow mete r<br />

were located in a shelter at the tower base a s<br />

was the power supply-temperature controllerfan<br />

controller. A Honeywell recorder and the<br />

URAS II, however, were located in the site<br />

building. Two 0 .004-inch copper-constanta n<br />

thermocouples were used for temperature<br />

sensing inside the cuvette and one 22-gage<br />

thermocouple was positioned in a shaded<br />

location outside .<br />

In addition to cuvette data, meteorologica l<br />

data were provided by instruments located o n<br />

the adjacent meteorological tower . Scholander<br />

bomb and pressure infiltrometer (Fry<br />

and Walker 1967) samples were routinel y<br />

taken to provide estimates of water potentia l<br />

and stomatal aperture. These data were correlated<br />

with cuvette and meteorological dat a<br />

collected during each trial . Sampling wa s<br />

carried out at cuvette level and illustrativ e<br />

values reported here represent averages-tw o<br />

to three twigs for Scholander pressures (Ps )<br />

and five to seven needles for stomatal infiltration<br />

pressure (Pstom) .<br />

During preliminary trials, the cuvette temperature<br />

was maintained within at least ±<br />

1 .5° C of ambient under conditions of high<br />

AI R<br />

INTAKE<br />

CUVETTE<br />

-1<br />

PUMPS<br />

MIXING --Lf<br />

RESERVOIR<br />

-r<br />

l~I<br />

_ 1<br />

FLO W<br />

METE R<br />

J 1<br />

1<br />

URASII<br />

FLO W<br />

METER<br />

MANIFOL D<br />

Figure 2 . <strong>Experimental</strong> system used during 1971 field trials .<br />

275


insolation (1 cal cm 2 min-1 ) and generally<br />

control was within ± 1° C . With the recent<br />

installation of aluminum heat sinks on th e<br />

cuvette floor, incorporation of "sensistor "<br />

(Texas Instruments) temperature sensors, an d<br />

additional electronic balancing of the system ,<br />

improvement of this performance is<br />

anticipated .<br />

Representative<br />

<strong>Experimental</strong> Results<br />

Summer data obtained with the prototyp e<br />

system agree with those of other investigators<br />

in suggesting a close relationship betwee n<br />

assimilation, stress, and stomatal behavior. To<br />

illustrate this, results of a single day's run<br />

have been included (fig. 3) .<br />

On August 3 and other clear days sampled ,<br />

light did not appear to limit photosynthesis ;<br />

maximum assimilation rates occurred between<br />

1030 and 1130 hours (PST) ; CO 2 uptake<br />

3<br />

2<br />

EXPT 7I-CR 2<br />

RUN 5 (BRANCH 2 )<br />

8/3/7<br />

37 1<br />

DF<br />

decreased between 1130 and 1230 but wa s<br />

accompanied by continued stress increase an d<br />

only small changes in stomatal aperture ; varying<br />

degrees of afternoon stomatal closure<br />

followed and coincided with further depression<br />

of assimilation and relief of stress .<br />

Acknowledgments<br />

The work reported in this paper was supported<br />

in part by National Science Foundatio n<br />

Grant No . GB-20963 to the Coniferous <strong>Forest</strong><br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution<br />

No . 44 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Bosian, G . 1933 . Transpirations-und Assimilationsbestimmungen<br />

an Pflanzen des Zentralkaiserstuhls<br />

. Z. Bot . 26 : 209-294 .<br />

°<br />

° CQ2 ASSIMILATI<strong>ON</strong><br />

(mg CQ2 dry g' hr -' ) °<br />

1 5<br />

10<br />

5<br />

Pstom<br />

`~_~<br />

Ps<br />

erg'°~<br />

[atmospheres)<br />

1 .5<br />

-°~-~ - 1.0<br />

0. 5<br />

0.5<br />

0.0<br />

3 0<br />

\ Re °2<br />

(cal cm miri 1 )<br />

10 _ °-0-0--0-0~<br />

CUVETTE TEM<strong>PE</strong>RATURE ('C )<br />

1 , 1 1 1<br />

06<br />

08 10 12 14 1 6<br />

TIME OF DAY<br />

Figure 3 . Representative data obtained during cuvette field trials .<br />

276


Bourdeau, P. F., and G. M. Woodwell. 1965.<br />

Measurements of plant carbon dioxide<br />

exchange by infrared absorption under con -<br />

trolled conditions in the field. In F. E .<br />

Eckardt (ed .), Methodology of plant<br />

ecophysiology, p . 283-289 . UNESCO .<br />

Fry, K. E., and R . B. Walker . 1967 . A pressure-infiltration<br />

method for estimatin g<br />

stomatal opening in conifers . Ecology 48 :<br />

155-157 .<br />

Hodges, J . D. 1965 . Photosynthesis in forest<br />

tree seedlings of the Pacific Northwest<br />

under natural environmental conditions .<br />

177 p. Ph.D . thesis on file, Univ . Wash . ,<br />

Seattle .<br />

Lange, O . L. 1962 . Eine "Klapp-Kiivette" zur<br />

C0 2 -Gaswechselregistrerung an Blattern<br />

von Freilandpflanzen mit dem URAS . Ber.<br />

Deut. Bot. Ges. 75 : 41-50 .<br />

Lerch, G. 1965 . On the problem of cuvetteclimate<br />

. In Bohdan Slavik (ed .), Water<br />

stress in plants, p. 291-293 . The Hague : Dr .<br />

W. Junk N .V. Publishers .<br />

Ritchie, G. A. 1969. Cuvette temperatures<br />

and transpiration rates . Ecology 50 :<br />

667-670 .<br />

Schulze, E . -13 . 1970 . Der CO2 -Gaswechsel der<br />

Buche (Fagus siluatica L.) in Abhangigkeit<br />

von den Klimafaktoren im Freiland . Flora<br />

159 : 177-232 .<br />

277


Aquatic Process Studies<br />

279


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium.<br />

Bellingham, Washington-March 23-24, 197 2<br />

Studying streams as a<br />

biological unit<br />

James R. Sedel l<br />

Department of Fisheries and Wildlife<br />

Oregon State University<br />

Abstract<br />

The case for viewing and studying streams from the standpoint of processors of materials and energy instea d<br />

of exporters from the forest is elaborated. The unique opportunity for stream and terrestrial biogeochemica l<br />

investigators to cooperate and be coinvestigators on the same stream systems is explored and specific programs<br />

are identified.<br />

The shift in emphasis from standing crop sampling and instantaneous ingestion and growth information t o<br />

process studies utilizing radioisotope material balance experiments and carbon flux experiments is explained .<br />

Short term experiments using radioisotopes as food markers are described and discussed as to their usefulness in<br />

determining the effect of food quality on ingestion rates and assimilation efficiencies .<br />

11<br />

The Stream as a<br />

Biological Unit<br />

The primary aquatic interface with th e<br />

terrestrial component occurs in small water -<br />

shed streams . There is a general consensu s<br />

that the stream itself probably contribute s<br />

little to the conservation of nutrients (compared<br />

to the total terrestrial system of whic h<br />

they are a part) since relative biomass level s<br />

are low and export is a pervasive feature o f<br />

streams, particularly during freshets . Suc h<br />

ideas, although partially correct, have helpe d<br />

to perpetuate out-moded views advanced by<br />

terrestrial ecologists around the turn of th e<br />

century concerning streams . The opening<br />

paragraph of a paper by Shelford and Edd y<br />

(1929) 1 is sadly appropriate in 1972 :<br />

Modern ecologists . . . have considered<br />

that the development of communitie s<br />

gives the clue to their dynamics and relations<br />

to each other. Most plant ecologists<br />

have usually assumed that there are n o<br />

permanent fresh water communities .<br />

This assumption is based upon negative<br />

evidence . Streams change their locations ,<br />

and it is essentially their abandoned posi-<br />

' V. E. Shelford and S . Eddy. Methods for the<br />

study of stream communities . Ecology 10 : 382-392 ,<br />

1929 .<br />

tions that become ponds and develo p<br />

into land communities . Streams are<br />

permanent as long as the existing climat e<br />

endures, and this is the condition under<br />

which land communities reach a climax .<br />

The state of the art of stream ecology ha s<br />

clearly upheld Shelford and Eddy's hypotheses<br />

that permanent stream communitie s<br />

exist, undergo successional development ,<br />

reach and maintain a relatively stable condition,<br />

and manifest seasonal and annual differences,<br />

i .e., streams are bona fide biological<br />

units. To continue further, streams are highl y<br />

evolved biological units. As Hynes (1970) has<br />

pointed out, almost every taxonomic group of<br />

invertebrates is represented in, on or near th e<br />

substratum of lotic environments . In contrast<br />

to lakes and ponds there are several group s<br />

which occur only in lotic systems and more<br />

which reach their maximum development an d<br />

diversity there . This is quite probably a consequence<br />

of the permanence of streams as compared<br />

with lakes and ponds . Rivers rarel y<br />

disappear so they are not evolutionary trap s<br />

(Hynes 1970) .<br />

The attitudes of terrestrial ecologists<br />

toward streams have not basically change d<br />

since the paper of Shelford and Eddy (se e<br />

footnote 1). A recent symposium at Orego n<br />

State University on forest land uses and<br />

281


stream environments (Krygier and Hall 1971 )<br />

indicates that attitudes toward streams b y<br />

policymakers and foresters may be changing .<br />

However, with the notable exceptions of<br />

Chapman (1966) and Lantz and Hall (personal<br />

communication), the analysis of the<br />

effects of logging on streams seldom consider s<br />

the stream directly from a biological point o f<br />

view .<br />

In perhaps the most complete study on a<br />

watershed to date, the Hubbard Brook <strong>Experimental</strong><br />

<strong>Forest</strong> in New Hampshire (Liken s<br />

et al. 1970, Bormann et al . 1969), the capacity<br />

for streams to alter and process the various<br />

kinds and forms of chemicals was not considered.<br />

The various forms in which nitroge n<br />

enters a stream (organic, both dissolved an d<br />

particulate, and inorganic) and their fate s<br />

have not been carefully explored for eithe r<br />

undisturbed or clearcut watersheds . On e<br />

might expect that more nitrogen is lost to th e<br />

terrestrial portion but retained in the stream<br />

portion of the forest ecosystem than Liken s<br />

and Bormann have shown . Indeed, Fishe r<br />

(1970), working on the same watershed, has<br />

shown that 80 percent or more of the particulate<br />

fraction that enters a small stream is<br />

processed in the stream . Kaushik and Hyne s<br />

(1971) and Triska (1969), in studies on th e<br />

fate and residence times for leaves in streams ,<br />

also indicate that processing and not export is<br />

the dominant process .<br />

Objectives of the Research on<br />

<strong>Andrews</strong> Watershed Streams<br />

The inclusion of streams in the Coniferou s<br />

Biome Study represents a unique opportunity<br />

for stream biologists and terrestrial biogeochemical<br />

investigators to cooperate on an d<br />

investigate the same stream systems, thu s<br />

exploring together long neglected problems .<br />

We have taken advantage of this opportunit y<br />

on the <strong>Andrews</strong> forest streams by emphasizing<br />

major terrestrial-aquatic couplings in our<br />

stream investigations such as : (1) obtainin g<br />

estimates of allochthonous inputs and exports<br />

both particulate and dissolved ; (2) documentation<br />

of successional changes in microflora<br />

and the identification of specific metaboli c<br />

activities carried out during leaf decomposition<br />

; (3) determining the significance of roo t<br />

return of nutrients via riparian vegetatio n<br />

with the aid of radioactive isotopes and mas s<br />

balance experiments; and (4) estimating th e<br />

importance of fish and amphibians as the to p<br />

carnivores in some of the small watershe d<br />

streams .<br />

The long-range objective of the Andrew s<br />

stream program is to elucidate the role of a<br />

stream in the functioning of the watershe d<br />

ecosystem. The inclusion of fish, which ar e<br />

certainly quantitatively significant in some o f<br />

the streams, makes it apparent that the othe r<br />

components and their relationships will b e<br />

defined as well . The determination of how th e<br />

relationships between productivity and structure<br />

at the various trophic levels are altered<br />

by various degrees of land use-in this instance,<br />

logging-will be a prime objective . Because<br />

of the relatively short-term nature of<br />

the research, logged vs . unlogged comparison s<br />

will be based primarily upon simultaneou s<br />

observations over a range of conditions rathe r<br />

than on the conventional before-and-after approach.<br />

The <strong>Andrews</strong> forest provides a wid e<br />

range of conditions from which to sample ; including<br />

virgin watersheds, watersheds wher e<br />

streamsides have been recently logged, and<br />

watersheds where streamside vegetation has<br />

grown after past logging. There are obvious<br />

problems in this comparative approach . We<br />

will have to look at a number of "undisturbed"<br />

streams to establish a sort of natura l<br />

variance in order to distinguish betwee n<br />

natural vs. imposed variation .<br />

The <strong>Andrews</strong> Approach<br />

to Stream Studies<br />

At present the emphasis is focused on th e<br />

role of the biota in the small streams in th e<br />

<strong>Andrews</strong> forest . We are presently determinin g<br />

the structure of the stream, its energy sources ,<br />

and its energy processing components . This<br />

descriptive work is both necessary and tedious<br />

and will continue into year 3 . The method s<br />

and approaches being used are standard for<br />

282


this type of stream work with some modifications<br />

to cover the peculiarities of our specifi c<br />

streams .<br />

A significant problem not being dealt with<br />

in year 2 is the dimension of the dissolved<br />

organic fraction (DOM) and the utilization of<br />

that fraction. The DOM from leaf leachate ,<br />

algal excretion, and soil solution represents a<br />

significant organic pool which turns ove r<br />

quite rapidly, even though a large amount o f<br />

rather refractory matter may be expected . K .<br />

W. Cummins (personal communication), 2<br />

attempting to model woodland streams in a<br />

deciduous forest, estimated that about 50 per -<br />

cent of the total DOM input was reflocculated<br />

into fine particulate organic matter (FPOM) .<br />

This occurred by flocculation around air<br />

bubbles, colloidal settling, and chemical precipitation<br />

. The other half was metabolized by<br />

microbes of which he estimated 50 percen t<br />

passed off as respiratory CO 2 and 50 percent<br />

went into production of microbial FPOM .<br />

The DOM compartment is a vital coupling between<br />

the dynamics of the stream biota an d<br />

such biome investigations as stream hydrolog y<br />

and biogeochemistry .<br />

Little is known concerning the man y<br />

changes in quantity and quality of DOM that<br />

are produced by the biological processes in<br />

and near streams . The functional role of DO M<br />

in stream metabolism is also understoo d<br />

poorly and that information which is availabl e<br />

is essentially circumstantial. Inferences hav e<br />

been drawn from physiological studies o f<br />

bacterial and algal cultures under laboratory<br />

conditions. Technological difficulties hav e<br />

delayed the transition from laboratory t o<br />

field investigations .<br />

The approaches planned, to tackle th e<br />

dynamics of DOM in forest streams, involv e<br />

an investigation of the processes of microbia l<br />

decomposition of allochthonous material ,<br />

measuring the quantity and quality of DO M<br />

involved, and examining the fluxing betwee n<br />

physical and biotic compartments in an d<br />

along the streams .<br />

The variety of stream conditions in th e<br />

2 K. W. Cummins . Narrative for a stream energy<br />

budget model . Unpublished manuscript on file a t<br />

Kellogg Biological Station, Michigan State University ,<br />

Hickory Corners . 5 p ., 1970 .<br />

<strong>Andrews</strong> watersheds provide excellent situations<br />

in which to measure the types an d<br />

amounts of DOM and hopefully develop a<br />

budget. Integral to the measurements will be a<br />

study of the functional microbial groups involved<br />

in the decomposition of the large<br />

organic material such as leaves, woody pieces ,<br />

and fish carcasses. The expertise of an investigator<br />

with knowledge of the biochemistry of<br />

microbial decomposition will be required .<br />

The necessary techniques for studyin g<br />

stream processes and conceptualizing streams<br />

in general have not as yet been adequately<br />

developed for coniferous systems . However ,<br />

major components with the forest strea m<br />

systems can be identified (fig . 1). Contrary to<br />

the opinion of others in the biome, th e<br />

periphyton and decomposer units can be use -<br />

fully approached as a "black-box ." The complexity<br />

can be used advantageously to measure<br />

disappearance of substrate and release o f<br />

products, and some understanding of ho w<br />

ecosystems work and respond to perturbations<br />

can be gained. Radioactive isotope<br />

experiments will allow us to look at coupled<br />

events and processes in the stream . The<br />

parameters will be determined by the kinetics<br />

of the isotope . We can refine the experiments<br />

to look at what features of the system deter -<br />

mine these parameters and how . Carbon-1 4<br />

and phosphorus-32 will be used in mass<br />

balance techniques developed by Saunder s<br />

(1969) and Saunders and Storch (1971 )<br />

(closed chambers, short-term experiments<br />

with C 14 ) and in situ P 32 techniques developed<br />

by Nelson et al . (1969) and Elwood and<br />

Nelson (personal communication) . 3 Isotopes<br />

of selected elements will be incorporated into<br />

biogenic material such as salmon carcasses an d<br />

the pathways and turnover rates of these elements<br />

following decomposition will be investigated<br />

.<br />

In addition to the decomposition studies<br />

already mentioned, an estimate of the large<br />

woody pieces (greater than 1mm) above, in ,<br />

and near the stream bed will be made . This<br />

will be done in cooperation with terrestrial<br />

biomass investigations on the <strong>Andrews</strong> forest .<br />

3 J. W. Elwood and D. J . Nelson . Measurement o f<br />

periphyton production and grazing rates in a strea m<br />

using a 32 P material balance method . Oikos (in press) .<br />

283


S, - 4ydrolo9i c<br />

S2 = I¢rr¢strIa l<br />

S3 = -A9uattc<br />

- p hysical -Processe s<br />

Figure 1 . Stream system categories and relationships .<br />

284


The primary production and energy source s<br />

estimations will utilize the aforementione d<br />

P32 material balance method . The techniqu e<br />

consists of computing a material balance o f<br />

P3 2 following a 30-60 min pulse release. Total<br />

standing crop and effective stream botto m<br />

area can be calculated by equating the quantities<br />

of P 3 2 per unit area of substrate on th e<br />

stream bottom and per unit weight of periphyton<br />

on these substrates to the total quantities<br />

of P 3 2 retained in the study reach of the<br />

stream .<br />

The retained P3 2 in the stream will be subsequently<br />

lost in both dissolved and particulate<br />

forms through three processes : (1) as P 3 2<br />

is replaced by stable phosphorus throug h<br />

metabolic turnover ; (2) periphyton containing<br />

P 3 2 may be sloughed from substrates and<br />

transported out of the study reach ; and<br />

(3) particulate P 32 released to the stream<br />

from primary and secondary consumers and<br />

drift of consumers .<br />

Rates of change in the various compartments<br />

can be estimated by monitoring th e<br />

P 3 2 in the periphyton and stream water over<br />

time. The production rate of periphyton an d<br />

the grazing rate of periphyton can then be<br />

estimated .<br />

Nutrients could be lost from the stream t o<br />

the riparian portion of the stream study re -<br />

search. The P32 technique allows one to<br />

measure this movement of nutrients from the<br />

aquatic to the terrestrial compartment .<br />

The rationale for carbon flux experiments<br />

are discussed in another paper in this symposium<br />

by Lighthart and Tiegs (1972) . We<br />

will not discuss this approach now, except t o<br />

say that we think the technique can b e<br />

adapted for use in stream research .<br />

An additional study in the year 3 progra m<br />

will be the role of mosses in the stream bed in<br />

fixing energy and cycling nutrients . Their<br />

associated invertebrate fauna make them an<br />

important compartment in the streams<br />

trophic structure .<br />

Studies on the production of benthic in -<br />

vertebrates will be completed . In the analysis<br />

of the data particular attention will have to be<br />

paid to "strategies of survival" used by the<br />

various functional groups of invertebrates .<br />

Population regulation must be expected to<br />

differ quite widely in response to the degre e<br />

of stability of the environment . Freshets in<br />

the <strong>Andrews</strong> forest occur frequently but unpredictably.<br />

Food supplies may disappea r<br />

almost entirely, and animals must be able t o<br />

withstand and recover from very great fluctuations<br />

in numbers . They must be adapted to a<br />

great lack of constancy .<br />

Special attention will be paid to the chironomid<br />

larvae, pupa and adults . This group<br />

which represents the greatest numbers, specie s<br />

and probably biomass of all of the aquati c<br />

insect groups, has been somewhat neglected i n<br />

the first 2 years . Using a foam gathering method<br />

the associated exuvia, pupae, and emergin g<br />

adult midges will be identified and quantified .<br />

W. P. Coffman (University of Pittsburgh) wil l<br />

be consulted on the identity of the midges .<br />

Feeding experiments with the major invertebrate<br />

species will determine ingestion an d<br />

assimilation rates as affected by the qualit y<br />

and quantity of food .<br />

Meaningful data concerning feeding habit s<br />

of selected consumers and the relative nutritional<br />

values of different constituents of th e<br />

periphyton are extremely rare . Gut analysis of<br />

insects are often misleading as to major sourc e<br />

of nutrition due to differential digestion an d<br />

assimilation rates . Sede11 4<br />

has reported a technique<br />

which allows one to determine if th e<br />

type of microflora on the natural substrate s<br />

affect the rates of ingestion of stream invertebrates.<br />

Substrates from the stream are eithe r<br />

differentially sterilized or are surface sterilized<br />

and then reinoculated with an algal ,<br />

bacterial, or fungal community or any combination<br />

of the three treatments . The treate d<br />

substrate is then soaked for 1 to 3 hours in a<br />

stream water solution of Co b 0 . The co" is<br />

adsorbed onto the aufwuchs and the specifi c<br />

activity of the food substance determine d<br />

(mg/cpm) . The Co b ° is a gamma emitter an d<br />

can be easily detected without killing th e<br />

animal. This being the case sensitive experimental<br />

designs can be tried which use th e<br />

individual as a block, thereby eliminatin g<br />

4 J. R. Sedell. Feeding rates and food utilization of<br />

stream caddisfly larvae of the genus Neophylax<br />

Trichoptera : Limnephilidae) using °Co and 4 C. In<br />

~ . J. Nelson (ed.), Symposium on radioecology : Proceedings<br />

of the Third National Symposium at Oak<br />

Ridge, Tenn. May 8-10, 1971 . (In press .)<br />

285


ingestion variation between individuals an d<br />

ingestion variation between periods . By coupling<br />

this method with C 14 labeling of different<br />

compartments of the aufwuchs one can<br />

determine assimilation, and the major source<br />

of nutrition or at least that food which i s<br />

most easily assimilated .<br />

Predator-prey experiments have been de -<br />

signed to examine the roles of macroinvertebrates,<br />

amphibians, and fish in the stream ,<br />

and their influence on the community composition<br />

and numbers of aquatic insects . Th e<br />

higher consumer portion of the study will als o<br />

relate the contribution of allochthonous an d<br />

autochthonous production to the elaboration<br />

of fish flesh and to determine how the relationship<br />

may be altered by logging or other<br />

land management practices .<br />

The role of amphibians in the consume r<br />

dynamics of the stream areas in which the y<br />

are located is being investigated by a member<br />

of the terrestrial consumer group . This effort<br />

will strengthen the aquatic-terrestrial couplin g<br />

studies .<br />

The <strong>Andrews</strong> stream program for years 3<br />

and 4 will be focused on aquatic-terrestria l<br />

couplings . Apart from viewing the stream as<br />

receiving the bulk of its energy from forest<br />

litter, the extent to which the stream delay s<br />

the export of minerals and nutrients or may<br />

return them to the forest is not known .<br />

Stream research during years 3 and 4 will be<br />

geared to provide some of these answers .<br />

Acknowledgments<br />

The author and stream program are indebted<br />

to J. R . Donaldson who has been a<br />

constant source of personal encouragemen t<br />

and an active proponent of a vigorous strea m<br />

program. Our present stream program is an<br />

expansion of one very ably organized over the<br />

past 2 years by J . D. Hall. Our approach to<br />

studying streams as ecosystems reflects i n<br />

many ways the ideas of K . W . Cummins, wh o<br />

has followed the program and given generously<br />

of his ideas since its inception .<br />

The work reported in this paper was sup -<br />

ported by National Science Foundation Gran t<br />

No. GB-20963 to the Coniferous <strong>Forest</strong><br />

Biome, U .S. Analysis of Ecosystems, International<br />

Biological Program . This is Contribution<br />

No . 45 to the Coniferous <strong>Forest</strong> Biome .<br />

Literature Cited<br />

Bormann, F . H., G . E. Likens, and J. S .<br />

Eaton . 1969 . Biotic regulation of particulate<br />

and solution losses from a forest ecosystem<br />

. Bioscience 19 : 600-610 .<br />

Chapman, D . W. 1966. The relative contributions<br />

of aquatic and terrestrial primary producers<br />

to the trophic relations of strea m<br />

organisms . In K . W . Cummins, C . A. Tryon ,<br />

Jr., and R . T. Hartmen (eds .), Organismsubstrate<br />

relationships in streams, p .<br />

116-130 . Spec . Publ. No. 4, Pymatuning<br />

Lab. Ecol., Univ. of Pittsburgh, Pittsburgh .<br />

Fisher, S . G. 1970 . Annual energy budget of a<br />

small stream ecosystem : Bear Brook, West<br />

Thorton, New Hampshire. 97 p . Ph.D .<br />

thesis on file, Dartmouth Coll., Hanover .<br />

Hynes, H. B. N . 1970. The ecology of running<br />

waters . 555 p. Toronto, Can . : Univ .<br />

Toronto Press .<br />

Kaushik, N . K., and H . B . N . Hynes . 1971 .<br />

The fate of the dead leaves that fall int o<br />

streams. Arch. Hydrobiol . 68 : 465-515 .<br />

Krygier, J. T., and J. D . Hall. 1971 . <strong>Forest</strong><br />

land uses and stream environment, proceedings<br />

of a symposium at Oregon State University,<br />

Oct. 19-21, 1970 . 252 p . Continuing<br />

Education Publication, Corvallis, Oreg .<br />

Lighthart, Bruce, and Paul E . Tiegs. 1972 . Exploring<br />

the aquatic carbon web . In Proceedings-research<br />

on coniferous forest ecosystems-a<br />

symposium, p . 289-300, illus .<br />

Pac. Northwest <strong>Forest</strong> & Range Exp . Stn . ,<br />

Portland, Oreg.<br />

Likens, G . E., F. H. Bormann, N . M . Johnson ,<br />

and others. 1970. Effects of forest cuttin g<br />

and herbicide treatment on nutrient budgets<br />

in the Hubbard Brook watershed ecosystem<br />

. Ecol . Monogr . 40 : 23-47 .<br />

Nelson, D. J., N. R. Kevern, J . L . Wilhm, an d<br />

N. A. Griffith. 1969. Estimates of periphyton<br />

mass and stream bottom area usin g<br />

phosphorus-32 . Water Res . 3 : 367-373 .<br />

286


Saunders, G . W. 1969 . Some aspect of feeding<br />

in zooplankton, p . 556-573 . In Eutrophication<br />

: causes, consequences, correctives .<br />

Natl . Acad . Sci., Washington, D .C .<br />

and T. A. Storch. 1971 . Coupled<br />

oscillatory control mechanism in a plank -<br />

ton system. Nature New Biol . 230 : 58-60 .<br />

Triska, F. J. 1970. Seasonal distribution of<br />

aquatic hyphomycetes in relation to the disappearance<br />

of leaf litter from a woodlan d<br />

stream . 189 p. Ph .D. thesis on file, Univ.<br />

Pittsburgh, Pittsburgh .<br />

287


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Exploring the<br />

aquatic carbon web<br />

Bruce Lighthart and Paul E . Tiegs l<br />

Western Washington State Colleg e<br />

Bellingham, Washington 9822 5<br />

Abstract<br />

An aquatic carbon web containing the six compartments dissolved inorganic carbon (DIC), phytoplank ton,<br />

zooplank ton, dissolved organic carbon (DOC), detritus, and chemoorganotrophic bacteria is discussed . Tentativ e<br />

methods are presented for measuring the pool size and kinetics about each compartment at one depth in the epiand<br />

hypo-limnions during the four seasons in Lake Washington, Lake Sammamish, Lake Chester Morse, an d<br />

Lake Findley.<br />

Introduction<br />

It is a primary concern to man to be able t o<br />

prepare predictive mathematical models of<br />

organic forms in aquatic systems if he is going<br />

to understand and ultimately develop the<br />

tools to wisely manage his water resources . To<br />

prepare such models, it is desirable to evaluate<br />

the pool size (standing stock) and flux (kinetics)<br />

of energy (Lindeman 1942) in all the<br />

fractions or compartments of the system .<br />

Models may be based on measurements o f<br />

energy or of several materials, e .g., carbon ,<br />

phosphorous, and nitrogen ; however, both<br />

approaches are fraught with technical difficulties.<br />

Of the materials named, carbon i s<br />

recommended because it is a major component<br />

of all organic matter and is cycle d<br />

through a relatively well known web in th e<br />

biosphere . In addition, it is possible to follo w<br />

its movement with c" and other simpl e<br />

analytical techniques .<br />

In the aquatic realm, the carbon web ma y<br />

be thought to form a cyclical system made u p<br />

of a series of interrelated compartments . I n<br />

this discussion the carbon web will be limite d<br />

to a consideration of carbon in the followin g<br />

compartments (fig . 1): (1) Dissolved inorgani c<br />

carbon (DIC), (2) Phytoplankton, (3) Zoo -<br />

plankton, (4) Detritus, (5) Dissolved organi c<br />

carbon (DOC), and (6) Chemoorganotrophi c<br />

bacteria (Stonier et al . 1963). The cycle is<br />

initiated when dissolved inorganic carbo n<br />

species such as bicarbonate (or other chemical<br />

species in the aquatic carbonate syste m<br />

(Stumm and Morgan 1970)) are fixed durin g<br />

the photosynthetic process into organic<br />

matter by the photosynthetic organism (primarily<br />

phytoplankton and secondarily photo -<br />

synthetic bacteria) . Phytoplankton form the<br />

first link in the carbon food chain and ar e<br />

either eaten by herbivorous zooplankton, die<br />

and become part of the detritus compartment,<br />

or excrete dissolved organic carbo n<br />

(DOC). As much as 80 percent of the algal<br />

photosynthate may be excreted as glycolic<br />

acid and other compounds into the wate r<br />

(Hutchinson 1957, Fogg 1963, 1965, Hoo d<br />

1970). Subsequently, microorganisms may<br />

take up the DOC transported by the turbulen t<br />

water and metabolize it to inorganic carbo n<br />

forms. Saunders (1957, 1971) and Saunder s<br />

and Storch (1971) have found that th e<br />

h eterotrophic bacteria and phytoplankto n<br />

form a coupled oscillating pair in which th e<br />

bacteria increase in activity during the day in<br />

response to the daylight production and re -<br />

lease of the algal photosynthate into the surrounding<br />

waters . The herbivorous zooplankton<br />

(and carnivorous zooplankton) also<br />

1 Bruce Lighthart is Director, Institute for Fresh- ,<br />

water Studies, and Assistant Professor, Department of<br />

Biology . Paul E . Tiegs is with Institute for Freshwater<br />

Studies .<br />

289


ATMOSPHERI C<br />

CARBO N<br />

DIOXIDE<br />

■<br />

DISSOLVED<br />

INORGANI C<br />

PH'TOPLAN KTQN ` (<br />

_ BACTERI A<br />

290


contribute to detrital compartment as a resul t<br />

of their fecal deposits or their death and decomposition.<br />

Ultimately DOC and detrital<br />

carbon are decomposed by bacteria or zoo -<br />

plankton feeding on bacteria/detritus mixtures,<br />

to dissolved inorganic carbon as a result<br />

of respiratory oxidative decompositio n<br />

processes .<br />

In nature the standing stock in each compartment<br />

is the dynamic equilibrium level<br />

attained as a result of input/output rate s<br />

about the compartment . The quality of th e<br />

standing stock will also vary as the input /<br />

output mechanisms change .<br />

With the advent of the carbon-14 tracer<br />

techniques, the transports of carbon to and<br />

from each compartment in an aquatic ecosystem<br />

is possible . For example, Steemann-<br />

Nielsen (1952) initiated such studies with C' 4<br />

bicarbonate to estimate the rate of primar y<br />

production in the ocean . Goldman (1963 )<br />

outlined the use of this method in fres h<br />

water .<br />

Sorokin (1966), Shuskina and Monako v<br />

(1969), and Johannes and Satomi (1967) use d<br />

labeled phytoplankton to determine the rat e<br />

at which zooplankton feed and process thei r<br />

waste . The production of dissolved organic<br />

carbon (DOC) by phytoplankton (Fogg 1965,<br />

Saunders and Storch 1971) and uptake by<br />

zooplankton and bacteria have been success -<br />

fully measured using carbon-14 tracers.<br />

Hobbie and Wright (1965) and Wright an d<br />

Hobbie (1966) refined the enzyme saturatio n<br />

technique of Parson and Strickland (1962) a s<br />

measured by a carbon-14 tracer, to evaluat e<br />

the pool size and potential flux of dissolve d<br />

organics through the bacterial (heterotrophic )<br />

compartment in the aquatic web . With<br />

carbon-14 labeled detritus Sorokin (1966 )<br />

measured assimilation by zooplankton . H e<br />

prepared the detritus by homogenizing radio -<br />

actively labeled algal cultures . Others hav e<br />

used autoclaved, labeled natural and culture d<br />

algal populations 2 (Bell and Ward 1970) .<br />

Labeled detritus may also be prepared, b y<br />

foam separating C-14 labeled dissolve d<br />

organics from the liquid phase (Baylor an d<br />

Sutcliff 1963) . However, there is evidenc e<br />

2 G . W. Saunders, personal communication .<br />

that this method will work only in sea water .<br />

For the first time, Saunders 3 recently combined<br />

the aforementioned methods to measure<br />

simultaneously all of the previously liste d<br />

carbon fractions and their fluxes in a "Compartment<br />

Analysis" scheme (Patten 1968 ,<br />

Atkins 1969). He has made this type of<br />

analysis at one location and one instant in<br />

time at several lakes .<br />

Compartment Analysis, in essence, is performed<br />

by establishing a series of known fractions<br />

or compartments in an experimenta l<br />

confine (e .g., glass carboy or plastic bag) and<br />

following a tracer such as carbon-14 as a func -<br />

tion of time as it proceeds through each compartment<br />

in the confined system . Measurement<br />

of the radioactive carbon as it passe s<br />

through each compartment can be used t o<br />

evaluate the rate of incorporation of carbo n<br />

into each compartment in succession, and b y<br />

isotopic dilution or total carbon conten t<br />

analysis, the pool size estimated . These data<br />

may be used to describe mathematically th e<br />

carbon change in compartments as a functio n<br />

of time, e .g .,<br />

d( q 1) _ - k l q l and<br />

dt<br />

d( q 2 )<br />

dt =k l q 1 -k 2 q 2<br />

where (4 1 and q 2 are the specific activities o f<br />

the first and second compartments and k l<br />

and k2 are the rate constants into each compartment.<br />

The compartments could very wel l<br />

be phytoplankton and zooplankton or DOC ,<br />

bacteria and zooplankton, etc. The specifi c<br />

activities rather than tracer amounts will b e<br />

used in mathematical models .<br />

It must be emphasized that the confinement<br />

of an aliquot of the sampled water during<br />

the compartment analysis will yield value s<br />

prevailing for (1) standing stock at the time o f<br />

sample confinement, and (2) uptake value s<br />

for that particular sample during the incubation<br />

period . Thus, it is very important whe n<br />

and where test water samples are chosen . For<br />

example, if lake water were collected for a<br />

compartment analysis of the carbon fraction s<br />

during maximum expected respiratory and<br />

3 Unpublished data .<br />

291


photosynthetic activity with respect to sea -<br />

son, time of day, and depth, one might expec t<br />

measurements to approach the maximu m<br />

values that these variables would attain in<br />

situ .<br />

Reviews and key papers concerning variou s<br />

aspects of the carbon cycle in the aquati c<br />

realm include the following : zooplankto n<br />

physiology (Corner and Coney 1968), primary<br />

productivity (Ryther 1963, Steemann -<br />

Nielsen 1963, and Strickland 1965), dissolve d<br />

organic carbon (Provisoli 1963, Duursma<br />

1965 ), detritus (Parsons 1963), bacteria<br />

(Wood 1965, Brock 1966, and Lighthart<br />

1969), and food chains (Riley 1963) . Also<br />

some detailed methods are given in IBP Hand -<br />

books 12 and 17, Primary Production i n<br />

Aquatic Environments and Secondary Productivity<br />

in Fresh Waters (Edmondson and<br />

Winberg 1971) .<br />

To this point, the discussion has been wit h<br />

what might be called a primary system, the<br />

components in the carbon web . This system is<br />

controlled by another [or] secondary system ,<br />

the physical and chemical environment . Som e<br />

of the environmental factors operating on th e<br />

primary system are listed in table 1 .<br />

Because factors in the secondary syste m<br />

regulate changes in the primary system, meas -<br />

urements of the primary system may allow u s<br />

to interpret the major controlling variables o f<br />

the water bodies .<br />

Table 1 .-Some environmental factors operating<br />

on the aquatic carbon we b<br />

Temperature<br />

Light intensity, quality, and duration<br />

pH<br />

Trace elements<br />

External inputs of organic matter ,<br />

i.e ., nutrients, antibiotics, and<br />

growth factors<br />

Phosphorou s<br />

Silica<br />

Calciu m<br />

Environmental confine s<br />

Water circulation<br />

Purpose<br />

It is proposed that, when possible, one station<br />

will be occupied on each of four lakes i n<br />

the Cedar River watershed : Lake Washington ,<br />

Lake Sammamish, Lake Chester Morse, an d<br />

Lake Findley, during the four seasons of th e<br />

year. While occupied, the flux and pool size s<br />

of six carbon compartments will be measured<br />

in the epi- and hypo-limnions . The purpose of<br />

this investigation is to determine whether the<br />

four lakes are distinguishable in terms of the<br />

fluxes and pool sizes of the six carbon compartments<br />

with the methods outlined below .<br />

Methods<br />

Carbon pools in each compartment will b e<br />

made by : (1) estimating the protoplasmic<br />

carbon pool from cell volume, and (2) from<br />

infrared spectrophotometric measurement o f<br />

oxidized cellular carbon . Carbon flux between<br />

compartment pools will be estimated by<br />

following the transfer of carbon-14 added to<br />

specific compartments .<br />

Carbon Pool Analysi s<br />

It is necessary to determine the "cold "<br />

carbon pool size, both to evaluate its standin g<br />

stock, and in calculating specific activities .<br />

Inorganic carbon in the lake water will b e<br />

measured either by determining the tota l<br />

alkalinity (American Public Health Association<br />

1971), pH, and temperature, and the n<br />

consulting the appropriate tables (Saunders e t<br />

al. 1962) for total available inorganic carbon ,<br />

or by direct measurement using an infrared<br />

CO 2 analyzer (Menzel and Vaccaro 1964) .<br />

The dissolved organic carbon in the water<br />

samples will be determined on the filtrate of a<br />

O.22µ Millipore filtered sample of the test<br />

water using the method of Menzel and Vaccaro<br />

(1964) . Carbonates will be driven off th e<br />

filtrate prior to analysis by sparging th e<br />

phosphoric-acid-acidified filtrate with nitrogen<br />

gas .<br />

The carbon content of the planktonic algae<br />

will be made either from dry weight estimation<br />

by calculation from cell counts an d<br />

29 2


volume measurement, realizing that the protoplasm<br />

has a specific gravity of 1 .1, and is<br />

80-percent water, the carbon content is 5 0<br />

percent of the dry weight (Lund 1965) ; or by<br />

regression with chlorophyll-a measuremen t<br />

(Steemann-Nielsen and Jorgenson 1968) ; or<br />

by both .<br />

(1)<br />

Algal carbon<br />

(protoplasmic volume) X<br />

(protoplasmic specific gravity )<br />

X (protoplasmic P .C. water<br />

content) X (P.C. dry weight<br />

carbon)<br />

Zooplankton carbon will be measured using<br />

the Menzel and Vaccaro (1964) method of<br />

individually picked or 15012 mesh net filtered<br />

samples .<br />

Bacterial carbon will be estimated as th e<br />

product of cell volume (assumed 0 .5 X 112 )<br />

times cell density (1 .1) times water content<br />

(80 percent) times the dry weight carbon con -<br />

tent (50 percent dry weight) of fluorescence<br />

(Strugger 1948) phase contrast microscope<br />

enumerated cells .<br />

Detritus carbon will be evaluated as the<br />

sum of the bacterial and algal dry weight, an d<br />

seston ash weights, subtracted from the total<br />

seston dry weight . The weight of the total<br />

seston will be determined via combustion o f<br />

quadruplicate 0.812 tared glass fiber filtere d<br />

and dried water samples less zooplankton . It<br />

is assumed that 50 percent of the volatil e<br />

seston is carbon (equation 2) .<br />

(2)<br />

Detritus carbon (seston D.W. - zooplankton) -<br />

(algas D .W. + bacterial D.W. +<br />

seston ash weight) X (P .C .<br />

seston D .W. carbon )<br />

Detritus carbon analysis will be attempted<br />

by infrared analysis of combusted total sesto n<br />

after the zooplankton have been remove d<br />

(equation 3) .<br />

(3 )<br />

Detritus carbon = (total seston carbon -<br />

zooplankton) - (algal carbo n<br />

+ bacterial carbon )<br />

Carbon Flux Analysis<br />

Carbon flux will be measured as the trans -<br />

Table 2.-List of compartment systems isolated for carbon flux measurements<br />

Bottle Added tracer Compartment system<br />

1 NaH 14 CO 3 Dissolved inorganic carbon<br />

Phytoplankto n<br />

Dissolved organic carbo n<br />

2 C-starch (U14 ) Dissolved organic carbo n<br />

with or without<br />

Bacteri a<br />

algal hydraulysate<br />

Dissolved inorganic carbon<br />

3 NaH 14 CO 3 Dissolved inorganic carbo n<br />

Phytoplankto n<br />

Zooplankto n<br />

4 <strong>Experimental</strong>ly Dissolved organic carbo n<br />

generated<br />

Bacteri a<br />

DOC-Cl 4<br />

Zooplankto n<br />

(Dissolved inorganic carbon) ?<br />

5 <strong>Experimental</strong>ly Detritu s<br />

generated<br />

Zooplankto n<br />

Detritus-C 14<br />

(Dissolved inorganic carbon)?<br />

29 3


18L SAMPLE LAKE WATER<br />

LIGH T<br />

ACIDIFIED TO pH 3<br />

WITH H 3 P04 AN D<br />

SPARGE WITH N 2<br />

1<br />

ADD ALIQUOT T O<br />

COCKTAIL AND COUNT<br />

Figure 2. Flow diagram of methods to be used to determine carbon flux from DIC to phytoplankton to DOC ,<br />

and detritus to zooplankton .<br />

294


fer of the indicated carbon-14 tracer labele d<br />

substrate through the isolated systems tabulated<br />

in table 2 . Each of the five systems will<br />

be isolated in either glass carboys, if it is a<br />

light-requiring system, or in blacked plasti c<br />

collapsible bottles, if a dark system. Th e<br />

bottles will be incubated in situ .<br />

The flux of carbon from DIC to DO C<br />

through the phytoplankton will be determined<br />

as the uptake during a daylight perio d<br />

of a carbonate spike added to a glass carboy<br />

of lake water incubated in situ . The zooplankton<br />

will be filtered out of this water with a<br />

0.1 mm mesh Nitex net and used in anothe r<br />

experiment. Samples will be withdraw n<br />

hourly from the submerged bottle through a<br />

tube to the water surface . The Millipore filtered<br />

(0 .22/1 porosity) samples will be used t o<br />

separate the phytoplankton from the filtrate .<br />

The filtrate containing the DOC c", after<br />

acidification and nitrogen sparging, will b e<br />

dried in a scintillation vial and counted afte r<br />

moistening . Algal respiration will also b e<br />

evaluated either as the slope of the phytoplankton<br />

carbon-14 loss immediately afte r<br />

darkening the bottle, or in separate classica l<br />

light/dark bottle measurements (fig. 2) .<br />

Upon completion of the foregoing experiment,<br />

a portion of that carboy of lake wate r<br />

will be used to evaluate the carbon flux fro m<br />

DOC generated by the phytoplankton in it, t o<br />

DIC via bacteria . This water will be cleared of<br />

organism and labeled DIC by first Millipor e<br />

filtration and then nitrogen sparging . Subsequently,<br />

fresh lake water , less its zooplankto n<br />

from the original sample site will be added to<br />

the " clean " hot-DOC containing water<br />

incubated in situ in the dark. Particulate (i .e . ,<br />

bacteria) and DIC containing activity in th e<br />

bottle will be monitored for a further 2 4<br />

hours (fig. 3) .<br />

The zooplankton initially filtered from th e<br />

water for the DIC to algae to DOC experiment<br />

(the first experimental description) wil l<br />

be added to a darkened plastic carboy containing<br />

sterile radioactive detritus so the<br />

carbon flux from the detritus to zooplankton<br />

and possibly to DIC compartments may b e<br />

measured . After 4 and 8 hours, one-half of<br />

the carboy will be filtered and the animal s<br />

counted numerically and radioactively. The<br />

three zooplankton data points so obtained<br />

will be graphed to evaluate the uptake rate<br />

(fig. 2) .<br />

The exact method of sterile detritus production<br />

awaits further experimentation to<br />

evaluate the chemical composition and<br />

optimum zooplankton feeding rate on<br />

detritus produced in several ways .<br />

The flux measurement of carbon fro m<br />

phytoplankton to zooplankton will be evaluated<br />

in a lighted glass carboy like the on e<br />

described for DIC to algae to DOC measurements<br />

except that the zooplankton will no t<br />

be removed from the water . Hourly phytoplankton<br />

samples, and 4- and 8-hour sample s<br />

of one-half the vessel will be filtered for<br />

subsequent numerical and radiological zoo -<br />

plankton counts . The mean specific activity<br />

of the phytoplankton food source for th e<br />

zooplankton may be evaluated as the mea n<br />

specific activity of algal carbon during th e<br />

experiment (fig . 4) .<br />

Finally, the flux of dissolved organic<br />

carbon through bacteria to zooplankton wil l<br />

be determined as the rate of carbon 1 4<br />

labeled DOC (i .e., starch and/or C 14 labeled<br />

algal protein hydraulysate) uptake by th e<br />

bacteria (particles passing through a 0 .1 Nitex net) incubated in situ in a<br />

darkened plastic carboy of lake water.<br />

Samples will be withdrawn hourly for th e<br />

bacterial-uptake and one-half the carboy filter<br />

after 4 and 8 hours for the zooplankto n<br />

measurements. Zooplankton will be filtere d<br />

out onto a 0 .1-mm mesh Nitex net (fig . 5) .<br />

All carbon-14 measurements will be made<br />

on a Packard Tri-Garb Model 3375 scintillation<br />

counter using the channel s ' ratio method .<br />

Schindler's (1966) method for phytoplankton,<br />

and Ward et al .'s (1970) method for zoo -<br />

plankton, detritus, and bacteria will be used .<br />

More details of proposed methods may be<br />

found in Edmondson and Winberg (1971) ,<br />

and Parsons and Strickland (1965) .<br />

The numerical methods that will be used i n<br />

the analysis of the data generated by thi s<br />

study will be presented in future communications,<br />

but are anticipated to include : (a) statistical<br />

evaluation of the differences betwee n<br />

flow and pool sizes of the biome test lakes as<br />

295


1500 m l<br />

LAKE WATER<br />

2 L FRO M<br />

PHYTOPLANKTO N<br />

EX<strong>PE</strong>RIMEN T<br />

Figure 3. Flow diagram of methods to be used to determine carbon flux from DOC to bacteria to DIC .<br />

296


18L LAKE WATER<br />

LIGH T<br />

50m1 HOURLY<br />

DURING DAYLIGHT<br />

1/2 CARBOY AT<br />

4 AND 8 HOUR S<br />

WASH GUTS 1 H R<br />

IN JUST TURBI D<br />

YEAST SUS<strong>PE</strong>NSI<strong>ON</strong><br />

NUMERICA L<br />

(AND/OR SEPARATE S<strong>PE</strong>CIES )<br />

AND RADIOACTIVE COUNT S<br />

ZOOPL AN KTO N<br />

Figure 4. Flow diagram of methods to be used to determine carbon flux from DIC to phytoplankton t o<br />

zooplankton .<br />

297


18L LAKE WATE R<br />

PRESERVE WIT H<br />

FORMALI N<br />

SPARGE WITH N 2<br />

10 MINUTES<br />

DRY ALIQUOT, AD D<br />

COCKTAIL AND COUN T<br />

D .O .C .<br />

NUMERICAL<br />

(AND OR SEPARATE S<strong>PE</strong>CIES )<br />

AND RADIOACTIVE COUNT S<br />

ZOOPLANKTO N<br />

Figure 5 . Flow diagram of methods to be used to determine carbon flux from DOC to bacteria to zooplankton .<br />

298


a function of time and space, (b) carbo n<br />

balances within the web and between the tw o<br />

test depths station, (c) carbon transfer efficiencies<br />

between compartments, (d) synthesis<br />

of our data together with other biome measurements<br />

that have small time and space<br />

resolutions, and (e) carbon flow and pool size<br />

input to mathematical models in preparation .<br />

Acknowledgments<br />

The authors want to thank Professor Freid a<br />

B. Taub for her active encouragement of this<br />

study, and Dr. G. W. Saunders, Jr., whos e<br />

work generated this scheme of methods out -<br />

lined in this paper .<br />

The work reported in this paper was sup -<br />

ported in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferous<br />

<strong>Forest</strong> Biome, U .S. Analysis of Ecosystems ,<br />

International Biological Program . This is Contribution<br />

No . 46 to the Coniferous <strong>Forest</strong><br />

Biome and Contribution No . 16, Institute for<br />

Freshwater Studies, Western Washington State<br />

College .<br />

Literature Cited<br />

American Public Health Association. 1971 .<br />

Standard methods for the examination o f<br />

water and waste waters . 874 p . Washington ,<br />

D .C .<br />

Atkins, G . L . 1969. Multicompartmen t<br />

models for biological sciences . 153 p .<br />

London: Methulen and Co . Ltd .<br />

Baylor, E . R., and W . H. Sutcliff, Jr . 1963 .<br />

Dissolved organic matter in seawater as a<br />

source of particulate food . Limnol .<br />

Oceanogr. 8(4) : 869-871 .<br />

Bell, R. K., and F . J. Ward, 1970 . Incorporation<br />

and organic carbon by Daphnia pulex .<br />

Limnol. Oceanogr. 15(3) : 713-726 .<br />

Brock, R. D. 1966. Principles of microbia l<br />

ecology. 306 p. Englewood Cliffs, N .J . :<br />

Prentice-Hall .<br />

Corner, E. D. S., and C . B. Coney . 1968 . Biochemical<br />

studies on the production of<br />

marine zooplankton . Biol. Rev. 4(3) :<br />

393-426 .<br />

Duursma, E . D. 1965. The dissolved organic<br />

constituents of sea water . In J. P. Rile y<br />

(ed.), Chemical oceanography, 433 p . Ne w<br />

York: Academic Press .<br />

Edmondson, W . T., and G . G . Winberg (eds .) .<br />

1971 . A manual of methods for the assessment<br />

of secondary productivity in fres h<br />

waters. IBP Handb . No . 17, 386 p . Philadelphia<br />

: F. A. Davis Company .<br />

Fogg, G. E. 1963 . The role of algae in organic<br />

production in aquatic environments. Brit .<br />

Phycol . Bull. 2(4) : 195-204 .<br />

1965. Algal cultures and phytoplankton<br />

ecology . 126 p. Madison : Univ .<br />

Wis . Press .<br />

Goldman, C. R. 1963. Measurement of primary<br />

productivity and limiting factors i n<br />

freshwater with carbon-14 . In M. S. Doty<br />

(ed .), Proc . Conf. on primary productivit y<br />

measurement, marine and freshwater . AE C<br />

Inf. T .I .O . F633, p . 103-113 .<br />

Hobbie, J . E., and R. T . Wright . 1965 . Bio -<br />

assay with bacterial uptake kinetic : glucose<br />

in freshwater. Limnol. Oceanogr. 10(3) :<br />

471-474 .<br />

Hood, D . W. 1970. Symposium on organi c<br />

matter in natural waters . Inst. Mar. Sci . ,<br />

Univ . Alaska Occas . Publ . 1, 625 p .<br />

Hutchinson, G . E . 1957 . A treatise o n<br />

limnology . Vol. I, 1015 p. New York : Joh n<br />

Wiley & Sons, Inc .<br />

Johannes, R . E., and M. Satomi. 1967. Measuring<br />

organic matter retained by aquatic<br />

invertebrates. J. Fish. Res. Board Can .<br />

21(11) : 2467-2471 .<br />

Lighthart, B . 1969. Planktonic and benthic<br />

bacteriovorous protozoa at 11 stations i n<br />

Puget Sound and adjacent Pacific Ocean . J .<br />

Fish. Res . Board Can . 26 : 299-304 .<br />

Lindeman, R . L. 1942. The trophic-dynami c<br />

aspect of ecology . Ecology 23 : 399-418 .<br />

Lund, J. W. G. 1965. The ecology of th e<br />

freshwater phytoplankton. Biol. Rev . 40 :<br />

231-293 .<br />

Menzel, D. W., and R. F. Vaccaro. 1964. The<br />

measurement of dissolved organic and particulate<br />

carbon in water . Limnol . Oceanogr .<br />

9(9): 138-142 .<br />

Parsons, T . R. 1963. Suspended organic mat -<br />

ter in seawater . In M. Seres (ed .), Progres s<br />

299


in oceanography, Vol . 1, p . 205-239. Ne w<br />

York: Pergamon .<br />

and J. D. H. Strickland. 1962 .<br />

On the production of particulate organic<br />

carbon by heterotrophic processes in se a<br />

water. Deep-Sea Res . 8 : 211-222 .<br />

Patton, B . C. 1968. Ecological modeling with<br />

analog and digital computers . Am. Inst .<br />

Biol . Sci., 40 p .<br />

Provasoli, L. 1963. Organic regulation of<br />

phytoplankton fertility . In M . N. Hill (ed .) ,<br />

The seas, p . 165. New York : Interscienc e<br />

Publ .<br />

Riley, G. A. 1963. Theory of food chain relationships<br />

in the ocean . In M. N . Hill (ed .) ,<br />

The seas, p . 438. New York : Interscience<br />

Publ .<br />

Ryther, J . H. 1963. Geographic variations i n<br />

productibility . In M. N. Hill (ed .), The seas,<br />

Nauk. USSR 18(4) : 955-958 .<br />

Sorokin, Y . I. 1966 . Carbon-14 method in the<br />

study of the nutrition of aquatic animals .<br />

Hydrobiology 51(2) : 209-224 .<br />

Stanier, R. Y., M . Doudoroff, and E . A .<br />

Adelberg . 1963. The microbial work . 75 3<br />

p. Englewood Cliffs, N .J . : Prentice Hall .<br />

Steemann-Nielsen, E . 1952. The use of radio -<br />

active carbon for measuring organic production<br />

in the sea. J. Conserv. Intern.<br />

Explor . 18 : 117-140 .<br />

1963. Productivity, definitio n<br />

and measurement . In M. N. Hill (ed.), The<br />

sea, p. 129. New York : Interscience Publ .<br />

and E. G. Jorgensen . 1968. The<br />

adaptation of plankton algae. I. General<br />

part. Physiol. Plant. 21 : 401-413 .<br />

Strickland, J. D. H. 1965. Production o f<br />

organic matter in the primary stages of the<br />

p. 347. New York : Interscience Publ .<br />

marine food chain . In J. P. Riley and G .<br />

Saunders, G . W. 1957 . Interrelations of dissolved<br />

organic matter and phytoplankton .<br />

Skirrow (eds.), Chemical oceanography, p .<br />

712 . New York : Academic Press .<br />

Bot. Rev. 23(6) : 389-410 .<br />

and T. R. Parsons . 1965 . A<br />

1971 . Coupled oscillations in<br />

the metabolism of bacteria in lakes .<br />

(Abstr.) Am. Soc. Microbiol . (Bacteriol .<br />

Proc.), p . 53 .<br />

and Thomas A . Storch . 1971 .<br />

Coupled oscillatory control mechanism in a<br />

planktonic system . Nature New Biol .<br />

230(10) : 58-60 .<br />

, F. B. Trama, and R. W. Bachmann.<br />

1962. Evaluation of a modified<br />

carbon-14 technique for shipboard estimation<br />

of photosynthesis in large lakes . Univ .<br />

Mich., Great Lakes Res. Div . Publ . No . 8, p .<br />

61 .<br />

Schindler, D . W. 1966 . A liquid scintillation<br />

method for measuring carbon-14 uptake i n<br />

photosyntheses. Nature (London )<br />

211(505) : 844-845 .<br />

manual of sea water analysis . Fish . Res . Bd .<br />

Can . Bull. 125, p. 203 .<br />

Strugger, S. 1948 . Fluorescence microscope<br />

examination of bacteria in soil . Can.J. Res .<br />

26 : 188-193 .<br />

Stumm, W ., and James J . Morgan. 1970 .<br />

Aquatic chemistry : an introduction emphasizing<br />

chemical equilibria in natural waters .<br />

583 p. New York : John Wiley & Sons, Inc .<br />

Ward, F . J ., B . Wong, and G . G . C. Robinson .<br />

1970 . A liquid scintillation procedure for<br />

determining the effect of size on self -<br />

adsorption of carbon-14 in Daphnia pulex .<br />

Limnol. Oceanogr. 15(4) : 648-652 .<br />

Wood, E. J. F . 1965 . Marine microbial<br />

ecology . 243 p . New York : Reinholt .<br />

Wright, R . T., and J. E. Hobbie . 1966 . Use of<br />

glucose and acetate by bacteria and algae in<br />

Shuskina, E. A., and A . V . Monakov . 1969 . aquatic ecosystems . Ecology 47(3) :<br />

The use of radiocarbon for the separation<br />

of planktonic animals into trophic levels .<br />

447-464 .<br />

300


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium.<br />

Bellingham, Washington-March 23-24, 197 2<br />

Dynamics of nutrient supply and<br />

primary production in<br />

Lake Sammamish, Washington<br />

Eugene B . Welc h<br />

an d<br />

Demetrios E . Spyridakis<br />

Department of Civil Engineering<br />

University of Washingto n<br />

Seattle, Washingto n<br />

Abstract<br />

Lake Sammamish, which lies about 12 miles east of Seattle, Washington, with moderate depth (mean 17 .7 m )<br />

and area (19.8 km 2 ), ranks third in productivity of the four Cedar River drainage lakes and can be classified<br />

mesotrophic. While diversion of over one-half the P from nearby Lake Washington during 1963-67 was followe d<br />

by reduction in winter mean P content and a rapid shift from eutrophy to mesotrophy (Edmondson 1970) ,<br />

mean winter P content and measured characteristics of plankton response have not changed in Lake Sammamis h<br />

following a diversion of similar magnitude . Annual nutrient budgets suggest a reduction in sedimented P sinc e<br />

diversion but little change in the quantity of P released from anaerobic sediment . P availability in the water<br />

column (winter mean content) appears to be controlled by Fe precipitation to a greater extent than in Lak e<br />

Washington. Experiments in situ show that N and P are equally limiting to summer phytoplankton productivity ,<br />

but as found in Lake Washington, P may be of more long-term significance.<br />

Introduction<br />

The supply of limiting nutrient is probabl y<br />

the most significant factor that determine s<br />

the trophic status of a lake . This relationshi p<br />

can be affected by several factors, particularl y<br />

those that control nutrient supply such as<br />

morphometric and hydrologic conditions .<br />

Vollenweider (1968) has related supply of N<br />

and P to trophic status of 20 lakes and foun d<br />

that the relationship was dependent on lake<br />

morphometry expressed as mean depth . This<br />

relationship illustrates that reasonably reliable<br />

predictions about the primary production and<br />

trophic status of lakes are possible with in -<br />

formation on the supply of limiting nutrient<br />

and lake morphometry . Such predictability i s<br />

valuable in the management of man's encroachment<br />

on lakes. Of still more value is<br />

knowledge of the response rate of trophi c<br />

status indicators to changes in nutrient<br />

supply .<br />

This paper presents preliminary findings on<br />

a nutrient supply budget for Lake Sammamish<br />

near Seattle, Washington, and ho w<br />

alteration in that supply by sewage diversio n<br />

has affected nutrient limitation and primar y<br />

production . The response in Lake Sammamish<br />

to nutrient diversion is compared to that in<br />

Lake Washington and mechanisms are hypothesized<br />

to explain the delayed recovery in Lak e<br />

Sammamish . Their trophic status is compare d<br />

to that of other study lakes in the Cedar River<br />

drainage, Findley Lake and Chester Mors e<br />

Lake, with the ultimate intent of refining predictive<br />

relationships among trophic status ,<br />

nutrient supply and morphometric factors i n<br />

lakes. Prediction of the rate of response o f<br />

trophic status indicators to change in nutrien t<br />

supply should also be enhanced by detaile d<br />

comparison of manipulated Lakes Sammamish<br />

and Washington .<br />

301


Methods<br />

The water column in Lake Sammamish wa s<br />

sampled at 2-week intervals in spring, summer,<br />

and early fall and monthly in winter during<br />

1970 and 1971 . Phytoplankton biomas s<br />

and productivity and nutrient content wer e<br />

determined in samples collected from severa l<br />

depths. Nutrient supply from surface water s<br />

was estimated by monthly sample collectio n<br />

and flow measurements from two major an d<br />

11 minor streams entering the lake .<br />

Primary productivity was determined i n<br />

situ according to procedures described b y<br />

Goldman (1961) . Water samples inoculate d<br />

with C 14 were incubated at four depths for 4<br />

hours and the results reported as integrate d<br />

productivity in the photic zone extrapolate d<br />

to daily rates assuming a 1 :1 relationship with<br />

incident light. Data are reported from on e<br />

centrally located station in each of the lakes .<br />

Methods of Strickland and Parsons (1968 )<br />

were followed for N, P, and Chlorophyll a<br />

(Chl a) analyses in water . Total and orthophosphate<br />

phosphorus were determined spectrophotometrically<br />

as a phosphomolybdat e<br />

complex. Reactive silicate was also deter -<br />

mined from a silico-molybdate complex .<br />

Nitrate and nitrite were determined spectrophotometrically<br />

following reduction in a<br />

cadmium-copper filled column and are reported<br />

together as NO 3 -N. Chl a was determined<br />

with a Turner Model 110 fluorometer .<br />

Cations were determined by atomic absorption<br />

techniques and anions by routine procedures<br />

(American Public Health Associatio n<br />

1971). All analyses except for total P wer e<br />

performed on filtered (0.451.t poresize) water<br />

samples .<br />

Surface sediments (surface to 10-cm depth )<br />

were collected with a Peterson dredge in the<br />

four lakes . Surface sediments in Lake Sammamish<br />

were sampled more extensively (2 6<br />

samples from 26 different depths) than those<br />

of the other three lakes where only four to si x<br />

samples were collected . Analyses for total C ,<br />

N, and P contents of air-dried sediments wer e<br />

performed with procedures of Baker (1970 )<br />

for C, Bremner (1960, 1965) for N, and<br />

Delfino et al. (1969) for P. The phosphomolybdate-ascorbic<br />

acid method of Strickland<br />

and Parsons (1968) was used for the determination<br />

of extracted P . Results are expresse d<br />

on an ovendried (104°C) basis .<br />

Bioassays to determine the limiting nutrient(s)<br />

were conducted in large (0 .21 m z X5m)<br />

plastic cylinders submerged in the lake for 7<br />

days. Nitrogen, P, C, and Si were added t o<br />

experimental bags separately and in combination<br />

. Phytoplankton response was determine d<br />

by daily measurements of productivity rat e<br />

and Chl a concentration . Significance of<br />

response was judged from results of analysi s<br />

of variance using a factorial design and<br />

Dunnett 's test at the 95-percent level of confidence<br />

(Steel and Torrie 1960) . Data are<br />

graphed as integrated values over time to indicate<br />

total production .<br />

Lake Sammamish<br />

Trophic Status<br />

Lake Sammamish is considered mesotrophic<br />

judging from measurements of phytoplankton<br />

productivity, biomass (Chi a), hypolimnetic<br />

oxygen deficit, and concentration s<br />

and loading of N and P . Guidelines for thes e<br />

characteristics are suggested in table 1 fo r<br />

judging the trophic status of a lake . With on e<br />

exception, values for these characteristics i n<br />

Lake Sammamish fall in between ranges typical<br />

of oligotrophy and those of eutrophy . Th e<br />

exception, mean winter ortho PO 4 -P, i s<br />

greater than the level often considered indicative<br />

of subsequent summer nuisance alga l<br />

blooms (Sawyer 1952). Of probably mor e<br />

significance than winter concentration, how -<br />

ever, is annual supply of P . In this regard ,<br />

Lake Sammamish lies clearly between safe<br />

and danger limits of eutrophication, her e<br />

construed to suggest oligotrophy an d<br />

eutrophy, respectively (Vollenweider 1968) .<br />

These ranges of measured characteristic s<br />

are at best only guidelines for judging the<br />

trophic status in lakes . Eutrophication o f<br />

lakes is a complex process and is affected b y<br />

climate, basin morphology and soil type . Differences<br />

in these factors could lead to considerable<br />

inconsistencies in judging trophi c<br />

status using the limited number of characteristics<br />

in table 1 . On the other hand, the<br />

302


Table 1.-Suggested criteria for judging trophic status of temperate<br />

lakes and respective values for Lake Sammamish<br />

Item Oligotrophic Eutrophic Lake Sammamish<br />

Chlorophyll a ug/l ' (growth season mean) 0-4 10-100 7 . 1<br />

Primary productivity 2 mgC/m 2 • day<br />

(growth season mean) 30-300 1,000-3,000 77 0<br />

Hypolimnetic 02 deficit 3 in mg 02 /cm 2 • day<br />

(mean rate) < .025 >.055 .05 3<br />

Ortho P04 -P in ug/14 (winter mean) >10 12 (total P, 26 )<br />

NO 3 -N in ug/l4 (winter mean) >300 14 2<br />

Total P annual supply s in g/m 2 • yr for<br />

mean depth 17 .7 m .26 .2 0<br />

Total N annual supplys in g/m 2 • yr for<br />

mean depth 17 .7 m 4 .00 3 .8 7<br />

' Based on data from several lakes in Canada and Unite d States .<br />

' Modified from Rhode (1969), including data from Schindler and Nighswander (1970) .<br />

3 After Mortimer (1941, 1942) .<br />

°Modified from Sawyer (1952) .<br />

5 After Vollenweider (1968) .<br />

Table 2.-Comparison of trophic status indicators (May-August means) in<br />

Lake Sammamish surface water with other lakes of the Ceda r<br />

River drainage, Western Coniferous Biome, Washington<br />

Lake Total P P04 -P N0 3 -N Si Chl a Productivity Secch i<br />

ug/l mg C/m 2 • day m<br />

Findley' 4 .9 1 .0 3 .0 76 0.3 370 1 5<br />

Chester Morse 5 .1 1 .0 16 .3 373 1 .6 520 7 . 3<br />

Sammamish 48 .0 7 .0 86 .0 1,100 7 .1 770 3 . 5<br />

Washington 2 18 .7 1 .1 56 .5 9 .5 1,070 2 .3<br />

July and August only because of earlier ice over .<br />

2W. T . Edmondson, personal communication .<br />

30 3


measurements of algal response and the tw o<br />

most commonly limiting nutrients probably<br />

represent a minimum list of variables most<br />

pertinent to the eutrophication process . To<br />

increase the general value of such guidelines ,<br />

the data base should be more extensive in<br />

addition to including sediment constituents .<br />

Cedar River Drainage Lakes<br />

Comparison of nutrient and plankton characteristics<br />

in the four lakes of the study are a<br />

(table 2) shows that Lake Sammamish is inter -<br />

mediate in trophic status. Data from May t o<br />

August show Lake Sammamish to be slightl y<br />

more eutrophic than Lake Washington based<br />

on mean nutrient concentrations, but less<br />

eutrophic based on algal density and productivity<br />

indices. Judging trophic status fro m<br />

nutrient content during the growing seaso n<br />

can be misleading . Lake Washington is actually<br />

more than twice as enriched as Lake Sammamish<br />

based on annual P supply (0 .48 versu s<br />

0.20 g/m2 ), which conforms to the differences<br />

in productivity .<br />

Of more interest in table 2 than the comparison<br />

of Lake Sammamish and Lake Washington<br />

is the striking contrast in the tota l<br />

series. Findley Lake is clearly oligotrophic ,<br />

while Lake Washington near sea level is apparently<br />

transitional between mesotrophy an d<br />

eutrophy . Lake Chester Morse and Lake Sammamish<br />

are intermediate in elevation an d<br />

trophic status, but nearly as widely separate d<br />

as are Findley and Washington (for other<br />

morphometric data, see Taub et al . 1972) .<br />

Data from these four lakes alone should pro -<br />

vide a substantial framework for prediction of<br />

trophic status from lake nutrient supply an d<br />

morphometry information, which is important<br />

to lake management .<br />

The two oligotrophic lakes are also widely<br />

different than the mesotrophic lakes in sediment<br />

characteristics and ionic composition .<br />

Table 3 summarizes data on major chemical<br />

ions for the four lakes . A graded sequence in<br />

water chemical composition from the lakes in<br />

the upper drainage to Lake Washington is<br />

readily apparent. A four- to tenfold increase<br />

in concentration is observed with most<br />

chemical parameters when comparing Findley<br />

and Chester Morse Lakes to Lake Sammamish<br />

and Lake Washington . These radical differences<br />

in chemical quality of the lake waters i n<br />

the Cedar River Drainage are primarily due to<br />

diversified human use and different geologic<br />

formation of the lake basins .<br />

Results of analyses for total C, N, and P i n<br />

samples of surface sediments are presented in<br />

table 4 . The larger C concentrations and the<br />

higher C/N ratios in the sediments of Findley<br />

and Chester Morse Lakes when compared<br />

with those values in the lakes of lower elevation<br />

appear to reflect trophic status and are<br />

most probably due to differences in allochthonous<br />

and autochthonous inputs . In th e<br />

two oligotrophic lakes, most of the organic C<br />

in sediments is derived from allochthonou s<br />

sources which are relatively resistant to mineralization,<br />

partially as a result of low N availa -<br />

Table 3.-Average summer (1971) chemical ion content in surfac e<br />

waters of Cedar River drainage lakes (Barnes 1972 )<br />

Lake HCO 3 S0 4 Cl Ca Mg Na K Specific conductanc e<br />

mg/1 micromhos/cm at 25° C<br />

Findley 9 .8 0 .4 0 .6 1 .1 0 .3 0.8


Table 4.-Average C, N, and P contents of surface sediments from<br />

Cedar River drainage lakes (Bauer 1971, Horton 1972) '<br />

Lake C 2 N P<br />

C/ N<br />

ratio<br />

N/P<br />

ratio<br />

percent mg/1<br />

Findley 8 .98 5 .55 1 .09 16 .18 5 .14<br />

Chester Morse 6 .12 3 .71 1 .56 16.41 2 .3 7<br />

Sammamish 5 .11 4 .82 1 .32 10.60 3 .6 5<br />

Washington 4 .22 3 .71 2 .13 11 .37 1 .74<br />

' Values are in terms of ovendried (104°C) sediment .<br />

2 The sediments from all four lakes contain less than 0 .1 percent CO3 - C on an ovendried basis .<br />

Table 5 .-Comparison of nutrient supply in Lakes Sammamish and Washington<br />

before and after diversion with special reference to Vollenweider' s<br />

(1968) nutrient supply limitations with respect to mean depth '<br />

Item<br />

Lak e<br />

Sammamish<br />

Lak e<br />

Washingto n<br />

Area (km2 ) 19 .8 87 .61 5<br />

Volume (km3 ) .350 2.88 4<br />

Maximum depth (m) 31 64<br />

Mean depth (m) 17 .7 32 . 9<br />

Flushing rate (year) 2 .2 3 . 0<br />

Prediversion annual total P supply (kg) 11,160 92,600<br />

Prediversion annual total P supply per surface area (g/m 2 ) .56 1 .0 6<br />

Percent total P income diverted 65 2 5 5<br />

Postdiversion annual total P supply (kg) 3,906 41,70 0<br />

Postdiversion annual total P supply per surface area (g/m 2 ) .20 .4 8<br />

Vollenweider's danger limit of P supply<br />

for respective mean depth (g/m 2 )<br />

.26 .4 2<br />

Prediversion annual inorg. N supply (kg) 49,100 246,10 0<br />

Prediversion annual total N supply per surface area (g/m 2 ) 3 4 .96 5 .6 3<br />

Percent inorg. N diverted 22 1 2<br />

Postdiversion annual inorg . N (NO3 -N) supply (kg) 38,298 216,56 8<br />

Postdiversion annual total N supply per surface area (g/m2 ) 3 3 .87 4 .3 3<br />

Vollenweider's danger limit of N supply<br />

for respective mean depth (g/m 2 ) 4 .00 6.0 0<br />

1 Emery, Moon, and Welch, unpublished data .<br />

2 Estimated on the basis of population equivalent nutrients diverted and prediversion annual income to the<br />

lake .<br />

3 Total N values are estimated by doubling inorg . N values .<br />

30 5


ility . Autochthonous sources are more significant<br />

in the mesotrophic lakes . The P conten t<br />

of all lake sediments was relatively high, with<br />

the highest occurring in Lake Washington .<br />

The smaller N/P ratios observed in the sediments<br />

of Lake Washington are undoubtedly<br />

due to the higher sediment P concentration .<br />

Nutrient Supply an d<br />

Primary Productio n<br />

The trophic status of a lake is probably<br />

most closely related to the supply of macro -<br />

nutrients N, P and C, which may not be easil y<br />

indicated by concentration . Although growth<br />

rate of phytoplankton is related to the concentration<br />

of a limiting nutrient, productio n<br />

is related to the supply rate of that nutrient<br />

(Dugdale 1967) . The trophic status in 20 well -<br />

known lakes in the world has been related to<br />

annual loading of N and P and mean depth<br />

(Vollenweider 1968) and with continued refinement<br />

such a relationship promises to be<br />

even more useful to management . As indicated<br />

earlier, Lake Washington is more productive<br />

than Lake Sammamish, and this tren d<br />

correlated with P supply or annual loading ,<br />

but not growing-season concentrations . Concentrations<br />

during complete mixing in th e<br />

winter more nearly represent available supply<br />

than do concentrations during the growin g<br />

season. Alteration in the nutrient supply<br />

should change productivity, but the rate o f<br />

that change may depend upon physical characteristics<br />

of the basin .<br />

Effect of Nutrient Diversion<br />

Sewage was diverted from Lake Washington<br />

r<br />

16 (1965 8 1970 )<br />

101- -0-- -0,s<br />

p \<br />

0 V I I } I I I I I I<br />

SEPT OCT NOV DEC IAN FEB MAR APL MAY JUN JUL AU G<br />

Figure 1 . A comparison of surface water total phosphorus for 1964-65 and 1970-71 in Lake Sammamish .<br />

Shaded regions represent differences between means for the period covered by the horizontal lines . Arrows<br />

to the right indicate winter-spring averages of surface water total phosphorus for Lake Washington fo r<br />

prediversion (1963) and postdiversion (1971) conditions (unpublished data-Emery, Moon, and Welch) .<br />

306


during 1963 to 1967 and from Lake Sammamish<br />

in 1968, by the Municipality o f<br />

Metropolitan_ Seattle . The diversion remove d<br />

about 55 percent of the annual external P<br />

supply and 12 percent of the inorganic N<br />

from Lake Washington . Phosphorus and N<br />

external supply into Lake Sammamish was<br />

reduced by about 65 and 22 percent, respectively.<br />

Although the N supply to Lake Washington<br />

is not greatly different than that t o<br />

Lake Sammamish, Lake Washington receive d<br />

about twice the supply of P than Lake Sammamish<br />

(1 .06 vs. 0.56 g/m 2 ) before diversio n<br />

and this difference is still maintained afte r<br />

diversion (0 .48 vs. 0.20 g/m 2 ) (table 5). Th e<br />

diversion brought the P supply to Lake Washington<br />

to near Vollenweider's danger limit fo r<br />

eutrophication and below the danger limit i n<br />

Lake Sammamish (table 5). The prediversio n<br />

N supply did not exceed the danger limi t<br />

nearly as much as did P in either lake so N<br />

diversion may be considered less significan t<br />

than P. According to the alteration in P sup -<br />

ply and if P is most significant in these lake s<br />

as Vollenweider's relationship shows, the n<br />

p h y t o p l ankton productivity and biomas s<br />

should have been reduced in both lakes .<br />

The mean winter (December to April) total<br />

P concentration in the surface waters of Lak e<br />

Washington decreased over 60 percent following<br />

sewage diversion (Edmondson 1970) .<br />

Although diversion was not complete unti l<br />

1967, winter P concentrations began gradually<br />

decreasing soon after the 4-year diversio n<br />

process was initiated in 1963 . In contrast,<br />

little difference can be seen in the 197 1<br />

winter mean P concentrations in Lake Sammamish<br />

3 years after diversion in 1968 (fig .<br />

1) .<br />

Phytoplankton biomass quickly responde d<br />

to the reduction in mean winter P0 4 -P content<br />

in Lake Washington as shown by Edmondson<br />

(1970) (fig. 2) . Chl a decreased in<br />

direct proportion to P0 4 -P, while the other<br />

macronutrients, C and N, varied independent<br />

of Chl a. A significant change in phytoplankton<br />

biomass, production or water clarity ha s<br />

not been observed in Lake Sammamish.' In<br />

I R. M. Emery, C. E. Moon, and E . B. Welch ,<br />

unpublished data .<br />

one respect this is gratifying because winter<br />

mean NO 3 -N and total P concentrations,<br />

which should indicate available supply, also<br />

have not changed . In another respect, the delayed<br />

responses of winter mean P content to<br />

diversion of over one-half the annual suppl y<br />

to the lake suggests that factors controllin g<br />

these winter levels in the two lakes are different<br />

in either kind or magnitude .<br />

Factors controlling winter P concentrations<br />

in Lake Sammamish are not yet understood ,<br />

but comparison of seasonal changes in total P<br />

and morphological characteristics between the<br />

two lakes offers a hypothesis . Winter P content<br />

remained high until the spring diato m<br />

pulse in Lake Washington following which a<br />

moderate decrease was observed (see footnot e<br />

1). In Lake Sammamish, total P content<br />

normally increased to peaks as high as 70 t o<br />

100 pg/1 following turnover in November . Instead<br />

of remaining high until the sprin g<br />

diatom pulse in April, as it does in Lake Washington,<br />

total P decreased during the winter in<br />

Lake Sammamish before the diatom pulse<br />

(fig. 1). The surface water in Lake Sammamish<br />

has been observed to become cloudy<br />

with particulate matter during turnover an d<br />

remain that way for a month or two before<br />

clearing. Phosphorus may be sorbed by this<br />

particulate matter and removed in shallower<br />

Lake Sammamish (mean depth 17 .7 m), while<br />

in deeper Lake Washington (mean dept h<br />

37 m), particulate matter from the bottom is<br />

not so readily mixed to the surface . In support<br />

of this, iron content during and followin g<br />

turnover is higher in Lake Sammamish than i n<br />

Lake Washington particularly in the hypolimnion<br />

. The lower residual P content in late<br />

winter in Lake Sammamish is undoubtedl y<br />

due to P sedimentation through interactio n<br />

with relatively greater amounts of iro n<br />

(Horton 1972, Shapiro et al . 1971) .<br />

Recovery rate in Lake Sammamish may b e<br />

slower than in Lake Washington because th e<br />

former had never attained the enrichment o r<br />

productivity level of the latter . Rate of recovery<br />

might have also been rapid in Lak e<br />

Sammamish if prediversion annual supply had<br />

been as great as that in Lake Washington .<br />

There may exist a control threshold level of P<br />

in the lake, above which alteration by manip -<br />

307


125<br />

•<br />

1933 1963 1964 1965 1966 1967 1968 1969<br />

Figure 2 . (Printed with permission of W. T. Edmondson and Science magazine-Copyright 1970 by th e<br />

American Association for the Advancement of Science .) Mean winter (Jan. to Apr.) values in surface water;<br />

summer (July and Aug .) values of Chl a in surface phytoplankton . The 1963 values, plotted as 100 percent ,<br />

were (in micrograms per liter) : P, 57 ; N, 428 ; and ChI a, 38 . Unconnected points show winter means (Jan .<br />

and Feb .) of bicarbonate alkalinity and free CO 2 in surface water (25 .3 and 3 .2 mg/liter in 1963) .<br />

ulation of external supply would be relativel y<br />

rapid while below that level the respons e<br />

would be very slow .<br />

Response to Short-Term Nutrient Change<br />

To relate external nutrient supply rate to<br />

production and trophic status in the lake, the<br />

nutrient of most significance must be known .<br />

Is phosphorus the most significant nutrient i n<br />

Lake Sammamish as has been the case i n<br />

Washington? This problem was approached i n<br />

two ways: (1) by observing the seasonal<br />

change in N/P ratio compared with productivity<br />

rate and biomass and (2) in situ experimental<br />

addition of macronutrients separately<br />

and in combination to plastic bags . Generally<br />

speaking, if the ratio of N :P by weight is<br />

308


greater than 7 .2:1 (16:1 by atoms) then P<br />

should be limiting further growth and its addition<br />

should increase production with th e<br />

reverse case true for N . If the ratio is aroun d<br />

7 .2 :1, then both nutrients would be require d<br />

to increase production. Of course, this can<br />

only be approximately true because th e<br />

atomic ratio of 16N :1P in algal cells is an<br />

average and interaction among nutrients ma y<br />

occur such that ratios different than thi s<br />

average may stimulate uptake of one or mor e<br />

nutrients .<br />

Lake Sammamish has one major phytoplankton<br />

outburst which occurs in the sprin g<br />

and is dominated by diatoms . In 1970, th e<br />

outburst occurred during May (fig . 3), whic h<br />

for unknown reasons was about 1 month later<br />

than normal. Productivity followed a similar<br />

trend as biomass (Chi a), but carbon assimilation<br />

(productivity) per unit biomass wa s<br />

noticeably greatest prior to the biomass peak .<br />

During this outburst, NO 3 -N decreased by<br />

more than 500 pg/l while ortho P04 -P decreased<br />

by 8 µg/l at most, a removal ratio of<br />

more than 60 :1 (fig. 4) . If this ratio of removal<br />

represents uptake by phytoplankton ,<br />

which it most likely does, then it is weighted<br />

heavily in favor of N relative to the expected<br />

requirements of cells. Inspection of figure 4<br />

suggests that prior to the outburst, P was i n<br />

shortest supply relative to needs, but after -<br />

ward, N actually decreased at the surface to a<br />

level less than P, and N should then have been<br />

limiting. The great change in N relative to P, if<br />

all a result of plankton uptake, suggests tha t<br />

the supply of P0 4 -P was much greater than<br />

indicated by the concentration decrease. This<br />

supply could have come from rapid recyclin g<br />

from the particulate phase and, thus, the relatively<br />

constant P0 4 -P content only represented<br />

the difference between supply and demand<br />

of P .<br />

The change in mean ratio of N :P in the<br />

photic zone is compared to surface productivity<br />

and biomass in figure 3 . Although it is<br />

clear from the ratio that P was in shortes t<br />

supply relative to needs prior to the outburst<br />

and that the uptake was still apparentl y<br />

weighted in favor of N, the mean ratio in th e<br />

photic zone following the outburst remaine d<br />

near the 7 .2 :1 required value . Thus, if th e<br />

average values in the photic zone best reflect<br />

the consequence of nutrient uptake by plank -<br />

ton, then both N and P would appear almos t<br />

1600 - 2 8<br />

I 24<br />

Q 1200<br />

- 80<br />

N E<br />

U<br />

rnE<br />

12<br />

800<br />

8<br />

4<br />

ti<br />

1<br />

- 0<br />

F<br />

M<br />

I I 1 I<br />

A M J J A 5 0 N D<br />

1970<br />

Figure 3 . Primary productivity and chlorophyll a in the surface water and mean N/P ratio in the photic zone o f<br />

Lake Sammamish during 1970 .<br />

309


8 0<br />

600<br />

500<br />

cn 400<br />

z<br />

z 300<br />

0<br />

Z 200<br />

100<br />

0<br />

J I- M A M J J A S<br />

1970<br />

Figure 4 . N0 3 -N and P0 4 -P (ortho) in the surface water of Lake Sammamish during 1970.<br />

0<br />

equally limiting relative to needs during th e<br />

summer .<br />

Nutrient limitation was also determined by<br />

adding N, P, C, and Si singly and in combination<br />

to plastic bags suspended in the lake .<br />

Nitrogen and P additions to bags increased th e<br />

concentrations from a mean of 3 pg/l N and 8<br />

µg/l P to 237 pg/l N and 21 pg/l P, a fina l<br />

ratio of about 11N :1P by weight. Silicon was<br />

added as the third macronutrient in 1970 that<br />

raised concentrations in the bags from a mean<br />

of 0.2 mg/l to 1 .3 mg/l, while C was added i n<br />

1971 that raised the concentration from 7 . 5<br />

mg/l to 16 .9 .<br />

Results of the two experiments are show n<br />

in figure 5 . Response is measured in terms o f<br />

Chl a averaged over 7 days and C 14 productivity<br />

averaged over the first 2 days in 197 0<br />

and the first 4 days in 1971 . In no case did<br />

the single addition of a nutrient significantly<br />

stimulate C 14 uptake or biomass accumulation.<br />

The phytoplankton response was significantly<br />

greater than the control only when N<br />

and P were added together. This suggests that<br />

the ratio of available N :P prior to addition of<br />

N or P must have been near the required rati o<br />

of the plankton cells . This is corroborated by<br />

the mean N:P ratio in the photic zone durin g<br />

summer being very close to the 7 .2:1 mean<br />

considered typical for phytoplankton cells .<br />

Interpretations of limiting nutrients fro m<br />

seasonal data and results of experiments must<br />

be used with caution to predict effects of<br />

nutrient change in a lake . The conclusion that<br />

N and P are equally limiting to phytoplankton<br />

in Lake Sammamish as a result of experimental<br />

additions is based on measure d<br />

changes in biomass . This is largely because the<br />

response is determined over several day s<br />

allowing ample time for biomass accumulation.<br />

Growth rate may or may not change i n<br />

response to nutrient addition or change de -<br />

pending upon where the raised limiting nutrient<br />

level is on the growth rate-nutrient concentration<br />

response curve . However, biomass<br />

will respond usually to an increased suppl y<br />

rate of limiting nutrient if other factors are<br />

not limiting growth. The change in the N : P<br />

ratio is observed after the net result of nutrient<br />

uptake and loss has occurred . This can b e<br />

taken to indicate that an increase or decreas e<br />

in the supply of limiting nutrient will potentially<br />

alter biomass accordingly if othe r<br />

environmental factors are optimum .<br />

310


2 .0<br />

1 .0<br />

0<br />

Si P-N N-Si P-Si P-N-Si Contr .<br />

(A) 1970 BAG EX<strong>PE</strong>RIMENT S<br />

r<br />

9<br />

6<br />

3<br />

0<br />

P N C P-N N-C P-C P-N-C Contr .<br />

(B) 1971<br />

BAG EX<strong>PE</strong>RIMENTS<br />

Figure 5. Response of phytoplankton to added nutrients in in situ experiments in Lak e<br />

Sammamish as measured by C 14 assimilation and Chl a content averaged over time .<br />

Significant differences (95-percent level) from the control are indicated by dotted line s<br />

(unpublished data-Emery, Moon, and Welch) .<br />

311


As the results of Edmondson (1970) sho w<br />

in figure 2, a continual reduction in winte r<br />

mean P04 -P content lead to a proportional<br />

long-term decrease in mean summer algal biomass.<br />

The winter means indicate the supply o f<br />

nutrient available for subsequent productio n<br />

in summer . This may not mean that in short -<br />

term experiments PO 4 -P was always limiting<br />

biomass production. In fact, Edmondson<br />

(1970) showed from the N :P ratio in surface<br />

water that prior to sewage diversion N wa s<br />

more limiting than P and as P began to de -<br />

crease more than N, then P became limiting .<br />

This indicates that relationships betwee n<br />

annual changes in supply of limiting nutrien t<br />

and mean biomass are more predictable than<br />

can be obtained from short-term experiments .<br />

Seasonal alterations in the N :P ratio, as occur<br />

in Lake Sammamish, can result in different<br />

nutrients limiting at different times as a resul t<br />

of interaction between biological uptake an d<br />

chemical factors. The nutrient that limits<br />

most of the time and is, therefore, most influential<br />

on lake production can be best indicated<br />

from annual data. Although N and P<br />

appear equally limiting in Lake Sammamis h<br />

during August, P supply is considered th e<br />

most critical factor controlling long-term production<br />

as has been the case in Lake Washing -<br />

ton. Although plankton biomass in Lake Sam -<br />

mamish has not responded to change i n<br />

external P supply, the springtime concentration,<br />

which also represents supply for summe r<br />

growth, has also not changed in contrast t o<br />

the situation in Lake Washington .<br />

Internal-External Nutrient Supply<br />

The internal supply of nutrients, principally<br />

from sediments, may also partly explai n<br />

the slow recovery in Lake Sammamish fro m<br />

diversion of part of the external supply . Th e<br />

total P budget in a 30 m deep water column<br />

of Lake Sammamish was calculated for the<br />

period of November 15, 1970, to Novembe r<br />

19, 1971, as a function of inflow to and outflow<br />

from the lake, and of seasonal variatio n<br />

in P concentration in the lake water colum n<br />

(table 6). These calculations were made in a<br />

manner similar to those employed by Vollenweider<br />

(1968) for data provided by Bachofe n<br />

(1960) on Lake Baldeggersee . The differences<br />

in the values of P concentration in the lak e<br />

water column at the onset and the end of th e<br />

stratified period, corrected for external P sup -<br />

ply, represent the relative contribution o f<br />

sedimented P in the water column . The onse t<br />

of stratification in Lake Sammamish wa s<br />

designated to coincide not with 0 2 depletio n<br />

in the hypolimnion but rather with the firs t<br />

Table 6.-Exchange of total P between water and deep sediments in Lake Sammamis h<br />

during one aerobic and one anaerobic period '<br />

Relative hypolimnion oxygen<br />

Aerobic period,<br />

Nov . 15, 1970-<br />

Anaerobic period ,<br />

May 21, 1971-<br />

conditions May 21, 1971 Nov. 19, 197 1<br />

(188 days) (182 days)<br />

Change in total P content o f<br />

representative column of water, g/m 2 -0 .870 +1 .07 0<br />

Lake retention from surface loading, g/m 2 + .134 + .26 5<br />

Algebraic difference<br />

= amount released by sediments, g/m 2 + .80 5<br />

or<br />

= amount taken up by sediments, g/m 2 -1 .00 4<br />

Mean daily rates of release or uptake, g/m 2 day - .006 + .004<br />

' From Moon (1972) .<br />

312


detectable increase in concentration of hypolimnetic<br />

total P (25-30 m from the lake water<br />

surface) over that observed in late spring.<br />

Thus, the increase in the hypolimnetic total P<br />

concentration from about 15 pg/1 in Ma y<br />

1971 to 25 pg/1 in June 1971 was taken t o<br />

indicate the onset of lake stratification .<br />

Similarly, the end of lake stratification coincided<br />

with the period at which maximu m<br />

changes in hypolimnetic total P concentration<br />

were observed : from about 80 pg/1 on November<br />

19, 1971, to 15 µg/1 on November 26 .<br />

A net P deposition in the sediments o f<br />

Lake Sammamish is indicated from the data<br />

in table 6 . For the period of May 21, 1971, to<br />

November 19, 1971, P was released by the<br />

sediments at a rate of 0 .004 g/m 2 -day . This is<br />

nearly three times greater than the corresponding<br />

external supply of 0.0015 g/m 2 -day<br />

for the same period . However, only a portion<br />

of the P internal supply from the sediment<br />

will find its way to the trophogenic layer of<br />

the lake. Normally, most of the sediment -<br />

released P will remain locked in the hypolimnion,<br />

and enrichment of the epilimnion by<br />

sediment phosphorus will depend on edd y<br />

diffusion and other mixing mechanisms conditioned<br />

by exposure, morphometric, an d<br />

hydrologic factors; or by the release of gase s<br />

from the sediments .<br />

It is difficult to appreciate the relative importance<br />

of this P internal supply to the<br />

trophic dynamics of the water column . Very<br />

few exact calculations of phosphorus internal<br />

supply have been made that would permi t<br />

accurate comparisons with these results . Internal<br />

P supply from sediments in Lak e<br />

Baldeggersee, a eutrophic lake in Switzerland ,<br />

of 0.009 to 0.01 g/m 2 -day (Vollenweider<br />

1968) is nearly double that observed in Lake<br />

Sammamish . This fact alone, however, canno t<br />

be employed to explain differences observe d<br />

in the trophic status of these two lakes . Sediment<br />

P release rates similar to the ones measured<br />

in the deep water column of Lake Sammamish<br />

would probably have a greater influence<br />

on the trophic status of small lakes or ,<br />

for that matter, the shallow waters of Lak e<br />

Sammamish . However, preliminary calculations<br />

of sediment P released in the shallower<br />

(10 m deep) waters of Lake Sammamish,<br />

where oxidized conditions predominate, indicate<br />

a P internal supply at least one-half that<br />

measured in deep water .<br />

The P uptake by Lake Sammamish sediments<br />

during the mixed period (November<br />

15, 1970, to May 21, 1971) was 0 .00 6<br />

g/m2 -day. Comparison of Lake Sammamish P<br />

budgets before and after sewage diversion<br />

indicates a nearly threefold decrease in P up -<br />

take by sediments in the period of November<br />

to May following autumn lake destratification.<br />

More striking than this observation is the<br />

fact that the rates of sediment P release fo r<br />

the period of May to November for the year s<br />

1964-65 and 1970-71 were identical, 0 .004<br />

g/m 2 -day. This fact will indicate that th e<br />

mechanism that controls the sediment- P release,<br />

primarily the anoxic conditions in the<br />

hypolimnion, has not changed appreciably<br />

following sewage diversion. The prevailing<br />

anoxic conditions in Lake Sammamish durin g<br />

late summer and fall bring about large in -<br />

creases in hypolimnetic Fe concentration s<br />

(Horton 1972), which upon autumn overturn<br />

cause rapid disappearance of P from the water<br />

column. Low total Fe concentrations in th e<br />

hypolimnion of Lake Washington (one-fifth<br />

to one-tenth of that in Lake Sammamish 2 )<br />

likely account for the low removal of P fro m<br />

Lake Washington following autumn overturn<br />

and maintenance of high winter P concentrations.<br />

In the winter months many more than the<br />

monitored 12 small creeks discharge int o<br />

Lake Sammamish ; however, their relative contribution<br />

to the P inflow is small when compared<br />

to the largest stream, Issaquah Creek ,<br />

which contributes more than 75 percent an d<br />

90 percent of the measured surface P inflow<br />

to the lake during the winter and summe r<br />

months, respectively. Contributions from<br />

ground and rainwater, or from direct soi l<br />

surface runoff, are not included in the calculations<br />

presented in table 6 . The only surfac e<br />

outflow from Lake Sammamish is through th e<br />

Sammamish River. Measurements of P out -<br />

flow from the lake were based on P measurements<br />

in that river .<br />

Although the deep water column total P<br />

2W. T. Edmondson, personal communication .<br />

313


udget in Lake Sammamish indicates a significant<br />

release of sediment P to the overlying<br />

waters, the relative contribution of internal P<br />

supply to the trophogenic layer, when compared<br />

to external supply is far from clear .<br />

Additional annual P budgets in Lake Sammamish<br />

and in other lakes of varying nutrient<br />

income, but with similar morphometric an d<br />

hydrological characteristics, will provide a<br />

more accurate delineation of the role of sediments<br />

in supplying P to the water column . In<br />

particular, measurements of P internal suppl y<br />

for Lake Washington, which is a lake that i s<br />

morphometrically a macroscale of Lake Sammamish,<br />

contains relatively smaller amount s<br />

of Fe and P in hypolimnetic waters during th e<br />

summer stagnation periods, and is mor e<br />

eutrophic than Lake Sammamish, will provid e<br />

a unique test case . External P supply that re -<br />

mains in the epilimnic and metalimnic layer s<br />

of Lake Sammamish, at a rate of 0 .001 5<br />

g/m 2 •day during the summer stagnatio n<br />

period, apparently provides a more readil y<br />

available source of P to the trophogenic laye r<br />

of the lake than does hypolimnetic P . A more<br />

accurate evaluation of the relative importance<br />

of internal P supply to Lake Sammamish productivity<br />

must await a better understanding<br />

of the mixing mechanisms .<br />

Acknowledgments<br />

Assistance was provided by graduate students<br />

in Civil Engineering working on individual<br />

research projects as part of the IBP program.<br />

In particular, these are R . S. Barnes, D .<br />

H. Bauer, R . M. Emery, G . R. Hendrey, M. A .<br />

Horton, and C . E. Moon .<br />

The work reported in this paper was sup -<br />

ported in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferous<br />

<strong>Forest</strong> Biome, U .S. Analysis of Ecosystems ,<br />

International Biological Program. Support was<br />

also furnished by an Environmental Protection<br />

Agency Research Fellowship, No .<br />

5-Fl-WP-26, a Public Health Service Training<br />

Grant and an Office of Water Resources Re -<br />

search Allotment Grant, No . A-045 Wash .<br />

This is Contribution No . 47 to the Coniferou s<br />

<strong>Forest</strong> Biome .<br />

Literature Cited<br />

American Public Health Association. 1971 .<br />

Standard methods for the examination o f<br />

water and wastewater . Ed . 13. New York .<br />

Bachofen, R. 1960. Stoffhaushalt and Sedimentation<br />

in Baldogger-und Hallwilersee . P .<br />

1-118 . Zurich : Juris-Vorlay .<br />

Baker, R. G . 1970. Initial reports of the dee p<br />

sea drilling project . Vol. IV, p . 746-748 .<br />

Washington, D.C . : U .S. Govt . Print. Office .<br />

Barnes, R. S. 1972 . Trace metal survey o f<br />

Lake Washington drainage from alpine<br />

regions to lowland lakes . 150 p . M .S. thesis<br />

on file, Univ . Wash., Seattle .<br />

Bauer, D. H. 1971 . Carbon and nitrogen in<br />

the sediments of selected lakes in the Lak e<br />

Washington drainage . 91 p. M .S. thesis on<br />

file, Univ. Wash., Seattle .<br />

Bremner, J. M. 1960. Determination of N in<br />

soil by the Kjeldahl method . J. Agric . Sci .<br />

55: 11-33 .<br />

. 1965. Total N . In C . A. Blac k<br />

(ed.), Methods of soil analysis, Part 2, p .<br />

1149-1178 . Madison, Wis . : Am. Soc .<br />

Agron .<br />

Delfino, J . J., G. G. Bortleson, and G . F . Lee.<br />

1969. Distribution of Mn, Fe, P, Mg, K, Na ,<br />

and Ca in the surface sediments of Lake<br />

Mendota . Environ . Sci. Technol. 3 :<br />

1189-1192 .<br />

Edmondson, W . T. 1970. Phosphorus, nitrogen<br />

and algae in Lake Washington afte r<br />

sewage diversion . Science 169 : 690-691 .<br />

Goldman, C . R . 1961 . The measurement o f<br />

primary productivity and limiting factors in<br />

freshwater with carbon-14 . In M. S. Doty<br />

(ed.), Proceedings of the conference on<br />

primary productivity measurements, marine<br />

and freshwater, p . 103-112 . U .S. At .<br />

Energy Comm ., Washington, D .C .<br />

Horton, M . A. 1972 . The role of the sediment s<br />

in the phosphorus cycle of Lake Sammamish.<br />

220 p. M .S. thesis on file, Univ .<br />

Wash., Seattle .<br />

Moon, C. E. 1972. The effect of wastewater<br />

diversion on the nutrient budget of Lake<br />

Sammamish . 160 p. M .S. thesis on file ,<br />

Univ . Wash., Seattle .<br />

Mortimer, C . H. 1941. The exchange of dissolved<br />

substances between mud and wate r<br />

314


in lakes . J. Ecol . 29 : 280-329 .<br />

. 1942 . The exchange of dissolved<br />

substances between mud and wate r<br />

in lakes . J. Ecol . 30 : 147-201 .<br />

Rhode, W. 1969. Crystallization of eutrophication<br />

concepts in northern Europe . In<br />

Eutrophication : causes, consequences, and<br />

correctives, p. 50-64 . Natl. Acad. Sci . ,<br />

Washington, D .C.<br />

Sawyer, C. N. 1952 . Some new aspects of<br />

phosphates in relation to lake fertilization .<br />

Sewage & Ind . Wastes 24 : 768-776 .<br />

Schindler, D . W. 1971. Carbon, nitrogen and<br />

phosphorus and the eutrophication of<br />

freshwater lakes . J. Phycol . 7 : 321-329 .<br />

and J. E . Nighswander. 1970 .<br />

Nutrient supply and primary production in<br />

Clear Lake, eastern Ontario . J. Fish . Res .<br />

Board Can . 27 : 2009-2036.<br />

Shapiro, J ., W . T. Edmondson, and D . E .<br />

Allison. 1971. Changes in the chemical<br />

composition of sediments of Lake Washing -<br />

ton, 1958-1970 . Limnol. & Oceanogr. 16 :<br />

437-452 .<br />

Steel, Robert G., and J. H. Torrie . 1960 .<br />

Principles and procedures of statistics. 48 1<br />

p. New York : McGraw-Hill .<br />

Strickland, J . D. H., and T. R. Parsons . 1965 .<br />

A manual of seawater analysis . Ed. 2 . Bull .<br />

Fish. Res . Board Can . No. 125, 203 p .<br />

Taub, F. B., R. L. Burgner, E . B. Welch, an d<br />

D. E. Spyridakis. 1972. A comparative<br />

study of four lakes . In Jerry F. Franklin, L .<br />

J. Dempster, and Richard H . Waring (eds .) ,<br />

Proceedings-research on coniferous forest<br />

ecosystems-a symposium, p. 21-32. Pac .<br />

Northwest <strong>Forest</strong> & Range Exp. Stn ., Portland,<br />

Oreg .<br />

Vollenweider, R . A. 1968. Scientific fundamentals<br />

of the eutrophication of lakes and<br />

flowing waters, with particular reference to<br />

N and P as factors in eutrophication . 182 p .<br />

Organ . Econ . Co-Op . & Dev . Dir. Sci. Aff . ,<br />

Paris .<br />

315


Proceedings-Research on Coniferous <strong>Forest</strong> Ecosystems-A symposium .<br />

Bellingham, Washington-March 23-24, 197 2<br />

Hydroacoustic assessment of<br />

limnetic-feeding fishes<br />

Richard E . Thorn e<br />

Fisheries Research Institut e<br />

University of Washingto n<br />

Seattle, Washington 9819 5<br />

Abstract<br />

Hydroacoustic techniques have been applied at the University of Washington to determine the number and<br />

biomass of limnetic fishes in order to evaluate their role in the productivity of lake systems. The lakes are<br />

surveyed with high frequency, high resolution portable echo sounders . The echo signals are recorded o n<br />

magnetic tape and analyzed by a special computer program. Information on size and species composition is<br />

obtained primarily by net sampling, but acoustic determination of size appears feasible in some cases .<br />

Introduction<br />

Determination of the numbers or biomas s<br />

of fishes has been a continual problem in<br />

fishery management . Traditional techniques<br />

based on catch-per-unit-effort and tagging<br />

experiments have many inadequacies. Th e<br />

problem of assessing fish populations in lake<br />

systems is further compounded by the fac t<br />

that fisheries in lake systems are generally<br />

limited, highly selective, or nonexistent, s o<br />

that catch statistics are of little value .<br />

Determination of the numbers and biomas s<br />

of limnetic-feeding fishes in lake systems i s<br />

essential to assess recruitment, growth, mortality<br />

rates, distribution patterns and inter -<br />

actions with other trophic levels . As part of<br />

International Biological Program studies o f<br />

the Coniferous <strong>Forest</strong> Biome at the Universit y<br />

of Washington, this problem has been<br />

attacked through the application of hydroacoustic<br />

assessment techniques .<br />

History of<br />

Acoustic Assessment<br />

of Fish Populations<br />

Echo sounders have been used since th e<br />

mid-1930's to study the distribution and relative<br />

abundance of fish populations, especiall y<br />

in marine environments . However, prior t o<br />

about 1960, hydroacoustic studies of fis h<br />

populations were essentially dependent on<br />

subjective interpretation of echogram records .<br />

During the last decade, a variety of electroni c<br />

devices for automated signal processing has<br />

been developed. These advances, combined<br />

with improved data acquisition systems an d<br />

increased understanding of acoustic principles,<br />

have resulted in a number of successfu l<br />

applications of acoustic techniques to fish<br />

population estimations both in marine and<br />

fresh waters (Truskanov and Scherbino 1966 ;<br />

Cushing 1968; Thorne 1970 ; Thorne and<br />

317


Woodey 1970 ; Thorne, Reeves, and Millika n<br />

1971 ; Moose, Thorne, and Nelson 1971) .<br />

Acoustic Characteristics<br />

of Fish<br />

An echo sounder produces a pulse of sound<br />

from its transducer in the form of a spherically<br />

spreading cone whose dimensions are<br />

dependent on the type and size of the transducer.<br />

The distribution of sound energy transmitted<br />

in different directions is described b y<br />

the directivity pattern function and is maxi -<br />

mum in the direction perpendicular to th e<br />

transducer surface, termed the acoustic axis .<br />

where Ie is the intensity of the reflecte d<br />

sound measured lm from the target ,<br />

and Ii is the intensity incident on th e<br />

target .<br />

Investigations into the relationship betwee n<br />

target strength and fish size indicate that, in<br />

general, the intensity of the echo from a fis h<br />

is proportional to its weight (Cushing et al .<br />

1963 ; Shishkova 1964) .<br />

Data-Acquisition System<br />

The equipment used on the lake studies a t<br />

the University of Washington includes an ech o<br />

sounder incorporated into a system by whic h<br />

target data is recorded on magnetic tape . A<br />

block diagram of the system is shown in<br />

figure 2 . The receiver-transmitter and the<br />

chart recorder is a Ross 200A Fineline ech o<br />

sounder with a frequency of 105 kHz and a<br />

transmitted pulse power of about 500 w . A<br />

transmitter pulse duration of 0.6 msec is<br />

generally used . The receiver amplifier include s<br />

a time-varied-gain circuit of 20 log R, where R<br />

represents depth . This circuit corrects for on e<br />

way spreading loss of signal intensity wit h<br />

depth .<br />

Figure 1 . Typical transducer directivity pattern .<br />

An example of a directivity pattern of a transducer<br />

is shown in figure 1 . A common way to<br />

describe the width of a sound beam is to us e<br />

the half value angle, that is, the angle at whic h<br />

the sound intensity has dropped to one-hal f<br />

(-3 in decibels) of the value it has on th e<br />

acoustic axis. Since the cross-sectional area of<br />

the cone increases with range or depth, the<br />

intensity of the sound within the cone correspondingly<br />

decreases in proportion to th e<br />

square of the depth .<br />

When sound is reflected by a fish target ,<br />

the intensity of the reflected sound is proportional<br />

to the incident sound intensity and is<br />

dependent on characteristics of the fish target.<br />

A measure of the magnitude of the echo<br />

from a fish is the target strength, TS, which i s<br />

defined as<br />

TS = 10 log (Ie/Ii)<br />

Chart<br />

recorde r<br />

Powe r<br />

supply<br />

interface<br />

amplifie r<br />

Receiver<br />

transmitte r<br />

Transduce r<br />

Top e<br />

recorder<br />

I<br />

Oscilloscop<br />

e<br />

Figure 2 . Block diagram of data-acquisition system .<br />

Two modifications of the echo sounder receiver<br />

were made so that it could be usable in<br />

the data collection system . A transistorized<br />

isolation amplifier, which prevents loading of<br />

the echo sounder receiver circuitry by associated<br />

equipment, was built and installe d<br />

318


within the receiver-transmitter unit . The isolation<br />

amplifier couples echo signal data an d<br />

synchronization pulses to the other components<br />

of the system . The second necessary<br />

modification to the receiver was the installation<br />

of a vernier type control that allows precise<br />

adjustment of the receiver gain .<br />

An interface amplifier is used to connec t<br />

the signal output of the echo sounder to th e<br />

input of the tape recorder. A direct connection<br />

between the two units is not possible be -<br />

cause of bandpass limitations of the tap e<br />

recorder. Target data from the output of th e<br />

echo sounder are of a 105 kHz frequency ,<br />

whereas the maximum frequency response of<br />

the tape recorder at unity gain is about 8 kH z<br />

at a tape speed of 3 3%4 IPS. The interface amplifier<br />

converts the 105 kHz output frequenc y<br />

of the echo sounder to a frequency of 5 kH z<br />

by the use of chopper and filter circuits .<br />

The transducer produces about an 8-degre e<br />

circular beam to the 3dB points . The transducer<br />

is generally mounted on a towin g<br />

vehicle suspended from the side of the boat .<br />

The towing vehicle can be used from smal l<br />

outboard vessels as well as large boats, thu s<br />

allowing complete portability of the system .<br />

Survey Design<br />

Surveys are primarily conducted at night ,<br />

since most limnetic-feeding fishes exhibit pronounced<br />

diel behavior patterns. The fish are<br />

generally dispersed in midwater at night bu t<br />

are either on the bottom or schooled i n<br />

deeper water during the day . The number and<br />

spacing of transects depends on the variability<br />

of the population and the degree of accurac y<br />

required. Studies of optimal survey procedures<br />

have not yet been carried out in detail .<br />

Transect design, especially in initial surveys ,<br />

has been governed primarily by the area to b e<br />

covered and the available time . During surveys<br />

of Iliamna Lake, a large sockeye-producing<br />

lake in Alaska, transects were spaced abou t<br />

six miles apart, while in Quinault Lake, a<br />

small sockeye-producing lake in Washington ,<br />

transects were spaced about 3/4 mile apart. In<br />

Lake Washington, where considerable data<br />

have been collected, survey coverage is about<br />

one transect per mile . The primary consideration<br />

in this design was the number of transects<br />

which could be run in a single night . If<br />

greater precision is required, the same transects<br />

are repeated a second night . During<br />

studies of the hake (Merluccius productus)<br />

population in Port Susan, Washington, w e<br />

found that about 40 transects with an average<br />

length of about 3,500 m over a 45 million m 2<br />

area were required to produce a precision o f<br />

about ±15 %, and an additional 40 transects<br />

reduced it to ±10 % (Thorne et al . 1971) .<br />

Basic Techniques<br />

Data Analysis<br />

Techniques of processing acoustic data fo r<br />

abundance estimation can be broken dow n<br />

into two types : echo counting techniques ,<br />

which count the numbers of individual tar -<br />

gets, and echo integration techniques, whic h<br />

relate fish density to integrated target voltage<br />

(Thorne and Lahore 1969) . An examinatio n<br />

of the theoretical variance associated with th e<br />

two techniques has been conducted by Moos e<br />

and Ehrenberg (1971) . Principles of echo integration<br />

and its application to fish populatio n<br />

estimates are described by Thorne (1970 ,<br />

1971), Thorne and Woodey (1970), an d<br />

Thorne, Reeves, and Millikan (1971) . Echo<br />

counting techniques depend on the ability t o<br />

resolve almost exclusively individual fish tar -<br />

gets, thus are limited to relatively low densities.<br />

However, echo counting techniques hav e<br />

the advantage that they may also provide dat a<br />

on size distribution of targets . Applications of<br />

hydroacoustic techniques to lake systems at<br />

the University of Washington have generally<br />

included a combination of the two techniques.<br />

Basic analysis of data is done by integration,<br />

but the integrators are calibrated an d<br />

size information derived by echo counting .<br />

Echo Countin g<br />

Echo counts are made by observation of<br />

specific depth intervals on an oscilloscope<br />

while the tape recording of a transect i s<br />

played back. The transecting boat speed and<br />

sounder pulse rate are such that nearly all fis h<br />

319


targets are insonified several times . The tru e<br />

number of fish is determined by countin g<br />

only the peak echo amplitudes from eac h<br />

series of returns from a single fish (corresponding<br />

to the location of the fish nearest<br />

the acoustic axis) . Counting errors are directl y<br />

proportional to the concentration of fish .<br />

When several fish are observed within th e<br />

counting stratum simultaneously, it is difficult<br />

to keep track of the various targets .<br />

Under these conditions an alternative counting<br />

technique can be used . Instead of counting<br />

every fish target, the number of targets<br />

within the stratum can be noted for randomly<br />

selected pulses, and a mean number of target s<br />

per pulse determined for the transect . This<br />

number can be directly compared with th e<br />

sampling volume of the cone to determine th e<br />

mean density of fish along the transect .<br />

Determination of the Sampling Volum e<br />

The sampling volume of the sounder con e<br />

can be approximated from the directivity pat -<br />

tern. However, the volume is also dependen t<br />

on the size and depth of the fish targets, th e<br />

transmitter power and receiver gain of th e<br />

sounder and the minimum threshold fo r<br />

counting. The sampling volume can be directly<br />

determined from the number of times a<br />

target remains within the sounder cone as th e<br />

boat passes over the fish at a known speed .<br />

The width of the path of a fish through th e<br />

sound cone is determined from the formula<br />

w =<br />

boat speed (meters/sec) times duration in cone (pulses)<br />

pulse rate of sounder (pulses/sec )<br />

where duration is the average number of time s<br />

an individual target is sounded upon . It can b e<br />

shown mathematically that the average length<br />

of parallel chords through a circle is 7r/4 time s<br />

the diameter. Thus the diameter of th e<br />

sounder cone at the depth of the fish target s<br />

is 4/7r times the average path width . The sampling<br />

volume of the cone on a single pulse i s<br />

then the cross-sectional area of the cone (irr 2 )<br />

at the mean depth times the depth interval .<br />

The volume of water surveyed along a transect<br />

is the diameter of the sampling cone a t<br />

the mean depth, times the depth interva l<br />

times the transect length .<br />

Integration Procedure<br />

Since echo counting from the oscilloscop e<br />

is time consuming and its application limited<br />

to lower density situations, basic data processing<br />

is generally done primarily by echo integration.<br />

Determination of integrated voltages<br />

from the data collected on magnetic tape is<br />

done by use of a special integration system<br />

utilizing a small general-purpose computer .<br />

This system, called the Digital Data Acquisition<br />

and Processing System (DDAPS), integrates<br />

voltages from fish targets within several<br />

depth intervals simultaneously and calculate s<br />

the fish abundance using input calibration an d<br />

target strength data (Moose, Green, an d<br />

Ehrenberg 1971) . A block diagram of the<br />

system is shown in figure 3 . An example of<br />

the DDAPS output is shown in figure 4 .<br />

Tap e<br />

playe r<br />

Teletyp e<br />

printer<br />

Digita l<br />

squarin g<br />

circui t<br />

Figure 3 . Block diagram of digital data acquisitio n<br />

Upper limit of<br />

depth interval<br />

system (DDAPS) .<br />

Integrated<br />

(voltage) 2<br />

Rectifie r<br />

and<br />

filte r<br />

PDP- 8<br />

Computer<br />

Fish density<br />

(N/m 3 )<br />

Numbe r of<br />

digital sample s<br />

6 3485 +0 .344950E-04 818 1<br />

12 6489 +13 .642376E-04 828 2<br />

18 38175 +0 .377970E-03 818 1<br />

24 24162 +0 .239227E-03 715 4<br />

30 3394 +0 .336040E-04 5 0<br />

36 0 +0 .000000E+00 0<br />

42 0 +0 .000000E+00 0<br />

48 0 +0 .000000E+00<br />

54 0 +0 .000000E+00<br />

Figure 4 . Example of DDAPS output .<br />

Analog t o<br />

digita l<br />

converte r<br />

Typically the average target strength is no t<br />

known precisely before processing . The density<br />

outputs of DDAPS are thus relative rathe r<br />

than absolute. Conversion to absolute density<br />

can be made either from measurement of tar -<br />

get strengths or by direct comparison of<br />

320


DDAPS output with determinations of absolute<br />

density by echo counting . An example<br />

regression of fish density from echo counts<br />

and integrated voltage is shown in figure 5 .<br />

3 6 9 12 1 5<br />

Fish targets per 1,000 m3<br />

Figure 5 . Relationship between integrated voltag e<br />

and fish density determined from ech o<br />

counts, Lake Washington, August 1971 .<br />

Sampling for Species and Size Compositio n<br />

The data processing methods described t o<br />

this point result in absolute estimates of the<br />

densities of fish targets within various dept h<br />

intervals along the various transects . These<br />

densities can be extrapolated over the respective<br />

volumes represented by the transects to<br />

obtain a total estimate of fish within the lake .<br />

The species and size composition is not derived<br />

directly from the acoustic data and must<br />

be obtained from net samples . Several types<br />

of nets have been applied at the University of<br />

Washington . In Lake Washington, which is<br />

accessible by large vessels, a program of sampling<br />

with a 10-ft Isaacs-Kidd midwater traw l<br />

has been conducted for several years . Development<br />

of midwater trawls for use from tw o<br />

outboard vessels is being conducted for lakes<br />

not accessible by large vessels . Vertical gil l<br />

nets of variable mesh size are also bein g<br />

applied .<br />

Size Determination from Acoustic Data<br />

As stated earlier, the target strength of a<br />

fish is a function of its size. Thus measurement<br />

of the size of the fish echoes gives an<br />

indication of the fish size . Unfortunately, the<br />

size of the fish echo is also dependent on th e<br />

orientation of the fish relative to the inciden t<br />

sound (its aspect) and upon its location with -<br />

in the sounder cone . The average effect of the<br />

directivity pattern can be sorted out, but ,<br />

even so, the variability associated with the<br />

relationship between echo amplitude and fish<br />

size is quite high, and the precision of size<br />

determination by acoustic methods is uncertain<br />

at this stage . It does seem feasible at least<br />

to determine the number of larger fish .<br />

Dawson 1 was able to estimate the number of<br />

adult sockeye in Lake Washington by counting<br />

the large echo targets . The technique<br />

probably can be used to determine the number<br />

of large fish such as squawfish and pea -<br />

mouth chub in lakes . This ability would be<br />

extremely valuable since the large fish are<br />

likely to be undersampled by the midwater<br />

trawls .<br />

Future Modifications<br />

The basic hydroacoustic techniques for fis h<br />

population assessment have been derived an d<br />

are being applied in several lake systems i n<br />

Washington, British Columbia, and Alaska, including<br />

IBP-funded studies of the Coniferou s<br />

<strong>Forest</strong> Biome conducted primarily in Lake<br />

Washington and Lake Sammamish . Future<br />

modifications of the technique will be directed<br />

primarily toward refinement of<br />

methods of size discrimination from acousti c<br />

data, and further automation of data processing.<br />

A program for determination of the distribution<br />

of target strengths is presently unde r<br />

development with funding from Sea Gran t<br />

and the National Marine Fisheries Service .<br />

1 J . J . Dawson . Estimation of the 1971 Lake Washington<br />

sockeye salmon escapement by means of a n<br />

echosounder . Informal Progr . Rep . to Wash . Stat e<br />

Dep . Fish ., Univ . Wash . Fish . Res. Inst ., Seattle, 7 p . ,<br />

1971 .<br />

321


This program will automatically calibrate th e<br />

integration program to absolute density, thu s<br />

eliminating the time-consuming echo countin g<br />

procedure . The data on target strength distribution,<br />

compared with the data from ne t<br />

sampling, will provide information on the precision<br />

and reliability which can be obtaine d<br />

from acoustic determination of size .<br />

Acknowledgments<br />

Contributions of Dr . R. L. Burgner, Dr . O .<br />

A. Mathisen, Mr . E. P. Nunnallee, and Mr . J .<br />

Dawson to the preparation of the manuscript<br />

are gratefully acknowledged .<br />

The acoustic assessment techniques de -<br />

scribed in this paper have been developed primarily<br />

by an interdisciplinary group under th e<br />

Sea Grant Marine Acoustics Program and sup -<br />

ported in part by National Science Foundation<br />

Grant No . GB-20963 to the Coniferous<br />

<strong>Forest</strong> Biome, U .S. Analysis of Ecosystems ,<br />

International Biological Program . This is Contribution<br />

No. 48 to the Coniferous <strong>Forest</strong><br />

Biome .<br />

Literature Cited<br />

Cushing, D . H. 1968 . Direct estimation of a<br />

fish population acoustically . J . Fish. Res .<br />

Board Can . 25(11) : 2349-2364 .<br />

F. R. H . Jones, R . B. Mitson ,<br />

and others. 1963. Measurements of the tar -<br />

get strength of fish . Radio & Electron. Eng .<br />

25 : 299-303 .<br />

Moose, P . H., and J. E . Ehrenberg . 1971 . Variance<br />

of the abundance estimate obtained<br />

with 4 fish echo integrator. J. Fish. Res .<br />

Board Can . 28(9) : 1293-1301 .<br />

, J. Green, and J . E . Ehrenberg .<br />

1971 . Electronic system and data processing<br />

techniques for estimating fish abundance.<br />

Inst. Elec. & Electron. Eng. Conf.<br />

Eng. Ocean Environ . Proc ., p. 33-36 .<br />

, R. E. Thorne, and M . O .<br />

Nelson . 1971 . Hydroacoustic technique s<br />

for fishery resource assessment . Mar . Tech .<br />

Soc . J. 5(6) : 35-37 .<br />

Shishkova, E . V. 1964. Study of acoustical<br />

characteristics of fish . In Modern fishin g<br />

gear of the world, vol . 2, p. 404-409. London:<br />

Fishing News Books, Ltd .<br />

Thorne, R . E. 1970. Investigations into th e<br />

use of an echo integrator for measuring<br />

pelagic fish abundance. 117 p . Ph .D . thesis<br />

on file, Univ . Wash., Seattle .<br />

. 1971 . Investigations into the re -<br />

lation between integrated echo voltage and<br />

fish density. J. Fish. Res. Board. Can.<br />

28(9) : 1269-1273 .<br />

and H. W. Lahore . 1969 . Acoustic<br />

techniques of fish population estimatio n<br />

with special reference to echo integration .<br />

Univ . Wash. Fish . Res . Inst. Circ. 69-10, 1 2<br />

p. Seattle .<br />

, J. E. Reeves, and A . E . Millikan.<br />

1971. Estimation of the Pacific hak e<br />

(Merluccius productus) population in Port<br />

Susan, Washington, using an echo integrator<br />

. J. Fish . Res . Board Can. 28(9) :<br />

1275-1284 .<br />

and J . C. Woodey. 1970. Stoc k<br />

assessment by echo integration and its<br />

application to juvenile sockeye salmon i n<br />

Lake Washington . Univ. Wash. Fish. Res .<br />

Inst. Circ . 70-2, 31 p . Seattle .<br />

Truskanov, M . D ., and M . N . Scherbino. 1966 .<br />

Determination of density of fish concentrations<br />

by means of hydroacoustics. Ryb .<br />

Khoz. (Fish . Econ .) 8 : 38-42 . (U .S. Dep .<br />

C o m m er., Joint Publ. Res. Serv. TT :<br />

66-34723 .)<br />

322 * GPO-796-161


------------- ------ -


The FOREST SERVICE et ; U :' S D partment of Agriculture<br />

is dedicated to the pti lg oaf: mel# ;ease management of the<br />

Natio'h's forest resoi.Mi-it for srrst it ids of wood, water<br />

forage, wildlife, a ecretq: :7h? gh forestry research ,<br />

cooperation with St s `and• 'pri v forest owners, an d<br />

management of the iortiil or sls'an¢! atonal Grasslands, i t<br />

strives - as directett N provide increasingl y<br />

greater service to a growtthfl

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!